Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

CHAPTER FOUR

4. BOUNDARY LAYER THEORY

4.1 INTRODUCTION
A fluid medium through which bodies such as aircraft or ships move exerts a resistance to motion on
the bodies. This resistance is called drag. The drag of a body results from two types of forces acting
on the body: shear forces and pressure forces. The drag, or surface resistance, from shear forces is
often called skin-friction drag and it is the only drag present for flow past a flat plate. In a pipe, this
is the only form of drag, and it results in pressure and energy losses along the length of the pipe. The
drag resulting from pressure forces is called form drag, which occurs when fluid flows by a bluff or
blunt shape such as sphere or cylinder. Bluff bodies are bodies that are tall in comparison to the
length in the direction of flow. Here, emphasis is placed on the mechanics of skin-friction drag.

Skin-friction drag is due to the viscous shearing that takes place between the surface and the layer of
fluid immediately above it. This occurs on surfaces of objects that are long in the direction of flow
compared to their height. Such bodies are called streamlined. A thin flat plate is an example of a
streamlined object. When a fluid flows over a solid surface, the layer next to the surface may become
attached to it (it wets the surface). This is called the “no slip condition”. The layers of fluid above
the surface are moving so there must be shearing taking place between the layers of the fluid. The
shear stress acting between the wall and the first moving layer next to it is called the wall shear
stress,  w . The result is that the velocity of the fluid increases with height y, for a typical boundary
layer, as shown in Figure 4.1.

Figure 4.1. Variation of velocity of fluid with height for a typical boundary layer.

In viscous fluid flow, the fluid particles in contact with the wall are brought to rest due to the drag
action of the wall. The fluid velocity increases from this zero value until it attains the value in the
main stream. The effect of the wall is to set up a shear force in the direction normal to the direction
of flow. Hence, a velocity gradient is set up in this direction, which gradually decreases until it attains
zero value. The region in the flowing fluid contiguous to the wall in which velocity gradient exists
as a result of the shear stress at the surface is called boundary layer. The boundary layer thickness
grows with distance from the leading edge. At some distance from the leading edge, it reaches a
constant thickness. It is then called a fully-developed boundary layer. The point where the velocity
gradient becomes zero and the velocity attains the same value as in the main stream of the fluid is
called the upper limit of the boundary layer. The general area of study of the flow pattern in the
boundary layer, as well as of the associated shear stress at the boundary, is called boundary layer
theory.

4.1.1 Flow Pattern in a Boundary Layer

The flow in the boundary layer can be laminar or turbulent. The flow pattern associated with
boundary layer can be visualized by analysing quantitatively the interaction between a fluid and the

1
surface of a thin, flat plate as the fluid (e.g., air) passes by the plate. Figure 4.2 illustrates such a
plate. When the fluid stream is interrupted by the plate, fluid passes over the top and underneath the
plate, so two boundary layers can be depicted in Figure 4.2 (one above and one below the plate).
However, Figure 4.2 shows what happens on one side only.

Figure 4.2. Boundary layer for flow past a flat plate.

In Figure 4.2, the fluid has a uniform velocity, U 0 , before it reaches the vicinity of the plate and first
comes into contact with the plate at point A called the leading edge. However, the fluid touching the
plate has zero velocity (the velocity of the plate), because of the no-slip condition characterizing
continuum flows. At this point A, the laminar boundary layer thickness,  l , is zero. Since a velocity
gradient exists between the fluid in thx e free stream and the fluid next to the plate, there is a shear
stress at the plate surface. As the fluid particles next to the plate pass the leading edge of the plate, a
retarding force (from the shear stress) begins to act on them. As these particles progress downstream,
they continue to decelerate. Moreover, these particles (because of their lower velocity) retard other
particles adjacent to them but farther out from the plate. Thus, the boundary layer becomes thicker,
or grows in the downstream direction. Thickening of the laminar boundary layer continues smoothly
in the downstream direction until the thickness is so great that the flow becomes unstable. The
boundary layer becomes turbulent; the streams do not remain parallel and the Reynolds number is in
the range 2  105 to 3  106. The transition from laminar to turbulent flow is not sharp; it takes place
over some length BC of the plate. Point B is called the critical point while the thickness of the
boundary layer at point B is called the critical thickness,  C , and the distance from the leading edge
to the critical point is called the critical distance, xC . The critical distance depends on the shape of
the leading edge, the degree of roughness of the surface of the wall, the velocity and the properties of
the fluid. In the turbulent boundary layer, eddies mix higher-velocity fluid into the region close to
the wall so the velocity gradient, U y , at the wall becomes greater than it is at the wall in laminar
boundary layer just upstream of the transition point. The turbulent region of the boundary layer
consists of three layers, the viscous sublayer (laminar sublayer), the buffer layer and the turbulent
layer. The laminar sublayer is a thin region next to the wall in which the flow is laminar: the velocity
gradient within this layer is linear. It is separated from the turbulent layer by the buffer layer, which
is a region of transition from laminar flow to turbulent flow. Compared to the laminar sublayer and
the turbulent layer, the buffer layer is so thin that it is ignored in the boundary layer analysis.

