You are on page 1of 129

Transilvania University of Brasov

Thermodynamics and Fluid Mechanics Department

Angel HUMINIC

FLUID MECHANICS
Theory and Applications

2007

CONTENTS
INTRODUCTION .....

1. FLUIDS .....

1.1 Description of fluids ..


1.2 Continuum Hypothesis. Concepts of homogenous and isotropic medium ..

3
5

1.3 Models for the fluid description ...

2. PROPERTIES OF FLUIDS

2.1 Pressure .

2.2 Density ....

10

2.3 Specific weight ...

11

2.4 Isothermal compressibility

12

2.5 Velocity of sound ...

13

2.6 Mach number ....

13

2.7 Viscosity .....

14

2.8 Applications of fluid properties

20

3. FUNDAMENTALS OF FLUID STATICS .

24

3.1 Forces on fluids .....

24

3.2 Eulers equation for fluids statics ...

25

3.3 Particular forms for the fundamental equation of statics .

27

4. FLUID FORCES ON SUBMERGED SURFACES .

33

4.1 Fluid forces on plane surfaces ......

34

4.2 Fluid forces on curved surfaces ..

36

4.3 Floating bodies ..

39

4.4 Static pressure measurement by manometers

40

4.5 Examples of fluid statics ..

44

5. IDEAL FLUID DYNAMICS ...

52

5.1 Description of the flow field ..

52

5.2 Acceleration in a flow field ...

55

5.3 Equations of the fluid motion ...

55

6. APLICATIONS OF THE BERNOULLIS EQUATION ..

62

6.1 Flow through small orifices ..

62

6.2 Time for a tank discharge ....

63

6.3 Flow over notches and weirs ...

64

6.4 Mixing of fluids. The ejectors ..

65

6.5 Fluid metering ....

67

7. APPLICATIONS OF THE MOMENTUM EQUATION ...

72

7.1 Hydrodynamic forces on a flat plate ...

72

7.2 Hydrodynamic force on a pipe nozzle

73

7.3 ENERGY FROM WIND. The Axial Momentum Theory Betzs Theory .

74

8. REAL (VISCOUS) FLUIDS FLOW .....................................................

78

8.1 Laminar and turbulent flows. Experiments of Reynolds ..

78

8.2 Velocity distribution on laminar and turbulent flows ....

79

8.3 Steady flow in pipelines. Head loss and Bernoullis augmented equation .. ...

81

8.4 Friction factor computation ..

83

8.5 Transient flow in pipelines. Water hammer ...

91

9. BOUNDARY LAYER THEORY

92

9.1 General description of the boundary layer

92

9.2 Boundary layer separation ..

96

9.3 Navier-Stokes equations .

98

9.4 The notion of resistance, drag, and lift ..

100

9.5 Applications of boundary layer theory

102

10. ROTODYNAMIC MACHINERY .

106

10.1 Introduction to rotodynamic machinery ..

106

10.2 Centrifugal pumps ..

106

10.3 Axial flow machinery ..

110

10.4 Introduction to hydraulic turbines .

112

11. DIMENSIONAL ANALYSIS

116

11.1 Introduction to dimensional analysis

116

11.2 Dimensions and units

116

11. 3 Dimensional homogeneity

117

11.4 Results of dimensional analysis

118

11.5 Buckingham's theorems

118

11.6 Common groups .

121

11.7 Scale models testing ..

122

12. REFERENCES ..

126

1. INTRODUCTION
As its name suggests, Fluid Mechanics is the branch of Applied Mechanics concerned with the
fluids. The analysis of the behaviour of fluids is based on the fundamental laws of Mechanics,
which relate continuity of mass and energy with force and momentum, together with their physical
properties.
Fluid Mechanics is involved in nearly all areas of Engineering either directly or indirectly, being one
of the oldest and broadest fields of Engineering. It encompasses aerodynamics, hydrodynamics,
gas dynamics, flows in turbomachineries, computational fluid dynamics (CFD), convection heat
transfer, acoustics, biofluids, physical oceanography, atmospheric dynamics, wind engineering a.o.
Fluid Mechanics has practical importance in almost all human activities, from meteorology (flow in
atmosphere) and, in fact, even astronomy (motion of interstellar gas) up to medicine (flow of fluids
inside human body). Modern design of aircraft, spacecraft, automobiles, ships, land and marine
structures, power and propulsion systems, or heat exchangers is impossible without a clear
understanding of the relevant Fluid Mechanics.
Also, fluids are involved in transport and/or conversion of the energy in almost its forms. The flow
problems are present if we study fuel combustion on one side and water evaporation on the other
side in a boiler, flow of heat-transporting medium in a nuclear reactor, in the blade system of a
turbine, in recycling and coolant pumps and also in the processes taking place in the cylinders of
the internal combustion engines and others.
The aims of this course are to introduce Fluid Mechanics and to establish its relevance in
engineering. It consists of some llectures which are presenting the concepts, theory and practical
applications. Worked examples will also be given to demonstrate how the theory is applied.

