Fluid Mechanics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 73

CHAPTER ONE

INTRODUCTION

Relevance and significance in engineering applications

Fluid mechanics is the study of fluids either in motion (fluid dynamics) or at rest (fluid statics) and
the subsequent effects of the fluid upon the boundaries, which may be either solid surfaces or
interfaces with other fluids. Both gases and liquids are classified as fluids, and the number of fluids
engineering applications is enormous: breathing, blood flow, swimming, pumps, fans, turbines,
airplanes, ships, rivers, windmills, pipes, missiles, icebergs, engines, filters, jets, and sprinklers, to
name a few. When you think about it, almost everything on this planet either is a fluid or moves
within or near a fluid.

The essence of the subject of fluid flow is a judicious compromise between theory and experiment.
Since fluid flow is a branch of mechanics, it satisfies a set of well-documented basic laws, and
thus a great deal of theoretical treatment is available. However, the theory is often frustrating,
because it applies mainly to idealized situations which may be invalid in practical problems. The
two chief obstacles to a workable theory are geometry and viscosity.

Fluids omnipresent
• Weather & climate
• Vehicles: automobiles, trains, ships, and planes, etc.
• Environment
• Physiology and medicine
• Sports & recreation
Many other examples!

Fluid Properties

Fluids are characterized by their properties such as viscosity  and density Foremost among the
properties of a flow is the velocity field V(x, y, z, t). In fact, determining the velocity is often
tantamount to solving a flow problem, since other properties follow directly from the velocity
field. In general, velocity is a vector function of position and time and thus has three components
u, v, and w, each a scalar field in itself:
V(x, y, z, t) = iu(x, y, z, t) + jv(x, y, z, t) + kw(x, y, z, t)
The use of u, v, and w instead of the more logical component notation Vx, Vy, and Vz is the result
of an almost unbreakable custom in fluid mechanics.

Viscosity
Recall definition of a fluid (substance that deforms continuously when subjected to a shear stress)
)
and Newtonian fluid shear / rate-of-strain relationship ( τ  μ
Reconsider flow between fixed and moving parallel plates (Couette flow)
ut=distance fluid
y
u=
h u(y)=velocity 

f=fl
profile u=

Newtonian fluid: τ  μ  μ
t
ut ut
tan   or   for small 
y y
u du du
therefore   i.e.,   = velocity gradient and τ μ
y dy dy
Exact solution for Couette flow is a linear velocity profile

U
u ( y)  y Note: u(0) = 0 and u(h) = U, satisfies no slip
h
U
τ μ = constant
h
Where; U/h = velocity gradient = rate of strain

 = coefficient of viscosity = proportionality constant for Newtonian fluid

   N m2  Ns
du m m m2
dy s

 m2
  = kinematic viscosity
 s
Flow Patterns
Fluid mechanics is a highly visual subject. The patterns of flow can be visualized in a dozen
different ways, and you can view these sketches or photographs and learn a great deal qualitatively
and often quantitatively about the flow.
Four basic types of line patterns are used to visualize flows:
1. A stream line is a line everywhere tangent to the velocity vector at a given instant.
2. A path line is the actual path traversed by a given fluid particle.
3. A streak line is the locus of particles which have earlier passed through a prescribed point.
4. A timeline is a set of fluid particles that form a line at a given instant.
The streamline is convenient to calculate mathematically, while the other three are easier to
generate experimentally. Note that a streamline and a timeline are instantaneous lines, while the
path line and the streak line are generated by the passage of time.

A streamline is a line everywhere tangent to the velocity vector at a given instant of time. (A
streamline is an instantaneous pattern.) For example, consider simple shear flow between parallel
plates. At some instant of time, a streamline can be drawn by connecting the velocity vector lines
such that the streamline is everywhere parallel to the local velocity vector. In this example,
streamlines are simply horizontal lines.

 A streakline is the locus of particles which have earlier passed through a prescribed point in
space. (A streakline is an integrated pattern.) For example, consider simple shear flow between
parallel plates. A streakline is formed by injecting dye into the fluid at a fixed point in space. As
time marches on, the streakline gets longer and longer, and represents an integrated history of
the dye streak. In this example, streaklines are simply horizontal lines.
A pathline is the actual path traversed by a given (marked) fluid particle. (A pathline is an
integrated pattern.) For example, consider simple shear flow between parallel plates. A pathline
is the actual path traversed by a given (marked) fluid particle. A pathline represents an integrated
history of where the fluid particle has been. In this example, pathlines are simply horizontal
lines.

A timeline is a set of fluid particles that form a line segment at a given instant of time. (A
timeline is an integrated pattern.) For example, consider simple shear flow between parallel
plates. A timeline follows the location of a line of fluid particles. A timeline represents an
integrated flow pattern, since the time line continually distorts with time, as shown in the sketch.
Notice the no-slip condition in action. The top of the time line moves with the top plate, i.e. at
velocity V to the right. The bottom of the timeline, however, stays in the same location at all
times, because the bottom plate is not moving.

Note: For steady flow, streamlines, streaklines, and pathlines are all identical. However, for
unsteady flow, these three flow patterns can be quite different.
Fluid Statics

Fluid either at rest or moving in a manner that there is no relative motion between adjacent
particles. No shearing stress in the fluid. Only pressure (force that develop on the surfaces of the
particles)
Pressure at a point N/m2 (Force/Area)

 
F  ma
 x y  z
 Fy  Py x z  Ps x s sin  Y:  
2
ay

 x y z  x y z
 Fz  Pz x y  Pz x s cos   2
az Z :  
2
az

y   s cos ; z   s sin

y : p y  ps  a y y
2
z
z : p z  ps  ( a z   )
2
What happen at a pt. ? x,  y, z  0

p y  ps
 p y  p z  p s  is arbitrarily chosen
p z  ps
Pressure at a pt. in a fluid at rest, or in motion, is independent of direction as long as there are no
shearing stresses present. (Pascal’s law)
The pressure of fluid at rest is depends on density of fluid, depth and gravity.
Chapter two
Integral Relations for A Control Volume

From consideration of hydrostatics, we now move to problems involving fluid flow with the
addition of effects due to fluid motion, e.g. inertia and convective mass, momentum, and energy
terms.
physical laws of fluid mechanics
Basic Conservation Laws:
Each of the following basic conservation laws is presented in its most fundamental, fixed mass
form. We will subsequently develop an equivalent expression for each law that includes the
effects of the flow of mass, momentum, and energy (as appropriate) across a control volume
boundary. These transformed equations will be the basis for the control volume analyses
developed in this chapter.
Conservation of Mass:
Defining m as the mass of a fixed mass system, the mass for a control volume V is given by:

m sys    dV
sys
The basic equation for conservation of mass is then expressed as:
The time rate of change of mass for the
dm  control volume is zero since at this point
0 we are still working with a fixed mass
dt sys system.
Linear Momentum:
Defining P sys as the linear momentum of a fixed mass, the linear momentum of a fixed mass
control volume is given by:
P sys = mV =  V  dV
sys
where: V is the local fluid velocity and dV is a differential volume element in the control
volume.
The basic linear momentum equation is then written as:

  
dP  d mV
F  dt sys

dt sys
Moment of Momentum:
Defining H as the moment of momentum for a fixed mass, the moment of momentum for a
fixed mass control volume is given by:
H sys =  r  V  dV
sys
where r is the moment arm from an inertial coordinate system to the differential control
volume of interest. The basic equation is then written as:
dH 

 Msys =  r  F  dt sys
Energy:
Defining E sys as the total energy of an element of fixed mass, the energy of a fixed mass
control volume is given by:

E sys   e  dV
sys

Where; e is the total energy per unit mass ( includes kinetic, potential, and internal energy ) of
the differential control volume element of interest.
The basic equation is then written as:
d E (Note: written on a rate basis)
Q  W  
d t  sys

It is again noted that each of the conservation relations as previously written apply only to
fixed, constant mass systems. However, since most fluid problems of importance are for open
systems, we must transform each of these relations to an equivalent expression for a control
volume which includes the effect of mass entering and/or leaving the system. This is
accomplished with the Reynolds Transport Theorem.
The Reynolds transport theorem
We define a general, extensive property (an extensive property depends on the size or extent of
the system), B sys , where:
B sys     d V
sys
Bsys could be total mass, total energy, total momentum, etc. of a system.
and B sys per unit mass is defined as  or 
dB
dm
Thus,  is therefore the intensive equivalent of B sys
Applying a general control volume formulation to the time rate of change of B sys , we obtain
the following (see text for detailed development):
dB  
    dV +   e  e Ve dA e    i  i Vi dA i

dt sys t cv Ae Ai

   
System rate Rate of B Rate of B
of change change of leaving c.v. entering c.v.
of B B in c.v.
 
transient term convective terms
where: B is any conserved quantity, e.g. mass, linear momentum, moment of momentum,
or energy.
We will now apply this theorem to each of the basic conservation equations to develop their
equivalent open system, control volume forms.

Conservation of mass equation


For conservation of mass, we have that: B = m and   1
From the previous statement of conservation of mass and these definitions, Reynolds Transport
Theorem becomes:
dm  
   dV +   e Ve  dA e    i Vi  dA i 0

dt sys t cv Ae Ai

or

  dV +  e Ve  dA e    i Vi  dA i  0
 t cv Ae Ai

  
Rate of change Rate of mass Rate of mass
of mass in c.v., leaving c.v., entering c.v.,
  
= 0 for steady-state m e m i
This can be simplified to
d m
   m e   m i  0
d t cv
Note that the exit and inlet velocities, Ve and Vi, are the local components of fluid velocities at the
exit and inlet boundaries relative to an observer standing on the boundary.
Therefore, if the boundary is moving, the velocity is measured relative to the boundary motion.
The location and orientation of a coordinate system for the problem is not considered in
determining these velocities.
Also, the result of Ve  dA e and Vi  dA i is the product of the normal velocity component
times the flow area at the exit or inlet, e.g.
Ve,n dA e and i,n i V dA
Special Case: For incompressible flow with a uniform velocity over the flow area, the previous
integral expressions simplify to:
    V d A   AV
m
Example: Water at a velocity of 7 m/s exits a
stationary nozzle with D = 4 cm and is directed toward
a turning vane with θ = 40˚, assume steady-state.

Determine:
a. Velocity and flow rate entering the c.v.
b. Velocity and flow rate leaving the c.v.
Solution

a. Find V1 and m 
Recall that the mass flow velocity is the normal component of velocity measured relative to the
inlet or exit area.
Thus, relative to the nozzle, V(nozzle) = 7 m/s and since there is no relative motion of point 1
relative to the nozzle, we also have V1 = 7 m/s ans.
From the previous equation:
    V d A   AV = 998 kg/m3*7 m/s*π*.042/4
m
m 1 = 8.78 kg/s ans.
b. Find V2 and m 2
Determine the flow rate first.
Since the flow is steady state and no mass accumulates on the vane:

m 1 = m 2 , m 2 = 8.78 kg/s ans.


