You are on page 1of 42

EMG 3207: Fluid Mechanics III

S.K. Musau

June, 2023

Contents
1 Introduction 3

2 Kinematics of Fluid Flow 4


2.1 Acceleration of a Fluid Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3 Incompressible Potential Flow 6


3.1 Potential Flow Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2 Rotation of a Fluid Element . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.3 The Velocity Potential and Stream Function . . . . . . . . . . . . . . . . . . 8
3.4 Potential Flow Nets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.5 Straight Line Flows and their Combination . . . . . . . . . . . . . . . . . . . 11
3.6 Flow Past a Cylinder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.6.1 Lifting Flow Over a Cylinder . . . . . . . . . . . . . . . . . . . . . . 15
3.7 Vortex Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

4 Compressible Flow 19
4.1 Basic Equations for Compressible Gases in Steady Flow Conditions . . . . . 19
4.1.1 Continuity Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.2 Momentum Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.3 Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.1.4 Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1
4.2 Isentropic Flow of a Perfect Gas . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2.1 Isentropic Flow in Converging Nozzle . . . . . . . . . . . . . . . . . . 25
4.2.2 Isentropic Flow in Convergent-Divergent Nozzle . . . . . . . . . . . . 26

5 Flows in Turbines and Pumps 27


5.1 Types of Rotodynamic Machines . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2 One-Dimensional Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2.1 Euler’s Turbine Equation . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.2.2 Application of Euler’s Equation on a Centrifugal Pump/Fan Impeller 30
5.2.3 Application of Euler’s Equation on an Axial Flow Machine . . . . . . 31
5.3 Isolated Blade and Cascade Considerations . . . . . . . . . . . . . . . . . . . 32
5.3.1 Isolated Blade Considerations . . . . . . . . . . . . . . . . . . . . . . 33
5.3.2 Cascade Considerations . . . . . . . . . . . . . . . . . . . . . . . . . 34
5.4 Energy Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

6 Boundary Layer Theory 37


6.1 Boundary Layer Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.2 Laminar Boundary Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.2.1 Equation of Motion of Fluid in a Laminar Boundary Layer . . . . . . 40
6.2.2 Flow Over a Flat Plate (The Blasius solution) . . . . . . . . . . . . . 41
6.3 Turbulent Boundary Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

7 Lift and Drag Forces 42

2
EMG 3207 Fluid Mechanics III Lecture Notes DeKUT

Course Outline
Kinematics of fluid element. Potential flow: Rotational and irrotational flows; circulation and
vorticity; stream functions and velocity potential functions. Potential flow nets. Superposi-
tion of rectilinear flows, source and sink. Vortex motion; free and forced vortex flow. Flow
past a cylinder. Pressure fields and lift forces. Compressibility effects in moving fluids: basic
equations for compressible gases in steady flow conditions; Mach number. One dimensional
isentropic flow in convergent and divergent nozzles. Flows in turbines and pumps; Degree
of reaction: impulse and reaction stages. Velocity triangles and utilization factors, losses
through stages and blade speed ratio. Concept of laminar and turbulent boundary layers.
Lift and drag considerations on bodies moving in a gas.

Prescribed text books


1. Douglas, J.F., Gasiorek J.M. & Swaffield J.A., (2001), Fluid Mechanics, Prentice Hall,
4th Ed.

2. Munson B.R., Young D.F. & Okiishi T.H. (1998) Fundamentals of Fluid Mechanics,
John Wiley and Sons, 3rd Ed.

3. Roberson J.A., Crowe C.T. & Elger D.F. (1999) Engineering Fluid Mechanics, John
Wiley and Sons, 9th Ed.

4. Bansal R.K. (1992) Fluid Mechanics and Hydraulic Machines, R.K. Laxmi Publications,
4th Ed.

5. Massey B.S. (1994) Mechanics of Fluids, Van Nostrand Reinhold (UK) Co. Ltd, 5th
Ed.

1 Introduction
This unit deals with the kinematics of fluid flow, compressible flow, fluid machinery and
boundary layer theory.

3
EMG 3207 Fluid Mechanics III Lecture Notes DeKUT

2 Kinematics of Fluid Flow


The kinematics of a flow describes the motion of the fluid without taking into account the
forces that cause this motion. The goal of kinematics is to describe the dependence of the
motion of the fluid elements on time for a given velocity field.
The flow of a fluid can be described by determining the position of every fluid particle at
every point in time. A particle’s change of position in time then yields its velocity and
acceleration. To differentiate between the different particles, we mathematically introduce
a particular coordinate system, fixed to the fluid particles but moving in space. We first
consider a family of surfaces with a = const, where a is given as some initial position as a
function of the spatial coordinates x, y, and z. We select two further families of surfaces b
= const and c = const such that a surface with a = const, a surface with b = const, and a
surface with c = const meet only at a single point. A fluid particle at this point of intersection
is then fully defined by the values of a, b, and c at a fixed but arbitrary time. A fluid particle
retains these fluid coordinates a, b, and c as its initial or rest position throughout its motion.
This means that each surface a = const, b = const, or c = const as an initial position is
always made up of the same fluid particles.
In order to determine the motion, i.e., the change in position of all fluid particles, the values
of the current position coordinates x, y, z of the particles have to be stated as functions of
time and of the fluid coordinates a, b, c of the initial position of the particle. We obtain

x = F1 (a, b, c, t), y = F2 (a, b, c, t), z = F3 (a, b, c, t) (1)

This system of equations (Eq. 1) is called Lagrange. Here we use fluid particle reference
frame. To fully describe the state of the flowing fluid, we need to know the pressure p
and, in the case of a compressible flow, the density ρ. In general, a simpler representation
that describes the flow state at every position and time more closely, without having
to consider each individual particle is used.
For steady flow, it is sufficient to state the magnitude and direction of the velocity
at each position in the space through which the fluid flows, and to make corresponding
statements about the pressure and, if necessary, the density.
For unsteady flow, this information (velocity, pressure and density at each position in space)

4
EMG 3207 Fluid Mechanics III Lecture Notes DeKUT

is stated at all time. Mathematically, the three velocity components u, v, w, the pressure p
and the density ρ are stated as functions of the spatial coordinates x, y, z and the time t.
For u, v, w we obtain the equations

u = f1 (x, y, z, t), v = f2 (x, y, z, t), w = f3 (x, y, z, t) (2)

This system of equations (Eq. 2) is called Euler. Here we use spatial fixed reference frame.
The system of equations 1 and 2 are known as the fundamental equations of kinematics.
Another representation of the instantaneous state of the flow of a fluid, is the use of stream-
lines. These run in the direction of the flow at all points; i.e., their tangents everywhere
have the direction of the velocity vector.
The differential equations of the streamlines read
dz w dz w dy v
= , = , = (3)
dy v dx u dx u