In both regions of the boundary layer, the velocity increases from zero at the plate and attains the free
stream velocity or undisturbed velocity, U 0 , at the outer edge of the boundary layer. The outer edge
itself is not sharp; it is usually taken as the point where the velocity is 99% of the bulk velocity, U 0 .
The dashed line is drawn so that u = 0.99 U 0 .

4.1.2 Shear-Stress Distribution Along the Boundary

It is the shearing force of the plate that decelerates the fluid to produce the boundary layer. Therefore,
the shear stress must be relatively large near the leading edge of the plate where the velocity gradient
is steep and that it becomes smaller as the velocity gradient becomes smaller in the downstream
direction. However, where the boundary layer becomes turbulent, the shear stress at the boundary
2
again becomes larger, consistent with the greater velocity gradient next to the wall in the turbulent
boundary layer, as depicted in Figure 4.3.

Figure 4.3. Shear-stress distribution on either side of the plate.

4.2 Boundary Layer Separation and Formation of Wakes

The drag on the fluid increases with the thickness of the boundary layer. Therefore, any phenomenon
that can cause increase in the thickness of the boundary layer is of interest to the engineer. Boundary
layer separation is one of such phenomena.

In normal situations, the flow in the boundary layer is in the forward direction, but the flow can be
reversed if the fluid pressure increases in the direction of flow (adverse pressure gradient). When the
fluid in the boundary layer enters a region of adverse pressure gradient, its momentum is depleted by
the adverse pressure gradient and the shearing forces. It may thus run out of momentum and its
velocity may be reversed. In order to avoid the backward flowing fluid, the boundary layer will be
forced to shift into the main stream. The point where this phenomenon first occurs is called
separation point. A zone of decelerated fluid is present behind the plate and large eddies (vortices)
called wakes are created downstream of the separation point since back flow cannot take place beyond
that point. The eddies consume large amount of mechanical energy. This separation of boundary
layers occurs when the change in velocity of the fluid flowing by an object is too large in direction or
magnitude for the fluid to adhere to the surface. Since formation of a wake causes large losses in
mechanical energy, it is often necessary to minimize or prevent boundary layer separation by
streamlining the objects or by other means (as discussed in Section 4.3).

Laminar boundary layers are more susceptible to separation in adverse pressure gradients than
turbulent boundary layers because momentum diffusion to the boundary in turbulent flow is higher
than in laminar flow. Momentum diffusion supplements that depleted by adverse pressure gradient
and shear forces.

The velocity profiles in separated flow and the directions of flow are illustrated in Figure 4.4.

Figure 4.4. Boundary layer separation.

Before separation, flow is in the forward direction and momentum transfer is toward the wall or the
U
boundary from the fluid. Velocity therefore increases with distance from the wall, i.e.,  0,
y y = 0
and the velocity profile is as shown between A and B. The velocity gradient decreases with distance
from the leading edge until the separation point B where the velocity gradient becomes zero, i.e.,

3
U
= 0 hence there is no forward or backward flow in the vicinity of the wall. Beyond the
y y=0

separation point, there is a reverse flow, and the streamlines bend away from the wall causing a steep
rise in the boundary layer thickness. So, the velocity gradient near the wall becomes less than zero,
U
 0 . This shifting of the forward flow from the wall to give way to the reverse flow is
y y = 0
described as separation. Separation causes rapid increase of the boundary layer thickness and
thereby increases frictional drag on the fluid and fluid energy loss.

