You are on page 1of 44

AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013

1. PROBLEM 1.B, Viscosity and Reynolds number


a.

3
  T  2 T0  S

0  T0  T  S
(1.1)


 T 

0  T0 
(1.2)

Equation(1.1) shows Sutherland’s expression which relates the dynamic viscosity of a gas to its temperature. In this
equation  is the dynamic viscosity, T the temperature and S the Sutherland’s constant with a value which is a gas
characteristic. Sutherland’s expression can be approximated by Equation (1.2) with a constant  valid, for a T near
T0 . The value of this constant for air is determined for three cases.

For T0  288K and S  111K :

The viscosity in standard atmosphere for T0  288K is 0  1.79 E  5 [1]. Calculating the viscosity for a
temperature of 273K (a value close to T0 ) gives:

3 3
 T 2 T  S  273  2 288  111
  0   0  1.79 E  5    1.72 E  5 kg ms
 
T0 T  S  288  273  111
Literature (Table A-2 in [2]) confirms this value. The exponent  can now easily be calculated by using the two
values for the temperature and their corresponding values for dynamic viscosity in Equation(1.2).


1.72 E  5  273 
 
1.79 E  5  288 
 1.72 E  5 
log  
  1.79 E  5   0.75
 273 
log  
 288 
For T  S :

3 3 3
  T  2 T0  S  T  2 S  T  2
  
0  T0  T  S  T0  S  T0 
3

2
For T  S :

1
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
3 3 1
1
  T  T0  S  T   T 
2 T  2 2
       
0  T0  T  S  T0   T0   T0 
1

2
b.
Isentropic relations [3] are used to investigate the behavior of Re for M 2 M1  100 :

1
 0    1 2   1
 1
 
M  (1.3)
2 


p2   2   T2   1
    (1.4)
p1  1   T1 
For this calculation, M 2  10 and M1  0.1 are used and for air   1.4 . Filling in Equations (1.3) and (1.4) for M 1
give:

1
 0  1.4  1 2 1.41
 1  1.005
1 
0.1 
2 

p1  1   1 
1.4

     0.993
p0   0   1.005 
 T1    p1  1.41.41
    0.993  0.998
 1

 T0   p0 
For M 2 , these calculations yield:

0 p T
 2020, 2  2.356 E  5, 2  0.0476
2 p0 T0

Combining the results gives:

2 0 1
   1.0051  4.926 E  4
 0 1 2020
p2 p0
  2.356 E  5  0.9931  2.372 E  5
p0 p1
T2 T0
  0.0476  0.9981  0.0476
T0 T1
The ratio of viscosity can now be calculated with Equation (1.2) and   0.75 .

2
  0.0476  0.1019
0.75

1
The velocity ratio U 2 U1 will be determined by calculating the ratio of speeds of sound and multiplying this by the
Mach numbers. Then the ratio of Reynolds numbers can be determined. It is found that there is a decrease in
Reynolds number for M  1 .

2
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
U 2 M 2  RT2 M 2 T2
   100  0.0476  21.82
U1 M 1  RT1 M 1 T1

Re2  2 U 2 1 1
  4.926E  4  21.82   0.1055
Re1 1 U1 2 0.1019

When this calculation is done for two low Mach numbers, an increase in the Reynolds number will be found. This is
obviously due to the increasing velocity U , associated with the Mach number. However, because of the Reynolds
number dependent on the mass flow U , eventually a decrease will be found. When investigating the ratio of
densities for M  1 using Equation(1.3), it can be concluded that a decrease in density can only happen for a
supersonic value. The resulting ratio can never be lower than one.

1
 0    1   1 1

1    1  0.2  0.4
  2 
Of course a decrease in speed of sound will reduce the velocity U for higher Mach numbers. But since the air density
will increase for lower Mach numbers, the net effect is that the Reynolds number experiences a maximum for a
value which is larger than M  1 .

c.
An expression can be derived for Re L for a given Mach number M and total pressure and temperature pt and Tt .
For this, the isentropic relations (Equations (1.3) and (1.4)), the approximated Sutherlands expression (Equation (1.2)
, with   0.75 ) as well as the perfect gas law (Equation (1.5)) are used.

p   RT (1.5)

Re U

L 
In which:

1
p    1 2   1
t  t 1  M 
RTt  2 
  1 2 
U  M  RTt  1  M 
 2 

T 
  0  
 T0 
0 is the viscosity of the gas corresponding to the total temperature and can be found in [2], Table A-2 or can be
calculated using either Equation (1.1) or (1.2). Furthermore,   1.4 and R  287 J kgK for air.

d.
Figure 1 shows Re L for pt  10 bar and Tt  300K , using both the Sutherland’s expression and the approximation.
The maximum value of Re L using Sutherland’s expression occurs at M  2.1 and is 220E 6 . The approximation
gives a Re L of 210E 6 at a Mach number of 1.8. A lower total pressure will result in a lower maximum Re L , and
vice versa. Of course, a change in total temperature does not have an influence in the Mach number at which the

3
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
maximum Re L occurs for the approximation. For the Sutherland’s expression however, the Mach number at
which the maximum occurs will be lower for a higher total temperature, and vice versa. The farther the total
temperature moves away from the used reference temperature of 288K , the less accurate the approximation is.
This is visible by the larger differences between the two plots for higher Mach numbers.

Re/L
2.50E+08

2.00E+08

1.50E+08 Sutherland's
expression
Re/L

approximation
1.00E+08

5.00E+07

0.00E+00
0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 2.7 3 3.3 3.6 3.9
Mach number

Figure 1, Re/L for pt  10bar and Tt  300K .

Note: the  omega of 0.75 is an approximation. There was a 15K difference between the reference temperature
and the chosen temperature. So a better approximation will be reached with an omega which is slightly higher.

4
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
2. PROBLEM 2.A, 2D Stagnation point flow

a.
Equation (2.1) gives the stream function  for the creeping flow around a circular cylinder with radius R in a
uniform flow field with velocity U  . In this equation, r is the radius from the center of the cylinder and  is the
angle measured from the stagnation point on the cylinder. This stream function is plotted in the left figure in Figure
2.

 R2 
  U  1  r sin 
r 2 
(2.1)

By use of a Taylor series expansion around the stagnation point, an approximation for the stream function can be
given in the form  Axy . Because only the area close to the stagnation point is considered, x  R and
y  r  R (local coordinate system) can be assumed. Small angle approximation gives sin    . The general
equation for the Taylor series expansion of function f  x  around point a can be seen in Equation(2.2).

f ' a
f a   x  a (2.2)
1!
The stream function is approximated at the stagnation point  r  R  :

    R    '  R  r  R 
 R3   R2 
 U   r  2  sin   U   1  2  sin   r  R 
 R   R 
 2U  sin   r  R 
2U 
 xy
R
Which gives A  2U  R . This approximation of the stream function is plotted in Figure 2 (the middle figure). The
top left figure in Figure 3 shows the error of this approximation compared to the stream function. It can be seen that
near the stagnation point, the solution is within 5% accuracy. When going a distance 0.05R up or down, the error will
be larger than 5%. At a distance of 0.1R left of the stagnation point, the error is larger than 10%.

b.
Equation (2.3) is the self-similar solution for a planar viscous stagnation point on a flat wall. In the previous section it
was derived that an approximation of the flow around the area of the stagnation point of a cylinder can be made
from the stream function. Here an approximation is made using the self-similar solution. This solution satisfies the
differential Equation(2.4). From the boundary conditions it is known that F  0  F '  0  0 . Numerically solving the
equation gives F ''  0  1.2326 . When solving for F ''' , a value of F '''  0  1 (and subsequently F ''''  0  0 ) are
found.

5
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
   AxF  
A (2.3)
y

F ''' FF '' 1  F ' 2  0 (2.4)


A Taylor series expansion of F   is made in order to approximate the stream function at the stagnation point (so
for small  ). Appling an expansion for the first 5 terms, and recognizing that many of the derivatives of F are zero,
yields:

1.2326 2 1 3
F       (2.5)
2 6
c.
Combining Equations (2.3) and (2.5) gives an approximation of flow around the stagnation point in the form
  Bxy 2 . In order to determine the value of B and the Reynolds number dependency, Re  U  R  is used.