1.1. FLUIDS
1.1.1 Description of fluids
We normally recognise three states of matter: solid, liquid and gas. Name fluids is a generalized
description meant to include both liquids and gases. There are two aspects of Fluid Mechanics,
which make it different to Solid Mechanics.
1. The nature of a fluid is much different to that of a solid
2. In fluids we usually deal with continuous streams of fluid without a beginning or end. In
solids we only consider individual elements.
3

A solid can resist a deformation force while at rest, this force may cause some displacement but
the solid does not continue to move indefinitely.
In contrast to solids the fluids do not have the ability to resist deformation. Because a fluid cannot
resist the deformation force, it moves, it flows under the action of the force. Its shape will change
continuously as long as the force is applied. The deformation is caused by shearing forces, which
act tangentially to a surface. Referring to the figure below, we see the force F acting tangentially on
a rectangular (solid lined) finite element ABCD. This is a shearing force and produces the (dashed
lined) rhombus element A'B'CD, defined by the angular displacement :
tg ( ) =

x
y

Fig 1.1 Angular displacement of a fluid element due to shear deformation


In consequence it can say that:
A fluid is a substance that deforms continuously, or flows,
when it is the subject to shearing (tangential) stresses (forces).
As will be discussed later on (see, Fluids Properties, Viscosity), the magnitude of the stress
depends on the rate of the angular deformation, in opposition as solids for what the magnitude of
the stress depends on the magnitude of the angular deformation. This definition implies a very
important point, which state that:
If a fluid is at rest there are no shearing forces acting.
At rest all forces must be perpendicular to the planes which they are acting.
When a fluid is in motion shear stresses are developed if the particles of the fluid move relative to
one another. When this happens adjacent particles have different velocities. If fluid velocity is the
same at every point then there is no shear stress produced: the particles have zero relative
velocity.

This is very well illustrated by the case of the rivers flow. At the bottom of the river, the velocity of
the water is zero and will increase toward the top of the river. This change in velocity across the
direction of flow is known as velocity profile and is shown graphically in the Figure 1.2:

Fig 1.2 Velocity profile in a river cross section


Because particles of fluid next to each other are moving with different velocities there are shear
forces in the moving fluid i.e. shear forces are normally present in a moving fluid. On the other
hand, if a fluid is a long way from the solid boundary, nearly to the top of the river in the presented
case, all the particles are travelling with the same velocity, the velocity profile would look closely
like in Figure 1.3.

Fig 1.3 Velocity profile in uniform flows


As presented in Figure 3 there will be no shear forces if all particles have zero relative velocity. In
practice we are concerned with flow past solid boundaries; aeroplanes, cars, pipe walls, river
channels etc. and shear forces will be present.
1.1.2 Continuum Hypothesis. Concepts of homogenous and isotropic medium
All fluids are composed of molecules discretely spaced and in continuous motion. In the definitions
used to describe fluids, this discrete molecular structure was ignored and the fluid was considered
as a continuum. This means that all dimensions in a fluid space are taken as large compared to the
molecular spacing.
In Fluid Mechanics, the continuum concept is expressed through that, for any point P ( x , y , z ) of a
fluid, we can associate a continuum description for each quantity (variable) that characterizes the
fluid, e.g. pressure, temperature, velocity etc., for a certain time t :
- pressure

p = p( x , y , z , t ) ,

- temperature

T = T ( x , y , z, t ) ,

- velocity

v = v ( x , y , z, t ) ,

and these function are continuous ones, consequently


they are derivative. The reason one uses the continuum hypothesis is that it allows us to use
differential calculus to analyze the properties of a fluid and its behaviour.
Continuous changes, or gradients, in physical properties and forces define mechanical problems,
and differential calculus is the mathematical tool that treats such gradients. Such a view is
reasonable from a modelling point of view as long as the mathematical model generates results
which agree with experiment. Practicaly we can assume that, the ratio of the mean free path length
of the molecules of a fluid is small, the continuum hypothesis is more accurate (see details about
Knudsen number in bibliography). Thus, the continuum concept is not always applicable. For
example, in the cases of highly compressible flows or flows in very dilute gases there will be
regions of discontinuity, such as the shock waves on the surfaces of an airplane which is flying
close to the sound speed, M = 1 (Mach: ratio between fluid velocity v and sound speed c , see
Lecture 2, Fluids Properties), see Figure 1.4.

Fig 1.4 Shock wave on airplane breaking the sound barrier


A fluid is considered homogenous if its density is constant for constant conditions of temperature
and pressure.
p,T = ct = ct

Off course, no real fluid is homogeneous in an absolute sense. Homogeneity must be specified
relative to a characteristic length, e.g. the size of the probe used to measure the properties of the
fluid in experiments.
A fluid medium can be considered as isotropic if it has the same properties in all directions around
a point.
The concepts of homogenous and isotropic medium are useful in study of Fluid Mechanics
because they allow us to consider that the Equations established for an elementary volume of fluid
(fluid particle) are applicable for entire fluid.