4.3 Boundary Layer Control


This involves procedures for producing desirable effect of preventing the boundary layer from
becoming turbulent and thus reducing the skin-friction drag.

One way to control the boundary layer, for example, on an air foil, is to shape the air foil in such a
way that the pressure distribution on its surface delays the onset of turbulence. But care must be taken
not to alter the air foil’s basic shape often thereby reducing its lift.

Another means of boundary layer control is to make the surface of the boundary such as an air foil
surface, somewhat porous and to apply a reduced pressure to the surface (reduced pressure or suction
is maintained inside the air foil) so that part of the boundary layer is drawn away through the porous
surface (part of the air of the boundary layer is sucked into the interior of the air foil). This keeps the
boundary layer thin, so it remains stable in its laminar state and a reduced skin-friction drag results.

Aircraft are not the only bodies that could benefit from boundary layer control, also, submerged
bodies such as submarines and torpedoes.

Another boundary layer control measure involves the use of a dilute polymer solution for drag
reduction in liquids. Another drag-reduction phenomenon associated with boundary layers is that of
the compliant surface. Studies reveal that porpoises have a special kind of skin (its outer surface is
very compliant) that delays the onset of turbulence in the boundary layer. Thus, they can swim faster
than any other animal of their size. Also, artificial surfaces have been developed to yield reduced
drag, but only limited success has been achieved.

4.4 Application of Momentum Equation to the Boundary Layer

It is the shearing force of the plate that decelerates the fluid to produce the boundary layer.
Consequently, there is momentum transfer in the direction transverse to the flow direction. Hence,
momentum equation can be applied to derive a relationship among all the interacting forces on the
wall.

Consider the control volume in Figure 4.5 in which the plate is arranged parallel to the x direction

and u is the component of the velocity vector U in that direction.

Figure 4.5. Forces on a control volume in the boundary layer.

4
H
Mass flow rate through side AB per unit width, m AB =  0
u dy (4.4.1)
 
u dy dx
H
Mass flow rate through side BC per unit width, m BC = 
x  0 
(4.4.2)
Mass flow rate through side DC per unit width, m  DC , is therefore given by:
  H
u dy dx
H
m DC = m AB + m BC =  u dy +
x  0
 (4.4.3)
0 
H
Rate of momentum transfer through side AB, FAB = 
0
u 2 dy (4.4.4)
Rate of momentum transfer through side BC is the product of mass through BC and the free stream
velocity, U o , i.e.,

FBC = U o   u dy dx
H
(4.4.5)
x  0 
Therefore, rate of momentum transfer through side DC is given by:
 
u 2 dy dx
H H
FDC = 
0
u 2 dy + 
x  
0 
(4.4.6)

 p  p
The net force in the x direction = pH −  p + x  H = − H dx and the shear force (retarding)
 x  x
= −  wdx . Since there is no acceleration, the sum of these forces is equal to the rate of change of
momentum of the system:
p   H   H
 u dy dx −  udy dx
H H
−H dx −  w dx = FDC − FAB − FBC =  u 2 dy +
x  0  u 2 dy − U 0
x  0
2

x 0  0 
p   H 2    H
−H dx −  w dx =   u dy dx − U 0   udy dx (4.4.7)
x x  0  x  0 
Since U 0 is not a function of x , Equation (4.4.7) reduces to:
 H p
w =  u (U o − u )dy − H (4.4.8)
x 0 x
Since there is no pressure variation along any transverse section and if the pressure drop per unit
length is small compared to the total pressure, we have:
w  H
u (U 0 − u )dy
x 0
= (4.4.9)

Beyond the upper edge of the boundary layer, there is no velocity variation in the transverse direction
so that u = U 0 , therefore H =  , hence, we have:
w  
u (U 0 − u )dy
x 0
= (4.4.10)

This is the momentum-integral equation of the boundary layer.