2.1 Acceleration of a Fluid Flow

The acceleration of a fluid flow, a is


dv d2 x
a= = 2,
dt dt
i.e the total derivative of a given velocity vector v = (u, v, w).
Consider the Euler system of equations (Eq. 2). The total derivative of the u component
u = f1 (x, y, z, t) of velocity vector is
∂u ∂u ∂u ∂u
du = · dt + · dx + · dy + · dz (4)
∂t ∂x ∂y ∂z
So the total time derivative of u is
du ∂u dt ∂u dx ∂u dy ∂u dz
= · + · + · + · , (5)
dt ∂t dt ∂x dt ∂y dt ∂z dt
with
dx dy dz
= u, = v, = w,
dt dt dt
from which, Eq. 5 reduces to
du ∂u ∂u ∂u ∂u
= +u · +v· +w· , (6)
dt ∂t
|{z} ∂x ∂y ∂z
| {z }
L C

5
EMG 3207 Fluid Mechanics III Lecture Notes DeKUT

where L is the local acceleration, i.e, local rate of change at a fixed position, and C is the
convective acceleration, i.e, convective spatial changes due to convection from place to
place.
Similarly, the acceleration in y and z directions are respectively given by

dv ∂v ∂v ∂v ∂v
= +u· +v· +w· (7)
dt ∂t ∂x ∂y ∂z
dw ∂w ∂w ∂w ∂w
= +u· +v· +w· (8)
dt ∂t ∂x ∂y ∂z
In general, the acceleration of a flow field, a, can be expressed as

dv ∂v ∂v ∂v ∂v ∂v
a= = +u· +v· +w· = + (v · ∇)v, (9)
dt ∂t ∂x ∂y ∂z ∂t
∂ ∂ ∂
with nabla operator ∇ = ( ∂x , ∂y , ∂z ) and v · ∇ the scalar product of the velocity vector v
and the nabla operator ∇.
For cartesian coordinate, this yield
     
du ∂u ∂u ∂u ∂u
a +u· +v· +w·
 x   dt   ∂t ∂x ∂y ∂z 
a =  ay = dv = ∂v ∂v ∂v ∂v (10)
+u· +v· +w·
     
dt ∂t ∂x ∂y ∂z

     
dw ∂w ∂w ∂w ∂w
az dt ∂t
+u· ∂x
+v· ∂y
+w· ∂z

In the case of a steady flow, all partial derivatives with respect to time vanish, so ∂/∂t =
0, while the substantial derivative with respect to time d/dt can indeed be nonzero when
convective changes occur. In unsteady flows both ∂/∂t 6= 0 and d/dt 6= 0.

3 Incompressible Potential Flow

3.1 Potential Flow Theory

A potential flow is frictionless and irrotational. The velocity of a potential flow is derivable
from a scalar velocity potential φ as
v = ∇φ (11)

Mathematically, the necessary and sufficient condition that a vector field which is single-
valued (in this case v) can be derived from a scalar potential function φ is that the curl of

6
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

the vector be zero (except at singular points), i.e

∇ × v = 0. (12)

The curl of velocity vector is known as vorticity (or rotation), ω = ∇ × v. If the


vorticity is zero, then the flow is term irrotational, i.e the rotation or angular velocity of
any infinitesimal element of the fluid is zero.
Potential flow occur outside the boundary layer, which forms near a surface. In the boundary
layer, the effects of viscosity (or friction) is dominant compared to the effects of inertia and
pressure. Outside the boundary layer the effects of viscosity are quite small. One of the most
important application of potential flow theory is in aerodynamics. For example, the flow
of air over an airplane is a problem in potential flow.

3.2 Rotation of a Fluid Element

For simplicity, we restrict the analysis to one cartesian component as shown below,

The z component of ∇ × v is

∂v ∂u
ζz = ∇ × v = − ,
∂x ∂y
i.e
∂ ∂
∂x ∂y ∂v ∂u
= − . (13)
u v ∂x ∂y

The angular velocity of line ∆x is

vx+∆x − vx
,
∆x
or  
vx+∆x − vx ∂v
lim = , (14)
∆x→0 ∆x ∂x

7
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

The angular velocity of line ∆y is

 
uy+∆y − uy
− ,
∆y
or  
uy+∆y − uy ∂u
lim − =− , (15)
∆y→0 ∆y ∂y
The average angular velocity of the square ∆x∆y, ωz is the average of angular velocity of
lines ∆x and ∆y,  
1 ∂v ∂u
ωz = − . (16)
2 ∂x ∂y
Therefore, comparing Eqs (13) and (16), results in

ζz = 2ωz ,

i.e, vorticity is two times rotation.


The rotation can be related to a line integral using Stoke’s theorem, which states
that ”the area sum of the rotation over a given area is equal to the integral of
the velocity (along a curve) integrated around a curve bounding the area”, i.e
Z I
ω.dA = v.dI.

The line integral is called circulation denoted by Γ


I Z
Γ = v.dI = ω.dA.

The concept of circulation is very important in the theory of lifting surfaces such as airfoils,
hydrofoils and blades of rotodynamic machines.

3.3 The Velocity Potential and Stream Function

For an incompressible steady flow (density is constant), the principle of mass conservation
(continuity) take the following form:
∇·v =0

or
∂vx ∂vy ∂vz
+ + = 0. (17)
∂x ∂y ∂z

8
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Stream function
A stream function, ψ, depends upon the position coordinates, i.e,

ψ = f (x, y),

and can be defined1 so that

∂ψ ∂ψ
vx = , and vy = − . (18)
∂y ∂x

The condition for potential flow Eq. (12) can be stated for a two dimensional flow (xy plane)
as
∂vy ∂vx
ωz = − = 0. (19)
∂x ∂y
Thus, in term of the stream function is
   
∂ ∂ψ ∂ ∂ψ
− − = 0, (20)
∂x ∂x ∂y ∂y

and
∂ 2ψ ∂ 2ψ
+ 2 = 0, (21)
∂x2 ∂y
This is the Laplace equation for the stream function, which must be satisfied for the flow
to be potential.

Velocity Potential
The condition of irrotational is a necessary and sufficient one for the velocity to be derivable
from a scalar velocity potential, φ, as

v = ∇φ, (22)

which in cartesian co-ordinate is

∂φ ∂φ ∂φ
vx = , vy = , vz = . (23)
∂x ∂y ∂z
1
The value of stream function represents the flow rate between a given streamline described by the stream
function and a referenced boundary. Therefore, its definition must ensures that the continuity equation is
satisfied.

9
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Therefore, for a flow that is both incompressible and irrotational, Eq. (17) and (22) can be
combined to yield
∇ · ∇φ = 0, ⇒ ∇2 φ = 0 (24)

or in cartesian coordinate for two dimensional flow (xy plane) is


   
∂ ∂φ ∂ ∂φ
+ = 0, (25)
∂x ∂x ∂y ∂y

so that
∂ 2φ ∂ 2φ
+ = 0, (26)
∂x2 ∂y 2
Therefore, a potential flow must also satisfy the Laplace equation for the velocity potential.

Relationship Between Stream Function and Velocity Potential


Comparing Eqs. (18) and (23), vx and vy can be equated to get

∂ψ ∂φ ∂φ ∂ψ
= and =− (27)
∂y ∂x ∂y ∂x

These equations are known as Cauchy-Riemann equations and they enabled the stream
function to be calculated if the velocity potential is known and vice versa in a potential flow.
Example

3.4 Potential Flow Nets

The fact that, for a potential flow, both the stream function and the velocity potential
satisfy the Laplace equation indicates that ψ and φ are interchangeable (Cauchy-Riemann
equations)and that the lines of constant ψ, i.e. streamlines, and the lines of constant φ,
called equipotential lines, are mutually perpendicular. This means that, if streamlines are
plotted, points can be marked on them which have the same value of φ and can be joined
to form equipotential lines. Thus a flow net of streamlines and equipotential lines is formed
(see figure below for illustration).