1.1.3 Models for the fluid description


It is necessary to emphasize that generally, fluids are very complicated objects of study and as a
consequence, it is almost always necessary to work with certain simplified fluid models. Such
models usually neglect properties not important in the mechanical context, as taste, smell, colour,
pH value and others. Also, in order to obtain mathematical simplifications, many models often
neglect even mechanical properties as long as they are of secondary consequences for a
phenomenon under study.
Common models for fluid are the following ones:
Newtonian fluids:

named after Sir Isaac Newton, who studied fluid motion, are fluids having
directly proportionality between stress and rate of angular deformation
starting with zero stress and zero deformation. In these cases, the constant
of proportionality is defined as the absolute or dynamic viscosity (see
Properties of fluids). The more common fluids like air and water are
Newtonian ones.

Non-Newtonian fluids:fluids having a variable proportionality between stress and deformation rate
(basically everything other than Newtonian fluid). A vast number of fluids
which are not commonly encountered but which are extremely important,
nevertheless, are non-Newtonian: plasmas, liquid crystals, foams a. o. Also,
some of the plastics behave like a solid and above a stress magnitude they
have fluidic behavior. Rheology is the subject that treats plastics and nonNewtonian fluids.
Ideal fluids:

fluids having no viscosity (Euler model). The effect of the shear stress
between the adjacent layers of fluid is neglected.

Light fluids:

Fluids having small mass density, as gasses and vapours. The effect of their
own weight can be neglected.

Incompressible fluids: fluids which have not significant variation of their volume with the pressure
changes, and as a consequence with constant mass density (Pascal model).
It is used in study of the liquids (heavy fluids).
When a subject is theoretically studied, is desirable to use the simplest possible model of fluid and
to apply a more complicated one only when is absolutely necessary.

2. PROPERTIES OF FLUIDS
A property is a characteristic of a substance which is invariant when the substance is in a particular
state. In each state the condition of the substance is unique and is described by its properties.
The properties outlines below are general properties of fluids, which are of interest in Engineering.
The symbol usually used to represent the property is specified together with some typical values in
SI units for common fluids. Values under specific conditions (temperature, pressure etc.) can be
found in reference books. V
2.1 PRESSURE, p
If a volume of fluid is isolated as a free body, the force system acting on the volume includes
surface forces over every element of area bounding the volume. In general, a surface force will
have components perpendicular and parallel to the surface. At any point, the perpendicular
component per unit area is called the normal stress. If this is a compression stress, it is called
pressure intensity, or simply pressure.
Thus, the pressure is defined as (pressure) force F per unit area A applied on a surface in a
direction perpendicular to that surface. Mathematically, it is expressed as:

F dF
=
,
A 0 A
dA

(2.1)

F
A

(2.2)

p = lim

in differential form, or simply:


p=

Also, in a fluid pressure may be considered to be a measure of energy per unit volume or energy
density by means of the definition of work. For a force exerted on a fluid, this can be seen from the
definition of pressure:
p=

F F d W Energy
=
=
=
A A d V Volume

(2.3)

Pressure is a scalar quantity, and its unit in the SI is [N/m 2 ] , or Pascal [Pa] science 1971, after
the name of French mathematician, physicist and philosopher Blaise Pascal (1623 1662). The SI
multiplies are:
the kilopascal, 1 kPa = 10 3 Pa, and
the megapascal 1 MPa = 10 6 Pa .
8

A non-SI multiple, still in use in many parts of the world, is the bar:
1 bar = 10 5 Pa .
Other units in use include the following:

the technical atmosphere having as symbol: at


1 at = 1

kgf
cm

N
N
= 9.80665 10 4 2 10 5 2 = 1 [bar ] .
m
m

(2.4)

the standard (physical) atmosphere ( atm or At ): is an established


constant and it is approximately equal to the air pressure at earth
mean sea level; the first one who measured the atmospheric pressure
was the Italian physicist and mathematician Evangelista Torricelli
(1608 1647). In 1643 he created a tube by 1 meter long, sealed at
the top end, filled it with mercury, and set it vertically into a basin of
mercury (see Figure 2.1); the column of mercury fell to about 70 cm,
leaving a vacuum above, this was the first barometer and the unit of
pressure torr (1 torr = 1 mmHg ) was named after him; as we now
know, the height of column fluctuated with changing atmospheric
pressure and the acceleration due to gravity on Earth's surface; the 4th
Resolution of the 10th General Conference of Weights and Measures stated that:
N
1 atm = 101325 2 = 760 [mmHg ]
m

(2.5)

Because pressure is commonly measured by its ability to displace a column


of liquid in a manometer, pressures are often expressed as a depth of a
particular fluid (see Figure 2.2). The most common choices are mercury
(Hg) and water (or alcohol); water is nontoxic and readily available, while
high density mercury allows for a shorter column (and so a smaller
manometer) to measure a given pressure. The pressure exerted by a
column of liquid of height h and density is given by the hydrostatic
pressure equation:
p = g h [column of fluid ] .
1 mm H 2O = 10 3

1 mm alc 803

kg
m

kg
m

9.81

9.81

m
s

m
s

10 3 m = 9.81

10 3 m 7.875

N
m2

N
m2

(2.6)

1 mm Hg = 133.322

N
m2

Accordingly with two previous shown situations, pressures can be measured:

relative to an absolute zero value (E. Torricelli) (called absolute pressure pabs ) or,

relative to the atmospheric pressure at the location of the measurement patm , called gauge
pressure pg or simply p ; an example of this is the air pressure in an automobile tire, which
might be said to be "2.2 bar, but is actually 2.2 bar above the atmospheric pressure; gauge
pressure is a critical measure of pressure wherever one is interested in the stress on
storage vessels and the plumbing components of fluidics systems:
p = pabs patm