Now: m 2 = 8.78 kg/s = (ρ A V)2
Since ρ and A are constant, V2 = 7 m/s ans.
Key Point: For steady flow of a constant area, incompressible stream, the flow velocity and
total mass flow are the same at the inlet and exit, even though the direction changes.
Energy equation and Bernoulli equation
Energy Equation (Extended Bernoulli Equation)
For energy, we have that
1
B  E =  e  dV and   e  u  V2  g z
cv 2
From the previous statement of conservation of energy and these definitions, Reynolds Transport
Theorem becomes
d E 
Q  W     e  d V   ee eVe d Ae   ei i Vi d Ai
d t  sys  t cv cv cv
After extensive algebra and simplification (see text for detailed development), we obtain
P1  P2 V22  V12
  z 2  z1  h f,1-2  hs
g 2g
    
Press. drop Press. drop Press. drop Press. drop Press. drop
from 1 – 2, due to due to due to due to
in the flow acceleration elevation frictional mechanical
direction of the fluid change head loss work by fluid

Note: this formulation must be written in the flow direction from 1 - 2 to be consistent with
the sign of the mechanical work term and so that hf,1-2 is always a positive term. Also note
the following:
 The points 1 and 2 must be specific points along the flow path,
 Each term has units of linear dimension, e.g. ft. or meters and z2 – z1 is positive for z2
above z1.
 The term hf,1-2 is always positive when written in the flow direction and for internal,
pipe flow includes pipe or duct friction losses and fitting or piping component
(valves, elbows, etc.) losses.
 The term hs is negative for pumps and fans, - hp ( i.e. pumps increase the pressure in
the flow direction) and positive for turbines, + ht (turbines decrease the pressure in the
flow direction).
 For pumps:
ws
hp  Where: ws = the useful work per unit mass to the fluid
g
Therefore: ws = g hp and W f  m ws  ρ Q g hp

Where: W f = the useful power delivered to the fluid

W f
and: W p  and p is the pump efficiency
p
The Bernoulli Equation
The Bernoulli equation is an equation that is closely related to the energy equation and is useful
in the analysis of many flows. The most common application is for steady, incompressible,
frictionless flow between two points along a stream line. For these conditions, Bernoulli’s
equation becomes
P1 V12 P2 V22
  g z1    g z 2  const
 2  2
where the constant is the same along a specified streamline. Different streamlines may have
different Bernoulli constants. Bernoulli’s equation can also be written as
P2  P1 V22  V12
  gz 2  z 1   0
 2
Students must be careful not to misuse Bernoulli’s equation. In particular, do not use the
Bernoulli equation for flows with any or all of the following:

1. friction between points 1 and 2


2. shaft work between points 1 and 2
3. heat transfer between points 1 and 2
4. significant compressibility effects between 1 and 2
Solved Problems
1. At low velocities the flow through a long circular tube, i.e. pipe, has a parabolic velocity
distribution (actually paraboloid of revolution).

  r 2 
u  u max 1    
 R 
 
i.e., centerline velocity

a) find Q and V
Q   V  ndA   udA
A A
2 R
 udA    u (r )rddr
A 0 0
R
= 2  u (r )rdr
0

dA = 2rdr

2
u = u(r) and not    d  2
0
R   r 2  1
Q = 2  u max 1    rdr u max R 2

=
0  R  2

1
V  u max
2
2. Water flowing through an 8-cm-diameter pipe enters a porous section, as in Fig. #, which allows
a uniform radial velocity Vw through the wall surfaces for a distance of 1.2 m. If the entrance
average velocity V1 is 12 m/s, find the exit velocity V2 if (a) Vw  15 cm/s out of the pipe walls;
(b) vw  10 cm/s into the pipe. (c) What value of vw will make V2  9 m/s?

Fig.#
Solution: (a) For a suction velocity of vw  0.15 m/s, and a cylindrical suction surface area,
Aw  2π (0.04) (1.2)  0.3016 m2
Q1  Qw  Q2
(12)(π )(0.08 )/4  (0.15)(0.3016)  V2 (π )(0.082 )/4 V2  3 m/s Ans. (a)
2

(b) For a smaller wall velocity, vw  0.10 m/s,


(12)(π )(0.082 )/4  (0.10)(0.3016)  V2 (π )(0.082 )/4 V2  6 m/s Ans. (b)

(c) Setting the outflow V2 to 9 m/s, the wall suction velocity is,
(12)(π )(0.082 )/4  (Vw )(0.3016)  (9)(π )(0.082 )/4
Vw = 0.05m/s = 5cm/s out.

3. The pipe flow in Fig.# fills a cylindrical tank as shown. At time t  0, the water depth in the
tank is 30 cm. Estimate the time required to fill the remainder of the tank.

Fig. #
Solution: For a control volume enclosing the tank and the portion of the pipe below the tank,
d
ρ dv + mout - min  0
dt
ρπ R 2 dh  (ρ AV ) − (ρ AV )  0
dt out In
2
dh  4 [998 π (0.12 )(2.5 − 1.9)]  0.0153 m/s,
dt 998(π )(0.752 ) 4
∆t  0.7/0.0153  46 s Ans.
3
.4. Water at 20C flows steadily at
40 kg/s through the nozzle in Fig. P#. If D1  18 cm
and D2  5 cm, compute the average velocity, in m/s,
at (a) section 1 and

(b) section 2.

Fig. #
Solution: The flow is incompressible, hence the volume flow Q  (40 kg/s)/(998 kg/m3) 
0.0401 m3/s. The average velocities are computed from Q and the two diameters:

V Q  0.0401  1.58 m/s Ans. (a)


1
A1 (π /4)(0.18)2

V Q  0.0401  20.4 m/s Ans. (b)


2 2
A2 (π /4)(0.05)
5. An incompressible fluid flows past an
impermeable flat plate, as in Fig. #, with a uniform
inlet profile u  Uo and a cubic polynomial exit
profile

3η − η 3 y
u ≈ Uo where η 
2 δ Fig. #
Compute the volume flow Q across the top surface of the control volume.
Solution: For the given control volume and incompressible flow, we obtain

5 3
 Q  Uo bδ − Uobδ , solve for Q  U obδ Ans.
8 8
6. Gasoline at 20°C is pumped through a smooth 12-cm-diameter pipe 10 km long, at a flow rate
of 75 m3/h (330 gal/min). The inlet is fed by a pump at an absolute pressure of 24 atm. The exit
is at standard atmospheric pressure and is 150 m higher. Estimate the frictional head loss hf, and
compare it to the velocity head of the flow V2/(2g). (These numbers are quite realistic for
liquid flow through long pipelines.)
For gasoline at 20°C, density, ρ = 680 kg/m3 or γ = (680)(9.81)=6670 N/m3
There is no shaft work;

7. A hydroelectric power plant (Fig.#) takes in 30 m3/s of water through its turbine and
discharges it to the atmosphere at V2 = 2 m/s. The head loss in the turbine and penstock system
is hf =20 m. Assuming turbulent flow,α =1.06, estimate the power in MW extracted by the
turbine.
Chapter three
Differential Relations for A Fluid Flow
Acceleration field

Velocity is an important basic parameter governing a flow field. Other field variables such as the
pressure and temperature are all influenced by the velocity of the fluid flow. In general, velocity
is a function of both the location and time. The velocity vector can be expressed in Cartesian
coordinates as

V = V(x,y,z,t)
= u(x,y,z,t) i + v(x,y,z,t)j + w (x,y,z,t) k

where the velocity components (u, v and w) are functions of both position and time. That is, u =
u(x, y, z, t), v = v(x, y, z, t) and w = w(x, y, z, t). This velocity description is called the velocity
field since it describes the velocity of all points, Eulerian viewpoint, in a given volume.

For a single particle, Lagrangian viewpoint, the velocity is derived from the changing position
vector, or

V = dr/dt = = dx/dt i + dy/dt j + dz/dt k

where r is the position vector (r = xi + yj + zk). Notice, the velocity is only a function of time
since it is only tracking a single particle.

A general expression of the flow field velocity vector is given by

V (r ,t)  iˆ ux, y, z, t   ˆj vx, y, z, t   kˆ wx, y, z, t 


One of two reference frames can be used to specify the flow field characteristics
Eulerian – the coordinates are fixed and we observe the flow field
characteristics as it passes by the fixed coordinates.
Lagrangian - the coordinates move through the flow field following individual
particles in the flow.
Since the primary equation used in specifying the flow field velocity is based on Newton’s second law,
the acceleration vector is an important solution parameter. In cartesian coordinates, this is expressed as

d V  V   V V  V   V
a   u v w    V  V
dt  t   x y  z   t
Total local convective
The acceleration vector is expressed in terms of three types of derivatives:
Total acceleration = total derivative of velocity vector
= local derivative + convective derivative of velocity vector
Example; Given the eulerian velocity-vector field

V  3 t iˆ  x z ˆj  t y 2 kˆ
find the acceleration of the particle.
For the given velocity vector, the individual components are
u = 3t v= xz w = t y2
Evaluating the individual components for the acceleration vector, we obtain

V
= 3 i + 0 j + y2 k
t
V V V
= zj = 2tyk = xj
x y z
Substituting, we obtain

dV
= ( 3 i + y2 k) + (3 t) (z j) + (x z) (2 t y k) + (t y2) (x j)
dt
After collecting terms, we have

dV
= 3 i + (3 t z + t x y2) j + ( 2 x y z t + y2) k ans.
dt
Boundary condition, Stream function, Vorticity and Irrotationality

Boundary Conditions for the Basic Equations


In vector form, the three basic governing equations are written as

Continuity:     V  0
t
dV
Momentum:    g   P   i j
dt
du
Energy:   P  V   k  T   
dt
We have three equations and five unknowns: , V, P, u, and T; and thus need two additional equations.
These would be the equations of state describing the variation of density and internal energy as functions
of P and T, i.e.
(P,T) and u = u (P,T)
Two common assumptions providing this information are either:
1. Ideal gas:  P/RT and du = Cv dT
2. Incompressible fluid:  = constant and du = C dT
Time and Spatial Boundary Conditions
Time Boundary Conditions: If the flow is unsteady, the variation of each of the variables (, V, P, u,
and T) must be specified initially, t = 0, as functions of spatial coordinates, e.g. x,y,z.
Spatial Boundary Conditions: The most common spatial boundary conditions are those specified at
a fluid – surface boundary. This typically takes the form of assuming equilibrium (e.g. no slip
condition – no property jump) between the fluid and the surface at the boundary.
This takes the form:
Vfluid = Vwall Tfluid = Twall
Note that for porous surfaces with mass injection, the wall velocity will be equal to the injection velocity
at the surface. A second common spatial boundary condition is to specify the values of V, P, and T at
any flow inlet or exit.
Example; For steady incompressible laminar flow through a long tube, the velocity distribution is given
by

 r 2 
v z  U 1  2  vr  0 v  0
 R 
where U is the maximum or centerline velocity and R is the tube radius. If the wall temperature is
constant at Tw and the temperature, T = T(r) only, find T(r) for this flow. For the given conditions, the
energy equation reduces to
2
d T k d  d T  d v 
 Cv v r  r    z 
dr r d r  d r   d r 
Substituting for vz and realizing the vr = 0, we obtain
2
k d  d T  d v z  4U 2  r 2
r        
r d r  d r   d r  R4
Multiply by r/k and integrate to obtain

dT  U 2 r3
  C1
dr k R4
Integrate a second time to obtain
 U 2 r4
T  C1 ln r  C2
4 k R4
Since the term, ln r, approaches infinity as r approaches 0, C1 = 0.
Applying the wall boundary condition, T = Tw at r = R, we obtain for C2

 U2
C2  Tw 
4k
The final solution then becomes
 U 2  r 4 
T r   Tw  
1  
4 k  R4 
The Stream Function
The necessity to obtain solutions for multiple variables in multiple governing equations presents an
obvious mathematical challenge. However, the stream function, , allows the continuity equation to be
eliminated and the momentum equation solved directly for the single variable, . The use of the stream
function works for cases when the continuity equation can be reduced to only two terms.
Steady, incompressible, plane, two-dimensional flow represents one of the simplest types of flow
of practical importance. By plane, two-dimensional flow we mean that there are only two velocity
components, such as u and v, when the flow is considered to be in the x–y plane. For this flow the
continuity equation reduces to

u v
 0
x y
We still have two variables, u and v, to deal with, but they must be related in a special way as
indicated. This equation suggests that if we define a function ψ(x, y), called the stream function,
which relates the velocities as

 
u , v
y x
then the continuity equation is identically satisfied:

         
2 2

      0
x  y  y  x  xy xy

Velocity and velocity components along a streamline


Another particular advantage of using the stream function is related to the fact that lines along
which ψ is constant are streamlines. The change in the value of ψ as we move from one point (x,
y) to a nearby point (x + dx, y + dy) along a line of constant ψ is given by the relationship:
 
d  dx  dy  vdx  udy  0
x y
and, therefore, along a line of constant ψ

dy v

dx u

The flow between two streamlines


The actual numerical value associated with a particular streamline is not of particular significance,
but the change in the value of ψ is related to the volume rate of flow. Let dq represent the volume
rate of flow (per unit width perpendicular to the x–y plane) passing between the two streamlines.
 
dq  udy  vdx  dx  dy  d
x y
Thus, the volume rate of flow, q, between two streamlines such as ψ1 and ψ2, can be determined
by integrating to yield:
2
q   d   2  1
1

In cylindrical coordinates the continuity equation for incompressible, plane, two-dimensional flow
reduces to

1   rvr  1 v
 0
r r r 
and the velocity components, vr and vθ, can be related to the stream function, ψ(r, θ), through the
equations;

1  
vr  , v  
r  r
For example, for 2-D, incompressible flow, continuity becomes
u  v
 0
x y
Defining the velocity components to be
 
u and v 
y x
which when substituted into the continuity equation yields

        
      0
 x   y   y   x 
and continuity is automatically satisfied.
Steady Plane Compressible Flow
In like manner, for steady, 2-D, compressible flow, the continuity equation is

 u    v   0
x y
For this problem, the stream function can be defined such that
 
u  and v
y x
As before, lines of constant stream function are streamlines for the flow, but the change in stream
function is now related to the local mass flow rate by
2 2
m 1 2    V  n  d A   d   2  1
1 1
Vorticity and Irrotationality
The concepts of vorticity and irrotationality are very useful in analyzing many fluid problems. The
analysis starts with the concept of angular velocity in a flow field.
Consider three points, A, B, & C,
initially perpendicular at time t, that
then move and deform to have the
position and orientation at t + dt.
The lines AB and BC have both
changed length and incurred angular
rotation d and d relative to their
initial positions.