10
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

3.5 Straight Line Flows and their Combination

The basic flow patterns in which the streamlines are straight lines are

• Rectilinear flow,

• Source flow, and

• Sink flow.

Rectilinear Flow

The convention for numbering the streamlines is that the stream function is considered
to increase to the left of the observer looking in the direction of flow as shown in the figure
above.
If the velocity of the rectilinear flow v is inclined to the x axis at an angle α, then its
components are
vx = v cos α and vy = v sin α.

The stream function is obtained by substituting the above expression into


∂ψ ∂ψ
dψ = dx + dy = vx dy − vy dx, (28)
∂x ∂y
which upon integration, yields
Z Z
ψ = v cos αdy − v sin αdx + constant. (29)

Since in a uniform flow v is constant and in a straight line flow α is constant, then Eq. (34)
becomes
ψ = vy cos α − vx sin α + constant. (30)

The constant of integration is made zero by choosing the reference streamline ψ0 = 0 to pass
through the origin, so that when x = 0 and y = 0, ψ = ψ0 = 0. Thus

ψ = v(y cos α − x sin α). (31)

11
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Since vx and vy are constant, then ∂vx /∂y and ∂vy /∂x are both zero and, therefore, the flow
is potential.
The velocity potential is obtained from
∂φ ∂φ
dφ = dx + dy = vx dx + vy dy. (32)
∂x ∂y
Therefore, Z Z
φ= v cos αdx + v sin αdy + constant, (33)

but if φ = φ0 = 0 at x = 0 and y = 0, then

φ = v(x cos α + y sin α). (34)

Examples

1. Derive an expression of stream function for a uniform, straight line flow in the direction
Ox, velocity u as shown in the diagram below.

2. Derive an expression of stream function for a uniform, straight line flow in the direction
Oy, velocity v as shown in the diagram below.

3. Derive an expression of stream function for a combined flow consisting of uniform flow
u = 10 m/s along Ox and uniform flow v = 20 m/s along Oy, as shown in the diagram
below.

Source and Sink Flow


A source flow is a flow in which the fluid flows radially outwards from a point, i.e, the
streamlines are straight lines emanating from a single point.
A sink flow is a flow in which the fluid flows radially towards a point, i.e, the streamlines
are straight lines directed toward a single point.
A sink flow is simply treated as a negative source flow and, thus, the mathematics of both
ma be explained by considering only source flow, as shown below

In a radial flow, since the velocity passes through the origin and is a function of θ only,

12
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

the tangential component of velocity does not exist and v = vr .


Consider a source of unit depth and let the steady rate of flow be q, known as the strength
of the source. Then, at any radius r, the radial velocity is given by

q
vr = (35)
2πr

Stream function, in terms of polar coordinate, is defined by

∂ψ 1 ∂ψ
vθ = − and vr = .
∂r r ∂θ

Therefore,
dψ = rvr dθ − vθ dr. (36)

Substituting the value of vθ and vr into Eq. (36), gives


 q   q 
dψ = r dθ = dθ. (37)
2πr 2π

Integrating,

ψ= + constant. (38)

If, ψ = ψ0 = 0 when θ = 0, the constant of integration becomes zero and


ψ= for a source, (39)


ψ=− for a sink. (40)

Combination of Rectilinear and a Source Flow
The simplest case of combining flow patterns is that in which a source is added to a uniform
rectilinear flow, as shown in the diagram below;

This is accomplished by additions of the stream functions of the two types of flow. The
stream function for a rectilinear flow parallel to the x axis is

ψR = uy = ur sin θ,

13
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and for a source is



ψS = .

Thus the stream function for the combined flow is

ψ = ψR + ψS = ur sin θ + . (41)

Example

3.6 Flow Past a Cylinder

A flow pattern equivalent to that of an ideal fluid passing a stationary cylinder, with its axis
perpendicular to the direction of flow, is obtained by combining a doublet and a rectilinear
flow, as shown in figure below;

NB: A doublet is obtained by combination of a source and a sink. Its stream function is
 m 
ψD = − sin θ,
2πr
where m = qds, is the strength of the doublet, q is the rate of flow per unit depth and ds is
the distance between them (sink and source).
Therefore, the resulting stream function for the flow past a cylinder is
 m 
ψC = ψD + ψR = − sin θ + ur sin θ,
2πr
or
  m 
ψC = ur − sin θ. (42)
2πr
Since the distance between the two stagnation points S is the diameter of the cylinder, i.e 2a,
where a is the radius, and the flow at stagnation point is zero, it follows that the streamline
passing through S is ψ0 = 0. Thus
  m 
ψ0 = ua − sin θ = 0, (43)
2πa

14
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

so that
m = 2πua2 .

Substituting to Eq. (42),


a2
 
ψC = u r − sin θ. (44)
r
Example

3.6.1 Lifting Flow Over a Cylinder

It occur when there is flow over a spinning cylinder (spinning about its axis at relatively high
rpm). The flow is synthesized by addition of non-lifting flow over a cylinder and a vortex
of strength Γ as shown above. The resulting streamlines pattern are not symmetrical about
the horizontal axis through point O. Hence the cylinder will experience a resulting finite
normal force acting upward (lifting force). However, the streamlines are symmetrical about
the vertical axis through O, and as a result the drag force will be zero.
Note that because a vortex of strength Γ has been added to the flow, the circulation about
the cylinder is now finite and equal to Γ.
It can be shown that, the coefficient of lift, cl is given by

Γ
cl =
RU∞

The lift per unit span, L is given by

1 2
L = ρ∞ U∞ Acl (45)
2

Here, ρ∞ is the density of fluid, U∞ is the flow velocity, and the plan form area A = 2R.
Therefore Eq. 119 can be rewritten as

1 2 Γ
L = ρ∞ U∞ 2R = ρ∞ U∞ Γ (46)
2 RU∞

15
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Equations 120 is called Kutta-Joukowski theorem and gives the lift per unit span for
a circular cylinder with circulation Γ. It is named after the German mathematician M.
Wilheim Kutta (1867-1944) and Russian physicist Nikolai E. Joukowski (1847-1921), who
independently obtained it in the first decade of 19th century (around 1900-1910).
Example
Consider the lifting flow over a circular cylinder with a diameter of 0.5 m. The free stream ve-
locity is 25 m/s, and the maximum velocity on the surface of the cylinder is 75 m/s(clockwise).
The free stream conditions are those at standard sea level conditions (ρ∞ = 1.22 kg/m3 ).
Calculate lift per unit span on the cylinder.