(2.7)

While pressures are generally positive, from the previous Equation we can see that there are
situations in which a negative pressure may be encountered if pabs < patm (vacuummetric
pressure).
O the basis of reaction to pressure, a subdivision of fluids into two main classes, either
compressible or incompressible, can be made. All gases and vapours are highly compressible.
Liquids by comparison are only slightly compressible. As we shall see, compressibility introduces
thermodynamic considerations into fluid flow problems. If incompressibility can be assumed, it is
much easier to describe the state of the fluid and its behaviour in motion. With some important
exceptions, liquids usually are treated as incompressible for practical purposes. Gases, on the
other hand, can be treated as incompressible only if the change in pressure is small throughout the
flow system.
2.2 DENSITY,
The density of a substance is the quantity of matter contained in a unit volume of the substance.

Mathematically, it is defined as the rate of mass per volume and has the SI unit kg / m 3 :

dm kg
.
dV m 3

(2.8)

A given amount of matter is said to have a certain mass that is treated as an invariant. Thus it
follows that density is a constant so long as the volume of the given amount of matter is unaffected
i.e., for a gas, so long as pressure and temperature conditions are the same, or generally for
homogeneous fluids. In this case:

m kg
.
V m 3

10

(2.9)

The previous formula is applicable also, for the average density of an matter. Generally, the
density of a substance depends on point of measurement, pressure p and temperature T . This is
holding good especially for gasses. Their density can be computed from the Equation of State,
applied for two conditions, one of them being known (as reference):

= 0
where:

p T0
p0 T

(2.10)

terms with index 0 are the reference parameters.

The Equation (2.9) express the absolute density. In order to facilitate the measurement of the
substance density, sometimes is used the relative density r , defined as the ratio of mass density
of a fluid to some standard (reference) mass density ref :
( fluid )r =

fluid
ref .

(2.11)

For liquids, which are treated generally as fluids with constant density, this reference is the
maximum density for water, apa = 1000 kg/m 3 . For gases, the reference is the air density in
standard conditions: 0air = 1.225 kg/m 3 for p0air = 760 mm Hg and T0air = 288.15 K .
The reciprocal of density is the specific volume, v .
v=

(2.12)

Associated with density, there can be defined another parameter, respectively the:
2.3 Specific weight,
Specific weight (or less often specific gravity) is defined as the rate of weight per volume and has

the SI unit N / m 3 :

dG N
dV m 3

(2.13)

For homogeneous fluids can be defined as the force exerted by gravity upon a unit volume of the
substance:

mg N
V m 3

(2.14)

The Relationship between and g can be determined by Newton's 2nd Law, since:

= g

11

(2.15)

2.4 ISOTHERMAL COMPRESSIBILITY, k AND BULK MODULUS OF ELASTICITY,


Compressibility is the measure of change of volume and density when a substance is subjected to
normal pressures. It is defined as relative change in volume (or density) for a given pressure
change:
k=

where:

dV 1
V0 dp

m2

.
N

(2.16)

V0 is the original volume.

The negative sign indicates a decrease in volume with an increase in pressure, as shown in Figure
below.

Fig. 2.3 Variation of volume in a cylinder due to pressure change


The reciprocal of compressibility is known as the bulk modulus of elasticity:

1
dp N
=
.
dV m 2
k
V0

(2.17)

If we take the derivative of Equation m = V = ct . , we find that:


dm = dV + V d = 0 dV = V d

d V d
=
,
V

(2.18)

and the Equations (2.16), (2.17) can be rewritten as:

k=

1 d m 2

.
dp N

(2.19)

dp N
.
d m 2

(2.20)

Compressibility data for liquids are usually given in terms of , as determined experimentally.
Theoretically, should depend on the manner or process in which the volume or density change
is effected, e.g., isothermally, adiabatically, etc. For the common gases (such as oxygen), these
two processes give (see also the thermodynamics properties of gases):

= p (for isothermal process);


12

(2.21)

= p (for adiabatic process);


where:

(2.22)

= c p / cv

ratio of specific heat at constant pressure to that at constant volume;

absolute pressure.

The effect of the process is smaller for liquids than for gases.
Associated with these two parameters, there can be defined another one, respectively the:
2.5 VELOCITY OF SOUND, c
It is given by the Equation:
c=

dp
=
d

1
m
(Newton).
=

k s

(2.23)

The magnitude of c is the velocity with which small-amplitude pressure signals will be transmitted
through a fluid of infinite extent or through a fluid confined by completely rigid walls. The density
change caused by an infinitely small pressure wave occurs almost frictionlessly and adiabatically.
For the liquids the speed of sound is usually determined from experimental values of and .
Using the adiabatic bulk modulus, Equation (2.22), the following Equation can be applied quite
accurately for the common gases:
c=

p m
.
s

(2.24)

2.6 Mach number, Ma


Mach number is a dimensionless measure of relative speed. It is defined as the speed of an object
relative to a fluid medium, v , divided by the speed of sound in that medium c ; as it is defined as a
ratio of two speeds, it is a dimensionless number:
Ma =

v
c

[] .