Fig.# Angular velocity and strain rate of two fluid lines


deforming in the x-y plane
We define the angular velocity z about the z axis as the average rate of counter-clockwise turning of
the two lines expressed as
1 d  d  
 z    
2  d t d t 
Applying the geometric properties of the deformation shown in Fig. # and taking the limit as
change in t approaches 0, we obtain

1 d v d u 
 z    
2 d x d y 
In like manner, the angular velocities about the remaining two axes are

1 d w d v  1 d u d w


 x      y    
2  d y d z  2 d z d x 
From vector calculus, the angular velocity can be expressed as a vector with the form
 = i  x + j  y + k z = 1/2 the curl of the velocity vector, e.g.

i j k
1 1   
  curl V 
2 2 x y z
u v w
The factor of 2 is eliminated by defining the vorticity, , as follows:  = 2  = curl V
Frictionless Irrotational Flows
When a flow is both frictionless and irrotational, the momentum equation reduces to Euler’s equation
dV
given previously by g - P = 
dt
As shown in the text, this can be integrated along the path, ds, of a streamline through the flow to obtain
V
 V2  V12   gz 2  z1  0
2 2
dP 1 2
 ds  
1 t 1  2
For steady, incompressible flow this reduces to
P 1
 V2  gz  constant along a streamline
 2
which is Bernoulli’s equation for steady incompressible flow
Velocity Potential
The condition of irrotationalty allows the development of a scalar function, , similar to the stream
function as follows:
If xV  0 then V = 
where  = (x,y,z,t) and is called the velocity potential function. The individual velocity components
are obtained from function with;
  
u v w
x y z
Orthogonality
For irrotational and 2-D flow, both  and  exist except at stagnation point and the streamlines and
potential lines are perpendicular throughout the flow regime. Thus for incompressible flow in the x-y
plane we have
    
u  and v 
y x x y
Examples of Plane Potential Flow
The following section shows several examples of the use of the stream function and velocity potential to
describe specific flow situations, in particular three plane potential flows as shown.

Fig. Three plane potential flows


Uniform Stream in the x Direction
For a uniform velocity flow represented by V = U i , both a velocity potential and stream function can be
defined. The flow field is shown in Fig. a and  and  are found as follows:
   
u U   and v 0 
x y y x
Integrating these expressions and setting the constants of integration equal to zero we obtain
 U y and  U x
Line Source or Sink at the Origin
In potential flow, a source is a point in a flow field which provides a point source of flow. For the case
where the z-axis is treated as a line source with fluid supplied uniformly along its length, a stream function
and velocity potential can be defined in 2-D polar coordinates starting with
Q m 1    1 
vr     v  0   
2  r b r r   r  r r 
Integrating and discarding the constants of integration as before, we obtain the equations for and
for a simple, radial flow, line source, or sink:
=m and  = m ln r
The flow field is shown schematically in Fig. b for a uniform source flow.
Line Irrotational Vortex

A 2-D line vortex is a purely circulating, steady motion with v = f(r) only and vr = 0. This flow is shown
schematically in Fig. c. For irrotational flow, this is referred to as a free vortex and the stream function
and velocity potential function may be found from
1   K   1
vr  0   and v   
r   r r  r r 
where K is a constant. Again, integrating, we obtain for  and 
 = - K ln r and  = K 
The strength of the vortex is defined by the constant K.

Worked Out Examples


Example 1. (Use of Continuity equation and Acceleration):
  1 
Consider the velocity field V  Axy  i  Ay 2  j in the xy plane, where A  0.25m 1  s 1 , and
2
the coordinates are measured in meters. Is this a possible incompressible flow field?
Calculate the acceleration of a fluid particle at point ( x, y )  (2,1) .

1. Statement of the Problem


a) Given
  1 
 Velocity field: V  Axy  i  Ay 2  j , where A  0.25m 1  s 1 .
2
b) Find
 Whether the velocity field is a possible incompressible flow field or not.
 Acceleration of a fluid particle at point ( x, y )  (2,1) .
2. System Diagram
It is not necessary for this problem.
3. Assumptions
 Steady state condition
 2 - D flow field problem
4. Governing Equations
 u v w
 Continuity Equation:    0
t x y z
u v
Incompressible 2 - D continuity equation:  0
x y
    
 DV V V V V
 Acceleration: a   u v w
Dt t x y z
2 - D problem:
Du u u u u
ax   u v w
Dt t x y z

Dv v v v v
ay   u v w
Dt t x y z

Since u  u ( x, y ) & v  v( x, y ) ,

u u
ax  u v
x y

v v
ay  u v
x y

5. Detailed Solution
  1 
Velocity field V  Axy  i  Ay 2  j shows the components to be:
2
u  u ( x, y )  Axy

1 2
v  v ( x, y )   Ay
2
The continuity equation, incompressible version of continuity equation in this case, must be
satisfied to have a valid incompressible flow field.
So, check the incompressible continuity equation:
u v    1 
  Axy   Ay 2   Ay   Ay   0  Continuity equation is satisfied.
x y x y  2 
Therefore, it can be said that there exists a possible incompressible flow field.
Acceleration is:
u u   1    1 
ax  u v  Axy  Axy   Ay 2   Axy  Axy  Ay    Ay 2   Ax 
x y x  2  y  2 

 1  1
 a x  A 2 xy 2   A 2 xy 2   A 2 xy 2
 2  2

v v   1   1    1   1 
ay  u v  Axy  Ay 2    Ay 2    Ay 2   0   Ay 2    Ay  
x y x  2   2  y  2   2 
1 2 3
ay  A y
2
Therefore, at ( x, y )  (2,1) , the acceleration is:

ax 
1 2 2 1
2 2

A xy  0.25m 1  s 1   2m  1m
2 2
 0.0625m / s 2

ay 
2 2
 2

A y  0.25m 1  s 1  1m   0.03125m / s 2
1 2 3 1 3

    
 a  a x  i  a y  j  (0.0625m / s 2 )  i  (0.03125m / s 2 )  j

6. Critical Assessment
Incompressible 2 - D continuity equation has been satisfied; therefore, the fluid flow is
incompressible in this flow field.
Example 3. (Test for incompressibility and irrotationality):
   
A flow is represented by the velocity field V  10x  i  10 y  j  30  k . Determine if the field is
(a) a possible incompressible flow and (b) irrotational.
1. Statement of the Problem
a) Given
   
 Velocity field: V  10x  i  10 y  j  30  k
b) Find
If the field is
 a possible incompressible flow
 irrotational
2. System Diagram
It is not necessary for this problem.
3. Assumptions
 Steady state condition
4. Governing Equations
 u v w
 Continuity Equation:    0
t x y z
u v w
Incompressible continuity equation:   0
x y z
     w v   u w   v u 
 Vorticity:   2    V  i     j     k   
 y z   z x   x y 
5. Detailed Solution
   
The velocity field is V  10x  i  10 y  j  30  k .

This shows the components to be:


u  10x
v  10 y

w  30
The continuity equation, incompressible version in this case, must be satisfied to have a
possible incompressible flow field. Checking the incompressible continuity equation:
u v w   
   10x    10 y   30  10   10  0  0
x y z x y z

It can be said that the flow field is valid (possible) since the incompressible continuity
equation is satisfied.
The next is to check the vorticity to see if the flow is rotational or irrotational.
  w v   u w   v u 
  i     j     k   
 y z   z x   x y 
w v  
x component:   30   10 y   0
y z y z

u w  
y component:   10 x   30  0
z x z x
v u  
z component:    10 y   10x   0
x y x y
    
Therefore,   i 0  j 0  k 0  0  This shows there is no rotation in the flow.

The flow is irrotational.


6. Critical Assessment
Incompressible continuity equation has been satisfied; therefore, the fluid flow is
incompressible in this flow field. Also the fluid has no rotation, as it satisfies the zero
vorticity requirement.

Chapter Four
Dimensional Analysis and Similitude
Dimensional Analysis
The analytically derived equations in engineering applications are correct for any system of units
and consequently each group of terms in the equation must have the same dimensional
representation. This is the law of dimensional homogeneity.

In many instances, the variables involved in physical phenomena are known, while the
relationship among the variables is not known. Such a relationship can be formulated between a
set of dimensionless groups of variables and the groups numbering less than the variables. This
procedure is called dimensional analysis. This procedure requires less experimentation and the
nature of experimentation is considerably simplified.

Buckingham pi theorem

 It states that if an equation involving k variables is dimensionally homogeneous, it can be


reduced to a relationship among  k  r  independent dimensionless products, where r is
the minimum number of reference dimensions required to describe the variable.
 The dimensionless products are frequently referred to as pi terms and the theorem is named
accordingly after famous scientist Edgar Buckingham (1867-1940). It is based on the idea
of dimensional homogeneity.
 Mathematically, if a physically meaningful equation involving k variables is assumed,
y  f  x1 , x2 .........., xk  (5)
such that the dimensions of the variables on the left side of the equation are equal to the
dimensions of any term on the right side of equation, then, it is possible to rearrange the
above equation into a set of dimensionless products (pi terms), so that
1     2 ,  3 ..........,  k r  (6)
where    2 ,  3 ..........,  k r  is a function of  2 through  k  r .
 The required number of pi terms is less than the number of original variables by r , where
r is determined by the minimum number of reference dimensions required to describe the
original list of variables. These reference dimensions are usually the basic dimensions
M , L and T (Mass, Length and Time) .
4.3. Non dimensionalisation of basic equations
Determination of pi Terms
Several methods can be used to form dimensionless products or pi terms that arise in dimensional
analysis. But, there is a systematic procedure called method of repeating variables that allows in
deciding the dimensionless and independent pi terms. For a given problem, following distinct steps
should be followed.
Step I: List out all the variables that are involved in the problem.
 The ‘variable’ is any quantity including dimensional and non-dimensional constants in a
physical situation under investigation. Typically, these variables are those that are
necessary to describe the “geometry” of the system (diameter, length etc.), to define fluid
properties (density, viscosity etc.) and to indicate the external effects influencing the
system (force, pressure etc.).
 All the variables must be independent in nature so as to minimize the number of variables
required to describe the system.
Step II: Express each variable in terms of basic dimensions.
Typically, for fluid mechanics problems, the basic dimensions will be either M , L and T or
F , L and T . Dimensionally, these two sets are related through Newton’s second law  F  m.a 
so that F  MLT 2 e.g.   ML3 or   FL4T 2 . It should be noted that these basic dimensions
should not be mixed.
Step III: Decide the required number of pi terms.
It can be determined by means of Buckingham pi theorem which indicates that the number of pi
terms is equal to  k  r  , where k is the number of variables in the problem (determined from
Step I) and r is the number of reference dimensions required to describe these variables
(determined from Step II).
Step IV: Select the number of repeating variables.
 Amongst the original list of variables, select those variables that can be combined to form
pi terms.
 The required number of repeating variables is equal to the number of reference dimensions.
 Each repeating variable must be dimensionally independent of the others i.e. they cannot
themselves be combined to form dimensionless product.
 Since there is a possibility of repeating variables to appear in more than one pi term, so
dependent variables should not be chosen as one of the repeating variable.
Step V: Formation of pi terms
Essentially, the pi terms are formed by multiplying one of the non-repeating variables by the
product of the repeating variables each raised to an exponent that will make the combination
dimensionless. It usually takes the form of xi x1a x2b x3c where the exponents a , b and c are
determined so that the combination is dimensionless.
Step VI: Repeat the Step V for each of the remaining non-repeating variables. The resulting set of
pi terms will correspond to the required number obtained from Step III.
Step VII: Checking of pi terms
Make sure that all the pi terms must be dimensionless. It can be checked by simply substituting
the basic dimension ( M , L and T ) of the variables into the pi terms.
Step VIII: Final form of relationship among pi terms
Typically, the final form among the pi terms can be written in the form of Eq. (6) where 1 would
contain the dependent variable in the numerator. The actual functional relationship among pi terms
is determined from experiment.
Illustration of Pi Theorem