Since U 0 is assumed to be independent of x, we can write:


w d  u  u 
dx 0 U 0
= 1 −  dy (4.4.11)
U 02
 U 0 

5
If the velocity of the free stream is assumed to vary with x, then the momentum equation has to be
derived from the equation of motion over a flat plate. The result as presented by Schlichting (1968)
is:
w H  dU 0 H
= u (U 0 − u )dy + (U 0 − u )dy
dx 0
(4.4.12)
 0 x

Equation (4.4.12) can be written in terms of the displacement thickness,  D , and the momentum
thickness,  M . The former is the amount by which the boundary will be displaced in the transverse
direction in inviscid flow so as to have the same mass flow rate as in the case of viscous flow.
Alternatively, the displacement thickness is a measure of the outward displacement of the streamlines
from the solid surface as a result of the reduced velocities within the boundary layer. If H is the width
H
of flow when the fluid is viscous, the corresponding mass flow rate is m v =  udy .
0

When the fluid is non-viscous, the area available for flow has to be reduced by an amount  D so as
to make the flow rate of the inviscid fluid equal to that of the viscous fluid. The mass flow rate in the
inviscid fluid is given by:
H H D
m i =  U 0 dy =  U 0 dy −  U 0 dy (4.4.13)
D 0 0

Since  is constant for incompressible flow and U 0 is also constant, the second term on the right
side is U 0 D . Equating m v and m i , we have:

  (U − u )dy
H
 DU 0 = 0 (4.4.14)
0

H  u 
D = 0
1 −

 dy
U 0 
(4.4.15)

Also, for momentum thickness, which is a measure of the momentum loss within the boundary layer
as a result of the reduced velocities within the boundary layer, we have:
u (U 0 − u )dy
H
 M U 02 =  0
(4.4.16)
H u  u 
M =  0 U0
 1 −

 dy
U 0 
(4.4.17)

Similarly, the energy thickness,  E , is given by:


u   u  
2
H
E = 0
1 −  
U 0   U 0  
 dy (4.4.18)
 
Hence, the values of the displacement and momentum thicknesses depend on the velocity profile in
the boundary layer.

Since the upper limit of integration is independent of x, the integration and differentiation operations
can be interchanged in Equation (4.4.12) to yield:
w
=
d
(U 02 M ) +  DU 0 dU 0 (4.4.19)
 dx dx
The momentum equation applies to both laminar and turbulent boundary layers in two-dimensional
incompressible flow.

6
4.4.1 Integral momentum balance for laminar boundary layer

The momentum equation provides a means of calculating the thickness of the boundary layer and the
drag on the fluid when the shear stress can be expressed as an explicit function of the velocity gradient.
Pohlhausen found a solution to Equation (4.4.11) by assuming a cubic velocity profile of the form:
u = a + by + cy 2 + dy 3 (4.4.20)
The constants a, b, c and d are evaluated from the following boundary conditions:
(i) At the wall or plate, there is no slip and therefore the fluid velocity there must be zero i.e., at
y = 0 , u = 0 hence a = 0.
(ii) At the edge of the boundary layer, the velocity equals the free stream velocity i.e., at
y =  , u = U 0  U 0 = b + c 2 + d 3 .
(iii) At the edge of the boundary layer, there is no momentum transfer (u = constant) i.e., at y =  ,
du du
= 0 = b + 2 c  + 3d  2 = 0 .
dy dy y =

(iv) At the wall, the shear stress is constant and therefore the gradient of momentum is zero i.e., at
d d 2u d 2u
y = 0, = 0 or = 0  = 2c = 0 so c = 0.
dy dy 2 dy 2 y = 0
3U 0 U0
Hence, a = 0 , b = , c = 0, d = − .
2 2 3
So, Equation (4.4.20) becomes:
3
u 3 y 1 y
=  −   (4.4.21)
U0 2   2  
du
Eliminating u from Equation (4.4.11) using Equation (4.4.20) and the definition  w =  , we
dy y=0

have:
3 U 0 d  2  3 y 1  y   3 y 1 y 
3 3

dx 0
= U0  −    1 − +    dy (4.4.22)
2   2  2      2 2    
Integrating, we have:
3 U 0
=
39 d
(U 02 ) (4.4.23)
2  280 dx
But U 0 is constant; Equation (4.4.23) reduces to:
140 
d  = dx (4.4.24)
13 U 0
Since δ = 0 at x = 0, we have, after integrating,
2 140 x
= (4.4.25)
2 13 U 0
x
 = 4 .64 (4.4.26)
U 0
Define Re x = U 0 x  , Equation (4.4.25) becomes:

7
4.64 x
 = (4.4.27)
Re x
 5 .0 x 
A comparison of this result with the result of the exact solution of Blasius   =  is within
 Re x 
 
93% accuracy. Hence, this is good enough for all practical purposes.