3.7 Vortex Motion

In a vortex flow, the streamlines are concentric circles, as shown in the figure below;

The characteristic of a vortex flow is that the radial component of velocity vr = 0 and
v = vθ . This is so because, there cannot be any flow across streamlines.
The are two fundamental types of vortex flow distinguished by the nature of flow; rota-
tional (forced) vortex and irrotational (free) vortex.
Free Vortex
Consider, in a free vortex flow, an element of fluid between streamlines ψ and ψ + dψ as
shown in the figure below;

Since it is irrotational, the vorticity and circulation across the stream must be equal to

16
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

zero. The circulation round the element, starting from A in the anticlockwise direction, is

ΓABCD = 0 − (vθ + dvθ )(r + dr)dθ + 0 + vθ rdθ, (47)

and , neglecting higher order terms (third order),

ΓABCD = −(vθ dr + rdvθ )dθ. (48)

Since the flow is irrotational, then

ΓABCD = −(vθ dr + rdvθ )dθ = 0, (49)

so that
vθ dr + rdvθ = 0, (50)

which can be integrated to give


rvθ = constant. (51)

This equation defines the relationship between the velocity and radius for a free vortex. It
shows that the velocity increases towards the centre of the vortex and tends to infinity when
radius tends to zero.
The constant in Eq. (51) can be determined by evaluating the circulation around the cen-
tre, i.e, along any of the concentric streamlines. The circulation around a circular streamline
is I
ΓC = vds = circumference × tangential velocity = 2πrvθ . (52)

The circulation is constant for any streamline (and the whole flow vortex field) since rvθ =
constant. It is used to measure the intensity of the vortex and is known as the vortex
strength. Thus Eq. (51) can be rewritten as
ΓC
rvθ = . (53)

The stream function can be obtained from

dψ = vr rdθ − vθ dr.

Since vr = 0, Z
ψ=− vθ dr (54)

17
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and, substituting for vθ from Eq. (53)


Z
ΓC ΓC
ψ=− dr = − ln r + constant. (55)
2πr 2π

The constant of integration is made zero by taking ψ = 0 at r = 1, so that,

ΓC
ψ=− ln r,

for anticlockwise rotation. The sign for the above expression becomes positive for clockwise
rotation.
Forced Vortex
Consider the same diagram used in the analysis of free vortex. The circulation around a
segmental element is
ΓABCD = −(vθ dr + rdvθ )dθ.

The area of the element is


A = rdθdr,

so that vorticity is given by


 
ΓABCD vθ dvθ
ζ= =− + , (56)
dA r dr

but rotation ω is the angular velocity, which at any radius r is related to the tangential
velocity vθ by

ω= ,
r
it follows that for a forced vortex the vorticity

ζ = −2ω,

and is constant for a given vortex.


The stream function is obtained by substituting vθ = rω into Eq. (54), which yields
Z
1
ψ = − ωrdr = − ωr2 + constant. (57)
2

But for ψ = 0 at r = 0,
1
ψ = − ωr2 ,
2

18
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

for anticlockwise direction.


Example
A two-dimensional fluid motion takes the form of concentric, horizontal, circular streamlines.
Show that the radial pressure gradient is given by

dp v2
=ρ ,
dr r

where ρ=density, v =tangential velocity, r =radius. Hence, evaluate the pressure gradient for
such a flow defined by ψ = 2 ln r, where ψ =stream function, at a radius of 2 m and fluid
density of 103 kg m−3 .

4 Compressible Flow
A flow is said to be compressible when the density is varying. For most fluids used in
engineering, density changes are negligible, and therefore, they are considered to be incom-
pressible. However, this assumption is not valid for gases, since large variations of density
can be produced as a result of the changes of pressure.
The compressibility of a fluid flow is defined by the Mach number, M, of the flow.

4.1 Basic Equations for Compressible Gases in Steady Flow Con-


ditions

For incompressible flows, the unknown parameters are usually pressure and velocity. For
compressible flows, density becomes an additional variable to be solved. Furthermore, sig-
nificant variations in fluid temperature may occur as a result of density or pressure changes.
We therefore, require four equations in order to solve the four unknown, i.e:

1. Continuity equation,

2. Momentum equation,

3. Energy equation,

4. Equation of state.

19
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Let us consider a control volume as shown below;

4.1.1 Continuity Equation

The conservation of matter requires that

Mass of flow through section 1 = Mass of flow through section 2,


   
∂ρ ∂A ∂v
ρAv = ρ + δs A+ δs v+ δs .
∂s ∂s ∂s
Expanding the above equation and ignoring high order terms, yields

∂v ∂A ∂ρ
ρA + ρv + Av = 0, (58)
∂s ∂s ∂s

or

(ρAv), (59)
∂s
which is the continuity equation in differential form. It can be integrated to get

ρAv = constant, ⇒ ρ1 A1 v1 = ρ2 A2 v2 . (60)

4.1.2 Momentum Equation

The forces acting on the body are pressure and body forces. Thus, the net force acting on
the control volume in direction of flow is given as
    
∂p ∂A ∂p δs
F = pA − p + δs A+ δs + δA p + − W cos α,
∂s ∂s ∂s 2

where the term  


∂p δs
p+ ,
∂s 2
is the average value of pressure acting on the curved surface of area δA.

20
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Conservation of momentum requires that the rate of change of momentum in a given


direction is equal to the net force in the same direction. Thus,

F = ma,
    
∂v ∂p ∂A ∂p δs
ρAδs = pA − p + δs A+ δs + δA p + − ρgAδs cos α, (61)
∂t ∂s ∂s ∂s 2
and ignoring higher order terms,

∂v ∂p
ρ = − − ρg cos α.
∂t ∂s

But
∂v ∂v ∂s ∂v δz ∂z
= =v , and cos α = = in the limit δs → 0.
∂t ∂s ∂t ∂s δs ∂s
Therefore,
∂v 1 ∂p ∂z
v + +g = 0. (62)
∂s ρ ∂s ∂s
Integration of Eq. (62) yields

v2
Z
1 ∂p
+ ds + gz = constant.
2 ρ ∂s

The second term in the above equation can only be integrated if the relationship between p
and ρ is known. For incompressible flow, ρ is constant and, thus

v2 p
+ + gz = constant,
2 ρ

which is the Bernoulli’s equation.

4.1.3 Energy Equation

From the first law of thermodynamics energy is conserved, i.e

∆Q − ∆W = ∆E. (63)

The total (internal, kinetic and potential) energy of a unit mass of fluid at point 1, E1 is
given by
v12
E1 = cv T1 + + gz1 ,
2

21
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and at point 2, E2
v22
E2 = cv T2 + + gz2 .
2
Therefore,
v22 v2
∆E = (cv T2 + + gz2 ) − (cv T1 + 1 + gz1 ). (64)
2 2
The work done is Z 2
∆W = pδV,
1

where δV is change in volume. For a constant pressure case,

∆W = p2 V2 − p1 V1 ,

and since V = 1/ρ and RT = p/ρ, then


p2 p1
∆W = − = RT2 − RT1 . (65)
ρ2 ρ1
Therefore, substituting Eqs (64) and (65) into (63), gives
v22 v2
∆Q = (cv T2 + + gz2 + RT2 ) − (cv T1 + 1 + gz1 + RT1 ), (66)
2 2
but cp = cv + R, so that
v22 v2
∆Q = (cp T2 + + gz2 ) − (cp T1 + 1 + gz1 ). (67)
2 2
The potential energy for gases is usually negligible and can be ignored. For adiabatic flow
(∆Q = 0),
v2 v12
   
cp T2 + 2 − cp T1 + = 0, (68)
2 2
or
v22 v2
cp T2 + = cp T1 + 1 , (69)
2 2
2
v
⇒ cp T + = constant. (70)
2