(2.25)

The Mach number is named after Austrian physicist and philosopher Ernst Mach (1838 1916).
The Mach number is commonly used both with objects travelling at high speed in a fluid, and with
high-speed fluid flows inside channels such as nozzles, diffusers or wind tunnels. At a temperature
of 15 degrees Celsius, Mach 1 is 340.3 m/s in the atmosphere. The Mach number is not a
constant, being temperature dependent.
According with Mach value, the following classification of fluid flows can be made:

Ma < 0.25

Subsonic incompressible flows;

0.25 < Ma < 0.8

Subsonic compressible flows;

0.8 < Ma < 1.2

Transonic flows; for Ma = 1 we have sonic flow;


13

1.2 < Ma < 5

Supersonic flows;

Ma > 5

Hypersonic flows.

2.7 VISCOSITY
Viscosity is the property of a fluid, due to cohesion and interaction between molecules, which offers
resistance to sheer deformation. Different fluids deform at different rates under the same shear
stress. Fluid with a high viscosity such as oil, deforms more slowly than fluid with a low viscosity
such as water. It is one of the most important properties of fluids due to it stand on the base of the
principle of fluids motion. The reciprocal of viscosity is fluidity.
2.7.1 Newton's Law of Viscosity
As mentioned before, fluids do not have the ability to resist deformation. Its shape will change
continuously as long a stress is applied. These deformations are caused by shearing forces, which
act tangentially to a surface. Referring to the figure below, we see the force F acting tangentially on
a rectangular (solid lined) finite element ABCD. This is a shearing force (stress) and produces the
dashed lined rhombus element A'B'CD.

Fig 2.4 Angular displacement of a fluid element due to shear deformation


A particle in point E (Figure 2.4) moves under the shear stress to point E' , in time t , on distance
x . The deformation that this shear stress causes is measured by the size of the angle , which is

know as shear strain:


tg ( ) =

x
y

(2.26)

In a solid, shear strain is constant for a fixed shear stress . In a fluid, increases as long as
is applied. It has been found experimentally that the rate of shear strain (shear strain per unit time)
is directly proportional to the shear stress. For small deformations we can write the shear strain as:

x
,
y

and rate of shear strain:

14

(2.27)

d dx 1
dv
=
=
.
dt
dt dy dy

(2.28)

where:
dv
dy

is the change in velocity with y, in the direction perpendicular to the layers,


or the velocity gradient.

Using the experimental results the proportionality between shear stress and rate of shear strain
was expressed by the following Equation (Sir Isaac Newton):

dv
,
dy

(2.29)

where:

is the constant of proportionality, known as the dynamic (molecular) viscosity.


Equation (2.29) is known as Newton's law of viscosity.

The relationship between the shear stress and the velocity gradient can be easelly understanding
from the following experiment, ilustrated below.

Fig 2.5 Laminar shear of fluid between two plates


Let consider two plates separated by a homogeneous fluid, spaced at a distance h . The two
plates are considered very large, with the area A , such that the edge effects may be ignored, and
that the lower plate is fixed. If a constant force F is applied to the upper plate, it will be put in
motion with the constant speed v . The first layer of fluid, atached on moving plate has same
velocity as the plate and this velocity will decrease in depth to v = 0 , velocity of the layer attached
on fixed plate. This decreasing of the velocity in depth is due to shear stress, , between layers of
fluid, that ultimately opposes to any applied force: = (F A ) . On the other hand, the applied force
is proportional to the velocity of the plate, F v , (higher velocity means higher force/stress) and
inversely proportional to the distance between the plates F y (lower distance means higher
force/stress). Combining these three relations results in the equation (2.29), expressed in terms of
shear stress.

15

2.7.2 Dynamic viscosity,


The dynamic viscosity is defined as the shear force, per unit area, (or shear stress), required to
drag one layer of fluid with unit velocity past another layer a unit distance away.

N s
or
dv m 2
dn

kg
m s

(2.30)

where:
n:

normal unitary vector.

Equation (2.30) shows that the velocity distribution in a fluid is a continuous field. Otherwise, the
shear stress becomes infinite for a finite velocity change between two adjoining points.
In Technical System the units of is P (Poise): 1 P = 1

g
. Typical values are:
cm s

water = 1.14 10 3 kg/m s ;


air = 1.78 10 5 kg/m s (for the standard atmosphere);
paraffinoil = 1.9 kg/m s ;
mercury = 1.552 kg/m s .
2.7.3 Kinematic viscosity,
Kinematic viscosity is defined as the ratio of dynamic viscosity to mass density:

m2

.
s

(2.31)

In Technical System the units of is St , Stokes (named after Sir George Gabriel Stokes, (1819 1903), an Irish mathematician and physicist, who at Cambridge made important contributions to
fluid dynamics, including the Navier-Stokes equations, optics, and mathematical physics):
1 St = 1

cm 2
.
s

Typical values are:

water = 1.14 10 6 m 2 / s ;
air = 1.46 10 5 m 2 / s (for the standard atmosphere);
paraffinoil = 2.375 10 3 m 2 / s ;
mercury = 1.145 10 4 m 2 / s .