The skin drag is due to the wall shear stress,  w , and this acts on the surface area (wetted area). The
drag force, F, is therefore F =  w  wetted area. The dynamic pressure, p D , is the pressure resulting
from the conversion of the kinetic energy of the stream into pressure, and is given by p D = U 02 2 .
The drag coefficient, C Df , is defined as:
drag force
C Df =
dynamic pressure  wetted area
F
C Df =
1
U 02  wetted area
2
 w  wetted area
C Df =
1
U 02  wetted area
2
Hence, the point skin friction coefficient, C fx , is defined as the ratio of the shear stress of the wall to
the dynamic pressure of the free stream, and is given by:
w 3 U 0 2 0.646
C fx = = = (4.4.28)
1 2 U 0 2
Re
U 02
x
2
Note that this is the same definition of the pipe friction coefficient, C f , and it is in fact the same
thing. It is used in the Darcy formula to calculate the pressure loss in pipes due to friction. However,
on a small area, the drag is dF =  w dA . If the body is not a thin plate and has an area inclined at an
angle  to the flow direction, the drag force in the direction of flow is  w dA cos  .

Figure 4.6. An inclined object to flow direction.

The drag force acting on the entire surface area is found by integrating over the entire area, that is:
F =   w cos  dA (4.4.29)
The mean skin friction coefficient over the whole length of the plate can be obtained by integrating
Equation (4.4.28) from 0 to L and dividing the result by L i.e.,
1 L 1.292
C fL =  C fx dx = (4.4.30)
L 0
Re L
The frictional drag on the plate per width is given by:

8
1  1 1.292 0.646 U 02 L
F f =  w L = C fL  U 02  L = U 02 L= (4.4.31)
2  2 Re L Re L
L
The total shearing force on the surface, Fs = 
0
 w Bdx , where Fs is the surface resistance produced
by viscous stresses on one side of the plate, B is the width of the plate and L is the length. This is
because the shear stress at the boundary,  w , varies along the plate, it is necessary to integrate this
stress over the entire surface to obtain the total shearing force on the surface.
1 0.323 U 0
 w = U 02 C fx = Re x (4.4.32)
2 x
which is used to obtain the local shear stress at any given section (any given value of x for the laminar
boundary layer).
L 0 .323 B U U0x
Fs =  0
dx = 0.646 BU 0 Re L (4.4.33)
0 x v
The exact solutions of the boundary layer equation for δ and C fL have been obtained by Blasius
(1913) as:
5x
 = (4.4.34)
Re x
1.328
C fL = (4.4.35)
Re L
where Re L is the Reynolds number based on the approach velocity and the length of the plate.

4.4.2 Integral momentum analysis for turbulent boundary layer

While the laminar boundary layer is a single layer of fluid in which the particles are all in laminar
flow, the turbulent boundary layer consists of two layers, a thin laminar sublayer next to the plate
above which lies the turbulent layer. However, the region is regarded as turbulent because the
thickness of the turbulent layer is very large compared to that of the laminar sublayer.