1.2326 2 1 3
   Ax   
2 6
1
The last term can be omitted since it will be very small for small  (for example   0.2 gives  0.23  0.001 ).
6
Substituting A , the Reynolds number and the coordinate scaling gives:

1.2326 U  R 2U  y 2
 x 2 Re
2 Re R R 2
U
  1.2326 2 2 Re xy 2
R
This can be written in a shorter form as   Bxy 2 with B  1.2326U  R 2 2 Re . This self-similar viscous solution
is not only a function of the undisturbed velocity and the radius of the cylinder, but also of the Reynolds number.
The approximation of this self-similar viscous solution is plotted in Figure 2 (right) with Re  1 . As is known, the
stream function is valid for creeping flow. Reynolds numbers up to 5 are considered creeping flow. For higher
Reynolds numbers separation will occur. This would indicate that the self-similar solution is more accurate for lower
Reynolds numbers. Figure 3 also shows the self-similar solution for Reynolds numbers of 0.1, 1 and 5. Indeed, it can
be seen that the stagnation point flow for Re  0.1 is most accurate. The solution for Re  5 is less applicable since
close to the stagnation point the error is in the range of 5-10%, but the error quickly increases when moving away
left from the stagnation point. It is also interesting to note that the error is approximately a variable of the x
coordinate, while for the stream function approximation the error is a function of both the x and y direction.

6
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
Stream function Stream function approximation Self-similar viscous solution
3 3 3

2 2 2

1 1 1

0 0 0

-1 -1 -1

-2 -2 -2

-3 -3 -3
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3

-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3

Figure 2, Stream function of the flow around a cylinder (left) and the stagnation point approximations (flow from left to right).

Stream function approximation Self-similar viscous solution, Re=1

0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
-0.05 -0.05
-0.1 -0.1
-0.15 -0.15
-0.2 -0.2
-0.25 -0.25
-1.5 -1.4 -1.3 -1.2 -1.1 -1 -1.5 -1.4 -1.3 -1.2 -1.1 -1

Self-similar viscous solution, Re=5 Self-similar viscous solution, Re=0.1

0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
-0.05 -0.05
-0.1 -0.1
-0.15 -0.15
-0.2 -0.2
-0.25 -0.25
-1.5 -1.4 -1.3 -1.2 -1.1 -1 -1.5 -1.4 -1.3 -1.2 -1.1 -1

Figure 3, Stream function approximation (top left) and self-similar viscous solutions of the stagnation point flow for several low Reynolds
numbers. Flow from left to right. Black: 0-1%, brown: 1-5%, red: 5-10%, yellow: 10-20%, white: >20% accurate.

7
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
3. PROBLEM 3.B, Self-Similar stagnation-point boundary layer with suction

a.
The Falkner-Skan equation describing the self-similar solution for a boundary layer flow is given in equation (3.1).
During this problem, the planar stagnation point flow with suction is considered. A planar stagnation point flow is
characterized by a value of   1 . The definition of the scaling coordinate  is given by Equation(3.2). The wall
suction velocity vw is incorporated using the appropriate scaling described by Equation (3.3).

f ''' ff ''  1  f '2   0 (3.1)

y U x
 (3.2)
x 2

vw 2U  x
fw   (3.3)
U 
The following boundary conditions apply:

f  0  f w , f '  0  0, f '     1
For a strong suction ( f w  1 ), the boundary layer thickness decreases and the gradients will increase. To analyze
this strong suction case it is convenient to stretch the  -coordinate according to the magnitude of the suction
strength during which Equation (3.1) will be transformed according to Equation (3.4). The stretched  -coordinate
becomes   f w .

1
f    f w     (3.4)
fw
Equation (3.4) and its derivatives can be substituted in Equation(3.1). During this process the earlier prescribed
boundary conditions will also change. The derivatives of Equation (3.4) are:

f '     '  
f ''    f w ''  
f '''    f w2 '''  
Substituting gives:

 1 
f w2 '''  f w    f w '' 1   '2  0
 fw 
f w2 ''' f w2 ''  '' 1   '2  0
1 1 1
 '''  '' 2
 '' 2  2  '2  0
fw fw fw

The boundary conditions become:

8
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
1
f  0  f w  f w  f w    0    0  0
fw
f '  0  0   '  0  0
f '   1   '   1

b.
The result of the former section can be approximated for strong suction ( f w  1 ). In order to do this the equation
can be rewritten as:

 '''  ''
1
f w2
 '' 1   '2   0 (3.5)

In equation (3.5) the term 1 f w2 is small and it will be further on indicated with a single small parameter  2 . The
velocity profile is given by  '  u U  and will be approximated using an asymptotic series:

    0     21    ... (3.6)


During this problem only the first terms 0 and 1 will be determined. Substituting equation (3.6) in Equation (3.5)
gives:


0 '''  21 ''' 0 ''  21 ''  2 0   21 0 ''  21 ''   1  0 '  21 '   0
2

0 ''' 0 ''  21 '''  21 ''  200 ''  401 ''  410 ''  611 ''  2 
 20 '2   40 ' 1 '  40 ' 1 '  61 '2  0
All equal powers of  can be collected after which all powers higher than 2 can be neglected:

0 ''' 0 ''  2 1 ''' 1 '' 00 '' 1  0 '2    4 01 '' 10 '' 20 ' 1 '    6 11 ''  61 '2   0

0 ''' 0 ''  2 1 ''' 1 '' 00 '' 1  0 '2   0 (3.7)

0 is found by recognizing  2  0 for strong suction and will yield the so called basic solution. This reduces
Equations (3.6) and (3.7) to:

  0 (3.8)
and

0 ''' 0 ''  0 (3.9)


A solution to this differential equation is:

0  e   C1  C2
0 '  e   C1 , C1  1
(3.10)
0 ''  e 
0 '''  e 
C1 is determined from the boundary conditions. Equations (3.10) satisfy the boundary conditions:

9
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
0 '  0   0
0 '     1

It will satisfy the boundary condition 0  0  0 when C2  1 . Instead the value C2  0 will be chosen. Now the

boundary condition 0  0  0 is not satisfied but, as will be clear later, the determination of 1 will be much easier
while it has no effect on the boundary layer properties.

The second term of the right hand side of Equation (3.6) is found by including by solving the  2 -term of Equation
(3.7). 0 is known and, conveniently, for very small and very large  the same differential equation as in Equation
(3.9) remains.

1 ''' 1 '' 00 '' 1  0 '2  0


1 ''' 1 ''  e     e   1   e   1  0
2

Resulting in:

1 ''' 1 ''  2e2   e  2e (3.11)


The right hand side of Equation (3.11) is plotted in Figure 4. Its value becomes zero for   0 and   . Thus, when
considering f w to be large, an approximate solution can be found by solving 1 ''' 1 ''  0 , which is the same
differential equation as was solved for 0 .

0.25

0.2

0.15

0.1

0.05

-0.05

-0.1
0 1 2 3 4 5 6
f
w

Figure 4, right hand side of Equation (3.11)

1  e     1
1 '  e   1
(3.12)
1 ''  e 
1 '''  e 
A minus one as integration factor may now be added to 1 so the equation also satisfies 1  0  0 . Combining
Equations (3.6), (3.10) and (3.12) gives:

10
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
   e       e    1
 2 

 '   e   1   2  e   1
(3.13)
 ''   e     2  e  
 '''   e     2  e  
The velocity profiles  ' are plotted in Figure 6 for f w  1 , f w  2 , f w  5 and f w  10 . The case of f w  0 is not
plotted because obviously it reaches infinity. The velocity profile and the other properties of the f w  10 (the best
approximation) case are plotted in Figure 5.

Stagnation point boundary layer approximation for fw=10


Approximated velocity profiles 7

fw=1 6 `
2
fw=2  ``
1.8 fw=5 5  ```

1.6 fw=10
4
1.4
3
1.2
`

1 2

0.8 1
0.6
0
0.4
-1
0.2

0 -2
0 1 2 3 4 5 6 0 1 2 3 4 5 6
fw fw

Figure 6, Approximated stagnation point velocity Figure 5, Stagnation point boundary layer
profiles for various wall suctions. approximation for strong suction.

c.
With the results from part b, approximate expression for the boundary layer properties can be found. Equations
(3.14) till (3.17) are approximations for the wall shear f ''  0    0 , scaled displacement thickness * , scaled
momentum thickness  and the shape factor H respectively. After inserting Equations (3.13) the boundary layer
properties can be found (see Equation (3.18) till (3.21)) for any value of f w . As can be seen in Figure 5, the boundary
layer properties settle at around   6 , so in these cases, the integrals are evaluated from zero till 6.

f ''  0  f w ''  0 (3.14)

 * U x  1

    1  f '  d   1   '  d (3.15)
x 2
*

0
fw 0

 U x  1

    f ' 1  f '  d    ' 1   '  d (3.16)
x 2 0
fw 0

11
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013

 * U x
* 
H  x 2  
*
(3.17)
  U  x 
x 2

 1 
f ''  0   f w  e   2 e  
 fw  (3.18)
1
f ''  0   f we   e 
fw

 1   e 
 1   2  e   1 d
1 
  *
fw 0
6
 *   e    2  e   1 d
1
(3.19)
fw 0

  *
1
fw

e 6  1   2  e 6  5  
0.9975 5.0025
fw

f w3


    e   1   2   e   1 1    e   1   2   e   1 d
1
fw 0

6

 e  e 2    2  1  3e   2e 2    4   e 2  2e   1 d


1
  

fw 0
(3.20)
 1 6 1 12  2 4 1 12  
 2  e  2 e     4  3e  e      2  2e  2 e  
1 9
  6 12 6

fw    
0.4975 4.0074 4.5050
   
fw f w3 f w5

 0.9975 5.0025 
 f 
 f w3 
H w
(3.21)
 0.4975 4.0074 4.5050 
  3
 
 f w f w f w5 

d.
The approximate boundary layer properties are calculated for f w  0 , f w  1 , f w  2 , f w  5 and f w  10 . From
the equations in part c, it becomes clear that the properties for no suction ( f w  0 ) are undefined since it involves a
division by zero. The approximation is derived for strong suction. Indeed with increasing f w , the error becomes
smaller. Only between f w  1 and f w  2 an increase in error for f ''  0  and H can be found. This may be due to
the larger error made when approximating the right hand side of the differential equation for 1 with zero.