In order to illustrate various steps in the pi theorem, following examples are considered.
Example 1 (Pressure drop in a pipe flow)
Consider a steady flow of an incompressible Newtonian fluid through a long, smooth walled,
horizontal circular pipe. It is desired to measure the pressure drop per unit length of the pipe
without the use of experimental data.
 According to Step I, all the pertinent variables involved in the experimentation of pressure
drop per unit length of the pipe, must be noted and should be expressed in the form,
pl  f  D,  ,  ,V  (7)
where D is the pipe diameter,  is the fluid density,  is the viscosity of the fluid and V
is the mean velocity at which the fluid is flowing through the pipe.
 Next step is to express all the variables in terms of basic dimensions i.e. M , L and T . It
then follows that
pl  ML2T 2 , D  L ,   ML3 ,   ML1T 1 and V  LT 1 (8)
 Third step is to apply pi theorem to decide the number of pi terms required. Since there are
five variables (including the dependent variable p ) and three reference dimensions, so
k  5 and r  3 . So, only two  k  r  pi terms are required.
 The repeating variables to be used to form pi terms (Step IV) need to be selected from the
list D,  ,  and V . It is to be noted that the dependent variable should not be used as one
of the repeating variable. Since, there are three reference dimensions involved, so we need
to select three repeating variable. These repeating variables should be dimensionally
independent i.e. dimensionless product cannot be formed from this set. In this case,
D,  and V are chosen as the repeating variables.
 Now, first pi term is formed between the dependent variable and the repeating variables. It
is written as,
1  pl Da V b  c (9)
Since this combination need to be dimensionless, it follows that
 ML   L   LT   ML 
b c
2
T 2 1 3
a
 M 0 L0T 0 (10)
The exponents a, b and c must be determined by equating the exponents for each of the
terms M , L and T i.e.
1  c  0 (For M )
2  a  b  3c  0 (For L) (11)
2  b  0 (For T )
The solution of this algebraic equations gives a  1; b  2; c  1 . Therefore,
p D
1  l 2 (12)
V
 The process is repeated for remaining non-repeating variables with other additional
variable    so that,
2  .Dd .V e . f (13)
and
 ML T   L   LT   ML 
e f
1 1 d 1 3
 M 0 L0T 0 (14)
Equating the exponents,
1  f  0 (For M )
1  d  e  3 f  0 (For L) (15)
1  e  0 (For T )
The solution of this algebraic equation gives d  1; e  1; f  1 . Therefore,

2  (16)
VD
 Now, the correct numbers of pi terms are formed as determined in step 3. In order to make
sure about the dimensionality of pi terms, they are written as,
pl D  ML T   L 
2 2

1    M 0 L0T 0
 V 2  ML3  LT 1 2

2 


 ML T   L   M
1 1
0 0
L T0
 V D  ML  LT   L 
3 1

 Finally, the result of dimensional analysis is expressed among the pi terms as,
D pl   
  (17)
V 2
 V D 
Significance of non-dimensional numbers in fluid flows
Dimensional Numbers in Fluid Mechanics
Forces encountered in flowing fluids include those due to inertia, viscosity, pressure, gravity,
surface tension and compressibility. These forces can be written as
dV  dV 
Inertia force  m.a   V   V V   V L
2 2

dt  ds 
du
Viscous force   A   A  V L
dy
Pressure force   p  A   p  L2
Gravity force  m g  g  L3
Surface tension force   L
Compressibilityforce  Ev A  Ev L2 ; where Ev is the Bulk modulus
The ratio of any two forces will be dimensionless. Inertia forces are very important in fluid
mechanics problems. So, the ratio of the inertia force to each of the other forces listed above leads
to fundamental dimensionless groups. These are,
1. Reynolds number  Re  : It is defined as the ratio of inertia force to viscous force.
Mathematically,
VL VL
Re   (1)
 
where V is the velocity of the flow, L is the characteristics length,  ,  and  are the density,
dynamic viscosity and kinematic viscosity of the fluid respectively. If Re is very small, there is
an indication that the viscous forces are dominant compared to inertia forces. Such types of flows
are commonly referred to as “creeping/viscous flows”. Conversely, for large Re , viscous forces
are small compared to inertial effects and flow problems are characterized as inviscid analysis.
This number is also used to study the transition between the laminar and turbulent flow regimes.
2. Euler number  Eu  : In most of the aerodynamic model testing, the pressure data are usually
expressed mathematically as,
p
Eu  (2)
1
V 2

2
where p is the difference in local pressure and free stream pressure, V is the velocity of the
flow,  is the density of the fluid. The denominator in Eq. (2) is called “dynamic pressure”. Eu
is the ratio of pressure force to inertia force and it is also called as the pressure coefficient C p . In
the study of cavitations phenomena, similar expressions are used where p is the difference in
liquid stream pressure and liquid-vapour pressure. The dimensional parameter is called “cavitation
number”.
3. Froude number  Fr  : It is interpreted as the ratio of inertia force to gravity force.
Mathematically, it is written as,
V
Fr  (3)
g .L
where V is the velocity of the flow, L is the characteristics length descriptive of the flow field
and g is the acceleration due to gravity. This number is very much significant for flows with free
surface effects such as in case of open-channel flow. In such types of flows, the characteristics
length is the depth of water. Fr less than unity indicates sub-critical flow and values greater than
unity indicate super-critical flow. It is also used to study the flow of water around ships with
resulting wave motion.
4. Weber number We  : The ratio of the inertia force to surface tension force is called Weber
number. Mathematically,
V 2 L
We  (4)

where V is the velocity of the flow, L is the characteristics length descriptive of the flow field,
 is the density of the fluid and  is the surface tension force. This number is taken as a index
of droplet formation and flow of thin film liquids in which there is an interface between two fluids.
For We 1 , inertia force is dominant compared to surface tension force (e.g. flow of water in a
river).
5. Mach number  M a  : It is the key parameter that characterizes the compressibility effects in a
fluid flow and is defined as the ratio of inertia force to compressibility force. Mathematically,
V V V
Ma    (5)
c dp Ev
d 
where V is the velocity of the flow, c is the local sonic speed,  is the density of the fluid and
Ev is the bulk modulus. Sometimes the square of the Mach number is called “Cauchy number”
 Ca  i.e.
V 2
Ca  M a2  (6)
Ev
Both the numbers are predominantly used in problems in which fluid compressibility is important.
When the M a is relatively small (say, less than 0.3), the inertial forces induced by fluid motion
are sufficiently small to cause significant change in fluid density. So, the compressibility of the
fluid can be neglected. However, this number is most commonly used parameter in compressible
fluid flow problems, particularly in the field of gas dynamics and aerodynamics.
6. Strouhal number  St  : It is a dimensionless parameter that is likely to be important in unsteady,
oscillating flow problems in which the frequency of oscillation is  and is defined as,
L
St  (7)
V
where V is the velocity of the flow and L is the characteristics length descriptive of the flow field.
This number is the measure of the ratio of the inertial forces due to unsteadiness of the flow (local
acceleration) to inertia forces due to changes in velocity from point to point in the flow field
(convective acceleration). This type of unsteady flow develops when a fluid flows past a solid
body placed in the moving stream.
4.5 Similitude
Modeling and Similitude
 A “model” is a representation of a physical system used to predict the behavior of the
system in some desired respect. The physical system for which the predictions are to be
made is called “prototype”.
 Usually, a model is smaller than the prototype so as to conduct laboratory studies and it is
less expensive to construct and operate. However, in certain situations, models are larger
than the prototype e.g. study of the motion of blood cells whose sizes are of the order of
micrometers.
 “Similitude” in a general sense is the indication of a known relationship between a model
and prototype i.e. model tests must yield data that can be scaled to obtain the similar
parameters for the prototype.
Theory of models
A given problem can be described in terms of a set of pi terms by using the principles of
dimensional analysis as,
1     2 ,  3 ,.......... n  (8)
This equation applies to any system that is governed by same variables. So, if the behavior of a
particular prototype is described by Eq. (8), a similar relationship can be written for a model of
this type i.e.
1m     2 m , 3m ,.......... nm  (9)
The form of the function remains the same as long as the same phenomenon is involved in both
prototype and the model. Therefore, if the model is designed and operated under following
conditions,
 2m   2
 3m   3
. (10)
.
 nm   n
then, it follows that
1  1m (11)
Eq. (11) is the desired “prediction equation” and indicates that the measured value of 1m obtained
with the model will be equal to the corresponding 1 for the prototype as long as the other pi
terms are equal. These are called “model design conditions / similarity requirements / modeling
laws”.