16

2.7.5 Viscosity in gases and liquids


Viscosity in gases arises principally from the molecular diffusion that transports momentum
between layers of flow. It is independent of pressure and increases as temperature increases.
In liquids, the additional forces between molecules become important. This leads to an additional
contribution to the shear stress. Thus, in liquids viscosity is independent of pressure (except at
very high pressure) and tends to fall as temperature increases (for example, water viscosity goes
from 1.79 cP to 0.28 cP in the temperature range from 0 C to 100 C).
The dynamic viscosities of liquids are typically several orders of magnitude higher than dynamic
viscosities of gases. The Sutherland's formula can be used to compute the dynamic viscosity of
various materials as a function of the temperature:
T +C T
= 0 0

T + C T0

2 kg

m s ,

(2.32)

where:

dynamic viscosity for standard atmosphere (at p0 and T0 );

Sutherland's constant (in the case of air C = 111 K ).

For water, the Poiseuilles Equation can be used to compute, with good accuracy, the kinematic
viscosity:

1.78 10 6
1 + 0.0337 t + 0.00022 t 2

m2
,

(2.33)

where
t:

is the temperature of water in Celsius degrees.

2.7.6 Classification of Newtonian / Non-Newtonian Fluids


Fluids obeying Newton's law where the value of is constant are known as Newtonian fluids. If
is constant the shear stress is linearly dependent on velocity gradient. This is true for most
common fluids. Fluids in which the value of is not constant are known as non-Newtonian fluids.
There are several categories of these, briefly outlined below. These categories are based on the
relationship between shear stress and the velocity gradient (rate of shear strain) in the fluid.
1 - Ideal fluid:

this is a fluid that is assumed to have no viscosity. This is a useful concept


when theoretical solutions are being considered - it does help achieve some
practically useful solutions;

2 Solids:

no rate of shear strain for shear stress;

3 Newtonian fluids: linear dependency between rate of shear strain and shear stress;

17

4 - Dilatant fluids:

viscosity increases with rate of shear e.g. quicksand.

5 - Pseudo-plastic:

no minimum shear stress necessary and the viscosity decreases with rate of
shear, e.g. colloidal substances like clay, milk and cement,

6 - Bingham plastic:

shear stress must reach a certain minimum before flow commences; above
this value of shear stress they have a Newtonian behaviour, not applicable
for real plastics.

Fig 2.6 Shear stress versus rate of shear strain


Each of these curves can be represented by the equation:
dv

dy

= A + B

(2.34)

where:
A , B and n are constants. For Newtonian fluids: A = 0 , B = , n = 1 .

2.8 THERMODYNAMIC PROPERTIES


2.8.1 Specific heat, c
Specific heat or specific heat capacity is the ratio of quantity of heat flowing into a substance per
unit mass, to the change in temperature. It is determined experimentally or can be computed using
molecular theory.
The specific heat for gases and vapours depends on how the change in state is effected: at
constant volume (or density), cv , or at constant pressure, c p . For a perfect gas, a gas which
obeys the equation of state (2.41), both are related through the relations:
c p = cv + R [J/kg K];
cp
cv

Thus:
cp =

= .

R ; cv =

(2.36)
R
.
1

where:
R [J/kg K]:

constant of the studied gas;

[-]

adiabatic constant for the gases; = 1.4 for air.

18

(2.35)

(2.37)

2.8.2 Specific internal energy, u


The specific internal energy, measured as energy per unit mass, is due to the kinetic and potential
energies bound into a substance by its molecular activity. It depends primarily on temperature. For
a perfect gas, it can be computed according to the relation:
du = cv dT [J/kg].

(2.38)

2.8.3 Specific enthalpy, h


The specific enthalpy represents the sum:
h=u+

[J/kg].

(2.39)

For a perfect gas the enthalpy depends of temperature only and can be computed according to the
equation:

p
dh = d u + = c p dT .

(2.40)

2.8.4 Equation of state for gases


For gases the pressure (as absolute value) p , the mass density , and the temperature T are
correlated. A useful approximation is the theoretical equation for perfect gasses (Clapeyron Mendeleev):
p V = mRT

= RT

R
M

(2.41)

where:
R [J/kg K]:

constant of the studied gas;

R = 8314.3 [J/kmolK]:

universal constant of the gasses;

M [kg/kmol]:

molar mass (molecular weight).

There are the following dependences between density and pressure for a gas:
Constant volume process - V = ct .

= ct = 0

(2.42)

Isothermal process - T = ct :
p

= ct =

p0

(2.43)

Adiabatic process - Q = 0
p

Polytrophic process (general process)


p

where:
n

= ct =

= ct =

p0

0
p0

0n

polytrophic exponent for the gases; n = 1.3 for air.


19

(2.44)

(2.45)

2.8 APPLICATIONS OF FLUID PROPERTIES


WORKED EXAMPLES
Exercise 1
For checking (or calibration) of manometers, the following experimental set-up (see Figure 2.7),
with a pressure pump, can be used:

Fig. 2.7
It consists on the piston 1 moving in the cylinder 2, through the whirling of the screw piston rod 3.
The pump is feed with oil from the tank 4, through the pipe 5, where are attached the manometers,
first one used as reference 6, and the tested one 7. Compute the number of revolutions of the
piston rod necessary that the reference manometer to show a pressure p = 75 kg / cm 2 . The
following are known:
d = 36 mm

inner diameter of the cylinder;

h = 0.8 mm

pitch of the screw piston rod;

V = 2000 cm 3

initial volume of oil;

k = 4.85 10 10 m 2 / N

oil compressibility.

