Vous êtes sur la page 1sur 47

Fluid Mechanics, 6th Ed.

Kundu, Cohen, and Dowling

Exercise 10.1. A thin flat plate 2 meters long and 1 meter wide is placed at zero angle of attack
in a low speed wind tunnel in the two positions sketched below.
drag on the plate in position #1
a) For steady airflow, what is the ratio: drag on the plate in position #2 ?
b) For steady airflow at 10 m/sec, what is the total drag on the plate in position #1?
c) If the air flow is impulsively raised from zero to 10 m/sec at t = 0, will the initial drag on the
plate in position #1 be greater or less than the steady-state drag value calculated for part b)?
d) Estimate how long it will take for drag on the plate in position #1 in the impulsively started
flow to reach the steady-state drag value calculated for part b)?

Solution 10.1. a) Here we need only consider the drag coefficient, CD, from the Blasius solution.
CD (L) =
(Drag per unit span on one side of the plate of length L) = 1.328
1
2
ρU 2 L ReL
where ReL= UL/ν. Therefore:
1.328 1
Total Drag on plate "j" = 2(span of plate "j") ⋅ ⋅ ρU 2 L where j = 1 or 2 , so
€ ReL 2
Drag on plate #1 span of plate #1 L1 L1 2
= ⋅ = = 2.
Drag on plate # 2 span of plate # 2 L2 L2 2 2
€ 1.328 1
b) Total Drag on plate #1 = 2(span of plate #1) ⋅ ⋅ ρU 2 L1
ReL1 2
€ 2(2m)1.328 1
= ⋅ (1.21kg /m 3 )(10m /s) 2 (1m)
(10m /s)(1m) (1.5 ×10 m /s) 2
−5 2

= 0.394 N. €
c) When the flow is impulsively started, a temporally-developing boundary layer is produced on
the top and bottom of plate #1. Initially, the shear stress is very high since: τ w ≅ µU(πνt)−1 2 so

the initial drag will be greater than the steady-state drag.
d) The flow will have reached steady-state when the temporally developing boundary layer skin
friction has reached the Blasius boundary layer skin friction everywhere on the plate. The
Blasius skin friction is lowest at the trailing edge, so an estimate for € the drag-relaxation time can
be based on matching skin friction coefficients at the plate's trailing edge between the
temporally-developing boundary layer and the Blasius solution. Therefore, with ReL= UL/ν:
4L L
(c f )temporal BL ≈ (c f )spatial BL at x = L –> U2 πνt ≈ 0.664
ReL
, or t ≈ 2
π (0.664) U
≈ 3.0 = 0.30s .
U
Interestingly, this result is independent of the fluid viscosity!

€ € €
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.2. Solve the Blasius equations (10.27) through (10.29) with a computer, using the
Runge–Kutta scheme of numerical integration, and plot the results. What value of f !! at η = 0
leads to a successful profile?

Solution 10.2. The Blasius equation is a non-linear third-order differential equation for f(η):
d3 f 1 d2 f
+ f = 0.
dη 3 2 dη 2
The boundary conditions are: df dη →1 as η → ∞, df/dη = f = 0 at η = 0. For a computer
solution using a Runge-Kutta integration scheme, this equation must be reduced to a set of three
first-order differential equations by defining:

g(η) = df dη , and h(η) = d 2 f dη 2 .
€ €
The above equation can then be written as three equations:
df dη = g(η) (A)
dg dη = h(η) (B)
€ €dh dη = − fh 2 (C)
subject to:

f(0) = g(0) = 0 and g(∞) = 1.
The set of three equations€ are readily integrated via the Runge-Kutta method starting

from η = 0 where the initial values of f and g are known. The initial value of h (= f !! at η = 0) is
not known but it can be found by trial and error (0.33205) by looking for the value of h(0) that
produces g = 1 at some suitably large value of η, perhaps 10 or 20.
A simple MatlabTM code that does this based on a trial & error value of h(0) is:

%Compute the Blasius Boundary Layer Profile


clear;
clc;
f0 = 0; %The first known boundary condition
g0 = 0; %The second known boundary condition
h0 = 0.33205; %The value adjusted by trial & error
eta_start = 0; %Starting point for eta
eta_end = 10; %End point for eta
%The next command invokes a Runga-Kutta integration scheme
[eta,f] = ode45(@Blasius,[eta_start eta_end],[f0 g0 h0]);

with the function

function df = blasius(eta,f)
% Solve the Blasius LBL profile equation.
% f(1)=stream function, f(2)=velocity, f(3)=velocity gradient
df = zeros(3,1);
df(1) = f(2); %The equivalent of (A)
df(2) = f(3); %The equivalent of (B)
df(3) = -0.500*f(1)*f(3); %The equivalent of (C)
end

The resulting plot of df/dη vs. η is:


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

$"
!#,"
!#+"
!#*"
!#)"
df u !#("
=
dη U
!#'"
!#&"
€ !#%"
!#$"
!"
!" %" '" )" +" $!"

η = y U νx


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.3. A flat plate 4 m wide and 1 m long (in the direction of flow) is immersed in
kerosene at 20°C, (v = 2.29 × 10 6 m2/s, ρ = 800 kg/m3) flowing with an undisturbed velocity of

0.5 m/s. Verify that the Reynolds number is less than critical everywhere, so that the flow is
laminar. Show that the thickness of the boundary layer and the shear stress at the center of the
plate are δ = 0.74 cm and τ0 = 0.2 N/m2, and those at the trailing edge are δ = 1.05 cm and τ0 =
0.14 N/m2. Show also that the total frictional drag on one side of the plate is 1.14 N. Assume that
the similarity solution holds for the entire plate.

Solution 10.3. ReL = UL/ν = (0.5m/s)(1m)/(2.29x106 m2/s) = 2.18x105 < Recr ~ 106. Thus, the
flow is expected to be laminar everywhere.
At x = 0.5 m, Rex = 1.09x105 so the 99% thickness from (10.30) and the shear stress from (10.31)
are:
δ99 = 4.9x Re1x 2 = 4.9(0.5) 1.09 ×10 5 = 0.742cm , and
τ 0 = 0.332 ρU 2 Re1x 2 = 0.332(800)(0.5) 2 1.09 ×10 5 = 0.201N /m 2 .
At x = 1.0 m, Rex = 2.18x105 so
€ δ99 = 4.9x Re1x 2 = 4.9(1.0) 2.18 ×10 5 = 1.05cm , and
€ τ 0 = 0.332 ρU 2 Re1x 2 = 0.332(800)(0.5) 2 2.18 ×10 5 = 0.142N /m 2 .
The total drag can be obtained from (10.33):
12 −3
CD = 1.33 Re€ x = 2.85 ×10 , so
1
€ D = ρU 2 (Area)CD = 0.5(800)(0.5) 2 (4)(1)(0.00285) = 1.14N .
2


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.4. A fluid with constant density and viscosity flows with a constant horizontal speed
U∞ over an infinite flat porous plate placed at y = 0 through which fluid is drawn with a constant
velocity Vs. For this flow the steady two-dimensional zero-pressure-gradient boundary layer
equations are (7.2) and (10.18) and the boundary conditions are u(y = 0) = 0, v(y = 0) = –Vs, and
u = U∞ for y → ∞ .
a) Assuming u depends only on y, determine u(y) in terms of ν, Vs, U∞, and y.
b) What is the wall shear stress τw? How does it depend on µ?
c)€What parametric change(s) decrease the boundary layer thickness?

Solution 10.4. a) If u = u(y), then ∂u/∂x = 0 = ∂v/∂y and this means that v is a most a function of
x. However, the boundary condition on v at y = 0 does not depend on x, thus v = const. = –Vs.
∂u ∂ 2u
Therefore, the horizontal momentum equation simplifies to: −Vs = ν 2 , which can be
∂y ∂y
integrated to find: −Vs y ν = ln(∂u ∂y ) +C , where C is a constant. This can be rearranged and
integrated again to find: u(y) = D + E exp(−Vs y ν ) . The boundary condition, u(0) = 0 requires D
+ E = 0, while u(∞) = U∞ sets D = U∞. Thus, the velocity€profile is: u(y) = U∞ [1− exp(−Vs y ν )]
€ shear stress is τ w = µ(∂u ∂y ) y= 0 = −µU∞ (−Vs ν ) = ρU∞Vs . It does not depend on µ!
b) The wall
€ layer thickness decreases when Vs increases, and ν decreases. The horizontal
c) The boundary
flow speed does not influence the thickness of this boundary
€ layer.

Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.5. A square-duct wind tunnel test section of length L = 1 m is being designed to
operate at room temperature and atmospheric conditions. A uniform air flow at U = 1 m/s enters
through an opening of D = 20 cm. Due to the viscosity of air, it is necessary to design a variable
cross-sectional area if a constant velocity is to be maintained in the middle part of the cross-
section throughout the wind tunnel.
a) Determine the duct size, D(x), as a function of x.
b) How will the result be affected if U = 20 m/s? At a given value of x, will D(x) be larger or
smaller than (or the same as) the value obtained in a)? Explain.
c) How will the result be affected if the wind tunnel is to be operated at 10 atm (and U = 1 m/s)?
At a given value of x, will D(x) be larger or smaller than (or the same as) the value obtained in
a)? Explain. [Hint: the dynamic viscosity of air (µ [N·s/m2]) is largely unaffected by pressure.]
d) Does the airflow apply a net force to the wind tunnel test section? If so, indicate the direction
of the force.

Solution 10.5. a) The duct length (1 m) and the flow speed (1 m/s) imply a Reynolds number of:
ReL = UL/ν = (1 m/s)(1 m)/(1.5x10–5m2/s) = 67,000, which much larger than unity but still in the
laminar boundary layer range.
If δ* is the displacement thickness of the boundary layer at position x, then conservation
of mass between the inlet of the duct and location x implies:
2
ρUD2 (0) = ρU ( D(x) − δ * ) .
For constant density and velocity in the duct, the pressure gradient must be zero. Therefore, the
Blasius solution for a flat plate laminar boundary layer can be used, and the above equation
reduces to:

D(x) = D(0) + δ * = D(0) + 1.72 νx U = 0.20m + (1.5 ×10−5 m 2 /s)x (1m /s)
= 0.20 + 0.00387 x,
where the final numbers provide D in meters when x is in meters.
b) If U is 20 m/s instead of 1 m/s and the flow remains laminar, the displacement thickness of the
boundary layer will be reduced by a factor of 20 ≈ 4.47. Thus, the required D(x) will be

D(x) = D(0) + 1.72 νx U = 0.20m + (1.5 ×10−5 m 2 /s)x (20m /s) = 0.20 + 0.000866 x .
Here, the requisite D(x) is smaller because the boundary layer will be thinner at the higher speed.
c) The factor of ten increase in pressure
€ causes a factor of 10 increase in density. Thus for
constant µ, ν = µ/ρ will decrease by a factor of 10. This implies:

D(x) = D(0) + 1.72 νx U = 0.20m + (1.5 ×10−6 m 2 /s)x (1m /s) = 0.20 + 0.00122 x .


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

So, the requisite D(x) will be smaller than the part a) result because the boundary layer will be
thinner at the higher pressure because of the drop in kinematic viscosity (or equivalently, the
increase in Reynolds number).
d) Yes, the airflow applies a force to the wind tunnel via wall shear stress. This force points to
the right.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.6. Use the control volume shown to derive the definition of the momentum
thickness, θ, for flow over a flat plate:
h
u" u% Drag force on the plate from zero to x x
ρU 2θ = ρU 2 ∫ $1− 'dy = = ∫ τ w dx
0 U
# U& unit depth into the page 0
The words in the figure describe the upper and lower control volume boundaries.
y! U!
zero shear stress, !
U! constant pressure, no through flow!
h > δ99!
ho!
no slip, τw ≠ 0!

x!