Flow similarity
In order to achieve similarity between model and prototype behavior, all the corresponding pi
terms must be equated between model and prototype. So, the following conditions must be met to
ensure the similarity of the model and the prototype flows.
1. Geometric similarity: A model and prototype are geometric similar if and only if all body
dimensions in all three coordinates have the same linear-scale ratio. It requires that the model and
the prototype be of the same shape and that all the linear dimensions of the model be related to
corresponding dimensions of the prototype by a constant scale factor. Usually, one or more of
these pi terms will involve ratios of important lengths, which are purely geometrical in nature. The
geometric scaling may also extend to the finest features of the system such as surface roughness
or small perterbance that may influence the flow fields between model and prototype.
2. Kinematic similarity: The motions of two systems are kinematically similar if homogeneous
particles lie at homogeneous points at homogeneous times. In a specific sense, the velocities at
corresponding points are in the same direction and are related in magnitude by a constant scale
factor. This also requires that streamline patterns must be related by a constant scale factor. The
flows that are kinematically similar must be geometric similar because boundaries form the
bounding streamlines. The factors like compressibility or cavitations must be taken care of to
maintain the kinematic similarity.
3. Dynamic similarity: When two flows have force distributions such that identical types of forces
are parallel and are related in magnitude by a constant scale factor at all corresponding points, then
the flows are dynamic similar. For a model and prototype, the dynamic similarity exists, when
both of them have same length-scale ratio, time-scale ratio and force-scale (or mass-scale ratio).
 For compressible flows, the model and prototype Reynolds number, Mach number and
specific heat ratio are correspondingly equal.
 For incompressible flows,
With no free surface: model and prototype Reynolds number are equal.
With free surface: Reynolds number, Froude number, Weber number and Cavitation
numbers for model and prototype must match.
In order to have complete similarity between the model and prototype, all the similarity flow
conditions must be maintained.
This will automatically follow if all the important variables are included in the dimensional
analysis and if all the similarity requirements based on the resulting pi terms are satisfied.
Model scales
In a given problem, if there are two length variables l1 and l2 , the resulting requirement based on
the pi terms obtained from these variables is,
l1m l2 m
 (12)
l1 l2
This ratio is defined as the “length scale”. For true models, there will be only one length scale and
all lengths are fixed in accordance with this scale. There are other ‘model scales’ such as velocity
V     
scale  m  v  , density scale  m    , viscosity scale  m    etc. Each of these scales
V       
is defined for a given problem.
Distorted models
In order to achieve the complete dynamic similarity between geometrically similar flows, it is
necessary to duplicate the independent dimensionless groups so that dependent parameters can
also be duplicated (e.g. duplication of Reynolds number between a model and prototype is ensured
for dynamically similar flows).
In many model studies, dynamic similarity requires the duplication of several dimensionless
groups and it leads to incomplete similarity between model and the prototype. If one or more of
the similarity requirements are not met, e.g. in Eq. 10, if  2 m   2 , then it follows that Eq. 11 will
not be satisfied i.e. 1  1m . Models for which one or more of the similar requirements are not
satisfied, are called “distorted models”. For example, in the study of open-channel or free surface
 Vl   V 
flows, both Reynolds number   and Froude number   are involved.
    gl 
Then, Froude number similarity requires,
Vm V
 (13)
gmlm gl
If the model and prototype are operated in the same gravitational field, then the velocity scale
becomes,
Vm l
 m  l (14)
V l
Reynolds number similarity requires,
m .Vm .lm  .V .l
 (15)
m 
and the velocity scale is,
Vm m  l
 . . (16)
V  m lm
Since, the velocity scale must be equal to the square root of the length scale, it follows that
3
 m   m  m   lm  2 3
      l  2 (17)
     l 
Eq. (17) requires that both model and prototype to have different kinematics viscosity scale, if at
all both the requirements i.e. Eq. (13) and (15) are to be satisfied. But practically, it is almost
impossible to find a suitable model fluid for small length scale. In such cases, the systems are
designed on the basis of Froude number with different Reynolds number for the model and
prototype where Eq. (17) need not be satisfied. Such analysis will result a “distorted model”.
Hence, there are no general rules for handling distorted models and essentially each problem must
be considered on its own merits.
Example: Water flows through a large valve having 0.8m diameter at a rate 85m 3/s. It is to be
tested in a geometrically similar model of 10cm diameter with water as working fluid. Determine
the required flow rate in the model.
Solution: In order to ensure the dynamic similarity, the Reynolds number must match for the model
and the prototype i.e.
 Re model   Re prototype
Vm .Dm Vp .Dp

m p
Since, water is used as the working fluid for the model and the prototype, so  m   p . Hence,
Vm D p

V p Dm
The flow rate for the model and prototype can be related as,
Qm  Dp    4.Dm  Dm
2

  
Q p  Dm    4.D p2  D p
Hence,
Dm  0.1 
Qm  Qp .  85     10.6 m s
3

Dp  0.8 
Chapter Five
Boundary Layer Concept

5.1 Reynolds number and geometry concept

Viscous internal flows have the following major boundary layer characteristics:
* An entrance region where the boundary layer grows and dP/dx ≠ constant,
* A fully developed region where:
• The boundary layer fills the entire flow area.
• The velocity profiles, pressure gradient, and ρw are constant;
i.e. they are not equal to f(x),
• The flow is either laminar or turbulent over the entire length of the flow,
i.e. transition from laminar to turbulent is not considered.
However, viscous flow boundary layer characteristics for external flows are significantly
different as shown below for flow over a flat plate:
-+
U• laminar to edge of boundary layer
free stream turbulent
y transition

(x)
turbulent
x laminar
xcr

Fig. Schematic of boundary layer flow over a flat plate


For these conditions, we note the following characteristics:
• The boundary layer thickness δ(x) grows continuously from the start of the fluid-surface
contact, e.g. the leading edge. It is a function of x, not a constant.
• Velocity profiles and shear stress τ are f(x,y).
• The flow will generally be laminar starting from x = 0.
• The flow will undergo laminar-to-turbulent transition if the stream wise dimension is greater
than a distance xcr corresponding to the location of the transition Reynolds number Recr.

• Outside of the boundary layer region, free stream conditions exist where velocity gradients
and therefore viscous effects are typically negligible.
As it was for internal flows, the most important fluid flow parameter is the local Reynolds
number defined as
 U x Ux
Rex  
 
where
ρ = fluid density ν = fluid dynamic viscosity
μ = fluid kinematic viscosity U = characteristic flow velocity
x = characteristic flow dimension
It should be noted at this point that all external flow applications will not use a distance from the
leading edge x as the characteristic flow dimension. For example, for flow over a cylinder, the
diameter will be used as the characteristic dimension for the Reynolds number.
Transition from laminar to turbulent flow typically occurs at the local transition Reynolds
number, which for flat plate flows can be in the range of
500, 000  Re cr 3,000, 00
With xcr = the value of x where transition from laminar to turbulent flow occurs, the typical
value used for steady, incompressible flow over a flat plate is
 U xcr
Re cr   500, 000

Thus for flat plate flows for which:
x < xcr the flow is laminar
x ≥ xcr the flow is turbulent
The solution to boundary layer flows is obtained from the reduced “Navier – Stokes” equations,
i.e., Navier-Stokes equations for which boundary layer assumptions and approximations have
been applied.
The boundary layer thickness, the following key results are obtained for the laminar flat plate
boundary layer:

Local boundary layer thickness


 x  5x
Re x

Local skin friction coefficient: 0.664


C fx 
(defined below) Re x
Total drag coefficient for length L ( integration of τw dA 1.328
over the length of the plate, per unit area, divided by 0.5 CD 
ρU 2 ) Re x

 w x  FD / A
where by definition C fx  and CD 
1
2
 U 2

1
2
 U 2

With these results, we can determine local boundary layer thickness, local wall shear stress, and total
drag force for laminar flow over a flat plate.

Turbulent Flow Equations


While the previous analysis provides an excellent representation of laminar, flat plate boundary
layer flow, a similar analytical solution is not available for turbulent flow due to the complex
nature of the turbulent flow structure. However, experimental results are available to provide
equations for key flow field parameters.
A summary of the results for boundary layer thickness and local and average skin friction coefficient
for a laminar flat plate and a comparison with experimental results for a smooth, turbulent flat plate
are shown below.
Laminar Turbulent

0.16 x
 x  5x
 x  
Re x Re1/x 7
0.664 0.027
C fx  C fx 
Re x Re1/7
x

1.328 0.031 for turbulent flow over


CD  C D  1/7 entire plate, 0 – L, i.e.
Re L Re L assumes turbulent flow in
the laminar region

w
C fx  1
2
 U 2 local drag coefficient based on local wall
where shear stress (laminar or turbulent flow
region)

and
 
CD = total drag coefficient based F/ A 1 L
   w  x  w dx
1
CD  1  U
2
A
 U
on the integrated force over the 2 2
length 0 to L 2 0

A careful study of these results will show that, in general, boundary layer thickness grows faster
for turbulent flow, and wall shear and total friction drag are greater for turbulent flow than for
laminar flow given the same Reynolds number. It is noted that the expressions for turbulent flow
are valid only for a flat plate with a smooth surface. Expressions including the effects of surface
roughness are available in the text.

Combined Laminar and Turbulent Flow

U• laminar to edge of boundary layer


free stream turbulent
y transition

(x)
turbulent
x laminar
xcr

Flat plate with both laminar and turbulent flow sections


For conditions (as shown above) where the length of the plate is sufficiently long that we have
both laminar and turbulent sections:
* Local values for boundary layer thickness and wall shear stress for either the laminar or
turbulent sections are obtained from the expressions for δ(x) and Cfx for laminar or
turbulent flow as appropriate for the given region.

* The result for average drag coefficient CD and thus total frictional force over the
combined laminar and turbulent portions of the plate is given by
(assuming a transition Re of 500,000)

0.031 0.031
CD  
Re1/7  
1/ 7
L 5x10 6
* Calculations assuming only turbulent flow can typically be made for two cases:
1. when some physical situation (a trip wire) has caused the flow to be turbulent from
the leading edge or
2. if the total length L of the plate is much greater than the length xcr of the laminar
section such that the total length of plate can be considered turbulent from x = 0 to
L. Note that this will over predict the friction drag force since turbulent drag is
greater than laminar.
With these results, a detailed analysis can be obtained for laminar and/or turbulent flow over flat
plates and surfaces that can be approximated as a flat plate.
Figure 7.6 in the text shows results for laminar, turbulent and transition regimes. Equations 7.48a
& b can be used to calculate skin friction and drag results for the fully rough regime.

 x 2.5
c f  2.87 1.58 log (7.48a)
  
 L 2.5
CD  1.89  1.62log (7.48b)
  
Equations 7.49a & b can be used to calculate total CD for combined laminar and turbulent flow
for transition Reynolds numbers of 5x105 and 3x106 respectively.

0.031 1440
CD   Re trans  5x10
5
Re1/7
L ReL
0.031 8700
C D  1/7  Re trans  3x10
6
Re L ReL
Example:
Water flows over a sharp flat plate 2.55 m long, 1 m wide, with U = 2m/s. Estimate the error in
FD if it is assumed that the entire plate is turbulent.
Water: ρ = 1000 kg/m3 ν = 1.02 E- 6m2/s

U L 2 m / s* 2.55 m
Reynolds number: Re L    5E 6  500, 000
 1.02E  6 m 2 / s
with Re cr  500,000  x cr  0.255m ( or 10% laminar)

a. Assume that the entire plate is turbulent

0.031 0.031
CD    0.003423
Re1/7
L 5x106 1/ 7

kg 2 m2
FD  0.5  U 2 CD A  0.5 1000 3  2 2  0.003423  2.55 m2
m s
This should be high since we have assumed that the entire
FD 17.46 N plate is turbulent and the first 10% is actually laminar.

b. Now consider the actual combined laminar and turbulent flow:

0.031 1440 0.031 1440


CD     6  0.003135
Re1/7
L ReL 5x10 6 1/7
 5x10 
Note that the CD has decreased when both the laminar and turbulent sections are considered.

kg 2 m2
FD  0.5  U 2 CD A  0.5 1000 3  2 2  0.003135  2.55 m2
m s
FD 15.99 N { Lower than the fully turbulent value}

17.46 15.99
Error  100  9.2% high
15.99
Thus, the effect of neglecting the laminar region and assuming the entire plate is turbulent is as
expected.
Question: Since xcr = 0.255 m, what would your answers represent if you had calculated the
Re, CD, and FD using x = xcr = 0.255 m?
Answer: You would have the value of the transition Reynolds number and the drag
coefficient and drag force over the laminar portion of the plate (assuming you used
laminar equations).
Pipe flow & Losses in pipe flow
Moody Chart
The fundamental difference between laminar and turbulent flow is that the shear
stress for laminar flow depends on the viscosity of the fluid whereas in case of
turbulent flow, it is the function of density of the fluid. In general, the pressure drop
p , for steady, incompressible turbulent flow in a horizontal round pipe of diameter
D can be written in the functional form as,
Pturb  f V , D, l ,  ,  ,   (15)
where V is the average velocity, l is the length of the pipe and  is a measure of the
roughness of the pipe wall. Similar expression can also be written for the case of
laminar flow in which the  term will be absent because the pressure drop in laminar
flow is found to be independent of pipe roughness i.e.
Plam  F V , D, l ,  ,   (16)
By dimensional analysis treatment, we can found that
 
 Pl    .V .D l 
1   , 
  .V 
2   D
2 lam
(17)
 
 Pl    .V .D l  
1    , , 
  .V 2    D D
2 turb
The only difference between two expressions in Eq. (17) is that the term  D  ,
which is known as the “relative roughness”. In commercially available pipes, the
roughness is not uniform; so it is correlated with pipe diameter and the contribution
 D  forms a significant value in friction factor calculation. From tests with
commercial pipes, Moody gave the values for average pipe roughness listed in Table
1.
Table 1: Average values of roughness for commercial pipes (Table 8.1; Ref. 1)
Material  (mm)
Riveted steel 0.9 – 9
Concrete 0.3 – 3
Wood 0.18 – 0.9
Cast iron 0.26
Galvanized 0.15
iron
Commercial 0.045
steel
Plastic and 0 (smooth
glass pipes)