12
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
Table 1, Stagnation point boundary layer properties of the complete Falkner-Skan equation, the approximation and its error.

fw  0 fw  1 fw  2 fw  5 f w  10

f ''  0 Complete equation 1.2326 1.8893 2.6701 5.3595 10.1939

Approximation - 2.000 2.5000 5.2000 10.1000

Error [%] - 5.8593 -6.3705 -2.9760 -0.9211

 *
Complete equation 0.6479 0.4593 0.3422 0.1810 0.0972

Approximation - -4.0049 -0.1265 0.1595 0.0948

Error [%] - -971.97 -136.98 -11.886 -2.5200

 Complete equation 0.2923 0.2150 0.1639 0.0892 0.0484

Approximation - -8.0149 -0.3930 0.0660 0.0457

Error [%] - -3827.8 -339.75 -26.007 -5.5580

H Complete equation 2.2162 2.1365 2.0876 2.0295 2.0091

Approximation - 0.4997 0.3220 2.4264 2.0734

Error [%] - -76.612 -84.574 19.063 3.1982

13
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
4. PROBLEM 4.E, Flat plate boundary layer with uniform suction – integral method

a.
The boundary layer integral momentum equation is given by:

U  d 2  w vw
  (4.1)
2 dx U  

The boundary condition for the momentum thickness at the beginning of the flat plate is  ( x  0)  0 . Using
coordinate scaling and introducing a shear correlation S , Equation (4.1) can be rewritten. For this rewritten
equation, the scaled momentum thickness has to satisfy  * (  0)  0 .

vw 2 * v  
x ,   w  , S  w
U   U 

2
  *    * 
d   v w  
U   vw   vw 
S
2   U   
d 2 
 vw 
U   2 vw 2 d *2 v 
 S  w *
2 vw  U  d 
2
 vw

d *2
 2S  *  (4.2)
d
The exact solutions for the flow near the leading edge and the developed boundary layer are described by the
Blasius solution and the suction profile solution respectively.  * can thus be determined for small and large values
of  , and compared with the behavior of Equation (4.2). First, the Blasius solution is tackled. Note that in the
derivation in the last step the minus disappeared. This is because vw is always negative (for suction) and the square
root of the square of vw on the right hand side makes it positive.

14
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
 
 0.66411
x U x

 *
vw 
 0.66411
 U  U 
2
U  2
vw vw
v 
*
v2
 w  0.66411 2w
U  U 
U  vw 1
 *  0.66411 
vw U  

1
 *  0.66411 (4.3)

 *   0   0
 *      
This result corresponds with the boundary condition  *   0 for Equation(4.2).  * will now be determined for the
suction profile. Note that the suction velocity profile is only dependent on the y-coordinate since  is only
dependent on y . The momentum thickness can be obtained using equation (4.4)[2]. A large y is defined as the
point where u  0.99U  . This is achieved at a distance from the wall   4.6  vw  [2].

u   yv
 1  e  ,   w
U 

Y 
u  u 
 
0
1 
U  U 
dy (4.4)

Solving the momentum thickness equation after inserting the expressions for  and  * gives:


4.6
  y   y 
vw vw vw

vw
*  
0


1  e 
  
 
1  1  e 
 dy


4.6

vw vw vw vw
y y 2 y
    1 e 
 1  2e 
e 
*
dy
vw 0

4.6

vw vw vw vw
y y 2 y

vw
*  
0
 e   2e   e 
dy

  v 4.6 v vw  v  v
 4.6 vw w 1  24.6 vw w    0  0 1  0
    e * w
2 e  e    e  2 e  e 
vw  vw v 2 v   vw v 2 v 
 w w  w w

v v  1 1
 *   w   e 4.6  2e 4.6  e 9.2  1  2  
 vw  2 2
15
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
1 1
 *  e4.6  e9.2   0.49 (4.5)
2 2
As expected also the momentum thickness is independent of  and thus a constant. Because it is known that  *
must be zero at   0 it becomes clear that the suction profile is only valid for large  when the boundary layer is
‘settled’ and does not change any more. This can be checked by inserting Equation (4.5) into Equation (4.2) and
recognizing that d *2 d  0 :

  1 1 
0  2  S   e 4.6  e 9.2    (4.6)
  2 2 
This is indeed a valid equation for a constant S .

b.
The integral momentum equation (4.2) and the expression for S are repeated here for convenience:

d *2
 2S  *  (4.7)
d

 w
S (4.8)
U 
It is clear from Equation (4.8) that the shear parameter S is not a constant because the momentum thickness 
grows with downstream distance. The solution to Equation (4.7) could be approximated by using a constant value for
the shear parameter. Values of S for small and large  could be obtained using the Blasius solution (Equation(4.3))
and the suction profile solution (Equation(4.5)). First, a shear parameter for small  is determined by differentiating
the square of Equation (4.3) and inserting the result in Equation (4.7):

2
1
  0.66411 
*2 2 2


d *2
 0.664112  2  S   * 
d

 *   0  0 , so:

0.664112
S  0.2205
2

The value of 0.2205 indeed corresponds with the results of the accurate finite-difference solution. A shear
parameter for large  can be obtained by inserting Equation (4.5) into Equation (4.7) and recognizing that for a full
developed boundary layer d *2 d  0 . Solving Equation (4.6) for S gives:

1 1
S  e 4.6  e 9.2   0.4899
2 2
The accurate finite-difference solution gives a value of S  0.4893 for   5 so it can be concluded that the value of
S according to the suction profile is a correct approximation. Figure 7 shows the numerically solved equation (4.7)
with S taken from the Blasius solution (red) and the suction profile solution (green). The green line coincides with the

16
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
asymptotic solution derived in the former question (see Equation (4.5)). The cyan colored line at  *  0 is the
2

asymptote for the Blasius solution at   0 .

*2
 versus  for a boundary layer with uniform suction
0.25

0.2

Finite difference solution


0.15 Blasius solution
Suction profile solution
*2

Adaptive S solution

0.1

0.05

0
0 1 2 3 4 5

Figure 7, Boundary layer approximations using different models

c.
From the result of question b it seems that it might be beneficial to introduce and adaptive value of S. In order to do
this, Thwaites expression (including wall suction velocity) is derived from the momentum integral equation and a
shape parameter  is introduced. Furthermore, it is assumed that S  A  B . It will turn out that Thwaites
expression is a function of both  and  * . Together with the definition of the shape parameter two equations with
two unknowns will remain, after which the coefficients A and B can be determined. The continuity and momentum
equations are [5]:

u v
 0 (4.9)
x y

1  u u u U  U 
  u v   U (4.10)
 y t x y t x

To obtain the momentum integral equation, the continuity equation is multiplied by u  U  . Subsequently this is
subtracted from the momentum equation which yields the following result:

1    U  
  U   u   u U   u   U   u   v U   u 
 y t x x y
This equation can be integrated from the wall till y   . Here it is important to recognize that there is a uniform
suction at the wall and that far away from the wall the shear is going to zero.

    
1    U  
   U   u  dy   u U   u dy   U   u  dy  v U   u  dy
y 0

 0 y t 0 x 0 x 0

In this equation the first and the last term can be evaluated and the displacement thickness and momentum
thickness can be substituted.
17
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
 
 u  u  u 
 *   1   dy,   1   dy
0  U  0
U  U 

w   U 
 U  *    U  * vwU 
 t x x
w  1 1  U  1 v
  * 2   * w
U  t U 
2
U  x x U  U
Steady flow is assumed and the shape factor H   *  can be recognized resulting in the Kármán integral relation:

w d  dU  vw
  2  H   (4.11)
U  dx
2
U  dx U 
Now Thwaites expression including the wall suction velocity is obtained by multiplying the Kármán integral relation
with U   . Further note that     and  d  0.5 2 . It can clearly be seen that the result becomes the
original Thwaites expression when vw  0 is inserted.

 w U  U  d U   dU  U  vw
  2  H   
U  
2
 dx  U  dx  U
 w  2 dU   2 dU   vw
  2  H  
U  2 dx  dx 

The first term is familiar, it is the same as the right hand side of Equation (4.8), the third term disappears since the
external flow is a constant over the domain ( dU  dx  0 ). Furthermore, the last term is equal to the scaled
momentum thickness.