Solution 10.6. The CV is stationary, so the integral form of the continuity equation is:
h
ρUho = ρ ∫ u(y)dy ,
0
where h is defined in the figure to be greater than the boundary layer thickness. The pressure is
the same everywhere so the integral form of the x-momentum equation is:
h x
€ −ρU 2 ho + ρ ∫ u 2 (y)dy = − ∫ τ w dx .
0 0

Here it is assumed that τw is positive (since µ and ∂u/∂y will be positive for the boundary layer
flow inside the CV), so the right-side integral represents the force the plate applies to the fluid in
the CV. Using the continuity equation, eliminate ho from the x-momentum equation:
h h x
−ρU ∫ u(y)dy + ρ ∫ u 2 (y)dy = − ∫ τ w dx .
0 0 0

Combine the two integrals, multiply by minus one, factor out ρU2, and recognize the definition
of the momentum thickness θ:
h ∞ x
2 u(y) # u(y) & 2
ρ ∫ u(y) (U − u(y)) dy = ρU ∫ %1− ( dy = ρU θ = ∫ τ w dx .
0 U U '
0
$ 0
Since u(y) = U for y > h, it is OK to extend the upper limit of the y-integration to +∞. The
integral of the shear stress on the right side of the last equation corresponds to the force the flow
exerts on the plate (per unit depth into the page). This force is positive and points in the
downstream direction (increasing x) and is therefore a drag force.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.7. Estimate the 99% boundary layer thickness on:


a) a paper airplane wing (length = 0.25 m, U = 1 m/sec),
b) the underside of a super tanker (length = 300 m, U = 5 m/sec), and
c) an airport run way on a blustery day (length = 5 km, U = 10 m/sec).
d) Will these estimates be accurate in each case? Explain.

Solution 10.7. Estimate the various thicknesses from the laminar zero-pressure-gradient
(Blasius) boundary layer solution: δ99 = 4.9 νx U .
a) δ99 = 4.9 (1.5 ×10−5 m 2 s–1 )(0.25m) (1ms−1 ) = 9.5mm
b) δ99 = 4.9 (1.0 ×10−6 m 2 s–1 )(300m) (5ms−1 ) = 3.8cm
c) δ99 = 4.9 (1.5 ×10−5 m€ 2 –1
s )(5000m) (10ms−1 ) = 0.42m
€ d) The estimate for part a) may be fairly accurate, but the estimates for b) and c) will not be
€ accurate. The downstream-distance-based Reynolds numbers in a), b), and c) are 1.7 ×10 4 ,
1.5 ×10 9 , and 3.3 ×10 9 respectively. At the larger two Reynolds numbers, the boundary layer

will be turbulent and its 99% thickness will be much greater than the Blasius boundary layer
estimate given above. €
€ €
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.8. Air at 20°C and 100 kPa (ρ = 1.167 kg/m3, ν = 1.5 × 10 5 m2/s) flows over a thin

plate with a free-stream velocity of 6 m/s. At a point 15 cm from the leading edge, determine the
value of y at which u/U = 0.456. Also calculate v and ∂u/∂y at this point. [Answer: y = 0.857 mm,
v = 0.384 cm/s, ∂u/∂y = 3012 s–1.]

Solution 10.8. From Table 10.1, η = y U νx = 1.4 when u/U = 0.456. Therefore,
y = 1.4 ν x U = 1.4 (1.5 ×10 −5 )(0.15) 6 = 0.857 mm.
At this wall normal distance, the Blasuis BL formula for v and Table 10.1 produce:
1 νU " € df % 1 1.5 ×10 −5 (6)
v= $− f + η ' = (−0.325 +1.4 ⋅ 0.456) = 0.00384 m/s.
2 x # dη & 2 0.15
Again using Table 10.1, the slope of the velocity profile is:
∂u ∂ (u /U) ∂η
=U = Uf !! U ν x = (6)(0.3074) 6 (1.5×10 −5 )(0.15) = 3012 s–1.
∂y ∂η ∂ y
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.9. An incompressible fluid (density ρ, viscosity µ) flows steadily from a large
reservoir into a long pipe with diameter D. Assume the pipe-wall boundary layer thickness is
zero at x = 0. The Reynolds number based on D, ReD, is greater than 104.
a) Estimate the necessary pipe length for establishing a parabolic velocity profile in the pipe.
b) Will the pressure drop in this entry length be larger or smaller than an equivalent pipe length
in which the flow has a parabolic profile? Why?

Solution 10.9. a) Fortunately, the exercise asks for an estimate. The flow in the entrance length
of a round pipe will accelerate on the pipe's centerline, and the inner wall of the pipe is curved.
Both of these features will cause the boundary layer growth inside the pipe to differ from that of
a Blasius boundary. However, the Blasius solution does account for diffusive boundary layer
growth, so it should be fine for producing an estimate of the entrance length L.
The presence of the pipe walls is felt on the pipe centerline when the wall boundary layer
has attained a thickness of D/2. Therefore, set δ99 ≈ D/2 and estimate δ99 from the Blasius
solution:
D/2 ≈ 5(νL/U)1/2 = 5(LD/ReD)1/2 ,
and solve for L to find: L ≈ 10–2D(ReD).
b) The pressure drop in the entry section will be larger because: 1) wall shear stresses are larger
when BL's are thin, and 2) the central portion of the flow must accelerate to accommodate the
displacement effect of the BL so the velocity at the edge of the boundary layer increases from the
x = 0 to x = L.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.10. A variety of different dimensionless groups have been used to characterize the
importance of a pressure gradient in boundary layer flows. Develop an expression for each of the
following parameters for the Falkner-Skan boundary layer solutions in terms of the exponent n in
Ue(x) = axn, Rex = Uex/ν, integrals involving the profile function f ! , and f !!(0) , the profile slope
at y = 0. Here u(x, y) = Ue (x) f ! ( y δ (x)) = Ue f !(η ) and the wall shear stress
τ w = µ (∂ u ∂ y ) y=0 = (µUe δ (x)) f !!(0) . What value does each parameter take in a Blasius
boundary layer. What value does each parameter achieve at the separation condition?
a) (ν Ue2 ) ( dUe dx ) , an inverse Reynolds number
b) (θ2/ν)(dUe/dx), the Holstein and Bohlen correlation parameter

(
c) µ )
ρτ w3 ( dp dx ) , Patel's parameter
d) (δ * τ w ) ( dp dx ) , Clauser’s parameter

Solution 10.10. a) Here Ue(x) = axn, so dU e dx = nax n−1; thus


ν dU e ν nax n−1 ν n
2
= n
=n = .
U e dx U e ax U e x Re x
The profile function does not€enter here. This parameter is zero for the Blasius boundary layer,
and is –0.0904/Rex at separation.
b) This time the profile function enters through the definition of the momentum thickness.
€ 2
θ 2 dU e 1 % ∞ u % u ( ( dU e δ 2 % ∞ ( 2 dU e
= '∫ '1− *dy * = ' ∫ f .(η)(1− f .(η))dη*
ν dx ν & 0 U e & U e ) ) dx ν &0 ) dx
δ2 % ∞ ( 2 dU e 1 % νx (% ∞ ( 2 nU e
= ' ∫ f .(η)(1− f .(η))dη* = ' *' ∫ f .(1− f .)dη*
ν &0 ) dx ν & U e )& 0 ) x
%∞ (2
= n' ∫ f .(1− f .)dη* .
&0 )
'∞ *2
This parameter is zero for the Blasius boundary layer, and is –0.0904 ) ∫ f "(1− f ")dη, at
(0 +
separation.

c) This parameter involves the shear stress and pressure gradient directly. Differentiating the
steady Bernoulli equation without a body force produces dp dx + ρU e dU e dx = 0 , so
€ 32
µ dp µ 1 # dUe & µ # ν x Ue & nUe2
= % − ρU ( = − % ( ρ
ρ %$ µUe f !!(0) ('
32 e
ρτ w3 dx ρ (µUe f !!(0) δ (x)) $ dx ' x

µρν 3 4 1 x3 4 # 1 & 2 µ1 4 1 1 1
= −n 32 % 32 34 (U e = −n 32 14
32
ρµ [ f !!(0)] x $ Ue Ue ' ρ [ f !!(0)] x Ue1 4
14

1 1
= −n 32
[ f !!(0)] Re1x 4
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

This parameter is zero for the Blasius boundary layer, and it goes to +∞ at separation since f "" =
0 at separation and for the Falkner-Skan laminar boundary layer solutions –n = +0.0904 > 0.
d) This parameter involves the profile function through the displacement thickness and through
the shear stress. Using the differentiated Bernoulli relation specified in part c) leads to:
δ * dp $ ∞ ' δ (x) $ dUe ' $∞ ' δ 2 (x) € nUe2
= & ∫ (1− u U ) dy ) & −ρUe ) = − & ∫ (1− f *) dη ) ρ
τ w dx % 0 ( µUe f **(0) % dx ( %0 ( µUe f **(0) x
$∞ ' ν x Ue Ue2 n $∞ '
= −n & ∫ (1− f ) dη )
* ρ =− & ∫ (1− f *) dη ).
%0 ( µUe f **(0) x f **(0) % 0 (
This parameter is zero for the Blasius boundary layer, and it goes to +∞ at separation since f "" =
0 at separation and for the Falkner-Skan laminar boundary layer solutions –n = +0.0904 > 0.


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Solution 10.11. Consider the boundary layer that develops as a constant density viscous fluid is
drawn to a point sink at x = 0 on and infinite flat plate in two dimensions (x, y). Here Ue(x) = –
UoLo/x, so set η = y νx |U e | and ψ = − νx |U e | f (η) and redo the steps leading to (10.36) to
find f """ − f "2 + 1 = 0 . Solve this equation and utilize appropriate boundary conditions to find
2
&1− αe− 2η ) 3− 2
f " =€
3( − 2η
+ − 2 where€ α= .
'1+ αe * 3+ 2

Exercise 10.11. Use the specified Ue to develop the appropriate expressions for ψ and h.
€ €ψ = νx |U | f (η) = νx U o Lo f (η) = νU L f (η) , and
e o o
x
y y U o Lo y
η= = = .
νx |U e | 2
νx U o Lo ν x
The Cartesian
€ velocity components are:
∂ψ df (η) ∂η UL 1 UL
u= = − νU o Lo = − νU o Lo f ' o o = − o o f ' , and
∂y
€ dη ∂y ν x x
∂ψ df (η) ∂η UL ( y+ νU o Lo
v =− = + νU o Lo = νU o Lo f ' o o * − 2 - = − ηf ' .
∂x dη ∂ x ν ) x , x
The€boundary conditions are:
f " →1 as η → ∞ (far from the wall the boundary layer velocity matches the free-stream),
f " → 0 as η → 0 (no slip at the wall), and
€ ηf # → 0 as η → 0 (no flow through the wall).