Now Eq. (17) can be simplified with reasonable assumption that the pressure drop
is proportional to pipe length. It can be done only when,
P l   
   Re ,  (18)
1
 .V 2 D  D
2
It can be rewritten as,
 l    .V 
2
P  f     (19)
 D  2 
where f is known as “friction factor” and is defined by,
  
f    Re ,  (20)
 D
Now, recalling the energy equation for a steady incompressible flow,
p1 V12 p V2
  z1  2  2  z2  hL (21)
 2g  2g
where hL is the head loss between two sections. With assumption of horizontal
 z1  z2  constant diameter pipe  D1  D2 or V1  V2  with fully developed flow,
p  p1  p2   hL   ghL (22)
From Eqs. (19) and (22), we can determine head loss as,
 l  V 
2
hL  f     (23)
 D   2g 
This is known as Darcy-Weisbach equation and is valid for fully developed, steady,
incompressible horizontal pipe flow. If the flow is laminar, the friction factor will

be independent on   and simply,
D
64
f  (24)
Re
The functional dependence of friction factor on the Reynolds number and relative
roughness is rather complex. It is found from exhaustive set of experiments and is
usually presented in the form of curve-fitting formula/data. The most common
graphical representation of friction factor dependence on surface roughness and
Reynolds number is shown in “Moody Chart” (Fig. 4). This chart is valid universally
for all steady, fully developed, incompressible flows.
The following inferences may be made from Moody chart (Fig. 4).
 For laminar flows ( Re  2100 ), f   64 Re  and is independent of surface
roughness
 At very high Reynolds number  Re  4000  , the flow becomes completely
turbulent (wholly turbulent flow) and is independent of Reynolds number. In this
case, the laminar sub-layer is so thin that the surface roughness completely
dominates the character of flow near the wall. The pressure drop responsible for
turbulent shear stress is inertia dominated rather than viscous dominated as found in
case of laminar viscous sub-layer. Hence, the friction factor is given by, f    D 
 The friction factor at moderate Reynolds number  2100  Re  4000  is indeed
dependent on both Reynolds number and relative roughness.
 Even for smooth pipes   0  , the friction factor is not zero i.e. there is always
head loss, no matter how smooth the pipe surface is. There is always some
microscopic surface roughness that produces no-slip behavior  thus f  0  on the
molecular level. Such pipes are called “hydraulically smooth”.

Fig. 4: The Moody chart: - friction factor as a function of Reynolds number and
surface roughness for round pipes (Pg. 477; Reference 1).
It must be noted that Moody chart covers extremely wide range of flow
parameters i.e. diameter of the pipes  D  , fluid density    , viscosity    and
velocities V  in non-laminar regions of the flow  4000  Re  108  that almost
accommodates all applications of pipe flows. In the non-laminar regions of fluid
flow, Moody chart can be represented by the empirical equation i.e.
1   D 2.51 
 2 log    (25)
 3.7 R f
f  e 
This equation is called “Colebrook formula” and is valid with 10% accuracy with
the graphical data.
Example 1
Water at 200C flows at a rate of 0.05 m3/s through a horizontal pipe of 0.1m diameter.
The pressure gradient is measured to be 2.5kPa/m along the length of the pipe.
Determine: (i) approximate thickness of the viscous sub-layer; (ii) approximate
centerline velocity; (iii) ratio of the turbulent to laminar shear stress at a point
midway between the centerline and pipe wall. Take density and kinematic viscosity
of water at 200C as 1000kg/m3 and 1.00410-6 m2/s respectively
Solution
(i) Irrespective of the flow to be laminar or turbulent, the shear at the wall can be
obtained from the equation as,
Dp 0.1 2.5  1000
w    62.5 N m 2
4l 4 1
Friction velocity is given by,
0.5
 
0.5
 62.5 
u  w 
*
   0.25m s
  1000 
Approximate thickness of viscous sub-layer is given by,
5 5 1.004 106
S    0.02mm
u* 0.25
(ii) The velocity of the flow is given by,
Q 4Q 4  0.05
V    6.37 m s
A D 2
  0.12
Reynolds number of the flow,
VD 6.37  0.1
Re    6.3  105
 1.004 10 6
Assuming the power-law velocity profile,
1
u  r n
 1   where n may be taken as 7
Vc  R 
Now,
1n
 r
R
Q  AV   udA  Vc  1    2 r  dr   R 2V
0
R
After integration,
Vc  n  1 2n  1
  1.224
V 2n 2
Thus, centerline velocity is given by, Vc  1.224V  7.8 m s
It may be noted that the centerline velocity for laminar flow is, Vc  2V  12.74 m s
(iii) The shear stress at a point midway between the centerline and the pipe wall i.e.
r  0.025m is
2 w r 2  62.5  0.025
   31.25 N m 2
D 0.1
This shear stress constitutes laminar and turbulent components i.e.
   lam   turb  31.25 N m2
The laminar shear stress is given by,
1 n
0.857
V  r  7.8  0.025 
 1.004 106 1000  
du n
 lam     c 1   1    0.04 N m 2
dr nR  R   7  0.05  0.05 
The ratio of turbulent to laminar shear stress is given by,
 turb    lam 31.25  0.04
   780
 lam  lam 0.04
Example 2 Air flows through a cast iron tube of 5mm diameter at an average velocity
of 60m/s. Check, whether the flow is laminar or turbulent and then determine the
pressure drop in a 0.1m section of the tube. If the flow is made laminar by
incorporating a very smooth tube, then determine the gain in the pressure drop. The
density and viscosity of air may be taken as 1.23kg/m3 and 1.8  10-5 N.s/m2
respectively.
Solution (i) Reynolds number of the flow is given by,
VD 1.23  60  0.005
Re    20500
 1.8 105
Since this value is well above 4000, so the flow is certainly turbulent. Thus,
f    D 
From the Table-1 and Moody chart, the friction factor for
 0.26
Re  20500 and   0.052 is 0.08
D 5
Thus,
 f  l   0.08  0.1 
p     V 2     1.23  60  3542.4kPa
2

 2  D   2  0.005 
64
(ii) Had the flow been laminar, f   0.003
Re

So, p   
f l  2  0.003  0.1 
  V    1.23  60  132.9kPa
2

 2  D   2  0.005 
Gain in pressure drop = 3542.4 -132.9 = 3409.5kPa
EXERCISES
1. Oil with density 900kg/m3 and viscosity of 0.00001m2/s flows at 0.2m3/s through
a cast iron pipe of 400m long and 300mm diameter. Determine; (i) head loss; (ii)
pressure drop if the pipe slopes down at 100 in the flow direction.
2. A liquid of density 950kg/m3 and viscosity of 0.00002m2/s flows through a 20cm
diameter pipe 100m long with a head loss of 8m. If the roughness ratio is 0.0002,
calculate the average velocity and flow rate.
3. A smooth brass pipeline of 80mm diameter and 1km long carries water at the rate
of 0.005m3/s: (i) calculate the head loss, wall shear stress, centerline velocity; (ii)
shear stress and velocity at 25mm from the centerline; (iii) thickness of laminar sub-
layer. Take the kinematics viscosity and density of water as 0.000002m2/s and
950kg/m3 respectively.
Chapter six
Compressible Flow
The simple definition of compressible flow is the variable density flows. In general, the density of
gases can vary either by changes in pressure and temperature. In fact, all the high speed flows are
associated with significant pressure changes.
Compressibility (κ): It is defined as the fractional change in the density of the fluid element per
unit change in pressure. One can write the expression for κ as follows;

Speed of sound
The so-called speed of sound is the rate of propagation of a pressure pulse of infinitesimal
strength through a still fluid. It is a thermodynamic property of a fluid.

Mach number
Mach number (Ma) is defined as a ratio of the speed of an object or flow relative to the speed of
sound in the medium through which it is travelling: Ma = vo/a
Where vo is the speed of the object and a is the speed of sound in the medium
Compressible Flow Regimes

In order to illustrate the flow regimes in a compressible medium, let us consider the flow over an
aerodynamic body (Figure below). The flow is uniform far away from the body with free stream
velocity (V∞) while the speed of sound in the uniform stream is a∞. Then, the free stream Mach
number becomes M ∞ (=v∞ /a ∞). The streamlines can be drawn as the flow passes over the body
and the local Mach number can also vary along the streamlines. Let us consider the following
distinct flow regimes commonly dealt with in compressible medium.
Subsonic flow: It is a case in which an airfoil is placed in a free stream flow and the local Mach
number is less than unity everywhere in the flow field (Fig-a). The flow is characterized by smooth
streamlines with continuous varying properties. Initially, the streamlines are straight in the free
stream, but begin to deflect as they approach the body. The flow expands as it passed over the
airfoil and the local Mach number on the top surface of the body is more than the free stream
value. Moreover, the local Mach number (M) in the surface of the airfoil remains always less than
1,

when the free stream Mach number (M∞) is sufficiently less than 1. This regime is defined as
subsonic flow which falls in the range of free stream Mach number less than 0.8 i.e. M∞ ≤0.8 .
Transonic flow: If the free stream Mach number increases but remains in the subsonic range close
to 1, then the flow expansion over the air foil leads to supersonic region locally on its surface.
Thus, the entire regions on the surface are considered as mixed flow in which the local Mach
number is either less or more than 1 and thus called as sonic pockets (Fig.-b). The phenomena of
sonic pocket is initiated as soon as the local Mach number reaches 1 and subsequently terminates
in the downstream with a shock wave across which there is discontinuous and sudden change in
flow properties. If the free stream Mach number is slightly above unity (-c), the shock pattern will
move towards the trailing edge and a second shock wave appears in the leading edge which is
called as bow shock. In front of this bow shock, the streamlines are straight and parallel with a
uniform supersonic free stream Mach number. After passing through the bow shock, the flow
becomes subsonic close to the free stream value. Eventually, it further expands over the airfoil
surface to supersonic values and finally terminates with trailing edge shock in the downstream.
The mixed flow patterns sketched in Figs.(b & c), is defined as the transonic regime.

Fig. #: Illustration of compressible flow regime:


(a) subsonic flow; (b & c) transonic flow; (d) supersonic flow; (d) hypersonic flow .

Supersonic flow: In a flow field, if the Mach number is more than 1 everywhere in the domain,
then it defined as supersonic flow. In order to minimize the drag, all aerodynamic bodies in a
supersonic flow, are generally considered to be sharp edged tipk. Here, the flow field is
characterized by straight, oblique shock as shown in Fig. (d). The stream lines ahead of the shock
the streamlines are straight, parallel and horizontal. Behind the oblique shock, the streamlines
remain straight and parallel but take the direction of wedge surface. The flow is supersonic both
upstream and downstream of the oblique shock. However, in some exceptional strong oblique
shocks, the flow in the downstream may be subsonic.

Hypersonic flow: When the free stream Mach number is increased to higher supersonic speeds,
the oblique shock moves closer to the body surface (Fig.-e). At the same time, the pressure,
temperature and density across the shock increase explosively. So, the flow field between the
shock and body becomes hot enough to ionize the gas. These effects of thin shock layer, hot and
chemically reacting gases and many other complicated flow features are the characteristics of
hypersonic flow. In reality, these special characteristics associated with hypersonic flows appear
gradually as the free stream Mach numbers is increased beyond 5.
As a rule of thumb, the compressible flow regimes are classified as below;
M <0.3 (incompressibleflow ) M <1
(subsonic flow )
0.8 <M <1.2 (transonicflow ) M >1
(supersonic flow )
M >5and above (hypersonic flow

Rarefied and Free Molecular Flow: In general, a gas is composed of large number of discrete atoms
and molecules and all move in a random fashion with frequent collisions. However, all the
fundamental equations are based on overall macroscopic behavior where the continuum
assumption is valid. If the mean distance between atoms/molecules between the collisions is large
enough to be comparable in same order of magnitude as that of characteristics dimension of the
flow, then it is said to be low density/rarefied flow. Under extreme situations, the mean free path
is much larger than the characteristic dimension of the flow. Such flows are defined as free
molecular flows. These are the special cases occurring in flight at very high altitudes (beyond 100
km) and some laboratory devices such as electron beams.