 2 dU   vw
S     (4.12)
2 dx 
The shear parameter S is a function of the shape factor  . Also it is clear that it is a function of the pressure
gradient (which is a function of dU  dx ) and the suction velocity at the wall (the last term). The shape factor again
is a function of the curvature of the velocity profile at the wall:

 2   2u 
   (4.13)
U   y 2  w
The velocity curvature at the wall is related to separation and thus the pressure gradient. However, it is known that
wall suction or blowing is also related to separation. This can be shown by taking the momentum equation (4.10),
 2u 1 
assuming steady flow and substituting  :
y 2
 y

 2u u u U 
  u  v  U
y 2
x y x
At the wall u  0 and v  vw , and assuming a pressure to be only a variable of the x-direction, this can be rewritten
to:

18
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
2
d u vw du 1 dp
  (4.14)
dy 2  dy  dx
Inserting this result in (4.13) gives a relation for  depending on the pressure gradient and the wall suction velocity:

 2  vw du 1 dp 
 
U    dy  dx  w

Both the shear parameter S and the shape parameter  are thus a function of the pressure gradient and the suction
velocity at the wall. Now a relation can be assumed such that S  A  B . Also the pressure gradient is zero
because dU  dx  0 . For a fully developed boundary layer, this reduces Equation (4.12) to:

S     A  B   * (4.15)

Now the constants A and B can be determined. For this, the results of the former question are used. Recall that at
  0 for the Blasius solution S  0.2205 and   0 . Which gives   0 . Filling in Equation (4.15) leads to:

0.2205  A  B  0  A
 
For the suction profile solution, S   *  0.4899 . From Equation (4.15) S     S  * , so the assumption    *
can be tried to calculate B.

0.4899  0.2205  B  0.4899


0.4899  0.2205
B  0.5499
0.4899
The resulting relation (with    * ) is S  0.2205  0.5499 * . This equation is plotted in Figure 8 along with the
accurate numerical results given in the assignment and shows quite a good agreement. The numerical calculation is
done again with this adaptive value of S . The result is given by the dotted black line in Figure 7. Although the shape
of this profile is not in very good agreement with the accurate solution, the starting and end points are coinciding.

*
 versus S
0.5

0.45

0.4

0.35

0.3

0.25
S

0.2

0.15

0.1

0.05 Approximate linear relation


Accurate numerical relation
0
0 0.1 0.2 0.3 0.4 0.5
*

*
Figure 8, S as function of θ

19
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
5. PROBLEM 5.D, Adiabatic flat plate boundary layer

a.
The Croco-Busemann relations tell that for a Prantdl number of one, one solution exists where the total enthalpy is
constant everywhere in the control volume. This is equivalent to a situation without heat transfer at the wall, so an
adiabatic flat plate. A self-similar boundary layer can be described by the following differential equations:

f ''' ff ''  0 (5.1)


and,

 '' Pr f  ' 2 Pr f ''2  0 (5.2)


with boundary conditions f  0  f '  0   '  0       0, f '     1 . Equation (5.1) is the Falkner-Skan
equation with   0 , so a flat plate boundary layer. From the energy equation it follows that for Pr  1 , the total
entropy at any point in the control volume must be a constant. This corresponds to a situation without any heat
transfer at the wall, thus an adiabatic flat plate boundary layer exists. Total entropy h is a function of the local
temperature and the local flow velocity (see Equation (5.3)). By recognizing that the entropy is a constant, the
temperature profile  can be related to the velocity profile f ' of the boundary layer.

U e2  u 2
h  he  (5.3)
2
The temperature profile  is defined as:

T    Te
   
1 2
Ue cp
2
Inserting Equation (5.3) and bearing in mind that f '  u U e gives:

 
c p  he  U e2  u 2   he 
1
  U e2  u 2
   
2
  1  f '2 (5.4)
1 2 Ue 2
Ue cp
2
It can be shown that this equation is indeed correct for Pr  1 by substituting the found expression for  in
Equation (5.2) and comparing the result with Equation (5.1). In order to do so,  must be differentiated twice.  ' is
found using the chain rule while  '' can be obtained by the product rule.

d  d  du
 ,u  f '
d du d
d
 2 u
du
du
 f ''
d

 '  2 f ' f '' (5.5)

 ''  2   f '  ' f '' f '  f ''  '   2  f '' 2  f ' f '''  (5.6)
Substitution in Equation (5.2) and inserting Pr  1 results in:

20
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
 2 f ''  2 f ' f '''   1  f  2 f ' f ''   2 1  f ''
2 2
0
2 f '  f ''' ff ''   0
f ''' ff ''  0
Which is the same equation as Equation (5.1).

b.
The temperature solution for Pr  1can be approximated using an asymptotic series expansion for  and with small
parameter   Pr  1 , see Equation(5.7). This equation can be substituted in Equation (5.2) with Pr  1 after which
equal powers of  can be collected.

    0     1    ... (5.7)

0 ''  1 '' f Pr 0 '  f Pr 1 ' 2 Pr f ''2  0


 0 '' f Pr 0 ' 2 Pr f ''2     1 '' f Pr 1 '   0
This results in two equations with which 0 and 1 can be determined:

0 '' f Pr 0 ' 2 Pr f ''2  0 (5.8)

1 '' f Pr 1 '  0 (5.9)


These equations have to satisfy the boundary conditions:

     f  0  f '  0  0
  0  f '     1
c.
Equation (5.8) is solved for 0 . First, the homogeneous equation 0 '' f Pr 0 '  0 is tackled. This equation has
the form ar 2  br  c  0 with coefficients a  1 , b  f Pr and c  0 so the solution will have the form
  C1er1  C2er2 . The roots r1 and r2 are calculated as  f Pr and zero respectively. The homogeneous equations
of equations (5.8) and (5.9) are the same, so the following solutions (the derivatives determined using the chain rule)
apply to both equations:

0,h  1,h  C1e  f Pr  C2


(5.10)
 '0,h   '1,h  C1 ff ' Pr e  f Pr
The boundary conditions f  0  f '  0   '  0       0, f '     1 and the knowledge that f  0 give
C2  0 and C1 undetermined since both f  0  and f '  0 are zero. For the time being a value of C1  1 is used. The
particular solution of Equation (5.8) can now be solved using the method of variation of parameters and assuming
C1  u1   , C2  u2   and u1 '   e f Pr  u2 '    0 .

0, p  u1e  f Pr  u2  1


0, p '  u1 ' e  f Pr  u1 f Pr e  f Pr  u2 ' 0  u1 f Pr e  f Pr
0, p ''  u1 f 2 Pr 2 e  f Pr  u1 ' f Pr e  f Pr
Substituting in Equation(5.8) to determine the values for u1 ' and u2 ' :

21
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
2
u1 f Pr e 2  f Pr 
 u1 ' f Pr e  f Pr 
 f Pr  u1 f Pr e  f Pr 
  2 Pr f '' 2

u1  f 2 Pr 2 e  f Pr   f 2 Pr 2 e  f Pr    u1 '   f Pr e  f Pr    2 Pr f '' 2

f ''2
u1 '  2
fe  f Pr
u2 ' can now be determined:

f ''2  f Pr
2 e
u1 ' e  f Pr fe  f Pr f ''2
u2 '     2
1 1 f
Integrating u1 ' and u2 ' gives the expression for the particular solution of Equation (5.8).

f ''2 f ''2
 2 fe f Pr d   2 f d
 f Pr 
1, p  e (5.11)

Substituting (5.10) and (5.11) in (5.7):

f ''2 f ''2
    C1e f Pr  e  f Pr  2  f Pr 
d   2 d   C1e  f Pr  (5.12)
fe f
d.
The recovery factor r is defined as [5]:

Tw  Te
r    0 
1 2
Ue cp
2
The recovery factor is a function of the Prandtl number. An approximation of this relation is given by r  Pr . It is
clear that for an adiabatic wall and constant total entropy in the control volume ( Pr  1 ) the recovery factor is one.
This can also be shown by solving Equation (5.12) for   0 . The two integrals are equal to zero resulting in an
expression for r which is a function of the Prandtl number only. Remember that   Pr  1 .

f ''2 f ''2
    e  f Pr  e  f Pr  2  f Pr 
d   2 d    e  f Pr 
fe f
  0  1  1  0  0    1  1  
For a Prandtl number of one both approximations can be calculated:

r  1 1
r  111  1
In Figure 9, both approximations of the recovery factor are plotted as a function of the Prandtl number. The figure
also shows the difference between the two approximations. At a Prandtl number of one, the difference is zero.