€ € ∂u ∂u 1 ∂p ∂ 2u
The horizontal boundary layer momentum equation is (10.9), u + v = − + ν 2 . The
€ € ∂x ∂y ρ ∂x ∂y
€ differentiated
€ Bernoulli equation, dp dx + ρ U e dU e dx = 0 , allows this to be simplified:
∂u ∂u dU e ∂ 2u ∂u ∂u U 2 L2 ∂ 2u
u + v = Ue + ν 2 or u + v = − o 3 o + ν 2 .
∂x ∂y dx ∂y x
∂€ ∂y x ∂y
where the second version € of the equation follows when U e (x) = –U o o/x. Substitute in the
L
velocity components in the final equation to find:
UL ∂ % U L ( νU o Lo ∂% UL ( U 2 L2 ∂2 % U L (
€− o o f # '− o o f #* − €ηf # '− o o f #* = − o 3 o + ν 2 ' − o o f #* .
x ∂x & x ) x ∂y & x ) x ∂y & x )
Perform the differentiations and simplify:
UL % UL UL η( νU o Lo UL UL 1 U 2 L2 UL UL 1
− o o f #' + o 2 o f # + o o f ## * + ηf # o o f ## o o = − o 3 o − ν o o o o 2 f ### ,
€ x & x x x) x x ν x x x ν x
− f #(+ f # + ηf ##) + ηf f# ## = −1− f ### or − f #(+ f # + ηf ##) + ηf f# ## = − f #2 = −1− f ###.
With a little rearrangement, the final equality then leads to the desired equation: f """ − f "2 + 1 = 0 .
This equation can be solved for f´ as follows. Multiply by f "" and integrate to find:

1€ 2 1 3
€ ( f "") − f " + f " + A = 0 .
2 3
The constant A can be evaluated by considering the limit η → ∞ . The first € boundary condition

provides f "(∞) = 1, and if f´ is constant then f ""(∞) = 0 . Thus, the last equation becomes


€ €
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

0 – 1/3 + 1 + A = 0 as η → ∞ ,
which requires A = –2/3. Therefore,
1 2 2 1 2 12
( f "") = − f " + f "3 , or f "" = ± [2 − 3 f " + f "3 ] ,
2 3 3 € 3
which can be separated:
df " df " df " 2
1 2
= 1 2
= 1 2
= dη .
€[2 − 3 f " + f "3 ] [(2 + f "€)(1− f ")(1− f ")] (1− f ")[2 + f "] 3
The possibility of the minus sign in front of the square root is dropped because η is always
positive along with 1− f # and 2 + f ". The final equality can be integrated with the substitution
2
F€ = 2 + f ":
2FdF 2dF 2
2
= 2
= dη
€ € (3 − F )F 3 − F 3
€ Now set F = 3 tanh(γ ) , so that dF = 3 (1− tanh 2 (γ )) dγ , thus
2dF 2 3 (1− tanh 2 (γ )) dγ 2 2 % F ( 2
∫ 3 − F 2 €∫ 3 1− tanh2 (γ ) = 3 γ = 3 tanh−1'& 3 *) = 3η + C
=

( )
€ % 2 + f $(
2 2
Backtracking all the way to f " produces: η+C = tanh−1' *.
3 3 & 3 )
€ 2 # 2&
The constant C can be evaluated from f "(0) = 0 : C = tanh−1% (
€ 3 $ 3'
€ $ η 3C ' $
2 η 2'
Inverting for f " yields: f " = 3tanh 2 & + ) − 2 = 3tanh & + tanh−1 ) − 2.
€ % 2 2 ( % 2 3 (
To reach the desired form, use the sum€formula for the hyperbolic tangent, and the definition of
the hyperbolic tangent. Then manipulate the exponentials and the square roots.
€ $ η 2 −η 2 '
2
$ '
2
− 2η
€ 2 & e −e 2 ) & 1− e 2 )
$ tanh η 2 + 2 3 ' + +
f " = 3&
( ) ) − 2 = 3& e
η 2
+e − η 2
3 ) − 2 = 3& 1+ e − 2η
3 ) −2
& + . ) & + . )
% (
&1+ tanh η 2 2 3 )
( ) &1+ - e
η 2
−e −η 2
0
2) &1+ - 1− e
− 2η
0
2)
&% , eη 2 + e−η 2 / 3 )( &% ,1+ e− 2η / 3 )(
$ 2 '2 2
&1− e
− 2η
+
3
1+ e−( 2η
)) $ 3 − 3e−
) − 2 = 3&

(
+ 2 1+ e− ) 2η '
) −2
= 3&
& 2) & 3 + 3e− 2η
+ 2 (1− e ) − 2η )
&%1+ e
− 2η
(
+ 1− e− 2η
) 3 )(
% (
2
$ + 3− 2. − '

$ 3+ 2− '
2 &1− - 0e )
= 3&
( 3 − 2 e−) 2η
) − 2 = 3&
, 3+ 2/ ) − 2.
& + 3− 2. − )
& 3+ 2+
% ( 3− 2 )e − 2η
)
( &1+ - 0e

)
&% , 3+ 2/ )(
The final equality is the desired form.


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.12. Start from the boundary layer equations, (7.2), (10.9), and (10.10), and
ψ = Ue (x)δ (x) f (η ) , where η = y δ (x) , with δ(x) unspecified, to complete the following items.
a) Show that the boundary-layer profile equation can be written:
f !!! + α ff !! + β (1− f !2 ) = 0 , where α = (δ ν ) d(Ueδ ) dx , and β = (δ 2 ν ) dUe dx .
b) The part a) equation will yield similarity solutions when α and β do not depend on x.
Therefore, assume α and β are constants, set Ue = axn, and show that n = β/(2α – β).
c) Deduce the values of α and β that allow the profile equation to simplify to the Falkner-Skan
profile equation (10.36).

Solution 10.12. a) The given stream function automatically satisfies the continuity equation, so
the next step is to substitute it into (10.9):
∂u ∂u dUe ∂2 u
u + v = Ue +ν 2 .
∂x ∂y dx ∂y
Here, the surface-normal boundary-layer momentum equation (10.10), –∂p/∂y = 0, and the
Bernoulli equation in the free stream have been used to replace the pressure gradient;
− (1 ρ ) (∂p ∂x ) = Ue ( dUe dx ) .
For the given form of the stream function the various velocity components and
derivatives are:
∂ψ 1 ∂ψ y y
u= = Ueδ f " = Ue f " , v = − = −Ue"δ f −Ueδ ! f +Ueδ f ! 2 = −Ue"δ f −Ueδ ! f +Ue f ! ,
∂y δ ∂x δ δ
∂u y ∂u 1 ∂2 u 1
= Ue! f ! −Ue f !! 2 δ ! , = Uee f !! , and 2
= Ue f """ 2 .
∂x δ ∂y δ ∂y δ
where a prime denotes differentiation of a function with respect to its argument. Equation (10.10)
is then reconstructed:
# y & # y &# 1& 1
Ue f !%Ue! f ! −Ue f !! 2 δ !( + % −Ue!δ f −Ueδ ! f +Ue f ! (%Uee f !! ( = UeUe! + νUe f !!! 2 ,
$ δ ' $ δ '$ δ' δ
and this can be simplified by performing the indicated multiplications on the left side, and
canceling equal and opposite terms:
δ! 1
UeUe! f !2 −UeUe! ff !! −Ue2 ff !! = UeUe! + νUe f !!! 2 .
δ δ
Multiply through by δ2/νUe to reach:
δ2 δ2 δδ ! δ2
Ue! f !2 − Ue! ff !! −Ue ff !! = Ue! + f !!! ,
ν ν ν ν
collect all the terms to one side of the equation,
δ2 δ
f + Ue! (1− f !2 ) + (δUe! +Ueδ !) ff !! = 0 ,
!!!
ν ν
and combine the terms in the last set of parentheses:
δ 2 dUe δ d
f !!! +
ν dx
(1− f !2 ) +
ν dx
(δUe ) ff !! = 0 , or f !!! + β (1− f !2 ) + α ff !! = 0 ,
where the given definitions, α = (δ ν ) d(Ueδ ) dx & β = (δ 2 ν ) dUe dx , have been used to reach
the final equation.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

b) When α and β are constants, their defining equations may be used to eliminate δ. Start with
the α-equation:
δ d(Ueδ ) δUe dδ δ 2 dUe Ue dδ 2
α= = + = +β .
ν dx ν dx ν dx 2ν dx
Using the two ends of this extended equality, substitute for δ2 from the definition of β to find a
non-linear equation for Ue:
U dδ 2 U d ! νβ $ U β d 2Ue
α= e +β = e # &+ β = − e +β .
2ν dx 2ν dx " dUe dx % 2 ( dUe dx )2 dx 2
Again using the two ends of this extended equality and simplifying leads to:
2
α − β " dUe % Ue d 2Ue
$ ' = − ,
β # dx & 2 dx 2
and this equation has power-law solutions: Ue = axn. Substituting this trial solution into this
equation produces:
α − β 2 2 2n−2 ax n α −β 1
an x =− an(n −1)x n−2 which simplifies to: n = − (n −1) .
β 2 β 2
Solving this final algebraic equation for n produces the desired relationship: n = β/(2α – β).
β ! n +1 $
c) The part b) result can be rearranged to find: α = # & . Substitute this into the profile
2" n %
equation from part a) to reach:
β # n +1 &
f !!! + β (1− f !2 ) + % ( ff !! = 0 ,
2$ n '
and this equation will match (10.36) when β = n. Thus, the values of α and β that produce the
Falkner-Skan profile equation are:
α = (n + 1)/2 and β = n.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.13. Solve the Falkner-Skan profile equation (10.36) numerically for n = –0.0904, –
0.654, 0, 1/9, 1/3, and 1 using boundary conditions (10.28) and (10.29) and the Runge–Kutta
scheme of numerical integration. Plot the results and compare to Figure 10.8. What values of f !!
at η = 0 lead to successful profiles at these six values of n?

Solution 10.13. The Falkner-Skan profile equation is a non-linear third-order differential


equation for f(η):
2
d 3 f n +1 d 2 f " df %
+ f − n$ ' + n = 0 .
dη 3 2 dη 2 # dη &
The boundary conditions are: df dη →1 as η → ∞, df/dη = f = 0 at η = 0. For a computer
solution using a Runge-Kutta integration scheme, this equation must be reduced to a set of three
first-order differential equations by defining:
f (η ) = f (η ) , f2 (η ) = df dη , and f3 (η ) = d 2 f dη 2 .
€1 €
The above equation can then be written as three equations:
df1 dη = f2 (η ) (A)
df2 dη = f3 (η ) (B)
df3 dη = −(1 2)(n +1) f1 f3 + nf22 − n (C)
subject to:
f1(0) = f2(0) = 0 and f2(∞) = 1.
The set of three equations are readily integrated via the Runge-Kutta method starting
from η = 0 where the initial values of f and g are known. The initial value of f3 (= f !! at η = 0) is
not known but it can be found by trial and error by looking for the value of f3 that produces f2 = 1
at some suitably large value of η, perhaps 10 or 20. These initial values of f !! are : 0.00, 0.1640,
0.33205, 0.51181, 0.75741, 1.232533.

A simple MatlabTM code that does this based on a trial & error value of h(0) is:

%Compute the Falkner-Skan Boundary Layer Profile


clear;
clc;
eta_start = 0;
eta_end = 10;

f0 = 0; %The first known boundary condition


g0 = 0; %The second known boundary condition
n = [-0.0904, -0.0654, 0, 1/9, 1/3, 1]; %Values of n
h0 = [0.00, 0.1640, 0.33205, 0.51181, 0.75741, 1.232533];
%Values adjusted by trial & error

figure; hold all;


for i = 1:length(n);
%The next command invokes a Runga-Kutta integration scheme
[eta, f] = ode45(@(eta, f) FS(eta, f, n(i)), [eta_start
eta_end],[f0 g0 h0(i)]);
xlswrite(['FS_data_n=' num2str(n(i)) '.xlsx'],[eta f])
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

x=(.5*(n(i)+1)).^.5*eta;
plot(x,f(:,2));
end
df u
=
dη U with the function

function df = FS( eta, f, n )


%UNTITLED Falkner-Skan LBL Profile
%f(1)=stream function, f(2)=velocity, f(3)=velocity gradient
%n=1;
df = zeros(3,1);
df(1) = f(2); %same as (A)
df(2) = f(3); %same as (B)
df(3) = -0.500*(n+1)*f(1)*f(3) + n*f(2)^2 - n; %same as (C)
end

η =ηyis:U νx
The resulting plot of df/dη vs.