The shear stress versus boundary layer thickness relationship for the turbulent boundary layer is
derived as follows. The shear stress at the wall can be shown from first principle to be:
1 dp
w = − rw (4.4.36)
2 dx
Defining friction factor, f, for the pipe flow as the ratio of dimensionless pressure drop to
dimensionless distance or pipe length, we have:
dp
1 v 2
f = 2 (4.4.37)
dx
D
where v is the mean flow velocity in the pipe.
Combining Equations (4.4.36) and (4.4.37), noting that D = 2rw , we have:

 w = f v
2

8
(4.4.38)
Blasius has shown that for 3,000  Re  10 ,000 ,

9
0.316 0.316
f = 1
= 1
(4.4.39)
Re 4
 vD  4

 
Using Equation (4.4.38) in Equation (4.4.37), we have:
0.316 v 2
w = 1
(4.4.40)
  vD  4 8
  
In pipe flow, the boundary layer develops from the entrance to the pipe at all points of the inside
periphery and somewhere downstream, all the boundary layers meet. At that point, the boundary
layer is said to be fully developed and the boundary layer thickness is equal to the pipe radius while
the centre velocity is equal to the free stream velocity. Since for turbulent flow, the ratio of the mean
velocity to the centre or maximum velocity is 1 to 1.235, the mean velocity is therefore U 0 1.235 .
Hence, Equation (4.4.40) becomes:
 (U 0 1.235 )
2
0.316
w = (4.4.41)
 (U 0 1.235 )(2 )   4
1
8
0 .023 U 02
w = (4.4.42)
(U 0 ) 4
1

Equations (4.4.38) to (4.4.42) apply to both pipes and flat plates.


The velocity distribution in the turbulent boundary layer is predicted from Prandtl’s (1904) equation:
1
u  y n
=  (4.4.43)
U0  
This equation is often referred to as Prandtl’s one-seventh power law since the value of n has been
established experimentally to be 7.
Combining Equations (4.4.11), (4.4.42) and (4.4.43), we have:
d   y  7  y  7 
1 1 2
   4

dx 0   
0.023   =   −    dy (4.4.44)
 U 0       
1
   4 7 d
0.023   = (4.4.45)
 U 0   72 dx
Integrating, we have:
1
   4 5
x  = 3.45 4 + k (4.4.46)
 U 0 
where k is a constant of integration.
Assume the flow is turbulent right from the leading edge, i.e.,  = 0 at x = 0 (practically this is not
usually true, the assumption only simplifies the problem), we have:
 0.376
= 1
(4.4.47)
x Re 5 x

Substituting  into Equation (4.4.42), we have:


U 02 0.0576
w = (4.4.48)
2 Re 15
x

10
1
but C fx =  w U 02 , we have:
2
0.0576
C fx = 1
(4.4.49)
5
Re x
1 L 0.072
C fL = 
L 0
C fx dx = 1
(4.4.50)
Re L 5
When  w from Equation (4.4.48) is integrated over the area of the boundary, the overall shearing
resistance is obtained thus:
U 02
Fs = C fl BL (4.4.51)
2
0.072 BL U 0
2

Fs = 1
(4.4.52)
Re L 5 2
where B is the width of the plate and L its length.
Schlichting (or Blasius) indicated that better agreement with experimental results is obtained when
the numerical coefficient is given a value of 0.074. Then
0.074
C fL = 1
(4.4.53)
Re L 5
Equations (4.4.47) to (4.4.53) are applicable in the Reynolds number range 5  10 5 to 10 7 . For higher
Reynolds numbers, the one-seventh power law is not precise and more refined analyses are required.
At Re  10 7 , C f is given by:
0.455 1700
C fL = − (4.4.54)
(log 10 Re L ) 2.58
Re L

= 0.079 Re −x 0.085 (4.4.55)
x
C fx = (2 log Re x − 0.65 )
−2.3
(4.4.56)

4.4.3 TWO-REGION BOUNDARY LAYER


The two-region boundary layer means a boundary layer over which the flow is laminar from the
leading edge to some point along the plate and turbulent from that point onwards. The drag coefficient
for such a boundary layer is taken as the mean of the expression for the drag coefficients for purely
laminar boundary layer and that for wholly turbulent boundary layer as follows:
1
L

C fL = (C fL )lbl . x c + (C fL )tbl . (L − x c ) (4.4.57)
where x c is the distance of the point of transition from the leading edge, L the length of the plate,
(C )fL lbl and (C fL )tbl are the drag coefficients for laminar and turbulent flows respectively.

1  1.328 0.074 
C fL =  xc + ( L − x c ) (4.4.58)
L  Re L Re 0L.2 

This equation can be used to calculate the total drag on the plate.
For a smooth plate, Re at the critical point is 500,000.

11

You might also like