22
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
Flat plate recovery factor difference between the asymptotic solution and Pr 1/2
2 0.3

1.8 0.2
1.6 0.1
1.4
0
1.2
-0.1
1
r

-0.2
0.8
-0.3
0.6

0.4 -0.4

0.2 r=Pr1/2 -0.5


Asymptotic solution
0 -0.6
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Prandtl number [-] Prandtl number [-]

Figure 9, approximated recovery factors and the difference between the two methods.

23
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
6. PROBLEM 6.D, Boundary layer calculations based on simple correlations
a.
The momentum thickness of a flat plate boundary layer can be approximated by a simple ‘method of quadrature’,
see Equation (6.1). Values for n , B and A , are specified for laminar and turbulent boundary layers and are tabulated
in Table 2.

d
dx
 Ren U eB   AU eB (6.1)

Table 2, values for n, A and B for the laminar and turbulent boundary layer.

Laminar Turbulent
n 1 1/6
A 0.45 0.0076
B 5 10/3
When a constant external velocity U e and a transition from laminar to turbulent at Re x  1.8  106 is assumed, the
growth of the momentum thickness d dx can be determined. Note that Re  U e  and Re x  U e x  .
d
dx
 Ren U eB   AU eB

AU B x  n Ax
x
1
  n B  AU eB dx  n e B  n n
Re U e 0 Re U e  U e
 n 1  nA A
n 1
 n n

x xU e Renx
 A
 n 1
x Re nx
d A
 n 1 n
dx Re x

The boundary layer growth is plotted in Figure 10. The discontinuity at Re x  1.8  106 is a result of the change of the
used coefficients n and A . It is indeed expected that the boundary layer growth will increase after the transition
point.

24
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
-3
x 10 Boundary layer growth for a flat plate flow
5
Boundary layer growth
4.5 virtual origin turbulent BL

3.5

3
d/dx

2.5

1.5

0.5

0
0 2 4 6 8 10
Rex 6
x 10
Figure 10, the boundary layer growth based on simple correlations with transition point at Rex=1.8E6.

b.
From the discontinuity in Figure 10, it is clear that the laminar and the turbulent boundary layer must be well
distinguishable. If the boundary layer is calculated by integrating the expression for d dx from x  0 till x  Re x
and the results are plotted this is indeed the case. Figure 11 shows the boundary layer development together with
an exponential fit of the turbulent part of the boundary layer.

Boundary layer development

16000

14000

12000

10000

8000

6000

4000
BL momentum thickness
2000 Turbulent BL fit
virtual origin turbulent BL
0
0 2 4 6 8 10
Rex 6
x 10
Figure 11, Boundary layer development according to simple correlations.

The ‘virtual origin’ of the turbulent boundary layer can be determined. This point is defined as the position where a
fully turbulent boundary layer would have started, such that the same momentum thickness would have been found
25
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
downstream of the transition point of the true boundary layer. So to determine this virtual origin the equation
describing the turbulent has to be known. The exponential curve fit of the turbulent part is found using MatLab and
given by Equation (6.2).

  aeb Re  ced Re
x x
(6.2)

Its coefficients are:

a  1.34  104
b  4.62  108
c  1.52  104
d  9.70  108

Solving for   0 gives:

 a  1.34  104 
ln    ln 
 c  1.52  104 
Re x    =0.88  106
bd 4.62  108  9.70  108
This virtual origin is indicated in Figure 10 and Figure 11 with a green circle.

c.
For a flat plate, the wall shear is always parallel with the undisturbed flow velocity. This means that any momentum
integral analysis for the boundary layer is valid for laminar flows, turbulent flows and combinations of both if the
upper boundary is the outer streamline [2]. The momentum thickness at downstream distance x is equal to the
summation of the total friction (the drag) from x  0 up till any downstream distance x . Since for a flat plate, the
distance from the beginning of the flat plate is always continuous, the momentum thickness must also be always
continuous, even when the flow turns from laminar into turbulent.

26
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
7. PROBLEM 7.D, The Clauser plot
a+b.
The wall shear stress in a turbulent boundary layer can be determined from a mean velocity profile using the Clauser
plot technique. A velocity model which is a function of the wall shear stress is taken and fitted to the mean velocity
profile. The velocity model used here is the wall of the law [2,5]:

u 1  yv* 
 ln  B (6.3)
v*    

v* Cf
 (6.4)
Ue 2

yU e
Re y  (6.5)

with   0.41 and B  5.0 . Combining equations (6.3), (6.4) and (6.5), gives the velocity profile as a function of
Re y distance:

u u v* v*  1  yv * U e  
 *   ln    B
Ue v Ue Ue     Ue  
u Cf  1  Cf  
  ln  Re y   B
Ue 2    2  
 
*
Note that the dependency on v is removed by multiplying with U e U e . The dynamic viscosity is chosen
  1.5E  5 m2 s (for air) and the external velocity is Ue  9.636 m s . The only unknown remaining is the skin
friction coefficient C f .Both the measured data from the assignment and the law of the wall are plotted in Figure 12.
In this figure, also the validity boundaries are shown [2]. The friction coefficient used for the law of the wall which
gives a good fit is C f  0.0028 .

27
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
Measured mean velocity profile of a turbulent boundary layer
1
Measured velocity profile
0.9 Law of the wall, Cf=0.0028

0.8 Validity boundaries

0.7

0.6
u/Ue [-]

0.5

0.4

0.3

0.2

0.1

0
3 4
10 10
Rey [-]

Figure 12, The Clauser plot.

c.
The strength of the wake component is the maximum difference between the measured velocity profile and the law
of the wall (within the validity range). For the current velocity profile and the law of the wall expression with
C f  0.0028 it is 0.012, or about 1%. This value is at a distance Re y  5782 from the wall, near the upper
boundary of the validity range.

28
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
8. PROBLEM 8.E, Turbulence modeling with the eddy viscosity

a.
For 2D turbulent boundary layer flow, turbulence is added to the momentum equation in the form of turbulent shear
 t . The definition of the turbulent shear is, according to White [2]:

u
 t    u ' v '  t
y

t is called the eddy viscosity. The equation can be divided by  for convenience.

t u
 u ' v '   t (7.1)
 y
An algebraic eddy viscosity model can be derived using a known mean velocity profile. So a function for  t can be
found for a known u y . The velocity profile which will be assumed here is given by Spalding’s law (Equation(7.2),
plotted in

Figure 13). But before this velocity profile can be used, the turbulent shear has to be resolved. For the inner layer of
the boundary layer, it is assumed that the turbulent shear stress distribution is also related to the velocity profile,
see Equation (7.3) [5]. In this equation  w is the wall shear stress. Substituting  t (Equation (7.3)) in Equation (7.1)
results in Equation (7.4).

   u    u   
2 3

y u e  B  e 1 u 
 u 
  (7.2)
 2 6 
 

t u 
 1  (7.3)
w y

 w  u   u
 1   t (7.4)
  y  y
It seems that nothing much changed, since the unknown turbulent shear stress is replaced by the unknown wall
shear stress. However, the wall shear stress is related to the friction velocity v * . The friction velocity is also used to
transform the coordinates u and y to u  and y  . From [2]:

u  v*
u  , y  y
v* 

w
v* 

Inserting these transformed coordinates and the expression for the friction velocity in Equation (7.4) gives the
following convenient equation expressing the eddy viscosity:

29
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
v * 2
 u 
 
u v * 2

1    t 
  y  y 
 u    t u 
1    
 y   y

1
 t  u    u  
 1
  y    y  
(7.5)

u  y  is obtained by differentiating Spalding’s law.

  u  
2
u   B   u  
 1 e e 1 u 

y   2 
 
Substituting in Equation(7.5):

1

 t   B   u
 e e  1   u 
 u 
 2   
   1   e  B  e u  1   u  
 u  
 2 

(7.6)
   2   2 
    
Figure 14 shows Equation (7.6) for the inner layer ( 0  y   50 ). The values   0.41 and B  5 are used according
to the assignment and White [2]. The viscous sublayer can be approximated by a linear expression. Figure 15 shows
the same figure multiplied by   1.51E  5 , which is approximately the viscosity of air.

Spalding`s law
25

20

15
+
u

10

0
-4 -2 0 2 4
10 10 10 10 10
y+

Figure 13, Spalding’s law

30
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
-4
Variation of eddy viscosity x 10 Variation of eddy viscosity of air
20

18

16

14 2
12
 t/ 

t
10

8
1
6

0 0
0 10 20 30 40 50 0 10 20 30 40 50
+
y y+
Figure 14, variation of eddy viscosity with y+ distance Figure 15, variation of eddy viscosity of air with y+ distance
b.
The Baldwin-Lomax turbulence model, for the viscous sublayer and the outer layer, also relates the eddy viscosity to
the velocity profile. It does so by introducing a function F.

u  

y

F  y 1  e A 
y  
The viscous sublayer is however not treated here, so F reduces to:

u
Fy (7.7)
y
Just like in part a, a velocity profile is assumed. In this case Cole’s wall-wake form is used:

U  u 2
v*
1
  ln   
 
1  3 2  2 3  (7.8)

In Equation (7.8), 3 2  2 3 is the wake function and   0.5 .  is a constant related to the pressure gradient. This
equation is differentiated in order to find the function F v * .