1"

0.9"

0.8"

0.7"

0.6"

0.5"

0.4"

0.3"

0.2"

0.1"

0"
0" 2" 4" 6" 8" 10"

The left most curve corresponds to n = 1, and the right-most curve corresponds to n = –0.0904
with the other monotonically arrayed in between. And – except for the aspect ratio and extent of
the horizontal axis – this plot is identical to Figure 10.8.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.14. By completing the steps below, show that it is possible to derive von Karman's
boundary layer integral equation without integrating to infinity in the surface-normal direction
using the three boundary layer thicknesses commonly defined for laminar and turbulent
boundary layers: i) δ (or δ99) = the full boundary layer thickness that encompasses all (or 99%) of
the region of viscous influence, ii) δ* = the displacement thickness of the boundary layer, and iii)
θ = momentum thickness of the boundary layer. Here, the definitions of the later two involve the
y=δ $
* u(x, y) ' y=δ
u(x, y) $ u(x, y) '
first: δ (x) = ∫ &1− )dy and θ (x) = ∫ &1− )dy , where Ue(x) is the flow
y= 0 % U e (x) ( y= 0 U e (x) % U e (x) (
speed parallel to the wall outside the boundary layer, and δ is presumed to depend on x too.
a) Integrate the two-dimensional continuity equation from y = 0 to δ to show that the vertical
d dU e
€ velocity at the edge of the boundary
€ layer is: v(x, y = δ ) = (U e (x)δ * (x)) − δ .
dx dx
b) Integrate the steady two-dimensional x-direction boundary layer momentum equation from y =
τ0 d 2 δ * (x) dU e2 (x)
0 to δ to show that: =
ρ dx
(
U e (x)θ (x) + ) 2 dx
.

d b(x ) # db & # da & b(x ) ∂f (x, y)
[Hint: Use Leibnitz’s rule ∫
dx a(x )
f (x, y)dy = % f (x,b) ( − % f (x,a) ( + ∫
$ dx ' $ dx ' a(x ) ∂x
dy to handle

the fact that


€ δ = δ(x)]
∂u ∂v
Solution 10.14. Start
€ with + = 0 , and integrate from y = 0 to δ to get:
∂x ∂y
δ
∂u
∫ ∂x dy +v(x, y = δ ) = 0
0
where v(x,y=0) =€0. Use Leibnitz’s rule to get the differentiation outside the integral:
δ
∂u d δ dδ
v(x, y = δ ) = − ∫ dy = − ∫ udy + U e (x)
€ 0 ∂x dx 0 dx
Add and subtract Ue(x) within the integral and rearrange the result:
d δ dδ
v(x, y = δ ) = −
dx
∫ (u − U e (x) + U e (x))dy + U e (x) .
dx
€ 0
δ
d d dδ
=− ∫
dx 0
(u − U e (x)) dy − (U eδ ) + U e (x)
dx dx
€ d# δ #
u & & dU e d dU e
= %U e (x) ∫ %1− (dy ( − δ = (U eδ * ) − δ
dx $ 0 $ U e (x) ' ' dx dx dx
€ ∂u ∂u 1 dp ∂ 2u
b) Start with u + v = − + ν 2 , multiply the continuity equation by u and add it to this
∂x ∂y ρ dx ∂y
2
€ ∂u ∂uv 1 dp ∂ 2u
equation to get: + =− + ν 2 . Use the Bernoulli equation substitution for the
∂x ∂y ρ dx ∂y
pressure
€ and integrate this equation from y = 0 to δ to get:
δ
∂u 2 dU τ
∫ ∂x dy + U e (x)v(x, y = δ ) = U eδ dxe − ρw
€ 0


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Use Leibnitz’s rule to get the differentiation outside the integral, substitute from part a) for the
vertical velocity at the edge of the boundary layer, and add and subtract uUe within the integral:
d δ 2 dδ %d dU e ( dU e τ w
dx 0
∫ ( u − uU e + uU e ) dy − U e2
dx
+ U e (x)' (U eδ * ) − δ
& dx dx )
* = U eδ
dx

ρ
Invoke the definition of the momentum thickness θ and then the definition of the displacement
thickness δ*:
& δ )
€ − d U 2θ + d (U ∫ ( u − U + U ) dy + − U 2 dδ + U (x)&( d U δ * − δ dU e )+ = U δ dU e − τ w
( ) ( )
e e e e e e
dx dx ' 0 * dx ' dx e dx *
e
dx ρ
d d dδ %d dU e ( dU e τ w
− (U e2θ ) + (−U e2δ * + U e2δ ) − U e2 + U e (x)' (U eδ * ) − δ * = U eδ −
dx dx dx & dx dx ) dx ρ
Now rearrange the result, and group like terms together:

τw d 2 d 2 * d dU e2δ dδ dU dU
ρ dx
= ( )
Ueθ +
dx
(
Ueδ − Ue ) dx
( )
U eδ * −
dx
+ U e2
dx
+ U eδ e + U eδ e
dx dx

*
The final four terms sum to zero, and the terms containing δ can be expanded to obtain:
2 * *
τw d 2 * dU e 2 dδ 2 dδ dU e d dU e
=
ρ dx
(
Ueθ + δ ) dx
+ Ue
dx
− Ue
dx
− δ *U e =
dx dx
( )
U e2θ + U eδ *
dx
,

and the final equality is identical to (10.43), von Karman's boundary layer integral equation.


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.15. Derive the von Karman boundary


y = h!
layer integral equation by conserving mass and Ue! Ue!
momentum in a control volume (C.V.) of width dx
and height h that moves at the exterior flow speed
Ue(x) as shown. Here h is a constant distance that is
comfortably greater than the overall boundary layer u(x,y)! !! u(x+dx,y)!
thickness δ.
p(x)! C.V.! p(x+dx)!

Solution 10.15. Use the integral laws as applied to a


moving control volume. Conservation of mass x! x+dx!
implies:
d$ h ' h
& ∫
dx ρBdy ) + ∫ ρ ([ u] x−dx / 2 − U e )e x ⋅ (−e x B)dy
dt % 0 ( 0
h
+ ∫ ρ([u] x +dx / 2 − U e )e x ⋅ (e x B)dy + ρ[(u − U e )e x + ve y ] y= h ⋅ e y Bdx = 0
0
The four terms above correspond to unsteady mass addition to the CV, mass flux on the left side
of the CV, mass flux on the right side of the CV, and mass flux on the top of the CV.
The CV is a constant size and the density is presumed to be constant, so the first term is
€ zero (the time derivative of a constant). Combine terms, perform the dot products, presume dx is
small, and divide by ρBdx to reach:
1 h h
∂u d %h (
∫ ([ u] x +dx / 2
− [ u] x−dx / 2 )
dy + [ v ] y= h
= 0 , or [ v ] y= h
= − ∫ dy = − ' ∫ udy * , (*)
dx 0 0 ∂x dx & 0 )
where the final equality follows from the fact that h is constant.
Now conserve horizontal momentum using the same CV.
d$ h ' h
€ & ∫
dx ρ uBdy ) + ∫ ρ [ u] x−dx / 2 ([ u] x−dx€ / 2 − U e )e x ⋅ (−e x B)dy
dt % 0 ( 0
h
+ ∫ ρ[ u] x +dx / 2 ([ u] x +dx / 2 − U e )e x ⋅ (e x B)dy + ρu[(u − U e )e x + ve y ] y= h ⋅ e y Bdx
0

= −τ 0 Bdx + ([ p] x−dx / 2 − [ p] x +dx / 2 ) hB


where τ0 is the wall shear stress, and p is the pressure. Simplify and combine terms, and divide
by ρBdx to reach:
d #h & 1 h 2 2 1 h
€ % ∫ udy ( +
dt $ 0 ' dx 0
∫ { }
[ u] x +dx / 2 − [ u] x−dx / 2 dy − U e ∫ {[u] x +dx / 2 − [u] x−dx / 2}dy + [uv ] y= h
dx 0
τ0 1
=− +
ρ ρdx
([ p]x−dx / 2 − [ p]x +dx / 2 )h
d dx d d
Use the chain rule on the total time derivative of the unsteady term, = = Ue , noting
dt dt dx dx
that horizontal velocity of the CV (= dx/dt) is Ue; and presume dx is small:

d #h & d h 2 d h τ 1 ∂p
U e % ∫ udy ( + ∫ u dy − U e ∫ udy + [ uv ] y= h = − 0 − h.
dx $ 0 ' dx 0 dx 0 € ρ ρ ∂ x


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Two terms on the left cancel, and at y = h, u = Ue and v is given by (*). In addition, the steady
Bernoulli equation implies: dp dx = − ρU e dU e dx . So, with these substitutions, the last equation
can be rewritten:
d h 2 d $h ' τ0 dU e τ0 dU e h
dx€0
∫ u dy − U e &∫
dx % 0
udy ) = −
ρ
+ U e
dx
h = −
ρ
+ U e
dx 0
∫ dy .
(
Rearrange this equation so the skin friction term appears by itself with positive sign. Then
manipulate the integrals to form the definitions of the momentum and displacement thicknesses.
τ0 d h 2 d &h ) dU e h

ρ
=−
dx 0
∫ u dy + U e ( ∫ udy + + U e
dx ' 0 dx 0
∫ dy
*
d & h u2 ) d & h u ) dU e h u dU e h
= − (U e2 ∫ 2 dy + + (U e2 ∫ dy + − U e ∫ dy + U e ∫ dy
dx ' 0 U e * dx ' 0 U e * dx 0 U e dx 0
d& h u & u) ) dU e h & u) d * dU e
= (U e2 ∫ (1− +dy = (U e θ ) + U eδ
2
(1− +dy + + U e
dx ' 0 U e ' U e * *

dx 0 ' U e * dx dx
.

The final equality is identical to (10.43), von Karman's boundary layer integral equation.


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.16. For the following approximate flat-plate boundary layer profile:
u %sin(πy 2δ ) for 0 ≤ y ≤ δ (
=& ) , where δ is the generic boundary layer thickness, determine:
U '1 for y > δ*
a) the displacement thickness δ*, the momentum thicknesses θ, and the shape factor H = δ * θ .
b) Use the zero-pressure gradient boundary layer integral equation to find: (δ x ) Re1x 2 ,
€ (δ * x )Re1x 2 , (θ x )Re1x 2 , c f Re1x 2 , and CD Re1L 2 for the approximate profile.
c) Compare these results to their equivalent Blasius boundary layer values.€

€ €
Solution € boundary layer velocity profile, u = sin$& πy ') = sin$& π ζ ') where
10.16.€a) The sinusoid
U % 2δ ( %2 (
y
ζ= , is a reasonably accurate approximate laminar boundary layer profile. The momentum
δ
thickness for this profile is:
δ 1 € π 2
u$ u' $ π '$ $ π '' 2δ
θ = ∫ &1− )dy = δ ∫ sin& ζ )&1− sin& ζ ))dζ = ∫ sin(ψ )(1− sin(ψ ))dζ
0 U
€ % U( 0
% 2 (% % 2 (( π 0
π 2 π 2
The integration is not too complicated, ∫ 0
sin(ψ )dψ = 1, and ∫ 0
sin 2 (ψ )dψ = π 4 , so
2δ & π ) & 4 − π )

θ= (1− + = δ( + = 0.1366δ
π ' 4 * ' 2π *
The displacement thickness for € theπ approximate profile
€ is:
δ 2
$ u' 2δ 2δ $ π ' $ π − 2 '
δ * = ∫ &1− )dy =
% € U( π 0
∫ (1− sin(ψ )) dψ = & −1) = δ&
π %2 ( % π (
) = 0.3634δ ,
0
so that the shape factor H = δ * θ = 2.66 . Substitution into the zero-pressure gradient (ZPG) Von
Karman boundary layer integral relationship produces:
τw dθ 2τ 2(πµU 2δ ) πν dθ ) 4 − π , dδ
€ = U2 , or c f = w2 = 2
= =2 =+ .
€ ρ dx ρU ρ U U δ dx * π - dx
where τ w = µ(∂u ∂y ) y= 0 = πµU 2δ has been obtained directly from the approximate sinusoidal
profile. Now use the fourth and sixth terms in the extended equality to form a differential
equation
€ for δ. €
€ πν & 4 − π ) d δ dδ % π 2 ( ν % 2π 2 ( ν ν
=( + , or δ =' * –> δ 2 = ' * x + const = 23.0 x + const .
Uδ ' π * dx dx & 4 − π ) U &4 − π )U U
Here the implicit assumption is that the boundary layer thickness is zero at x = 0 so the constant
ν 2
drops out. Thus: δ = 23 x = 4.80 ⋅ x ⋅ Re−1 x . Using the results above for the other length
€ U
€ €
scales and parameters produces:
θ 12
b) Rex = 0.654 (0.664)
x
€*
δ
Re1x 2 = 1.743 (1.721)
x
€ δ 12
Rex = 4.80 (5.0)
x


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

πν 1 2 πν
c f Re1x 2 =
Rex = 2
Re1x 2 = 0.654 (0.664)
Uδ U ⋅ 4.80xRe−1x

1 L 0.654 0.654 ν L dx 2 ⋅ 0.654 1 2


CD Re1L 2 = ∫ 12
dx ⋅ Re12
L = ∫ ⋅ Re1L 2 = ReL = 1.309 (1.328)
L 0 Rex L U 0 x Re1L 2
€ c) The Blasius boundary layer results are listed above at the right in parentheses. The sinusoid
velocity profile results are all within 4% of the Blasius values.

Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.17. An incompressible viscous fluid with kinematic viscosity ν flows steadily in a
long two dimensional horn with cross sectional area A(x) = Aoexp{βx}. At x = 0, the fluid
velocity in the horn is uniform and equal to Uo. The boundary layer momentum thickness is zero
at x = 0.
a) Assuming no separation, determine the boundary layer momentum thickness, θ(x), on the
lower horn boundary using Thwaites method.
b) Determine the condition on β that makes the no-separation assumption valid for 0 < x < L.
c) If θ(x = 0) was nonzero and positive, would the flow in the horn be more or less likely to
separate than the θ(x = 0) = 0 case with the same horn geometry?

Solution 10.17. a) As stated above, use Thwaites method to estimate the boundary layer
momentum thickness. Start with conservation of mass to determine the U(x): U(x)A(x) = U o Ao ,
so U(x) = U o Ao A(x) = U o exp{−βx} . Thus, the Thwaites equation becomes:
0.45ν x 5 0.45ν +6 β x x 5 −5 β x!
θ 2 (x) = ∫ U ( x !)d x ! = e ∫ Uo e € dx!
U 6 (x) 0 Uo6 0

€ $
0.45ν +6 β x e −5 β x
1 ' 0.09ν +6 β x
=
Uo
e & + )=
% −5β 5β ( βUo
e (1− e−5 β x )

θ 2 dU 0.09 +6 βx
b) First compute λ = = e (1− e−5 βx )(−βU oe− βx ) = −0.09(e +5 βx −1) . Separation will
ν dx βU o
have occurred if λ < λseparation = –0.090. Therefore, λ > λseparation at x = L is required to avoid
separation. This means:
−0.09(e +5 βL −1) > λseparation .

Convert this requirement to a condition on β:
λ 1 % λseparation ( 0.139
e +5 βL −1 < − separation , or β < ln&1− )≈
€ 0.09 5L ' 0.09 * L
c) If the boundary layer starts with a non-zero momentum thickness it is thicker and therefore
more likely to separate in an adverse pressure gradient.
€ €
y
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling
b
Exercise 10.18. The steady two-dimensional velocity potential for a source of strength q located
a distance b above a large flat surface located at y = 0 is:
x
q xsep 2
φ (x, y) =
2π ( 2 2
ln x + (y − b) + ln x + (y + b) 2
)
a) Determine U(x), the horizontal fluid velocity on y = 0.
b) Use this U(x) and Thwaites method to estimate the momentum thickness, θ(x), of the laminar
boundary layer that develops on the flat surface when the initial momentum thickness θo is zero.
x ξ 5 dξ x 6 (x 2 + 4b 2 )
[Potentially useful information: ∫ 0 2 = ]
(ξ + b 2 ) 5 24b 4 (x 2 + b 2 ) 4
c) Will boundary layer separation occur in this flow? If so, at what value of x/b does Thwaites
method predict zero wall shear stress?
d) Using solid lines, sketch the streamlines for the ideal flow specified by the velocity potential

given above. For comparison, on the same sketch, indicate with dashed lines the streamlines you
expect for the flow of a real fluid in the same geometry at the same flow rate.

! ∂φ $ q ! x x $ q! x $
Solution 10.18. a) U(x) = # & = # 2 2
+ 2 2&
= # 2 2&
" ∂ x %y=0 2π " x + (y − b) x + (y + b) %y=0 π " x + b %
b) Use the Thwaites integral without the θo-term.
6 6
2 0.45ν x 5 0.45νπ 6 ( x 2 + b 2 ) x q 5 " ξ %5 0.45νπ ( x 2 + b 2 ) x ξ 5dξ
θ (x) = 6
U (x) 0
∫ U (ξ )dξ = q6 x 6
∫ π 5 $# ξ 2 + b2 '& dξ = qx 6
∫ (ξ 2 + b2 )5
0 0
Evaluate the integral using the given information:
0.45νπ (x 2 + b 2 )6 x 6 (x 2 + 4b 2 ) 0.45νπ (x 2 + b 2 )2 2
θ 2 (x) = 6 4 2 2 4
= 4
(x + 4b 2 )
qx 24b (x + b ) 24b q
c) Compute the correlation parameter λ from the results of part b):
θ 2 dU 0.45π (x 2 + b 2 )2 2 2 q
" b 2 − x 2 % 0.45 " x 2 %" x 2 %
λ= = (x + 4b ) $ ' = $ 4 + '$1− ' .
ν dx 24b 4 q π # (x 2 + b 2 )2 & 24 # b 2 &# b 2 &
Clearly when x > b, λ becomes negative and its magnitude increases with increasing x, therefore
at some x > b boundary layer separation will occur. Within Thwaites method, λ = λsep = –0.090
predicts the point, xzs, where the wall shear stress is zero; therefore, use the prior result and the
quadratic formula to find
0.45 # x zs2 &# x zs2 &
λsep = % 4 + 2 (%1− 2 ( →
24 $ b '$ b '
12
x zs $ 3 9 + 4(4 − 24 λsep 0.45) '
= &− + € ) = 1.35 .
b &% 2 2 )(

d) The streamlines for the ideal flow correspond to a
point source above a flat surface. The real flow
streamlines must include boundary layer separation at
€ a value of x that is of order b. Furthermore, the
thickness of the separated flow region will cause
upward shifts in the ideal flow streamlines.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.19. A fluid-mediated particle-deposition process requires a laminar boundary layer


flow with a constant shear stress, τw, on a smooth flat surface. The fluid has viscosity µ and
density ρ (both constant). The flow is steady, incompressible, and two-dimensional, and the flat
surface extends from 0 < x < L. The flow speed above the boundary layer is U(x). Ignore body
forces.
a) Assume the boundary layer thickness is zero at x = 0, and use Thwaites’ formulation for the
shear stress, τ w = (µU θ ) l( λ) with λ = (θ 2 ν )( dU dx ) , to determine θ(x) and U(x) in terms of λ,
ν = µ ρ , x, and τ w µ = constant. [Hint: assume that U θ = A and l( λ) are both constants so
that τ w µ = Al( λ) .]
b) Using
€ the Thwaites integral (10.50) and the results of part a), determine λ.

c) Is boundary layer separation a concern in this flow? Explain with words or equations.
€ € € €

Solution 10.19. a) For Thwaites method, the shear stress is provided by the correlation
τ w = (µU θ ) l( λ) . Thus, the simplest way to achieve constant τw is for U to be proportional to θ,
i.e. U = Aθ where A is a constant, and l(λ) = constant, where λ = (θ 2 ν )( dU dx ) = constant.
Eliminating U from this system of equations leaves:
€ τw

= Al( λ ) , and λ = A(θ 2 ν )( dθ dx ) . (1,2)
µ €
Integrate (2) to find: λνx = Aθ 3 3 where the constant of integration is zero because θ = 0 at x =
0. Now use equation (1) to eliminate A in the relationship for θ. Thus, the solution falls into an
implicit form involving λ and l(λ). €
€ 13 13
€ θ (x) = ( 3λl( λ)νµx τ w ) and U(x) = Aθ (x) = ( 3τ w λνx µ l ( λ))
2 2 2

b) Put the part a) results into the Thwaites integral:


x
23 0.45ν 53
(3λl(λ)νµx τ w ) = 2 6 3 ∫ ( 3τ w2 λνξ µ 2 l 2 ( λ )) dξ .
€ €(3τ w λνx µ 2l 2 (λ)) 0
Perform the integration:
23 0.45ν 53 x 8 3 0.45 3νx
(3λl(λ)νµx τ w ) = 63 (3τ 2
w λν µ 2 l 2 ( λ ))
83
=
8 ( 3τ w2 λνx µ 2 l 2 ( λ))1 3
.
€ (3τ w2 λνx µ 2l 2 (λ))
Use the two ends of the equality and simplify to find:
23 13 0.45 0.45
(3λl(λ)νµx τ w ) (3τ w2 λνx µ 2 l 2 (λ)) 3νx , or λ =
= 3νxλ = = 0.05625
8 8

c) Boundary layer separation is not a concern here because λ and τw are both positive and
constant, so the separation condition λ = –0.090 is not approached.
€ €
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.20. The steady two-dimensional potential for incompressible flow at nominal
horizontal speed U over a stationary but mildly wavy wall is: φ (x, y) = Ux −Uε exp (−ky) cos ( kx ) ,
where kε << 1. Here, ε is the amplitude of the waviness and k = 2π/Λ, where Λ = wavelength of
the waviness.
a) Use the potential to determine the horizontal velocity u(x, y) on y = 0.
b) Assume that u(x, 0) from part a) is the exterior velocity on the wavy wall and use Thwaites’
method to approximately determine the momentum thickness, θ, of the laminar boundary layer
that develops on the wavy wall when the fluid viscosity is µ, and θ = 0 at x = 0. Keep only the
linear terms in kε and ε/x to simplify your work.
c) Is the average wall shear stress higher for Λ/2 ≤ x ≤ 3Λ/4, or for 3Λ/4 ≤ x ≤ Λ.
d) Does the boundary layer ever separate when kε = 0.01?
e) In 0 ≤ x ≤ Λ, determine where the wall pressure is the highest and the lowest.
f) If the wavy surface were actually an air-water interface, would a steady wind tend to increase
or decrease water wave amplitudes? Explain.
U! y!
ε!
x!
Λ/2! Λ!

Solution 10.20. a) The horizontal velocity u(x,y) is obtained from:


u = ∂φ ∂x = (∂ ∂x )[Ux − Uε exp(−ky ) cos( kx )] = U + Ukε exp(−ky ) sin( kx ) .
when evaluated on y = 0, this becomes: u(x, y = 0) = U (1+ kε sin(kx)) .
b) Evaluate the Thwaites integral with θ = 0 at x = 0 to find:
0.45ν x 5 0.45ν x
5
θ 2€= 6 ∫ U ( x $)d x $ = 6 ∫ (1+ kε sin(kx )) dx
$ $
U (x) 0 U (1+
€ kε sin(kx)) 0
x
0.45ν 0.45ν x
≈ (1− 6kε sin(kx)) ∫ (1+ 5kε sin(kx &)) dx & = (1− 6kε sin(kx))[ x & − 5ε cos(kx &)] 0
U 0 U
€ 0.45ν 0.45νx
= (1− 6kε sin(kx))[ x − 5ε cos(kx) + 5ε] ≈ (1− 6kε sin(kx) + 5(ε x)(1− cos(kx))
U U
'0.45νx *1 2
€ Thus, θ ≈ ) (1− 6kε sin(kx) + 5(ε x)(1− cos(kx)), .
( U +
€ c) First compute the correlation parameter:
θ 2 dU 0.45x
λ= = (1− 6kε sin(kx) + 5(ε x)(1− cos(kx))(Uk 2ε cos(kx)) .
€ ν dx U
This form can be made easier to look at if it is rearranged and the second order terms are
dropped:
λ ≈ 0.45(kx)(kε)cos(kx)(1− 6kε sin(kx) + 5(ε x)(1− cos(kx)) ≈ 0.45(kx)(kε)cos(kx) .