Mach Waves
Consider an aerodynamic body moving with certain velocity (V) in a still air. When the pressure
at the surface of the body is greater than that of the surrounding air, it results an infinitesimal
compression wave that moves at speed of sound (a). These disturbances in the medium spread out
from the body and become progressively weaker away from the body. If the air has to pass
smoothly over the surface of the body, the disturbances must ‘warn’ the still air, about the approach
of the body. Now, let us analyze two situations: (a) the body is moving at subsonic speed V <a; M
<1; (b) the body is moving at supersonic speed V >a; M >1.
Case I: During the motion of the body, the sound waves are generated at different time intervals
(t) as shown below. The distance covered by the sound waves can be represented by the circle of
radius at ,2at, 3at.....soon. During same time intervals (t), the body will cover distances represented
by, Vt , 2Vt , 3Vt.......soon . At subsonic speeds (V <a; M <1), the body will always remain inside
the family of circular sound waves. In other words, the information is propagated through the
sound wave in all directions. Thus, the surrounding still air becomes aware of the presence of the
body due to the disturbances induced in the medium. Hence, the flow adjusts itself very much
before it approaches the body.
Case II: Consider the case, when the body is moving at supersonic speed V >a; M >1. With a
similar manner, the sound waves are represented by circle of radius at ,2at 3at.......soon after
different time (t) intervals. By this time, the body would have moved to a different location much
faster from its initial position. At any point of time, the location of the body is always outside the
family of circles of sound waves. The pressure disturbances created by the body always lags behind
the body that created the disturbances. In other words, the information reaches the surrounding
air much later because the disturbances cannot overtake the body. Hence, the flow cannot adjust
itself when it approaches the body. The nature induces a wave across which the flow properties
have to change and this line of disturbance is known as “Mach wave”. These mach waves are
initiated when the speed of the body approaches the speed of sound (V =a; M =1). They become
progressively stronger with increase in the Mach number.

Fig. #: Spread of disturbances at subsonic and supersonic speeds.


Some silent features of a Mach wave are listed below;

- The series of wave fronts form a disturbance envelope given by a straight line which is
tangent to the family of circles. It will be seen that all the disturbance waves lie within a
cone, having a vertex/apex at the body at time considered. The locus of all the leading
surfaces of the waves of this cone is known as Mach cone.

- All disturbances confine inside the Mach cone extending downstream of the moving body
is called as zone of action. The region outside the Mach cone and extending upstream is
known as zone of silence. The pressure disturbances are largely concentrated in the
neighborhood of the Mach cone that forms the outer limit of the zone of action.
- The half angle of the Mach cone is called as the Mach angle (µm) that can be easily
calculated from the geometry.
Sin(μm) = at/Vt = a(2t)/V(2t) = a(3t)/V(3t) …. = a/V = 1/M ; μm = sin-1(1/M)

Fig. #: Illustration of a Mach wave.


Normal shock wave
A normal shock wave is one of the situations where the flow properties change drastically in one
direction. The shock wave stands perpendicular to the flow as shown in Figure below. The
quantitative analysis of the changes across a normal shock wave involves the determination of
flow properties.
All conditions of are known ahead of the shock and the unknown flow properties are to be
determined after the shock. There is no heat added or taken away as the flow traverses across the
normal shock. Hence, the flow across the shock wave is adiabatic (q =0).

Fig: Schematic diagram of a standing normal shock wave.


A shock wave (or simply "shock") is a type of propagating disturbance. Like a normal wave, a
shock wave carries energy and can propagate through a medium or, in special cases, through a
field such as the electromagnetic field in the absence of a physical medium. Shock waves are
characterized by a sudden change in the characteristics of the medium (such as pressure,
temperature, or speed) as a positive step function. The corresponding negative step is an
expansion wave. An acoustic shock wave travels through most mediums at a higher speed than a
normal wave.
Unlike solitons (another kind of nonlinear wave), the energy of a shock wave dissipates
relatively quickly with distance. Additionally, the companion expansion wave from a shock
approaches, and eventually merges with the shock, partially cancelling it out. Thus the sonic
boom associated with the passage of an aircraft is the sound wave resulting from the degradation
and merging of the shock wave-expansion wave pair produced by the passage of a supersonic
aircraft.

Shock Waves and Expansion Waves


For some back pressure values, abrupt changes in fluid properties occur in a very thin section of
a converging–diverging nozzle under supersonic flow conditions, creating a shock wave.
Normal shock waves: The shock waves that occur in a plane normal to the direction of flow. The
flow process through the shock wave is highly irreversible.
Fanno line: Combining the conservation of mass and energy relations into a single equation and
plotting it on an h-s diagram yield a curve. It is the locus of states that have the same value of
stagnation enthalpy and mass flux.
Rayleigh line: Combining the conservation of mass and momentum equations into a single
equation and plotting it on the h-s diagram yield a curve

The h-s diagram for flow across a normal shock.


Chapter seven
Introduction to 2D-Potential Flow Theory

Plane potential flow


Potential Function ():

 
 Definition: u and v 
x y
v u
 Characteristic: It always satisfies the irrotationality (i.e.,  z   0)
x y
 Physical meaning:  = constant is a potential line

Streamline and potential line are orthogonal to each other


Potential Flow:

 Governing equation:  2  0 or  2  0
Inviscid flow: Euler’s equations of motion
Flow fields in which the shearing stresses are zero are said to be inviscid, nonviscous, or
frictionless. for fluids in which there are no shearing stresses the normal stress at a point is
independent of direction:

 p   xx   yy   zz
For an inviscid flow in which all the shearing stresses are zero, and the normal stresses are replaced
by −p, the Navier-Stokes Equations reduce to Euler’s equations

 V 
 g  p      V   V 
 t 
In Cartesian coordinates:

p  u u u u 
 gx    u v  w 
x  t x y z 
p  v v v v 
 gy    u v  w 
y  t x y z 
p  w w w w 
 gz    u v w 
z  t x y z 
The Bernoulli equation derived from Euler’s equations
The Bernoulli equation can also be derived, starting from Euler’s equations. For inviscid,
incompressible fluids, we end up with the same equation
p V2
  gz  const
 2
It is often convenient to write the Bernoulli equation between two points (1) and (2) along a
streamline and to express the equation in the “head” form by dividing each term by g so that

p1 V12 p2 V22
 z   z
 2g 1  2g 2
The Bernoulli equation is restricted to the following:

 inviscid flow
 steady flow
 incompressible flow
 flow along a streamline
The Irrotational Flow and corresponding Bernoulli equation

If we make one additional assumption—that the flow is irrotational  V  0 —the analysis of


inviscid flow problems is further simplified. The Bernoulli equation has exactly the same form at
that for inviscid flows:

p1 V12 p2 V22
 z   z
 2g 1  2g 2
but it can now be applied between any two points in the flow field, not limited to applications
along a streamline.

Various regions of flow: (a) around bodies;


(b) through channels
The Velocity Potential
For an irrotational flow:

 w v   u w   v u 
 V    i     j   k  0
 y z   z x   x y 

So we have

w v u w v u
 ,  , 
y z z x x y
It follows that in this case the velocity components can be expressed in terms of a scalar function
 (x, y, z, t), called velocity potential, as
  
u , v , w
x y z
In vector form:

V  
The velocity potential is a consequence of the irrotationality of the flow field, whereas the stream
function is a consequence of conservation of mass. It is to be noted, however, that the velocity
potential can be defined for a general three-dimensional flow, whereas the stream function is
restricted to two-dimensional flows.
For an incompressible flow we know from the conservation of mass:

 V  0
and therefore for incompressible, irrotational flow, it follows that

 2  0
The velocity potential satisfies the Laplace equation.
In Cartesian coordinates:

 2  2  2
 2  2 0
x 2
y z
In cylindrical coordinates:

1     1   2
r   0
r r  r  r 2  2 z 2
Some Basic, Plane Potential Flows
For potential flow, basic solutions can be simply added to obtain more complicated solutions
because of the major advantage of Laplace equation that it is a linear PDE. For simplicity, only
plane (two-dimensional) flows will be considered. Since we can define a stream function for plane
flow,

 
u , v
y x

If we now impose the condition of irrotationality, it follows

u v

y x
and in terms of the stream function

       
    
y  y  x  x 

 2  2
 0
x 2 y 2
Thus, for a plane irrotational flow we can use either the velocity potential or the stream function—
both must satisfy Laplace's equation in two dimensions. It is apparent from these results that the
velocity potential and the stream function are somehow related. It can be shown that lines of
constant (called equipotential lines) are orthogonal to lines of constant ψ (streamlines) at all
points where they intersect. Recall that two lines are orthogonal if the product of their slopes is −1,
as illustrated by this figure

Along streamlines ψ=const:


dy v

dx along  const u
Along equipotential lines = const

 
d  dx  dy  udx  vdy  0
x y

dy u

dx along  const v
Uniform flow at angle α with the x axis

Velocity potential:   U  x cos   y sin  

Stream function:   U  y cos   x sin  

Velocity components: u  U cos  , v  U sin 


Source or sink (m > 0 source; m < 0 sink)

m
Velocity potential:   ln r
2
m
Stream function:   
2
m
vr  , v  0
Velocity components:
2 r
Free vortex (Γ > 0 counterclockwise; Γ < 0 clockwise)


Velocity potential:   
2

Stream function:    ln r
2

vr  0, v 
Velocity components:
2 r
Doublet (with strength k=ma/π)
K cos 
Velocity potential:  
r
K sin 
Stream function:  
r
K cos  K sin 
Velocity components: vr   , v  
r2 r2
To Solve Potential Flow Problems:

 Superposition of Elementary Flows


o Basic elementary flows:
 Uniform flow
 Free vortex
 Source/Sink
 Doublet
 Method of Image

Superposition of Plane-Flow solutions


Superposition:
For example: Flow over a circular cylinder = Uniform flow + Doublet

Uniform flow:  uniform  U  y cos  x sin 

 sin
Doublet:  doublet  
r

Flow over a circular cylinder:    uniform   doublet

Superposition of Basic, Plane Potential Flows


Source in a Uniform Stream—Half-Body
Flow around a half-body is obtained by the addition of a source to a uniform flow.
The flow around a half-body: (a) superposition of a source and a uniform flow; (b) replacement of
streamline ψ = πbU with solid boundary to form half-body.

m
Velocity potential:   Ur cos   ln r
2
m
Stream function:   Ur sin   
2
m
vr   , v  U sin 
Velocity components:
2 r
Rankine Ovals
Rankine ovals are formed by combining a source and sink with a uniform flow.

The flow around a Rankine oval: (a) superposition of source–sink pair and a uniform flow; (b)
replacement of streamline ψ = 0 with solid boundary to form Rankine oval.

m
Velocity potential:   Ur cos    ln r1  ln r2 
2
m  2ar sin  
Stream function:   Ur sin   tan 1  2 2 
2  r a 
12
 ma 
l   a2 
 U
Body half length:

h2  a 2 2 Uh
Body half width: h tan
2a m
Method of Image:

 Used to simulate ground effects


Solution Procedure:
Step 1: Draw image vortices so that resultant velocity normal to wall is zero
Image vortices are constructed as:

 Same distance below the wall


 Opposite rotation

ccw
original
a
vortex
opposite wall same
rotation distanc
a
image
cw vortex

Step 2: Find induced velocity at location B (point of interested) by all vortices (original +
images)

K 
 v   ,  > 0 if +
r 2 r

For example, VB induced by vortex 1 (original vortex):

Magnitude Sign
V
r x – comp: V cos (+)
1 y – comp: V sin (+)


V B ,1  (V cos , V sin )
B
wall
Step 3: Find stream function  at any location (x, y)
 
   ln r   ln r 2 where  > 0 is in the counter-clockwise direction
2 4
Example:
A positive line vortex with strength is located at a distance (x, y) = (a, 2a) from the corner.
1) Compute the total induced velocity at point B, where (x, y) = (2a, a).
2) Find the stream function  at any location (x, y).
y

(a, 2a)
B = (2a, a)

Examples of the Principle of Superposition


1. Source Plus a Uniform Flow:
Let us superpose a source flow at the origin and a uniform flow along the x-axis to realize how
the superposition principle works.