Solution:
It is recommended to convert all the quantities in the International System of Units (if is necessary):
d = 36 mm = 36 10 - 3 m = 0.036 m ;
h = 0.8 mm = 0.8 10 - 3 m = 0.0008 m ;

V = 2000 cm 3 = 2000 10 6 m 3 = 0.002 m 3 ;


k = 4.85 10 10
p = 100

kg f
cm

m2
;
N

= 100

9.81 N
10

20

= 9.80 10 6

N
m2

Through the whirling of the screw piston rod, the piston is moving by a distance l = n h (number of
revolutions x pitch). The oil will be compressed in the cylinder and pipe, due to an increasing of
pressure p , from 0 (initial gauge pressure shown by manometer) to p = 100 kg / cm 2 .
Using Equation of the isothermal compressibility,
V
= k p ,
V

(2.46)

and replacing the volume variation with:


V = l

d2
4

=nh

d2
4

(2.47)

it found that:
n h d2
4Vkp
=kpn=
4V
h d2
n=

4 2 10 3 4 ,85 10 10 9.81 10 6
0.8 10 3 0.036 2

(2.48)

= 11.7 rovolutions

Exercise 2
A plate having the surface S = 0.8 m 2 and mass m = 2 kg is sliding on an inclined plane with the angle

= 30 which is covered with an oil film by thickness = 2 mm . Oil mass density is = 0.9 kg/dm 3
and the kinematic viscosity = 0.4 stokes . Compute the velocity of the plate in uniform motion.

Fig. 2.8
Solution:
S = 0 ,8 m 2 ;
m = 2 ,0 kg ;

= 30 ;

= 2 mm = 2 10 -3 m = 0.002 m ;
= 0 ,9 kg/dm 3 = 0 ,9 10 3 kg / m 3 = 900 kg / m 3 ;
= 0 ,40 stokes = 0,40 cm 2 / s = 0 ,40 10 4 m 2 / s .

21

Due to the action of the tangential component of the weight


GT = G sin , plate is sliding on the inclined plane with the velocity
v , which is constant in uniform motion (when viscous friction force is

equal with GT ).
From the Newtons Equation of tangential stress in a fluid:

=
where: =

Fig. 2.9

GT
v
=

(2.49)

is the dynamic viscosity.


m g sin
v
m g sin
= v =
S

S
v =

2 9 ,81 2 10 3 sin 30
0 ,9 10 3 0 ,4 10 4 0,8

(2.50)

= 0.681 m/s

Exercise 3
Determine the effect of the water temperature on magnitude of sound speed knowing the density and the
modulus of elasticity:

water = 1000 kg/m 3 and

water = 1.914 10 9 N/m 2 for t water = 4 C ;

water = 999.26 kg/m 3 and

water = 2.020 10 9 N/m 2 for t water = 20 C .

Solution:
Using the Newtons Equation for the sound speed in fluid:

c=

dp
=
d

(2.51)

it found that:

cwater =

1.914 10 9
= 1388 m/s if t water = 4 C ;
1000

cwater =

2.020 10 9
= 1422 m/s if t water = 20 C .
999.26

In consequence, the sound speed in water is increasing with the temperature rise.

SELF-ASSESSMENT EXERCISES
Exercise 4
A piston is moving with the constant velocity v = 0.1 cm / s in a pressure cylinder having the inner
diameter D = 50 mm and length l = 10 cm , which is filled with a liquid with the modulus of

22

elasticity = 2 10 4 daN / cm 2 . Compute the time and the displacement x [mm] of the piston if the
pressure in cylinder increase from zero to p = 200 bar . Make a sketch.
Exercise 5
The velocity distribution of a viscous liquid flowing over a fixed plate is given by the Equation:
u = 0.68 y y 2

where;
u

is velocity in [m/s]

is the distance from the plate in [m].

What are the shear stresses at the plate surface and at y = 0.34 m , if the dynamic viscosity of the
fluid is = 0.9 N s m 2 ? Draw the variation = ( y ) for y = ( 0 0.34) m .
Exercise 6
In a fluid the velocity measured at a distance of 50 mm from the boundary is 1 m/s . The fluid has
dynamic viscosity 2 Pas and relative density 0.8 , to the water density. What is the velocity
gradient, and shear stress at the boundary assuming a linear velocity distribution? What is the
kinematic viscosity of the fluid in St (Stokes)? Make a sketch.
Exercise 7
Compute the sound speed in air for a temperature t air = 20 C , for an adiabatic dependence
between density and pressure. The molar mass of the air is M air = 28.96 kg/kmol and the
adiabatic exponent is = 1.4 . The universal constant of the gasses is R = 8314.3 [J/kmolK] .
Exercise 8
Explain why the viscosity of a liquid decrease while that of a gas increases with a temperature rise.