4.1.4 Equation of State

It gives the relationship between pressure, density and temperature for a pure substance.
The perfect gas law is given as
p = ρRT. (71)

22
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

4.2 Isentropic Flow of a Perfect Gas

The flow of compressible fluid in a duct is affected by area change, friction, heat transfer,
electromagnetic, etc. If all these parameters are included in the equations of motion, then
a lot of mathematical complexities arise. An isentropic flow is where there is no heat
transfer involved and the entropy remains constant. Therefore, the effect of area alone is
considered.
Using continuity equation (60) and momentum equation (62), it is possible to show that
dA
dp(1 − M 2 ) = ρv 2 , (72)
A
and
dv dA
(M 2 − 1) = , (73)
v A
where M is the Mach number given by
v
M= ,
c
where c is the velocity of sound and is defined by
 
2 ∂p
c = .
∂ρ isentropic
From Eqs (72) and (73), if M<1 (subsonic flow), then the term (1 − M 2 ) is always
positive and (M 2 − 1) always negative and thus an increase in area is always accompanied
by an increase in pressure and a decrease in velocity.
If M>1 (supersonic flow), then the term (1 − M 2 ) is always negative and (M 2 − 1) always
positive and thus an increase in area is always accompanied by an decrease in pressure and
a increase in velocity. These are illustrated in the figure below:

Therefore, a flow cannot be accelerated from rest to a velocity greater than the velocity
of sound (M > 1) in a converging duct regardless of the pressure difference imposed. To
achieve a supersonic speed, a converging-diverging nozzle is used.

23
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Stagnation Conditions and Use of Tables


The stagnation conditions define a reference state for compressible flows. These are condi-
tions where the flow velocity is zero. In most cases, the stagnation condition is reservoir or
atmosphere. The fluid must be accelerated to attain actual state or decelerated from the
actual state to stagnation.
From energy view point, stagnation enthalpy is defined as the energy required to bring
isentropically or adiabatically, the fluid to rest.
From conservation of energy between stagnation conditions and any point in the flow
(subscript 0 denote stagnation conditions)

v2
h0 = h + .
2

For a perfect gas


v2
cp T0 − cp T = .
2
Therefore,
v2 v2
 
T0 = +T =T +1 ,
2cp 2cp T
but
γR v
cp = , γRT = c2 , and M = ,
γ−1 c
so that
v2 v2
     
γ−1
T0 = T +1 =T +1 ,
2γRT /(γ − 1) 2 c2
Therefore,  
T0 γ−1 2
= 1+ M . (74)
T 2
For a given value of γ (specified gas), it is possible to tabulate T0 /T or T /T0 against M.
The ratio of pressures and densities between the stagnation condition and any point in
the flow can be derived using equation of motion and continuity equation as
  γ
p0 γ − 1 2 γ−1
= 1+ M . (75)
p 2
 1
 γ−1
ρ0 γ−1 2
= 1+ M . (76)
ρ 2
Therefore, for a given gas it is also possible to tabulate p0 /p and ρ0 /ρ against M.

24
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

The mass flow rate ṁ becomes


p
ṁ = ρAv = ρAM γRT ,

but ρ = p/RT , hence √ √


pAM γRT pAM γ
ṁ = = √ .
RT RT
Since,
γ
− γ−1 − 1

 p 
γ−1 2 γ−1 2 2
p = p0 1+ M , and T = T0 1 + M ,
2 2
then,
√   γ+1
p0 AM γ γ − 1 2 2−2γ
ṁ = √ 1+ M . (77)
RT0 2
We can denote the the conditions when sonic conditions are achieved by superscript ∗. For
an isentropic flow p0 and T0 are constant.
A can be directly related to M by choosing a reference area for which M = 1 and denote
the area as A∗ . Since by conservation of mass, ṁ is the same at all section. Thus,
√   γ+1 √   γ+1
p0 AM γ γ − 1 2 2−2γ p0 A∗ γ γ − 1 2−2γ
√ 1+ M = √ 1+ ,
RT0 2 RT0 2
or " γ+1
# 2−2γ
A 1 1 + γ−1
2

= γ−1 (78)
A M 1 + 2 M2
Again, for a given gas it is also possible to tabulate A/A∗ against M. For a given value
of A/A∗ , there are two possible solution for M. One solution is subsonic and the other is
supersonic, as illustrated in the following diagram:

4.2.1 Isentropic Flow in Converging Nozzle

Assume a fluid in a large reservoir is to be discharged through a converging nozzle as shown


below:

25
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

For a constant reservoir pressure pr , it is desired to determine the rate of mass flow
through the nozzle as a function of back pressure pb imposed on the nozzle. The are several
possibilities of flow depending on the magnitude of the back pressure:

1. For pb > pr , the flow is reversed.

2. For pb = pr , there will be no flow in the nozzle.

3. For pb < pr , the flow is introduced into the nozzle.

4. pb can be reduced further, therefore accelerating the velocity at the nozzle exit until
the exit velocity equals the velocity of sound, i.e M = 1 at the exit.

NB: The flow cannot be accelerated beyond sonic conditions as observed earlier in a con-
verging nozzle.
Changes in conditions downstream are communicated to the reservoir through a wave
traveling through the fluid at the velocity of sound. For subsonic flow, changes are commu-
nicated faster than the velocity of the fluid and the reservoir adjusts itself by sending more
fluid. Since the nozzle is able to adjust itself to the back pressure, pexit = pb for subsonic
flow.
When M = 1 at the exit plane, the flow is said to be choked and reduction in pb cannot
result into any increase in ṁ.

4.2.2 Isentropic Flow in Convergent-Divergent Nozzle

Assume a fluid in a large reservoir is to be discharged through a converging-diverging nozzle


as shown below:

When pb equal to the reservoir pressure, then the pressure distribution in the nozzle is
uniform and there will be no flow induced in the nozzle.

26
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

By reducing the pressure slightly, flow is introduced and is subsonic in both converging
and diverging section. The pressure will decrease in the converging section and increase in
the diverging section.
Further reduction in back pressure causes the velocity to increase at the throat and reach
a sonic condition. Now the pressure wave in the subsonic flow which was communicating
with the reservoir to tell it to release more flow can not pass the throat. Thus any further
decrease in back pressure is not communicated to the reservoir and the mass flow remain the
same. The nozzle is said to be choked. When this happens, there are two possibilities:

• Either the flow continues to accelerate to supersonic flow and thus in the diverging
section the pressure drops resulting in what we call the Laval nozzle.

• Or the flow decelerates so that the flow in the diverging section is subsonic and the
pressure increases.