 1 2 
u  U   v *   ln   
  
1  3 2  2 3  

 1 2   y
2
 y  
3

u  U   v *    ln y  ln     1  3    2    
        
 
du  11  2  y y2  
 v *     0   6 2  6 3 
dy  y      

du  1 12    2  
 v *     (7.9)
dy       
Substituting (7.9) in (7.7) results in:

31
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
 1 12      2
F   v *       
     (7.10)
F 1 12 2
 
v*  
 3 
This equation is plotted in Figure 16. Also the maximum value of this equation is indicated here. The maximum
values are of interest because these are the parameters that are of importance for the modeling constants of the
Baldwin-Lomax turbulence model. These values are analytically obtained by taking the derivative of Equation (7.10)
and setting it equal to zero.

4.5 X: 0.67
Y: 4.607

4
F/v*

3.5

2.5

2
0 0.2 0.4 0.6 0.8 1

Figure 16, F/v* of the Baldwin Lomax turbulence model and its maximum

d F v * 12
d


 2  3 2   0

36 2 24
   0
 
36 24
a ,b  ,c  0
 
b  b2  4ac 24   24  
2

1,2       
2a  72    72
1 1 2
1,2    1  0,2 
3 3 3
Obviously2  2 3 is the correct coordinate of the maximum value. Inserting 2 in Equation (7.10) gives the
maximum value of F v * .

1 12 2
F
 
v * max  
2  23 
6  2   2  
2 3
F 1
          4.607
v * max 0.41 0.41   3   3  
With the extrema known, the modeling constants Ckleb and Ccp can be determined.

32
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
ymax U  *
Ckleb  , Ccp 
 ymax Fmax

Ckleb is evaluated easily by substituting ymax  max . This may be done because  is only a scaling parameter so the
maximum value of y is found for the maximum value of  .

max 2
Ckleb   max 
 3
A bit more effort has to be done to obtain Ccp . The boundary layer thickness  * must first be determined by
evaluating the definition of  * . The boundary layer thickness is dependent on the shape of the velocity profile. For
this again Cole’s wall-wake expression is used.


 u 
 *   1  dy
0  U 
 1  
 1 2   y     
 2 3

 U   v *    ln y  ln     y
 *   1   1  3    2 
     dy
 U           
0
    
v*

  y
2
 y 
3

U  0
*   ln y  ln   2   1  3    2   dy
      
v*      
*     ln      ln   2          0  
U     2  
v*     v 
*
*        1   
U  
2
 2   U 
This result can be inserted in the expression for Ccp .

U v* v * 1   
Ccp  1    
ymax Fmax U  Fmax max

Ccp 
1

1  0.5  1.1912
4.607 0.41   2 3
The calculated modeling constants are tabulated in Table 3 together with the values given by Granville [2]. The
accuracy with respect Granville’s constant is also calculated.

Table 3, Modeling constants of the Baldwin-Lomax turbulence model

Granville present result accuracy

Ckleb 0.5906 2/3 +12.8%

Ccp 1.2432 1.1912 -4.2%


c.
Eddy viscosity modeling can also be done by the K  L viscosity model by Prandtl. The Prandtl eddy viscosity
equation is dependent on a turbulent length scale L and, the turbulent kinetic energy K and a constant C .

33
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
t  C K L (7.11)
The ratio between the turbulent kinetic energy and the turbulent shear stress is approximately a constant:

u ' v '
 0.30 (7.12)
K
The mixing length model with mixing length [2] (Equation (7.13)) can be used to determine the constant C in
equation (7.11). For this the turbulent length scale is assumed to be equal to the mixing length, so L  .

u
t  2
(7.13)
y
From White [2] an additional equation to relate the turbulent shear to the velocity profile is used:

u
u ' v '   t (7.14)
y
Inserting (7.13) in (7.11) and rearranging gives:

y
t  C K t
u
1 u
C t
K t y
Substituting Equation (7.14) and using relation (7.12)

1 u ' v ' u ' v '


C t   0.30  0.5477 (7.15)
K t  t K

d.
The turbulence dissipation  is modeled with equation (7.16). The modeling constant CD can be determined when
the production P of turbulence is assumed to be equal to the dissipation, so P   . The production of turbulence is
related to an increase in energy due to shear. It is taken from the two dimensional turbulence kinetic-energy
equation for boundary layers (Equation (7.17), [2]). The turbulent shear stress is defined as   u ' v ' .

3
K 2
  CD (7.16)
L

 t u
P (7.17)
 y
Combining the two equations and substituting the shear stress yields:

3
K 2 u
CD  u ' v '
L y
Again we can rearrange the equation, assume L  and substitute equation (7.14).

34
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
u L
CD  u ' v '
y K 3 2
y
t
u u
CD  u ' v '
y K
3
2

t
t
u ' v ' u ' v '
CD  u ' v '
t 3
K 2

  u ' v ' 
1
2

CD   u ' v '  t
2

t K 2
3

3
 u ' v '  2
3
CD     0.3 2  0.1643
 K 

e.
The K   eddy viscosity model relates the eddy viscosity to the kinetic energy and the turbulence dissipation
(Equation (7.18)).

K2
 t  C (7.18)

The equation for the K  L eddy viscosity and the turbulence dissipation (Equations (7.11) and (7.16) respectively)
and the modeling constants can be used to obtain C .

K2
C K L  C 3
K 2
CD
L
3
LK 2 K
C  CCD  CCD
L K2
1 3
C  0.3 2  0.3 2  0.32  0.09

35
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
Bibliography

[1] Wikipedia, 2013. Viscosity. [online] Available at: <http://en.wikipedia.org/wiki/Viscosity#Air> [Accessed 23


April 2013].

[2] White, F.M., Viscous Fluid Flow. Third edition. New York: McGraw-Hill.

[3] Anderson, J. 2006. Fundamendals of Aerodynamics. 5th edition. New York: McGraw-Hill.

[4] Anderson, J. 2004. Introduction to Flight. 5th edition. New York: McGraw-Hill.

[5] Oudheusden, van B.W., 2012. AE4120 Viscous Flows, lecture slides.

36
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
MATLAB Code
close all;
clear all;
clc;
%%
% Viscous Flows Assignment, Robin Vermeij, 1357263
% 27-06-2013
%%
%select problem to solve (1 till 8)
problem = 3;

switch problem
case 1
%% PROBLEM 1.B - VISCOSITY AND REYNOLDS NUMBER
case 2
%% PROBLEM 2.A - 2D STAGNATION POINT FLOW
U_inf = 1; %Undisturbed velocity
R = 1; %Cylinder radius
Re = 1; %Reynolds number
Re_h = 5; %High Reynolds number
Re_l = 0.1; %Low Reynolds number

%Create meshgrid
[r,t] = meshgrid( R:0.01:3*R , deg2rad(-180):deg2rad(0.1):deg2rad(180));
%Convert to cartesian coordinates
x = r .* cos(t);
y = r .* sin(t);

%Calculate stream function


psi = U_inf .* ( 1 - (R.^2 ./ r.^2) ) .* r .* sin(t);

%Calculate approximation of stream function at stagnation point


%(correct for local coordinate system, so x becomes x-R)
A = 2 * U_inf / R;
psi_appr = A .* (x-R) .* y;

%Calculate viscous (self-similar) solution for stagnation point


%B = (1.2326 - ((x-R).*A.*Re) ./ (3*U_inf*R)) .* 0.5 .* A .* sqrt(Re*2);
B = 1.2326 * sqrt(2) * U_inf * sqrt(Re) / R^2;
psi_ss = (x-R).^2 .* y .* B;
psi_ss_h = (x-R).^2 .* y .* B ./ Re .* Re_h; %High Re
psi_ss_l = (x-R).^2 .* y .* B ./ Re .* Re_l; %Low Re

%Calculate differences
psi_perc = 100 ./ psi .* psi_appr - 100;
psi_perc_ss = 100 ./ psi .* psi_ss;
psi_perc_ss_h = 100 ./ psi .* psi_ss_h;
psi_perc_ss_l = 100 ./ psi .* psi_ss_l;

%delete invalid rows (Inf and NaN)