Then note that τ w = µ(U θ ) l( λ) and that l(λ) increases monotonically with increasing λ.
Therefore, the shear stress will be highest where λ is highest. Now switch to considerations of


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

wavelength. Given that cos(kx) < 0 for Λ/2 ≤ x ≤ 3Λ/4 and that cos(kx) > 0 for 3Λ/4 ≤ x ≤ Λ, τw
must be higher for 3λ/4 ≤ x ≤ λ.
d) For separation, set λsep = −0.090 ≈ 0.45(kx)(kε)cos(kx) . This can be solved numerically for kε
= 0.01 to find kx ≈ 21.6, so separation is expected to occur when x is large enough.
e) Use the steady flow Bernoulli equation with the reference condition taken far above the wavy
surface:
1 €2 1
2 2
( 2
)
2
ρU + P∞ = ρ U 2 [1+ kε exp(−ky)sin(kx)] + U 2 [ kε exp(−ky)cos(kx)] + P(x) .
Evaluate on y = 0 and keep only the linear terms in kε to find:
1
P(x) − P∞ = ρU 2 (−2kε sin(kx) + ...) .
€ 2
Thus, the minimum pressure occurs where sin(kx) = 1: kx = π/2 or x = Λ/4 (the wave crest), and
the maximum pressure occurs where sin(kx) = –1: kx = 3π/2 or x = 3Λ/4 (the wave trough).
f) Based on the results
€ provided above, a steady wind will increase the wave amplitude by at
least two mechanisms. The boundary layer shear tends to preferentially pull surface water away
from wave troughs and toward wave crests (a shear stress effect), and the wind-induced pressure
tends to suck wave crests up and push wave troughs down.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.21. Consider the boundary layer that develops in stagnation point flow:
U e (x) = U o x L .
a) With θ = 0 at x = 0, use Thwaites method to determine δ*(x), θ(x), and cf(x).
b) This flow also has an exact similarity solution of the full Navier Stokes equations. Numerical
€ evaluation of the final nonlinear ordinary differential equation produces: c f Rex = 2.4652 ,
where Re x = U e x ν = U o x 2 Lν . Assess the accuracy of the predictions for cf(x) from the
Thwaites method for this flow.

€ & 6
€ Solution 10.21. a) The Thwaites equation is: θ 2 = 0.45ν
x
Uo )
6
U e (x) 0
5
e
2
∫ U (ξ )dξ + θo (' U (x) +* . Here, θ = 0
e
at x = 0 and U e (x) = U o x L , so only the integral term contributes:
2 0.45ν L6 x 5 ξ 5 0.45ν L νL
6 ∫
θ = 6
U o 5 dξ = , or θ = 0.2739 .
U o x 0€ L Uo 6 Uo
€ θ 2 dU e θ 2 U o 0.45
Now compute the correlation parameter λ = = = = 0.075
ν dx ν L 6
The wall shear
€ stress is obtained from: €
µU e µU o x L & U o3 )1 2
τw = l( λ) = (0.3267) = 1.193µx( 3 + ,
θ € 0.2739 νL U o 'νL *
where the numerical value for l(λ) has been obtained by interpolation in the tabulation of laminar
boundary layer functions.
12
€ τw 2(1.193)µx % U o3 ( νL
Thus, the skin friction coefficient is: c f = 1 = 2 2 2 ' 3*
= 2.386 .
2
ρU e ρU o x L &νL ) Uo x 2
2
νL νL
Similarly, δ * = θH( λ) = 0.2739 (2.36) = 0.6464 where again the value of H(λ) has been
Uo Uo
obtained by interpolation. €
b) If c f Rex = 2.4652 is the exact answer, then the Thwaites calculation is acceptably accurate
(3%
€ low) for most engineering purposes.


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.22. A laminar boundary layer develops on a large smooth flat surface under the
influence of an exterior flow velocity U(x) that varies with downstream distance, x.
a) Using Thwaites method, find a single integral-differential equation for U(x) if the boundary
layer is to remain perpetually right on the verge of separation so that the wall shear stress, τw, is
zero. Assume that the boundary layer has zero thickness at x = 0.
γ
b) Assume U(x) = U o ( x L) and use the result of part a) to find γ.
c) Compute the boundary layer momentum thickness θ(x) for this situation.
d) Determine the extent to which the results of parts b) and c) satisfy the von Karman boundary
layer integral equation, (10.43), when τw = 0 by computing the residual of this equation.

Interpret the meaning of your answer; is Von Karman’s equation well satisfied, or is the residual
of sufficient size to be problematic?
e) Can the U(x) determined for part b) be produced in a duct with cross sectional area
−γ
A(x) = Ao ( x L) ? Explain your reasoning.

Solution 10.22. a) When τw is zero, then λsep = λ ≡ (θ 2 ν ) dU dx . Now plug this into the
Thwaites integral:
0.45ν x 5 νλsep 0.45 dU x 5
θ 2 (x) = 6 ∫ U ( x $)dx $ = , or λ sep = 6 ∫ 0U ( x #)dx # .
U (x) 0 € dU dx U (x) dx
γ −1 5γ +1
γ 0.45 γU o $ x ' U o5 L $ x ' 0.45γ
b) Plugging in U(x) = U o ( x L) produces: λsep = 6 6γ & ) & ) = ,
U o ( x L) L % L ( 5γ + 1 % L ( 5γ + 1
and€for λsep = –0.090, this equation produces: € γ = –0.10. [Note that this value is with in 0.01 of
the exact Falkner-Skan exponent value, –0.0904 for a laminar boundary layer that is perpetually
on the€verge of separation.]

c) Evaluate the Thwaites integral with θo = 0 using γ from part b):
2 0.45ν x 5 0.45ν U o5 L ' x *
−0.5+1
0.90νL ' x *
1.1
$ 0.90νL '1 2 $ x ' 0.55
θ = 6
U (x) 0
∫ U ( x $)dx $ = U 6 x L −0.6 −0.5 + 1)( L ,+ = U )( L ,+ , θ (x) = &% U )( &% L )( .
o( ) o o

τw d 2 dU
d) Use = (U θ ) +Uδ * but set τw to zero and recognize that δ * = 3.55θ from the
ρ dx dx
€ Thwaites shape-factor tabulation. Evaluate the two right side€terms of the equation:
# −0.2 12 0.55 & −0.65 12
d d % 2 # x & # 0.90νL & # x & ( U o2 # x & # 0.90€νL &
dx
(U θ ) = dx %U o %$ L (' %$ U (' %$ L (' ( = 0.35 L %$ L (' %$ U ('
2

$ o ' o
1 2 12
* dU # x &−0.1 # 0.90νL & # x & 0.55 # Uo # x & &
−1.1 −0.65
U o2 # x & # 0.90νL &
Uδ = U o % ( 3.55% ( % ( %% −0.1 % ( (( = −0.355 % ( % (
dx $ L' $ Uo ' $ L ' $ L $ L' ' L $ L ' $ Uo '


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

−0.65 12
U 2 " x % " 0.90νL %
Thus, we see that there is a residual = 0.005 o $ ' $ ' which is only 1.4% of either
L # L & # Uo &
of the terms. This residual is acceptable given the approximate nature of Thwaites method.
−γ
e) For an ideal two-dimensional duct with A(x) = Ao ( x L) , the upper and lower walls would
+0.1
only diverge in proportion to € ( x L) . However, the boundary layer displacement thickness
+0.55
grows like ( x L) . Thus, adjustments would have to be made in the duct geometry to account
for boundary layer growth. Such € adjustments would involve greater divergence of the upper and
lower walls, but increased€ wall divergence is likely to produce boundary layer separation. Thus,
−0.10
the part-b) flow speed U(x) = U o ( x L) cannot be produced in a duct with cross sectional area
€ −0.10
A(x) = Ao ( x L) when viscous laminar boundary layers are present on the walls of the duct.



Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.23. Consider the boundary layer that develops on a cylinder of radius a in a cross
flow
a) Using Thwaites method, determine the momentum thickness as a function of ϕ, the angle from
the upstream stagnation point (see drawing).
b) Make a sketch of cf versus ϕ.
c) At what angle does Thwaites method predict vanishing wall shear stress?

Solution 10.23. The potential for 2D ideal flow around a cylinder is: φ (r,θ ) = Ur cosθ (1+ a 2 r 2 ) .
For the specified geometry, θ = π − ϕ and cosθ = −cos ϕ so φ (r,ϕ ) = −Ur cos ϕ (1+ a 2 r 2 ) . Thus,
the exterior velocity, Ue, outside the boundary layer that develops on the cylinder’s surface, is
%1 ∂φ ( €
given by: U e = ' * = [2U sin ϕ ] x= aϕ = 2U sin( x a) , so the Thwaites integral becomes:
& r ∂ϕ
€ )r= a,x= aϕ € €
x ϕ
2 0.45ν 0.45aν
6 ∫ 32U sin (ξ a) dξ =
5 5
θ = 6
2U sin ϕ 0
∫ sin 5 (ϕ ')dϕ '.
(2U sinϕ ) 0
€ initial condition term has been ignored since Ue(0) = 0. The integral of sin5(ϕ) may be
The
looked up or performed via sin5(ϕ) = (1-cos2ϕ)2sin(ϕ) and an appropriate integration variable
switch. The result is:

2 0.45aν & 8 2 3 1 5 ) 0.225a 2 & 8 2 3 1 5 )
θ = 6 ( − cos ϕ + cos ϕ − cos ϕ + = 6 ( − cos ϕ + cos ϕ − cos ϕ + ,
2U sin ϕ '15 3 5 * Resin ϕ '15 3 5 *
where Re = Ua/ν (U is the velocity used in all the non-dimensional parameters in this problem).
dU e 2U 0.03cosϕ
b) Here
dx
=
a
cos ϕ , so for λ, we get: λ = 6
sin ϕ
( )
8 −15cosϕ + 10cos 3 ϕ − 3cos5 ϕ . Now,

evaluate the skin friction coefficient:
−1 2
µU e 1 νU e l( λ)sin 4 ϕ ( 8 2 3 1 5 +
cf = l( λ) 1 = 2l( λ ) 2 = 8.43 * − cosϕ + cos ϕ − cos ϕ - .
€ θ 2
ρU 2 €U θ Re ) 15 3 5 ,
Before plotting make a table of cf values.
ϕ λ S(λ) c f Re
€ ––––––––––––––––––––––––––––––––––––––––––––
1° 0.075 0.33 0.11
15° 0.074 0.32 1.7
30° 0.072 0.32 € 3.1
60° 0.59 0.31 4.4
90° 0.0 0.22 2.5
100° -0.060 0.11 1.0
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

103° -0.089 ~0.0 ~0.0

5
c f Re
4.5

3.5

2.5

1.5

0.5

0
0 20 40 60 80 100 120
angle (degrees)

The maximum surface shear stress is felt at ϕ ≈ 60°.


c) The angle where Thwaites methog predicts a vanishing shear stress is ϕ ≈ 103°.
Unfortunately, this value is not terribly useful as a prediction of the actual separation point (ϕs ≈
80°-85°) because boundary layer separation causes significant changes to the outer flow around
the cylinder that invalidates a portion of this computation. For this reason, reliable prediction of
flows with separation remains a difficult task.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.24. An ideal flow model predicts the following surface velocity for the suction (i.e.
the upper) side of a thin airfoil with chord c placed in a uniform horizontal air stream of speed
Uo :
15
Ue (x) = 2Uo [ x c ] exp (−x c) .
a) Assuming that x is the coordinate along the foil's suction surface, use Thwaites method to
estimate the momentum thickness θ(x) of the laminar boundary layer that develops on this
surface.
b) Using the results of part a) show that the correlation parameter λ is given by:
0.45 " 5x c 5x %" 5x %
λ= 2$
e −1− '$1− '
125 ( x c ) # c &# c&
c) Does Thwaites method predict boundary layer separation in the range 1/5 < x/c < 1?
d) If a laminar boundary layer is predicted to separate from the surface of this airfoil, suggest at
least two changes that could be made to the foil that would tend to prevent separation.
1.4"
1.2"
1"
Ue (x) 0.8"
Uo 0.6"
Uo!
!
!
0.4"
0.2"
!
!
c!
0"
0" 0.2" 0.4" 0.6" 0.8" 1"
x/c!