5 Flows in Turbines and Pumps


Fluid machinery (add or subtract energy from the fluid) can generally be classified into:
rotodynamic and positive displacement machines. Examples of machines that add energy to
the fluids are pumps, fans and compressors, while machines that extract energy from fluid
are turbines.
In rotodynamic machines, there is a free passage of fluid between the inlet and outlet of
the machine without any intermittent sealing taking place. All rotodynamic machines have
a rotating part called a runner, impeller or rotor.
In positive displacement machines, fluid is drawn or forced into a finite space bounded
by mechanical parts and is then sealed in it by some mechanical means. The fluid is then
forced out or allowed to flow out from the space and the cycle is repeated.
In this section we are going to looked at the fluid flow in rotodynamic machines and
the relationships between the rate of fluid flow and the difference in total head across the
impeller.

27
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

5.1 Types of Rotodynamic Machines

The direction of fluid flow in relation to the plane of impeller rotation distinguishes different
classes of rotodynamic machines. This include:

1. Axial flow machines - The flow is perpendicular to the impeller, i.e the flow is along its
axis of rotation.

2. Centrifugal/radial - The fluid approaches the impeller axially but it turns at the ma-
chine inlet so that the flow through the impeller is in the plane of the impeller rotation.

3. Mixed flow machines - The flow is partly axial and partly radial.

5.2 One-Dimensional Theory

The real flow through an impeller is three-dimensional, i.e the velocity of the fluid is a
function of three positional coordinates. The velocity distribution, therefore is very complex
and depend upon the number of blades, their shapes and thickness and the width of the
impeller.
The one dimensional theory simplify the problem by making the following assumptions:

1. The blades are infinitely thin and the pressure difference across them is replaced by
imaginary body forces acting on the fluid and producing torque.

2. The number of blades is infinitely large, so that the variation of velocity across blade
passages is reduced and tends to zero (axisymmetrical flow), i.e

∂v
= 0.
∂θ

3. There is no variation of velocity in the meridional plane, i.e across the width of the
impeller. Thus,
∂v
= 0.
∂z

The result of these assumptions is that the flow velocity is a function of radius alone, i.e
v = f (r).

28
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

5.2.1 Euler’s Turbine Equation

This is the general expression for the energy transfer between the impeller and the fluid based
on the one-dimensional theory.
Consider a centrifugal impeller rotating with constant angular velocity ω, as shown below;

At inlet, the fluid moving with an absolute velocity v1 enters the impeller through a
cylindrical surface of radius r1 and may make an angle α1 with the tangent at that radius.
At outlet, the fluid leaves the impeller through a cylindrical surface of radius r2 , absolute
velocity v2 and may make an angle α2 with the tangent at that radius.
From Newton’s second law applied to angular motion,

Torque = Rate of change of angular momentum = ṁvw2 r2 − ṁvw1 r1 = ṁ(vw2 r2 − vw1 r1 ),

where ṁvw1 r1 and ṁvw2 r2 are the angular momentum entering and leaving the impeller per
second, respectively.
The work done in unit time is given by:

Work done per second, Et = torque × angular velocity = ṁ(vw2 r2 − vw1 r1 )ω.

But u1 = ωr1 and u2 = ωr2 . Hence on substitution,

Work done per second, Et = ṁ(u2 vw2 − u1 vw1 ).

The unit is joules per second or watts.


The specific energy (energy per unit mass) is

Es = (u2 vw2 − u1 vw1 ),

and the Euler’s head E is given by

E = (1/g)(u2 vw2 − u1 vw1 ).

29
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

But vw1 = v1 cos α1 and vw2 = v2 cos α2 , so that

E = (1/g)(u2 v2 cos α2 − u1 v1 cos α1 ).

Using cosine rule,


2
vr1 = u21 + v12 − 2u1 v1 cos α1 ,
1
u1 v1 cos α1 = (u21 + v12 − vr1
2
).
2
Similarly,
1
u2 v2 cos α2 = (u22 + v22 − vr2
2
).
2
Substituting these expressions, then
1
(u22 + v22 − vr2
2
) − (u21 + v12 − vr1
2

E=( ) ,
2g
or
v22 − v12 u22 − u21 vr1
2 2
− vr2
E= + + ,
2g 2g 2g
| {z } | {z } | {z }
A B C
where A is the increase of kinetic energy of the fluid in the impeller, B is the energy used
in setting the fluid in circular motion about the impeller axis (forced vortex) and C is the
regain of static head due to a reduction of relative velocity in the fluid passing through the
impeller.

5.2.2 Application of Euler’s Equation on a Centrifugal Pump/Fan Impeller

Consider a centrifugal pump/fan impeller, as shown below;

Since the flow at inlet and outlet is through cylindrical surfaces and the velocity compo-
nents normal to them are vf 1 and vf 2 , the continuity equation applied to inlet and outlet
is
m = ρ1 πr1 b1 vf 1 = ρ2 πr2 b2 vf 2 ,

30
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and for incompressible flow,


πr1 b1 vf 1 = πr2 b2 vf 2 .

At inlet the following assumptions are made:

1. The absolute velocity is radial. Therefore, v1 = vf 1 and vw1 = 0.

2. The blade angle at inlet β1 is such that the blade meets the relative velocity tangentially.
This assumption is known as ’no shock’ condition.

At outlet the following assumptions are made:

1. The fluid leaves the impeller with a relative velocity tangential to the blade at outlet.
Thus, β20 = β2 . Thus from the velocity triangle at outlet,
u2 − vw2
cot β2 = ,
vf 2
so that
vw2 = u2 − vf 2 cot β2 .

Therefore applying these assumption in the Euler’s equation

E = (u2 /g)(u2 − vf 2 cot β2 ). (79)

The total amount of energy transferred by the impeller is, thus

Et = ṁgE = ṁu2 (u2 − vf 2 cot β2 ). (80)

5.2.3 Application of Euler’s Equation on an Axial Flow Machine

Consider an axial flow machine, as shown below;

The changes in velocity from inlet to outlet take place at the same radius and, hence

u1 = u2 = u = ωr.

31
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

Since the flow area is the same at inlet and outlet,

vf 1 = vf 2 = vf .

It is obtained from
m = ρvf π(R22 − R12 ).

The following assumptions are made with regard to velocity triangles:

1. There is no prewhirl at inlet and, hence, α1 = 900 , vw1 = 0, and v1 = vf .

2. The blade is set at an angle such that it meets the relative fluid velocity tangentially.

3. At the outlet, the relative velocity leaves the blade tangentially.

Thus from the velocity triangle at outlet,

vw2 = u − vf cot β20 ,

which, on substitution into the Euler’s equation, gives

E = (u/g)(u − vf cot β20 ). (81)

This equation applies to any particular radius r and is not necessarily constant over the
range from R1 to R2 . However, the blade can be twisted so that the increase in u with radius
is counterbalanced by an equal decrease in vf cot β20 and, hence the equation is constant
throughout.
Exercise

5.3 Isolated Blade and Cascade Considerations

In these approaches, the number of blades is considered to be finite. For this method we
define:

1. Pitch - The distance between the adjacent blades s.

32
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

2. Blade solidity, σ - The ratio of the blade chord c to the pitch. It is the measure of the
closeness of blades.
c
σ=
s

If the blades are close to one another, the flow between them may be treated as conduits.
Whereas if the blades are very far apart, they must be treated as bodies in external flows.