[row column] = size(psi_perc);
psi_perc(ceil(row/2),:) = [];
psi_perc(:,1) = [];
psi_perc_ss(ceil(row/2),:) = [];
psi_perc_ss(:,1) = [];
psi_perc_ss_h(ceil(row/2),:) = [];
psi_perc_ss_h(:,1) = [];
psi_perc_ss_l(ceil(row/2),:) = [];
psi_perc_ss_l(:,1) = [];
x_perc = x;
x_perc(ceil(row/2),:) = [];
x_perc(:,1) = [];
y_perc = y;
y_perc(ceil(row/2),:) = [];
y_perc(:,1) = [];

% mirror x matrices to make sure flow goes from left to right.


x = -x;
x_perc = -x_perc;

%Define isolines for stream function


iso = -3:0.3:3;
% define error ranges
iso_val = [0 1 5 10 20];

% Plot figures
figure('OuterPosition',[200 300 1500 600]);

%Stream function
37
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
subplot(1,3,1)
contourf(x,y,psi,iso)
colormap(bone)
caxis([-3 3])
colorbar('Location','SouthOutside')
axis('square');
title('Stream function')

%Stream function approximation


subplot(1,3,2)
contourf(x,y,psi_appr,iso)
axis('square');
caxis([-3 3])
colorbar('Location','SouthOutside')
title('Stream function approximation')

%Self-similar viscous approximation


subplot(1,3,3)
contourf(x,y,psi_ss,iso)
axis('square');
caxis([-3 3])
title('Self-similar viscous solution')
colorbar('Location','SouthOutside')

%Plot close up and show validity


figure('OuterPosition',[200 50 1000 1000]);
area = [-1.5*R -.99*R -.25*R .25*R];

%Stream function approximation


subplot(2,2,1)
contourf(x_perc,y_perc,abs(psi_perc),iso_val)
colormap(hot)
caxis([0 15])
axis(area)
title('Stream function approximation')

%Self-similar viscous approximation


subplot(2,2,2)
contourf(x_perc,y_perc,abs(psi_perc_ss),iso_val)
caxis([0 15])
axis(area)
title(['Self-similar viscous solution, Re=',num2str(Re)])

%Self-similar viscous high Re approximation


subplot(2,2,3)
contourf(x_perc,y_perc,abs(psi_perc_ss_h),iso_val)
axis(area)
caxis([0 15])
title(['Self-similar viscous solution, Re=',num2str(Re_h)])

%Self-similar viscous low Re approximation


subplot(2,2,4)
contourf(x_perc,y_perc,abs(psi_perc_ss_l),iso_val)
axis(area)
caxis([0 15])
title(['Self-similar viscous solution, Re=',num2str(Re_l)])
case 3
%% PROBLEM 3.B - SELF-SIMILAR STAGNATION-POINT BOUNDARY LAYER WITH SUCTION
f_w = [0 1 2 5 10]; % Wall suction
eta = 0:0.01:6; % Stretching coordinate
f2 = zeros(length(f_w),1);
f2_comp = [1.2326 1.8893 2.6701 5.3595 10.1939]';
eta_delta = zeros(length(f_w),1);
eta_delta_comp = [0.6479 0.4593 0.3422 0.1810 0.0972]';
eta_theta = zeros(length(f_w),1);
eta_theta_comp = [0.2923 0.2150 0.1639 0.0892 0.0484]';
H_comp = [2.2162 2.1365 2.0876 2.0295 2.0091]';
phi0 = zeros(length(f_w),length(eta));
phi1 = zeros(length(f_w),length(eta));
phi2 = zeros(length(f_w),length(eta));
phi3 = zeros(length(f_w),length(eta));

%Calculate approximated boundary layer


phi0_0 = exp(-eta) + eta;
phi1_0 = exp(-eta) + eta - 1;
phi0_1 = -exp(-eta) + 1;
phi1_1 = phi0_1;
phi0_2 = exp(-eta);
phi1_2 = phi0_2;
phi0_3 = -exp(-eta);

38
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
phi1_3 = phi0_3;

for i=1:length(f_w)
phi0(i,:) = phi0_0 + (1 / f_w(i)^2) * phi1_0;
phi1(i,:) = phi0_1 + (1 / f_w(i)^2) * phi1_1;
phi2(i,:) = phi0_2 + (1 / f_w(i)^2) * phi1_2;
phi3(i,:) = phi0_3 + (1 / f_w(i)^2) * phi1_3;
end

A = ones(size(eta));

%Integrate to calculate de values for question d.


for i=1:length(f_w) %Calculate for every f_w
f2(i) = f_w(i) .* phi2(i,1);
% eta_delta(i) = trapz(eta,1-phi1(i,:)) ./ f_w(i);
eta_delta(i) = (trapz(eta,A) - trapz(eta,phi1(i,:)) ) ./ f_w(i);
eta_theta(i) = trapz( eta , phi1(i,:) .* (1 - phi1(i,:)) ) ./ f_w(i);
end

f2
f2_err = 100 ./ f2_comp .* f2 - 100
eta_delta
eta_delta_err = 100 ./ eta_delta_comp .* eta_delta - 100
eta_theta
eta_theta_err = 100 ./ eta_theta_comp .* eta_theta - 100
H = eta_delta ./ eta_theta
H_err = 100 ./ H_comp .* H -100

figure()
plot(eta,phi0(5,:))
hold on
plot(eta,phi1(5,:),'r')
plot(eta,phi2(5,:),'g')
plot(eta,phi3(5,:),'k')
grid on
xlabel('\etaf_w')
title('Stagnation point boundary layer approximation for f_w=10')
legend('\phi','\phi`','\phi``','\phi```','Location','NW')

figure()
plot(eta,phi1(2,:),'r')
hold on
plot(eta,phi1(3,:))
plot(eta,phi1(4,:),'k')
plot(eta,phi1(5,:),'g')
axis([0 eta(end) 0 max(phi1(2,:))*1.1])
grid on
xlabel('\etaf_w')
ylabel('\phi`')
title('Approximated velocity profiles')
legend('f_w=1','f_w=2','f_w=5','f_w=10','Location','NW')

%Calculate right hand side of equation for phi_1 (which is assumed zero)
rhs = -2.*exp(-2*eta)-eta.*exp(-eta)+2.*exp(-eta);
% rhs2 = exp(-eta).*(eta+1);
figure()
plot(eta,rhs)
% hold on
% plot(eta,rhs2,'c')
xlabel('\etaf_w')
case 4
%% PROBLEM 4.E - FLAT PLATE BOUNDARY LAYER WITH UNIFORM SUCTION, INTEGRAL METHOD

xi_acc = [ 0
0.1
0.2
0.3
0.4
0.5
0.7
1.0
1.4
2.0
3.0
4.0
5.0 ];

%Import accurate numerical results for flat plate boundary layer with
%uniform suction

39
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
t_acc = [ 0
0.1747
0.2293
0.2657
0.2930
0.3147
0.3475
0.3816
0.4113
0.4395
0.4652
0.4787
0.4864];

S_acc = [ 0.2205
0.3008
0.3290
0.3487
0.3640
0.3766
0.3961
0.4172
0.4364
0.4555
0.4736
0.4835
0.4893];

S_bl = 0.66411^2 / 2; %S according to blasius solution calculated at xi = 0


S_sp = -.5*exp(-9.2)-exp(-4.6)+.5; %S according to suctin profile, not calculated yet
S_ad = S_bl; %Adaptive S, starts with blasius S.
S_ad = 0;

% integrate numerically over 0 < xi < 5


maxxi = 5; %end coordinate
minxi = 0; %begin coordinate
dxi = 0.0001; %stepsize
xi = minxi:dxi:maxxi; %xi-coordinate

%Blasius solution variables


t_bl = zeros(1,length(xi)); %theta for blasius solution
dt2dxi_bl = zeros(1,length(xi)-1); %change of theta per step

%suction profile variables


t_sp = ones(1,length(xi));
t_sp(1) = S_sp; %initial condition
dt2dxi_sp = zeros(1,length(xi)-1);

%Adaptive S variables
t_ad = zeros(1,length(xi));
dt2dxi_ad = zeros(1,length(xi)-1);

%solve differential equation


for i = 1:length(dt2dxi_bl)

%Blasius solution:
dt2dxi_bl(i) = 2 * ( S_bl - t_bl(i) );
t_bl(i+1) = sign(dt2dxi_bl(i)) * sqrt(abs(dxi * dt2dxi_bl(i))) + t_bl(i);

%Suction profile:
dt2dxi_sp(i) = 2 * ( S_sp - t_sp(i) );
t_sp(i+1) = dxi * sign(dt2dxi_sp(i)) * sqrt(abs(dt2dxi_sp(i))) + t_sp(i);

%Adaptive S:
S_ad = 0.2205 + 0.5499 * t_ad(i);
% S_ad = 0.2205 + 0.2750 * 2* t_ad(i);
dt2dxi_ad(i) = 2 * ( S_ad - t_ad(i) );
t_ad(i+1) = dxi * sign(dt2dxi_ad(i)) * sqrt(abs(dt2dxi_ad(i))) + t_ad(i);

end

S_plot = [0.2205 0.2205+0.5499*t_sp(end)];