Solution 10.24. a) The boundary layer is launched from a stagnation point, so the basic Thwaites
formula is:
x
0.45ν x 5 2 0.45ν
θ2 = 6 ∫ U e dx , or in this case: θ = 6 6 6 5 −6 x c ∫
2 5Uo5 ( x" c) e−5 x" c dx" .
Ue 0 2 Uo ( x c) e 0

Simplify, change to an integration variable ξ = x´/c, and use the hint:


x c
0.45ν ce−6 x c # (1+ 5ξ ) &
x c
2 0.45ν ce−6 x c
65 ∫
θ = ξ exp (−5ξ ) d ξ = 65 %
− exp (−5ξ )( .
2Uo ( x c) 0 2Uo ( x c) $ 25 '0
0.45ν ce−6 x c " (1+ 5 x c) −5x c 1 % 0.45ν ce x c " 5x c
Evaluate: θ2 = 65 $
− e + '= #e − (1+ 5x c)%& .
2Uo ( x c) # 25 25 & 50Uo ( x c)6 5
b) The correlation parameter is λ = (θ2/ν)(dUe/dx), and the missing ingredient from part a) is:
dUe (x) "1 −4 5 1 5 1 % −x c
= 2Uo $ [ x c ] − [ x c ] 'e .
dx # 5c c&
θ 2 dUe (x) 0.45ce x c " 5x c &1 −4 5 1 5 1 ) −x c
So, λ= = 65#
e − (1+ 5x c )$% 2Uo ( [ x c ] − [ x c ] + e , or
ν dx 50Uo ( x c ) ' 5c c*
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

0.45 "#e 5x c − (1+ 5x c)$%(1− 5 ( x c)) which is the desired form.


λ= 2
125 ( x c)
c) From part b), λ = 0 at x/c = 1/5, and λ = (0.45/125)(e5 – 6)(–4) = –2.05 at x/c = 1.
d) Yes, the value of λ at x/c = 1 is less than λsep = –0.090. Thus, the Thwaites method predicts
that a laminar boundary layer will separate downstream of x/c = 1/5 and upstream of x/c = 1.
Four possible changes to the foil that might delay or prevent boundary layer separation
are as follows. (i) The surface could be roughened or a boundary layer trip could be added to
stimulate transition to a turbulent boundary layer. (ii) The thickness of the foil could be reduced
so that the pressure gradient predicted by ideal flow for its upper surface is less adverse aft of x/c
= 1/5. (iii) Compressed air could be jetted downstream along the foil's surface to speed up the
slow moving portion of the boundary layer thereby keeping it attached. (iv) Suction could be
applied to the aft portion of the foil to remove the lower (slow moving) portion of the boundary
layer thereby keeping it attached.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.25. An incompressible viscous fluid flows steadily in a large duct with constant
cross sectional area Ao and interior perimeter b. A laminar boundary layer develops on the duct’s
sidewalls. At x = 0, the fluid velocity in the duct is uniform and equal to Uo, and the boundary
layer thickness is zero. Assume the thickness of the duct-wall boundary layer is small compared
to Ao/b.
a) Calculate the duct-wall boundary layer momentum and displacement thicknesses, θ(x) and
δ*(x) respectively, from Thwaites’ method when U(x) = Uo.
b) Using the δ*(x) found for part a), compute a more accurate version of U(x) that includes
boundary layer displacement effects.
c) Using the U(x) found for part b), recompute θ(x) and compare to the results of part a). To
n
simplify your work, linearize all the power-law expressions, i.e. (1− bδ * Ao ) ≅ 1− nbδ * Ao .
d) If the duct area expanded as the flow moved downstream, would the correction for the
presence of the sidewall boundary layers be more likely to move boundary layer separation
upstream or downstream? Explain.

Solution 10.25. a) The Thwaites’ results with a constant velocity and zero momentum thickness
at x = 0 are readily obtained:
2 0.45ν x 5 0.45νx
θ = 6
Uo 0
∫ U o dx =
Uo
, or θ = 0.671 νx U o .

Here, λ is zero because dU/dx = 0 so the shape factor is 2.61 which means that:
δ * = 2.61⋅ θ = 1.75 νx U o .
b) At any downstream € displacement effect of the boundary layer will
distance x in the duct, the

decrease the effective flow area of the duct. Therefore a simple mass balance between the duct
entrance (x = 0) and the location x requires:
ρU o€Ao = ρU(x)( Ao − bδ * ) , or U(x) = U o (1− bδ * Ao )
−1

c) Use the revised U(x) found in part b) and Thwaites’ method to find:
x x
0.45ν −5 0.45ν
θ2 = −6 ∫ U 5
o (1− bδ *
Ao ) dx ≈
U
(1− 6bδ *
Ao ) ∫ (1+ 5bδ * Ao )dx
U o (1−€bδ Ao ) 0
6 *
€ o 0
x
0.45ν 0.45ν
=
Uo
( *

0
(
1− 6bδ Ao ) ∫ 1+ 5b(1.75 νx U o Ao dx = ) Uo
(1− 6bδ * Ao )( x + 10bδ * x 3Ao )

€ 0.45νx 0.45νx

Uo
(1− 6bδ * Ao + 10bδ * 3Ao ) =
Uo
(1− 8bδ * 3Ao )

€ where linearization has been used to make the integration and algebra easier. Here, we see that
the revised value of θ(x) will be smaller than that estimated in part a).
d) The displacement effect of the boundary layer acts to increase the flow speed. This means
€ that dU/dx will be larger when the boundary layer’s displacement effect is accounted for, thus the
correlation parameter λ will be larger and further from its “separation” value of –0.090.
Therefore, the predicted location of separation will be further downstream in an expanding duct
when the boundary layer displacement effect is accounted for. In the present situation, λ is zero
without accounting for the boundary layer displacement effect but λ becomes positive because
dU/dx is nonzero and positive. Thus, when the boundary layer displacement effect is included:
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

dU(x) d −1 d U ob dδ * U ob *
= U o (1− bδ Ao ) ≈ U o (1+ bδ Ao ) =
* *
= δ ,
dx dx dx Ao dx 2Ao x
and
θ 2 (x) dU(x) 0.45x dU(x) 0.45 bδ *
λ≡
ν dx

Uo
(1− 8b δ *
3Ao)
dx

2
(1− 8b δ *
3Ao)
A0
>0 !


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.26. Water flows over a flat plate 30 m long and 17 m wide with a free-stream
velocity of 1 m/s. Verify that the Reynolds number at the end of the plate is larger than the
critical value for transition to turbulence. Using the drag coefficient in Figure 10.12, estimate the
drag on the plate.

Solution 10.26. ReL = UL/ν = (1.0m/s)(30m)/(1.0x106 m2/s) = 3.0x105 > Recr ~ 106. Thus, the
flow is expected to be turbulent over most of the plate. From Figure 10.12, CD ≈ 0.003 so that:
1
D = ρU 2 (Area)CD = 0.5(10 3 )(1.0) 2 (17)(30)(0.003) = 765N .
2


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.27. A common means of assessing boundary layer separation is to observe the
surface streaks left by oil or paint drops that were smeared across a surface by the flow. Such
investigations can be carried out in an elementary manner for cross-flow past a cylinder using a
blow dryer, a cylinder 0.5 to 1 cm in diameter that is ~10 cm long (a common ball-point pen),
and a suitable viscous liquid. Here, creamy salad dressing, shampoo, dish washing liquid, or
molasses should work. And, for the best observations, the liquid should not be clear and the
cylinder & liquid should be different colors. Dip your finger into the viscous liquid and wipe it
over two thirds of the surface of the cylinder. The liquid layer should be thick enough so that you
can easily tell where it is thick or thin. Use the dry one third of the cylinder to hold the cylinder
horizontal. Now, turn on the blow dryer leaving the heat off and direct its outflow across the
wetted portion of the horizontal cylinder to mimic the flow situation in the drawing for Exercise
10.23.
a) Hold the cylinder stationary, and observe how the viscous fluid moves on the surface of the
cylinder and try to determine the angle ϕs at which boundary layer separation occurs. To get
good consistent results you may have to experiment with different liquids, different initial liquid
thicknesses, different blow-dryer fan settings, and different distances between cylinder and blow
dryer. Estimate the cylinder-diameter-based Reynolds number of the flow you’ve studied.
b) If you have completed Exercise 10.23, do your boundary layer separation observations match
the calculations? Explain any discrepancies between your experiments and the calculations.

Solution 10.27. a) Using a dark shampoo on a white ball-point pen, the 3rd author of this
textbook found that the separation point occurred on the upstream side of the cylinder near ϕ =
90°. The cylinder diameter was 8 mm and flow speed was probably about 10 m/s. Thus, an
estimate of the Reynolds number is:
UD (10m / s)(0.008m)
ReD = ≈ ≈ 5,000
ν 1.5 ×10−5 m 2 / s
b) The blow-drier-and-ball-point-pen separation point results, ϕs near to but less than 90°, match
commonly quoted experimental values for the separation angle (ϕ ≈ 80°-85°). However, they do
not match the Thwaites-method-calculated
€ location of zero shear stress (ϕ ≈ 103°). The primary
difference between the experiments and the Thwaites calculation is the separated flow in the
wake of the cylinder. The experiment includes the separated flow while the potential flow
surface velocity that is input to the Thwaites calculation does not include the separated flow. In
this case the cylinder’s wake makes important changes to the surface flow on the cylinder so the
experimental and calculated boundary layers develop under different flow fields and therefore
separate at different points. In general, Thwaites method is only successful in predicting whether
or not boundary layer separation will occur. Once boundary layer separation has occurred, a
theory that accounts for flow in the separation zone is needed.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.28. Find the diameter of a parachute required to provide a fall velocity no larger
than that caused by jumping from a 2.5 m height, if the total load is 80 kg. Assume that the
properties of air are ρ = 1.167 kg/m3, ν = 1.5 × 10-5 m2/s, and treat the parachute as a
hemispherical shell with CD = 2.3. [Answer: 3.9 m]

Solution 10.28. The fall velocity from a 2.5 m height is [2gh]1/2 = [2(9.81)(2.5)]1/2 = 7.0 m/s.
At steady state, the drag on the parachute equals the load, so that D = mg = (80)(9.81).
Thus, using CD = 2.3, the parachute's cross sectional area should be
D 80(9.81)
Area = 1 2
= = 11.93m 2 , or a diameter = [4(11.93)/π]1/2 = 3.90 m.
CD 2 ρU 2.3(0.5)(1.167)(7) 2


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.29. The boundary layer approximation is sometimes applied to flows that do not
have a bounding surface. Here the approximation is based on two conditions: downstream fluid
motion dominates over the cross-stream flow, and any moving layer thickness defined in the
transverse direction evolves slowly in the downstream direction. Consider a laminar jet of
momentum flux J that emerges from a small orifice into a large pool of stationary viscous fluid
at z = 0. Assume the jet is directed along the positive z-axis in a cylindrical coordinate system.
In this case, the steady, incompressible, axisymmetric boundary layer equations are:
1 ∂ (RuR ) ∂w ∂w ∂w 1 ∂p ν ∂ & ∂w )
+ = 0 , and w + uR =− + ( R +,
R ∂R ∂z ∂z ∂R ρ ∂z R ∂ R ' ∂R *
where w is the (axial) z-direction velocity component, and R is the radial coordinate. Let r(z)
denote the generic radius of the cone of jet flow.
a) Let w(R,z) = (ν z) f (η) where η = R z , and derive the following equation for f:
€ η €
ηf # + f ∫ ηfdη = 0 .
η


b) Solve this equation by defining a new function F =

∫ ηfdη . Determine constants from the
R = r(z )


boundary condition: w → 0 as η → ∞, and the requirement: J = 2πρ ∫ w 2 (R,z)RdR = constant.
R= 0
c) At fixed z, does r(z) increase or decrease € with increasing J?
[Hints: i) the fact that the jet emerges into a pool of quiescent fluid should provide information
about ∂p/∂z, and ii) f (η) ∝ (1+ const ⋅ η 2 )−2 , but try to obtain this result without using it.]