5.3.1 Isolated Blade Considerations

In this case, the assumption is that the blades are very far apart, i.e s → ∞ and σ → 0.
In section 3.6.1, it was shown that lift is dependent on circulation. From Eqs (119) and
(120),
1
Γ = U∞ Acl . (82)
2
Consider the circulation around a single blade as shown below;

I I B I C I D I A
ΓABCD = vds = vds + vdl + vds + vdl. (83)
A B C D

But, I C I A I B I D
vdl = − vdl, vds = −vw1 s1 , and vds = vw2 s2 .
B D A C

Therefore, the circulation around a blade, Γb = ΓABCD is given by

Γb = vw2 s2 − vw1 s1 . (84)

Now, consider circulation around two blades, as shown below:

33
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

I I B I C I D I E I F I E
ΓACDF = vds = vds + vds + vdl + vds + vds + vdl. (85)
A B C D E F

But,
I D I A I B I F I C I E
vdl = − vdl, vds + vds = Γb , and vds + vds = Γb ,
C F A E B D

so that,
Γ = ΓACDF = Γb + Γb = 2Γb . (86)

And for number of blade, say z;

Γ = zΓb = z(vw2 s2 − vw1 s1 ), (87)

but zs1 = 2πr1 and zs2 = 2πr2 , so that

Γ = 2π(vw2 r2 − vw1 r1 ). (88)

However, Euler’s Equation is

E = (ω/g)(vw2 r2 − vw1 r1 ). (89)

Comparing Eqs (88) and (89), we obtain


Γ Eg
= . (90)
2π ω
or
ωΓ ωzΓb
= . E= (91)
2πg 2πg
This equation may be used in conjunction with Kutta-Joukowsky’s Eq. (82).

5.3.2 Cascade Considerations

In this case, the assumption is that the blades are very close, i.e the solidity σ is significant.
The geometrically identical blades are arranged at the same distance from one another and
positioned in the same way with respect to the direction of flow to form a cascade.
A straight cascade (applied in an axial flow impeller) is where the blades are arranged
along a straight line. While circular cascade (applied in a centrifugal impeller) is where
the blades are arranged around the circumference.
Cascade can also be classified according to the way the operate as follows:

34
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

1. Impulse Cascade - when velocities upstream and downstream of a cascade are the
same in magnitude, so that the change in momentum is due to change in direction.
From Bernoulli’s equation there will be no pressure difference between upstream and
downstream side of the cascade.

2. Reaction Cascade - The pressure difference exists due to absolute velocities upstream
and downstream not being the same.

Straight Cascade
Consider a stationary straight cascade of height Z, as shown below:

Let the upstream fluid velocity v1 making an angle α1 with the line of the cascade be
deflected so that the downstream velocity v2 makes an angle α2 . The deflection is

ε = α2 − α1 .

The fluid velocities v1 and v2 are resolved into components parallel and normal to the cascade,
vt and vn , respectively. Applying Bernoulli’s equation, the pressure difference is given by

1
p1 − p2 = ρ(v22 − v12 ). (92)
2

Applying continuity, the mass flow through the cascade is

ṁ = sZρ1 vn1 = sZρ2 vn2 .

It implies that for an incompressible flow (density is constant)

vn1 = vn2 = vn .

Thus, the change in velocity is entirely due to change of tangential velocity component, i.e

v22 = vt2
2
+ vn2 , and v12 = vt1
2
+ vn2 ,

35
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

⇒ v22 − v12 = vt2


2 2
− vt1 .

Therefore, Eq. (92), reduces to


1 2 2
p1 − p2 = ρ(vt2 − vt1 ) = ρvt (vt2 − vt1 ), (93)
2
where vt = (1/2)(vt1 + vt2 ) is the mean tangential velocity. The force acting on a single blade
of the cascade in the direction perpendicular to it is

Fn = sZ(p1 − p2 ) = sZρvt (vt2 − vt1 ) = ρZvt Γb .

The rate of change of momentum across the cascade is again due to a change in vt and,
therefore, gives rise to a force in the direction of cascade, Ft , i.e

Ft = ṁ(vt2 − vt1 ) = sZρvn (vt2 − vt1 ) = ρZvn Γb .

The resultant force on the blade is


q q
F = Ft2 + Fn2 = ρZΓb (vt2 + vn2 ),

and it act at an angle β given by


Fn ρZvt Γb vt
cot β = = = . (94)
Ft ρZvn Γb vn
But vt1 = vn cot α1 and vt2 = vn cot α2 . Therefore,
vn
vt = (cot α1 + cot α2 ) ,
2
which on substituting into Eq. (94), result in
1
cot β = (cot α1 + cot α2 ) . (95)
2
β is equal to β∞ , defined as the mean direction of flow and obtained by superposition of the
inlet and the outlet velocity triangles, as shown below:

The force F being perpendicular to the mean direction of flow is the lift on the blade.
Exercise

36
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

5.4 Energy Losses

There are two main source of losses in hydraulic machine:

1. The velocities in the blade passages and at the impeller outlet are not uniform owing
to the presence of blade and the real flow being three dimensional. This results in a
diminished velocity of whirl component and, hence, reduces Euler’s head.

2. There are losses of energy due to friction, separation and wakes associated with the
development of boundary layers.

6 Boundary Layer Theory


Boundary layer is a thin viscous region between the solid body and fluid. This arises from
”no slip condition” at the solid boundary, i.e, zero relative velocity between fluid and wall.
The fluid velocity ought to resume its normal value at a very short distance from the wall.
This result in a large velocity gradient in this region and hence friction plays a significant
role. For the vast region of flow field away from the body, the velocity gradients are relatively
small, and frictions plays virtually no role. This natural division of the flow into regions;
one where friction is much more important than the other, was recognized by the famous
German fluid dynamist Ludwig Prandtl in 1904.

6.1 Boundary Layer Properties

Consider the viscous flow over a flat plate as shown below:

The viscous effects are contained within a thin layer adjacent to the surface, the thickness
is exaggerated for clarity. In the diagram, vw = 0 is the flow velocity at the surface which is
zero (no slip condition), Tw is the wall temperature (the temperature of the fluid immediately
at the surface is equal to temperature of surface), δ is the velocity boundary-layer thickness,

37
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and δT is the thermal boundary-layer thickness.