+  ???

UF Source, q
Since we do not know what the resulting flow pattern will be we sum up stream functions of a
uniform flow and a source flow at the origin first.

    UF  Source

q  Ur sin   q
 Uy  [ y  r sin  ]
2 2

1  1  q 
 Vr    Ur cos   
r  r  2 

V     U sin 
r
To determine how superposition works, we need to visualize the flow first. A convenient way to
achieve this is the determination of stagnation points in the flow, and the equation of the
stagnation streamline. Remember the stagnation points are the points where fluid velocity
components are zero. A streamline that passes through a stagnation point is called a stagnation
streamline.
The stagnation streamline can represent the shape of a solid surface since, by definition, no fluid
can cross a streamline (since flow is always tangent to it). Thus, stagnation streamlines have
particular importance in the study of ideal flows.
For the present superposition, the stagnation points are determined from the solution of all (rs,s)
values that satisfy:

1 q 
Vr s   Urs cos s    0
rs  2 
Vs  U sin s  0
If we satisfy the second equation first, we find

s  0 , or, 
q
If s  0 , Vr  0  U rs cos 0  0
2
q
 rs  
2U
Since “q” for a source and “U” are positive quantities, the above equation is absurd (rs cannot be
q
< 0). However, if s   , Vr  0  rs  which is a possible solution. Thus, the flow
2U
has only one stagnation point located at:

q
rs  s  
2U

To determine the equation of the stagnation streamline,  is set to a constant equal to its value at
the stagnation point.

qs q
    s  U rs sin s   [by substitution]
2 2
Thus the plot of the stagnation streamline is given by all values of (r,) satisfying:

q q
U r sin   
2 2
q 
or, U r sin   (1  )
2 
For illustration purposes, suppose our free stream U  1 m / s, q  6.28 m 2 / s . This flow will
have the stagnation point and the stagnation streamline as shown.

Note that as   0 , the body approaches y = 3.14 asymptotically. The velocity Vr  0 and
V  0 at the stagnation point. But as we move away from the stagnation point the velocity
picks up gradually on the stagnation streamline (since ideal flows slip on a solid surface). Thus,
this flow is pictured as a flow over a semi-infinite body (called a Rankine-halfbody).
One interesting point to note in this analysis is that the point of singularity (origin) is trapped
outside the body. Therefore although a source flow is singular, the fluid flow over the resulting
body has no points of discontinuity in the flow field.

1 m/s y

3.14 m (as 0)


+  x
3.14 m (as 0)
q =6.28 m2/s

Suppose we are interested in determining what is the pressure at the point B. For this, we must
first calculate the velocity at the point B(r  1.57 m,   90) from the expressions of Vr and
V. Then apply the inviscid Bernoulli equation between two points ( and B) on the stagnation
streamline. If the pressure at “” is p, Bernoulli equation gives:

1 1
p   U 2  p B  U B
2

2 2
U B  VrB  V B
2 2 2
where, at the point B.

Plane flow past closed-body shapes


Flow around a Circular Cylinder
A doublet combined with a uniform flow can be used to represent flow around a circular cylinder.
The flow around a circular cylinder

K cos 
Velocity potential:   Ur cos  
r
K sin 
Stream function:   Ur sin  
r
Velocity components:

 a2   a2 
vr  U 1  2  cos  , v  U 1  2  sin 
 r   r 
Doublet Plus Uniform Flow

The ideal flow over a circular cylinder may be shown to be a superposition of a uniform flow
and a doublet “” at the origin with its axis along the “-x” axis. To show this, first set up the
superposition formula for .

   UF   Doublet
qs q
 U rs sin s  
2 2
1  1      
 Vr    Ur cos   cos     U  2
cos 
r  r  2r   2r 
      
V     U sin   sin     U  2
sin 
r  2 r 2
  2 r 
Determination of Stagnation Points: To find (rs,s) set Vr  V  0 and solve simultaneously. If
V  0 , sin  s  0 or,  s  0 or  . Check if both of these are valid solutions for stagnation

points. Since V  0 cannot yield U   0 because r becomes imaginary, we solve the
2r 2
values of rs by setting Vr  0 .

  
U   0  rs   rs 
2

2rs
2
2U 2U
Note that the rs values cannot be negative therefore the negative root discarded. Thus, there are 2
     
feasible stagnation points in the flow,  ,0  and  ,   . These points must also be
 2U   2U 
used to obtain the equation for stagnation streamline.


    s  U rs sin s  sin s  0
2rs
Therefore the equation of stagnation streamline is derived when (r,) = 0. Thus
   
U rs sin  s  sin s  0 , or  U r   sin   0 . Since sin   0 for any arbitrary point,
2rs  2r 
  
 Ur    0 or,
 2r 
The boxed equation shows that r is constant (  and U are constant for this steady flow). Thus it
describes the equation of a circle (or circular cylinder in 3-D). If we choose to represent the
radius of the cylinder by R,   2UR 2 for the flow. This relationship may be used for design.
Suppose you want to construct a study flow over a circular cylinder of R = 1 m, in a force stream
U = 1 m/s, the strength of the doublet you would use to construct this flow must be
  2  6.28 m 3 / s .

Evaluation of Pressure Field:

y
Stagnation streamline
B(r,)
“” 
x
To obtain the pressure field for a steady flow over a circular cylinder, we apply Bernoulli
equation between 2 points on a stagnation streamline as shown.
2
p  U 2 p VB
    (neglecting z between “” and “B”)
 2  2
p  p
  U 2  VB
2

VB  Vr B  V B
2 2 2
,

where Vr  0 [   2U r 2 ] and, V  2U sin 


B B

p  p  U 2  4U 2 sin 2 
 
 2
 p  p 
or, Cp   1  4 sin 2  ; The above expression, C p  p  p  is the definition of
1 2 1
U 2
U 2
2
pressure coefficient on a body, which is the ratio of gauge pressure and the flow-stream’s
dynamic pressure.
y
We notice that the plot of Cp is
symmetric about the “y”-axis.

1
90
0 18
  Cp (0) = 1
 0

- Cp (90) = -3
3
Cp starts at a value of 1 at the front stagnation point, crosses over the  axis at   30 , reaches
a peak value of –3 at   90 , crosses over the  axis at   150 and ends at C p  1 at
  180 . Since Cp is a measure of the fluid gauge pressure, the net pressure force in the x-
direction (which we define as the drag force on the body) must be zero [More formally,
2

F    ( p dA) sin  , which may be shown to be zero by substituting for the pressure field
D
0

and integrating]. This is the famous D’Alembert’s Paradox in fluid mechanics.


We know why the drag is zero. The real pressure curve around a circular cylinder is not
symmetric about the y-axis. However, we shall discuss these implications again during
discussions of flow separation in the chapter on boundary layers. For the present case, this
symmetry is justifiable since the flow is inviscid.
To summarize, we have presented two super position examples. In each case, the flow stream
functions are formed first. Then velocity components are computed.
Next, the stagnation points and equation of stagnation streamline are found. (Note: The
stagnation streamline represents the body shape over which the superposed flow results.) Then
pressure is found on the body surface by the application of Bernoulli equation. Lastly, the body
pressure may be integrated to evaluate forces on the body.
Worked Out Examples
(Ideal Flows)
Example 1. (Superposition of Source + U.F.):
Consider the flow field formed by combining a uniform flow in the positive x direction and a
source located at the origin. Obtain expressions for the stream function, velocity potential, and
velocity field for the combined flow. If U = 25 m/s, determine the source strength if the
stagnation point is located at x = -0.5 m. Plot the stagnation streamline. Evaluate the locations of
the branches of the stagnation streamline far downstream.
1. Statement of the Problem
a) Given
 Flow field formed by combining a uniform flow in the positive x direction and a
source located at the origin.
 U = 25 m/s
 Stagnation point is located at x = -0.5 m.

b) Find
 Expressions for the stream function (), velocity potential (), and velocity field for
the combined flow.
 Source strength if U = 25 m/s and the stagnation point is located at x = -0.5 m.
 Plot the stagnation streamline.
 Evaluate the locations of the branches of the stagnation streamline far downstream.
2. System Diagram
Flow Field is a superposition of Uniform Flow + Source Flow = Flow past a half-body

3. Assumptions
 Steady state condition
 Incompressible fluid flow
 Inviscid fluid flow
 2 - D problem
4. Governing Equations
 Stream Function Definition
1 
Vr 
r 


V  
r

 Velocity Potential Definition (2 - D case)



Vr  
r

1 
V  
r 

 Elementary Plane Flows


 Uniform Flow (positive x direction)
  Ur sin( )
  Ur cos( )
 Source Flow (from origin)
q
 
2
q
   ln(r )
2
5. Detailed Solution (Note: Stream function, Velocity Potential, Velocity field are all required)
Since this is an elementary plane flow problem, the stream function and the velocity potential
of the flow field can be obtained by superposition technique:
Flow field = Uniform flow + Source flow

 Stream Function
q
  Ur sin( )  
2

 Velocity Potential
 q   
   Ur cos( )  
q
ln(r )  Ur cos( )  ln(r )
 2   2 
Velocity field can be obtained using either stream function definition or velocity potential
definition. Here, use velocity potential definition:

    
lnr   U cos 
q q 1
Vr      Ur cos 
r r   2  2 r

1  1     1
 Ur cos  2 lnr   r  Ur sin  0  U sin
q
V   
r  r    
    1  
 er   U sin  e
q
 V  Vr  er  V  e  U cos  
 2 r

Source Strength

Stagnation point is located at x = -0.5 m  Vr = V = 0 at r = 0.5 m &  = .

 U cos  
q 1 q 1 q
Vr  U cos   U   0
2 r 2 0.5 
V  U sin  U sin   0

q
 U   0  q  U  q  25 m2/s

Stagnation streamline
Now, the stream function can be written as
q U  
  Ur sin( )    Ur sin    U r sin  
2 2  2

and the stream function value ss at the stagnation point is

    
 ss  U r sin    U 0.5sin     U
 2  2 2

  
Stagnation streamline is   U r sin     ss  U
 2 2

        
 U r sin    U  r sin    r sin    r 
 2 2 2 2 2 2 2 sin 

Locations of the branches of the stagnation streamline far downstream


Stagnation streamline equation was (See above)

 
r sin 
2

r sin (in polar coordinate system) is equivalent to y (in rectangular coordinate system).

 
Thus, y 
2

Branches of the stagnation streamline far downstream occur when   0 & 2 .

  
y  lim 
 0 2 2

  
y  lim 
  2 2 2

Therefore, the locations of the branches of the stagnation streamline far downstream is at


y .
2

6. Critical Assessment
Since this is a superposition of an elementary plane flow problem, it doesn't include
viscosity. It gives an exact solution to a real case problem as long as the viscous effects are
small. In other words, if we are only interested in the ideal pressure flow field, this should be
the approach. But viscosity must be included to obtain a more realistic solution.
Reference
Textbook:
Frank M. White, Fluid Mechanics with Student CD (McGraw-Hill
Series in Mechanical Engineering), Oct 17, 2006.

References:
1. Yunus A. Cengel and John Cimbala, Fluid Mechanics, Jan 31, 2005.
2. Robert L Mott, Applied Fluid Mechanics SI Version, May 31, 2006.
3. Iain G. Currie, Fundamental Mechanics of Fluids, Third Edition (Mechanical Engineering
(Marcell Dekker)), Dec 12, 2002.
4. Donald F. Young, Bruce R. Munson, Theodore H. Okiishi, and Wade W. Huebsch, A Brief
Introduction to Fluid Mechanics, Jan 22, 2007.
5. Bruce R.Munson, et al, Fundamentals of Fluid Mechanics, 2005.
6. Krishnamachar, P & Manohar, M, Fluid Mechanics I, 4Th Edition, 2004.
Krishnamachar, P & Manohar, M, Fluid Mechanics II, 2nd Edition, 2004. Attendance

You might also like