% Plot solutions
figure()
plot(xi_acc,t_acc.^2,'LineWidth',1.5)
hold on
plot(xi,t_bl.^2,'r','LineWidth',1.5)
plot(xi,t_sp.^2,'g','LineWidth',1.5)
plot(xi,t_ad.^2,'k:','LineWidth',1.5)

40
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
plot(xi,zeros(1,length(xi)),'c')
grid on
xlabel('\xi')
ylabel(['\theta','^*^2'])
title('\xi versus \theta^*^2 for a boundary layer with uniform suction')
legend('Finite difference solution', 'Blasius solution', 'Suction profile solution','Adaptive S solution')

figure()
plot([t_bl(1) t_sp(end)],S_plot)
hold on
plot(t_acc,S_acc,'r')
axis([0 0.5 0 0.5])
xlabel('\theta^*')
ylabel('S')
title('\theta^* versus S')
grid on
legend('Approximate linear relation','Accurate numerical relation','Location','SE')
case 5
%% PROBLEM 5.D - ADIABATIC FLAT PLATE BOUNDARY LAYER
Pr = 0:0.01:2; %Prandtl number Pr
e = (Pr - 1); %small paramater epsilon
r = Pr.^.5; %recovery factor r

T0 = exp(0);
T1 = 1;
T = T0 + e .* T1;

figure()
subplot(1,2,1)

%Calculate difference
d = r - T;

plot(Pr,r)
hold on
plot(Pr,T,'r');
grid on
xlabel('Prandtl number [-]')
ylabel('r')
title('Flat plate recovery factor')
legend('r=Pr^{1/2}','Asymptotic solution','Location','SE')

subplot(1,2,2)
plot(Pr,d)
grid on
xlabel('Prandtl number [-]')
% ylabel('')
title('difference between the asymptotic solution and Pr^{1/2}')
case 6
%% PROBLEM 6.D - BOUNDARY LAYER CALCULATIONS BASED ON SIMPLE CORRELATIONS
Rex = linspace(0,1E7,100000);
Rex_trans = 1.8E6;
dtdx = zeros(length(Rex),1);
t = zeros(length(Rex)-1,1);

A_lam = 0.45;
A_tur = 0.0076;
B_lam = 5;
B_tur = 10/3;
n_lam = 1;
n_tur = 1/6;

for i=1:length(Rex)
if Rex(i) < 1.8E6
dtdx(i) = sqrt(A_lam / Rex(i)^n_lam);
else
dtdx(i) = (A_tur / Rex(i)^n_tur)^(1 / (n_tur+1));
end

end

%remove infinite first value


dtdx(1)=[];

%Integrate to determine the momentum thickness


for i=2:length(Rex)-1
t(i) = trapz( Rex(1:i), dtdx(1:i));
end

%Make a matrix only containing the values for theta for the turbulent part

41
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
%for curve fitting
turID = find(Rex > Rex_trans); %find indices of turbulent part
f = fit(Rex(turID)',t(turID-1),'exp2');

%Calculate turbulent boundary layer


t_tur = f.a .* exp(f.b .* Rex) + f.c .* exp(f.d .* Rex);

%Calculate virtual origin turbulent boundary layer


t_tur_start = -(log(-f.a/f.c))/(f.b-f.d);

figure()
plot(Rex(2:end),dtdx)
hold on
plot(t_tur_start,0,'og')
grid on
xlabel('Re_x')
ylabel('d\theta/dx')
title('Boundary layer growth for a flat plate flow')
axis([0 Rex(end) 0 0.005])
legend('Boundary layer growth','virtual origin turbulent BL','Location','NE')

figure()
plot(Rex(2:end),t')
hold on
plot(Rex,t_tur,'r:')
hold on
plot(t_tur_start,0,'og')
xlabel('Re_x')
ylabel('\theta')
title('Boundary layer development')
legend('BL momentum thickness','Turbulent BL fit',...
'virtual origin turbulent BL','Location','SE')
grid on
axis([0 Rex(end) 0 1.1*max(t)])
case 7
%% PROBLEM 7.D - THE CLAUSER PLOT
%Import velocity profile
U = [1 0.25 0.203
2 0.50 0.339
3 1.00 0.456
4 1.50 0.508
5 2.00 0.541
6 2.50 0.563
7 3.00 0.579
8 3.50 0.593
9 4.00 0.605
10 4.50 0.616
11 5.00 0.626
12 6.00 0.643
13 7.00 0.657
14 8.00 0.670
15 9.00 0.682
16 10.00 0.692
17 12.00 0.712
18 14.00 0.730
19 16.00 0.748
20 18.00 0.766
21 20.00 0.784
22 22.00 0.801
23 24.00 0.818
24 26.00 0.838
25 28.00 0.856
26 30.00 0.874
27 32.00 0.892
28 34.00 0.910
29 36.00 0.928
30 38.00 0.944
31 40.00 0.959
32 42.00 0.973
33 44.00 0.984
34 46.00 0.993
35 48.00 0.998
36 50.00 1.000
37 52.00 1.000];

%Convert [mm] to [m]


U(:,2) = U(:,2) .* 1E-3;

U_e = 9.636; %external flow velocity [m/s]


v = 15E-6; %dynamic viscosity [m^2/s]

42
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
kappa = 0.41; %von Karman constant
B = 5; %Integration constant
valrange = [35 350]; %Validity range of law of the wall, White p.417

%Calcualte Re_y = yU_e/v


U(:,4) = U(:,2) .* U_e ./ v;

%Initial skin friction coefficient


C_f = 0.0028;

%Calculate v*/U_e from initial skin friction coefficient


vu_e = sqrt(C_f/2);
%Calculate law of the wall velocity profile
U(:,5) = vu_e .* ( (1 / kappa) .* log( ( U(:,2) .* U_e .* vu_e) /v) + B);

%Convert validity range to correct units by dividing by v*/U_e


valrange = valrange ./ vu_e;

%Determine the wake component of the velocity profile within the validity
%range
i = 1;
while U(i,4) < valrange(1)
i = i + 1;
end

j = 1;
while U(j,4) < valrange(2)
j = j + 1;
end
j = j - 1;

wakecomp = U(i:j,3) - U(i:j,5);

wakecompmin = min(wakecomp);
wakecompmax = max(wakecomp);

if abs(wakecompmax) > abs(wakecompmin)


I = find(wakecomp == wakecompmax);
else
I = find(wakecomp == wakecompmin);
end

disp(['Strength of wake component = ', num2str(wakecomp(I)),' at heigth ', num2str(U(I,2)),'mm, Re_y = ',
num2str(U(I,4)) ])

%Plot velocity profile on semi logarithmic scale


figure()
semilogx(U(:,4),U(:,3))
hold on
semilogx(U(:,4),U(:,5),'r')
hold on
plot([valrange(1) valrange(1)],[0 U(end,3)],'k:')
plot([valrange(2) valrange(2)],[0 U(end,3)],'k:')
plot(U(I,4),U(I,3),'o')
axis([0 U(end,4) 0 U(end,3)])
ylabel('u/U_e [-]');
xlabel('Re_y [-]')
title('Measured mean velocity profile of a turbulent boundary layer')
legend('Measured velocity profile',['Law of the wall, C_f=', num2str(C_f)],'Validity
boundaries','Location','NW')
case 8
%% PROBLEM 8.E - TURBULENCE MODELING WITH THE EDDY VISCOSITY
du = 0.001;
u = 0:du:25;
y = zeros(1,length(u));
k = 0.41; %Von Karman constant kappa
B = 5; %Integration constant B

%Calculate Spalding's law


y = u + exp(-k*B) .* (exp(k.*u) - 1 - k.*u - (k.*u).^2 / 2 - (k.*u).^3 / 6);

%Calculate du/dy
dudy = du ./ diff(y);
n = (1 - dudy) .* (dudy).^-1;

figure()
semilogx(y,u)
% plot(y,u)
xlabel('y^+')
ylabel('u^+')

43
AE4120 Viscous Flows Robin Vermeij, 1357263 27-6-2013
title('Spalding`s law')

figure()
plot(y(1:end-1),n)
xlabel('y^+')
ylabel('\nu_t/\nu')
title('Variation of eddy viscosity')
axis([0 50 0 20])
grid on

figure()
plot(y(1:end-1),n.*1.51E-5)
xlabel('y^+')
ylabel('\nu_t')
title('Variation of eddy viscosity of air')
axis([0 50 0 3E-4])
grid on

eta = 0:0.01:1; %scaled y-coordinate


pi = 0.5;
Fv = (1 / k) + ((12 .* pi) ./ k ) .* ( eta.^2 - eta.^3 );

figure()
plot(eta,Fv)
xlabel('\eta')
ylabel('F/v*')
grid on

%% End switch function


end

44

You might also like