δ is defined as the distance above the wall where u = 0.99ue . Here ue is velocity at the outer
edge of the boundary layer. For the flow over a flat plate ue = v∞ . For a body of general
shape, ue is the velocity obtained from an inviscid flow solution evaluated at the body surface.
δT is defined as the distance above the wall where T = 0.99Te . Here Te is temperature at the
outer edge of the thermal boundary layer. For the flow over a flat plate Te = T∞ . For a body
of general shape, Te is obtained from an inviscid flow solution evaluated at the body surface.
At any x station, the variation of u between y = 0 and y = δ, i.e, u = u(y) is defined as the
velocity profile within the boundary layer. This profile is different for different x stations.
Similarly, the flow temperature will change above the wall, ranging from T = Tw at y = 0 to
T = 0.99Te at y = δT . At any given x station, the variation of T between y = 0 and y = δT ,
i.e, T = T (y) is called temperature profile within the boundary layer.
The relative thickness of δ and δT depend on the Prandtl number, Pr. Various research have
shown, that if P r = 1, then δ = δT ; if P r > 1, then δT < δ; if P r < 1, then δT > δ.
For air at standard conditions, P r = 0.71, hence the thermal boundary layer is thicker than
the velocity boundary layer, as shown in the above diagram.
The consequent of velocity gradient at the wall is the generation of shear stress at the wall,
i.e  
∂u
τw = µ , (96)
∂y w
 
∂u
where ∂y
is velocity gradient at y = 0 (at the wall).
w
Similarly, the temperature gradient at the wall generate heat transfer at the wall,
 
∂T
qw = −k , (97)
∂y w
 
where ∂T ∂y
is the temperature gradient at y = 0 (at wall).
w

Definitions

1. Displacement thickness, δ ∗
Consider point y1 above the boundary layer as shown below

38
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

The actual mass flow (per unit depth) across the vertical line connecting y = 0 and
y = y1 , mB is Z y1
mB = ρudy (98)
0

If there were no boundary layer, so that the speed would be ue everywhere, then the
mass flow, mideal , would be Z y1
mideal = ρe ue dy (99)
0

The difference between the two, will bring a ”missing mass flow”, i.e given as

Missing mass flow = ρe ue δ ∗ (100)

Missing mass flow = mideal − mB


Therefore, Z y1

ρ e ue δ = (ρe ue − ρu)dy (101)
0

Z y1  
∗ ρu
⇒δ = 1− dy δ ≤ y1 → ∞
0 ρ e ue

2. Momentum thickness , θ
Consider the same diagram for analysis of displacement thickness.
The momentum flow across a small segment dy, MB is given by

MB = udm = ρu2 dy (102)

The same elemental mass, outside the boundary layer has a momentum, M ideal given
by
Mideal = ue dm = ρuue dy (103)

The total decrement in momentum flow across the vertical line from y = 0 to y = y1
result in a missing momentum given as
Missing momentum = ρe u2e θ, i.e
Z y1
ρe u2e θ = (ρuue − ρu2 )dy (104)
0
Z y1  
ρu u
⇒θ= 1− dy δ ≤ y1 → ∞
0 ρ e ue ue

39
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

3. Energy thickness, δ E
Assignment 1: show that
Z y1  2 !
ρu u
δE = 1− dy (105)
0 ρe ue ue

6.2 Laminar Boundary Layer

The flow in a boundary layer can be laminar or turbulent. As far as boundary layer is
concerned, the transition from laminar to turbulent occurs at Reynold numbers above 105
based on mean fluid velocity and distance measured from the entry to the duct. In this
section we shall consider properties of the laminar boundary layer.

6.2.1 Equation of Motion of Fluid in a Laminar Boundary Layer

For simplicity, we shall consider two-dimensional flow along a plane portion of the surface.
This plane is taken as xz-plane, with x-axis in the direction of flow. The velocity distribution
is independent of z, and velocity has no z-component. We also assume that the flow is steady.
The exact Navier-Stokes equations and continuity equations are then
 2
∂ vx ∂ 2 vx

∂vx ∂vx 1 ∂p
vx + vy =− +ν + , (106)
∂x ∂y ρ ∂x ∂x2 ∂y 2
 2
∂ vy ∂ 2 vy

∂vy ∂vy 1 ∂p
vx + vy =− +ν + , (107)
∂x ∂y ρ ∂y ∂x2 ∂y 2
∂vx ∂vy
+ = 0. (108)
∂x ∂y
Here vx and vy denotes the velocity components in x and y directions, respectively, p is the
pressure and ν is the kinematic viscosity.
Since the boundary layer is very thin, it is clear that the flow in it takes place mainly parallel
to the surface, i.e, the velocity component vy is small compared to vx .
The velocity varies rapidly along y-axis, an appreciable change in it occurring at distances of
the order of the thickness δ of the boundary layer. Along x-axis, the velocity varies slowly, an
appreciable change in it occurring only over at distances of the order of length l characteristic
to the problem (the dimension of the body). Hence y-derivatives of the velocity are large
in comparison to x-derivatives. Therefore, ∂ 2 vx /∂x2 may be neglected in comparison with

40
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

∂ 2 vx /∂y 2 . Also comparing Eqs. 106 and 107, it can be seen that derivative ∂p/∂y is small in
comparison with ∂p/∂x, i.e it can be assume that ∂p/∂y = 0. This implies that pressure is
a function of x only, i.e p(x) and instead of writing ∂p/∂x, it is written as a total derivative
dp(x)/dx. This derivative can be be expressed in terms of velocity of the main stream U (x).
Since we have potential flow outside the boundary layer, Bernoulli’s Equation, p + (1/2)ρU 2
= constant, hold, whence (1/ρ)dp/dx = −U dU/dx.
Thus we obtain the equations of motion in the laminar boundary layer in the form of Prandtl’s
Equation:
∂ 2 vx
 
∂vx ∂vx 1 ∂p dU
vx + vy −ν =− =U , (109)
∂x ∂y ∂y 2 ρ ∂x dx
∂vx ∂vy
+ = 0. (110)
∂x ∂y

6.2.2 Flow Over a Flat Plate (The Blasius solution)

Consider the incompressible, two-dimensional flow over a flat plate at 00 angle of attack as
shown below:

For such flow density is constant, µ is constant and dp/dx = 0 (because the inviscid flow
over a flat plate at 00 angle of attack yields a constant pressure over the surface). Energy
equation is not needed to calculate the velocity field for an incompressible flow.
Hence the boundary layer equations, reduces to

∂ 2 vx
 
∂vx ∂vx
vx + vy −ν =0 (111)
∂x ∂y ∂y 2

∂vx ∂vy
+ = 0. (112)
∂x ∂y
The independent variables (x,y) are transformed to (ξ, η), where

yp √
ξ = x, η= Rex , and ψ(x, y) = νxU f (η) (113)
x

41
EMG 2307 Fluid Mechanics III Lecture Notes DeKUT

and u = ∂ψ/∂y and v = −∂ψ/∂x (stream function definitions). A similarity solution can be
obtained as
2f 000 + f f 00 = 0 (114)

The wall shear stress, τw is generally expressed in terms of nondimensional local skin friction
coefficient, cf as
τw
cf = (115)
(1/2)ρU 2
which can be shown to be
0.664
cf = √ (116)
Rex
where Rex is the local Reynolds number. The total drag force on the top surface of entire
plate is the integrated contribution of τw (x) from x = 0 to x = c. Letting Cf denote the skin
friction drag coefficient, we obtain
Z c
1
Cf = cf dx, (117)
c 0

which can be shown to be


1.328
Cf = √ (118)
Rec
where Rec is the Reynolds number based on the total plate length c.
It can also be shown that for laminar flow, the boundary thickness δ is given by
δ
= 5.48Re−0.5 (119)
x
Example

6.3 Turbulent Boundary Layer


−1/5
Cf = 0.074Rel (120)

It can also be shown that for turbulent flow, the boundary thickness δ is given by

δ = 0.37xRex−1/5 (121)

7 Lift and Drag Forces

42

You might also like