Solution 10.29. a) Start with the given boundary layer equations and use w(R,z) = (ν z) f (η)
where η =€R z , ∂ ∂R = (1 z) ∂ ∂η and ∂ ∂z = (−η z) ∂ ∂η in the continuity equation to find:
1 R ∂w 1 R ( νη ν + ν η
uR = − ∫ R dR = − ∫ R* − 2 f ' − 2 f - dR = ∫ η 2 f ' + ηf dη
R 0 ∂z R0 ) z z , €R 0
( )
€ Now start
€ assembling terms of
€ the momentum equation:
∂w ν ' ν νη * ν2 2 ∂w ' ν η 2 *ν
w (
= f ⋅ )− 2 f − 2 f , = − 3 f + ηff , uR ) - ( )
= ) ∫ η f % + ηf dη, 2 f % , and
€ ∂z z ( z z + z ∂R (ηz 0 +z
2
ν ∂ $ ∂w ' ν ∂ $ ν ' ν $ ν ν ' ν
&R ) = & R 2 f *) = & 2 f * + R 3 f **) = 3 ( f * + ηf **) .
R ∂ R % ∂ R ( R ∂ R % z ( ηz % z z ( ηz
€ If the fluid far from the jet is still, then ∂p/∂z€ = 0 in the boundary layer approximation because
∂p/∂R = 0 too. Reassemble the momentum equation:
ν2 'ν η *ν ν2
€ − 3 ( f 2 + ηff %) + ) ∫ (η 2 f % + ηf ) dη, 2 f % = 3 ( f % + ηf %%)
z (ηz 0 +z ηz
Divide out the common factor of ν 2 z 3 :
&1 η 2 ) 1
−( f + ηff $) + ( ∫ (η f $ + ηf ) dη+ f $ = ( f $ + ηf $$)
2

€ 'η 0 * η
This is a single equation€for f without z or R appearing; thus the assumption of a similarity
solution is successful. Multiply both sides by η, expand terms and integrate by parts.
& 2 η ) η
€ 2 2
−ηf − η ff $ + f $(η f − 2 ∫ ηfdη+ + f $ ∫ ηfdη = f $ + ηf $$
' 0 * 0


Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

η
2
Combine common terms to find: −ηf − f $ ∫ ηfdη = f $ + ηf $$. This compact form can be further
0

d $ '
η η
d
& f ∫ ηfdη) = f * ∫ ηfdη + ηf , and
2
simplified by noting that: (ηf #) = ηf ## + f # . Thus, one
dη % 0 ( 0 dη
€ η
integration with respect to η yields: ηf # + f ∫ ηfdη = 0 . Here, the constant must be zero based
0
on an evaluation€of the equation at η = 0. €
η

ηfdη , then F " = ηf , and F "" = ηf " + f , so the above equation becomes:
b) If F =
1 1 €
F "" − F " + F "F = 0 , or η 2 F ## + η(F −1) F # = 0 . This last equation is equidimensional so the
η η
2
€ € t € d d 2 d d2 d
substitution η = e will lead to simplification. Here, η = , and η = − , so the
dη dt dη 2 dt 2 dt
€ 2
€ d F dF
equation becomes: 2
+ (F − 2) = 0 . This equation is autonomous; thus we can let
dt dt
€ d 2 F dy dF
dF € the last remnant € of the momentum equation
= y(F) so that = = yy " , and
dt dt 2 dF dt
becomes:€yy " − (F − 2)y = 0 . Therefore when y ≠ 0, y " = −F + 2 , or y = A + 2F − F 2 /2 . Now go
back to the definition of F to determine the constant A:
2
dF dF 1 η


€ y(F) =
dt


= A + 2F − F 2 /2 = A + 2 ∫ ηfdη +
€ €
η
[
2
∫ ]
ηfdη .
2
When η = 0, then F = 0, so A = 0. Thus, y(F) = 2F − F 2 = η dF dη . Separate and integrate
dη dF dF dF
using a partial fraction decomposition: = 2 =− + to reach:
€ 2η F − 4F 4(F − 4) 4F
1 $ F ' € 12 $ F '1 4 4C 4η 2
ln η1 2 + B = ln& ) . Now exponentiate: C η = & ) and solve for F: F = .
4 %4 −F( %4 − F( (1+ C 4η 2 )
1€ 8C 4
Recall that F " = ηf , so f = F # = 4
4 2 2 . The constant C can be evaluated from:
η (1+ C η )
R = r(z) 2
€ J R = r(z)
2 €ν 2 2
r(z )/ z
2 2€
r(z)/ z
64C 8η
= ∫ w (R,z)RdR = ∫ 2 f (η)RdR = ν ∫ ηf dη = ν ∫ 4 2 4

2πρ R = 0 R= 0 z 0 0 (1+ C η )
€ integral may be performed after a change of variable γ = 1+ C 4η 2 , and by noting that the
The €
upper limit corresponds to η → ∞. Thus:
∞ 4
€ J = 32C 4ν 2 ∫ γ −4 dγ = 32C ν 2 , so C 4 = 3J 2 = 3 Re 2 , where Re = J ρ
2πρ 0 3 64 πν€ ρ 64 π ν
2
Therefore, the final solution is: w(R,z) =
ν ( 3 8π )Re
z (1+ ( 3 64 π ) Re 2 (R /z) 2 ) 2
€ c) From the result of part b),€r(z) will decrease with increasing € J (increasing Re). The laminar jet
narrows as its Reynolds number increases.

Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.30. A simple realization of a temporal boundary layer involves the spinning fluid in
a cylindrical container. Consider a viscous incompressible fluid (density = ρ, viscosity = µ) in
solid body rotation (rotational speed = Ω) in a cylindrical container of diameter d. The mean
depth of the fluid is h. An external stirring mechanism forces the fluid to maintain solid body
rotation. At t = 0, the external stirring ceases. Denote the time for the fluid to spin-down (e.g. to
stop rotating) by τ.
a) Case I: h >> d. Write a simple laminar-flow scaling law for τ assuming that the velocity
perturbation produced by the no-slip condition on the container’s sidewall must travel inward a
distance d/2 via diffusion.
b) Case II: h << d. Write a simple laminar-flow scaling law for τ assuming that the velocity
perturbation produced by the no-slip condition on the container's bottom must travel upward a
distance h via diffusion.
c) Using partially-filled cylindrical containers of several different sizes (drinking glasses and
pots & pans are suggested) with different amounts of water, test the validity of the above
diffusion estimates. Use a spoon or a whirling motion of the container to bring the water into
something approaching solid body rotation. You'll know when you're close to solid body rotation
because the fluid surface will be a paraboloid of revolution. Once you have this initial flow
condition set-up, cease the stirring or whirling and note how long it takes for the fluid to stop
moving. Perform at least one test when d & h are several inches or more. Cookie or bread
crumbs sprinkled on the water surface will help visualize surface motion. The judicious addition
of a few drops of milk after the fluid starts slowing down may prove interesting.
d) Compute numbers from your scaling laws for parts a) and
b) using the viscosity of water, the dimensions of the
containers, and the experimental water depths. Are the scaling
laws from parts a) and b) useful for predicting the
experimental results? If not, explain why.
(The phenomena investigated here have some important
practical consequences in atmospheric and oceanic flows and
in IC engines where swirl and tumble are exploited to mix the
fuel charge and increase combustion speeds.)

Solution 10.30. For all simple unsteady diffusion problems, the length scale of “diffusion-
penetration” is proportional to the square root of the product of the diffusion constant and time.
For momentum diffusion in fluid flows, ν is the diffusion constant. In the following, let τ be the
time it takes for the swirling fluid to come to rest.
a) If h >> d, it should be possible to ignore the bottom of the container. Therefore, the "stopping"
signal must penetrate a distance of d/2 inward from the walls of the container. So, for an
approximate unsteady diffusion analysis, we have: d/2 ≈ ντ , or τ ≈ d2/4ν .
b) If d >> h, it should be possible to ignore the walls of the container. Therefore, the "stopping"
signal from the no-slip boundary condition on the bottom of the container must penetrate upward
a distance of h from the bottom of the container. So, for an approximate unsteady diffusion

analysis, we have: h ≈ ντ , or τ ≈ h2/ν
c) For all standard-kitchen-size containers with water depths of a several cm, the experimental
stopping time was less than a minute.

Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

d) Using the formulae from part a) or b), a length scale of 5 cm (~two inches), and the kinematic
(0.05m) 2
viscosity of water: τ ≈ = 2500 seconds ≈ 40 minutes!
1.0 ×10−6 m 2 /sec
The expectation here is that the estimated and actual times would match within a factor of two or
so. Clearly something is wrong with the simple diffusion analysis for these simple swirling
flows. The milk drops in the swirling water show that the flow is very three-dimensional and
€ of the flow with the bottom and sides of the pan or cup is very important. In
that the interaction
particular, the viscous boundary layer on the bottom of the container causes inward radial flow
along the bottom and the fluid stops rotating because of a phenomenon called Ekman pumping.
Fluid Mechanics, 6th Ed. Kundu, Cohen, and Dowling

Exercise 10.31. Mississippi River boatmen know that when rounding a bend in the river, they
must stay close to the outer bank or else they will run aground. Explain in fluid mechanical terms
the reason for the cross-sectional shape of the river at the bend.

Solution 10.31. The Reynolds number based on mean channel width or radius of curvature is
very large. Thus, the riverbed boundary layer is thin compared with the width or depth of the
river, and most of the flow is inviscid. The primary turning flow around the bend is governed by
∂p ∂r = ρvθ2 r , but v decreases rapidly to zero in the thin boundary layer on the riverbed. Thus,
θ

∂p ∂r in the boundary layer, which is largely the same as that in the inviscid flow, is not
balanced by ρvθ2 r in the boundary layer since [v ]boundary layer < [v ]inviscid flow. Here, ∂p ∂r is too
θ θ

large in the boundary layer flow for it to remain parallel to the flow above it. The excess

pressure gradient pushes the boundary layer flow toward the lower pressure which occurs at the

riverbank on the inside of the river bend. This induces cross-flow in the boundary layer that
€ boundary layer fluid toward the inner riverbank, and causes a downward
moves € secondary flow
at the outer riverbank. Together these effects set up the secondary flow pattern shown below. In
the Mississippi River, the water transports silt which is scoured from the outer bank and
deposited on the inner bank. This transport accounts for the cross section of the river channel.
This secondary flow phenomenon also explains why small river meanderings tend to grow over
time as the river erodes its outer bank and deposits silt on its inner bank. The "goose necks" of
the San Juan River is southeastern Utah provide an extreme (and impressive) example of river
meandering.

r
!"#!$

boundary
layer

Vous aimerez peut-être aussi