23

3. FUNDAMENTALS OF FLUID STATICS


FUNDAMENTAL EQUATION OF FLUID STATICS
Characteristic for the Fluid Statics is that the fluids are at rest, with no relative motion between fluid
elements, so there are no shear forces (or shear stress) acting on a fluid. In consequence real
fluids at rest can be treated as ideal ones, having no viscosity.
The fundamental problems of Fluid Statics are the following:

determination of the distribution of pressure in a homogeneous fluid;

computation of the fluid forces on submerged surfaces;

pressure measurement using manometers

3.1 FORCES ON FLUIDS


Two types of forces are acting on a fluid, having a mass m in a
volume V bounded by a surface S (see Figure 3.1):

mass forces and

surface forces.

In the following, only the external forces will be considered, due to the
internal ones are balancing each other.

Fig. 3.1

The elementary (differential) mass forces acting on a fluid element is


expressed by the Equation:
r
r
r
dFm = fm dm = fm dV
r
where: fm

(3.1)

specific mass force (unitary mass force); it has dimension of acceleration; generally,
r
r
for fluids on the gravity field of Earth fm = g (gravity acceleration).

The resultant mass force is:


r
r
Fm = fm dV
V

(3.2)

The surface forces represent the result of the interaction of fluids with surrounding environment
(other substances, either solids or fluids), through the surface S :
r
r
(3.3)
dFS = fS dS
r
specific surface force (unitary surface force); it depends of surface position vector
where: fS

24

r
r
r , normal unitary vector on surface n (pointed outwardly from the surface) and
r
r r r
time t : fS = fS ( r , n , t )

Generally, a surface force will be at an angle to the surface, thus having both normal and
r
r
tangential components: dFn , respectively dF :
r
r
r
r
(3.4)
dFn = dFS cos = fS dS cos = fn dS ,
r
r
r
r
(3.5)
dF = dFS sin = fS dS sin = f dS ,
r
normal unitary surface force; it defines the pressure p on the surface dS :
where fn
r
r
fn = p n
r
f
tangential unitary surface force; it defines the shear stress on the surface dS ;
3.2 THE EULERs EQUATION FOR FLUID STATICS
The equation for fluid statics can be obtained from the equilibrium of the forces on an arbitrary
elementary volume of fluid:
r
r
Fm + Fp = 0 .

(3.6)

r
If the fluid is at rest the tangential component of the surface force is zero dF = 0 , thus the

resultant surface force is:


r
FS =

r
p n dS .

(3.7)

In this way Equation (3.6) becomes:


r
r
Fm + FS = 0

dV +

p n dS = 0 .
S

(3.8)

Using first Gauss theorem:


r

p n dS = p dV .

(3.9)

The vector differential operator is (in Cartesian coordinates Oxyz ):


r
r
r
+k
+ j
=i
z
y
x

(3.10)

r r r
where: i , j , k are the unitary vectors on Ox , Oy and Oz directions.

From Equations (3.6) and (3.7):


r

f m dV p dV = 0

(3.11)

If the volume V 0 , then previous Equation reduces to:


r
r
r
1
1
fm p = 0 fm p = 0 fm grad p = 0

25

(3.12)

Equation (3.12) represents the Eulers equation for fluid statics in vectorial form. In Cartesian
coordinates it can be writing as:
1 p
=0
x
1 p

=0
y
1 p

=0
z

fmx
fmy
fmz

(3.13)

r
where fmx , fmy , fmz are the scalar components of fm .
r
r
Multiplying the equation (3.12) by dr and taking into account that p dr is the total differential of

pressure:
r p
p
p
p dr =
dx +
dy +
dz = dp ,
x
y
z

(3.14)

we obtain the Euler s Equation of fluid statics (sum of two differential continuous functions):
r
r 1
1
fm dr = dp dp + d = 0 ,
(3.15)

where ( x , y , z ) is the potential of the mass forces (the potential energy due to the mass forces).
If the components of unitary mass forces are known, then:
f mx =

U
U
U
, f my =
, fmz =
,
z
y
x
r

= d = fm dr = (fmx dx + fmy dy + fmz dz ) ,


r

(3.16)

By integration, Equation (3.15) leads to:


dp

+ = ct. .

(3.17)

The Equation (3.17) is known also as the fundamental Equation of fluid statics. It describes the
pressure field in fluids, either compressible or incompressible.
The first term,

dp

, represents the potential of the pressure forces. In order to solve this integral,

the relationship between pressure and density must be know for gasses.
For fluids having the density as function of pressure, the equipotential surfaces = ct represent
also surfaces of constant pressure, such of those which are correspond to the interfaces to the
fluids. As consequence, the free surfaces of fluids are surfaces of constant pressure. The shape
r
and orientation of these surfaces depend on the unitary mass force fm . They are normal
r
(perpendicular) on fm .

26

3.3 PARTICULAR FORMS FOR THE FUNDAMENTAL EQUATION OF STATICS


3.3.1 Light fluids
For fluids having small mass density, as gasses and vapours, the effect of their own weight can be
neglected. We can consider that the potential of mass forces is null: d = 0 . From the Equation
(3.15) we obtain that:
1

dp = 0 p = ct . ,

(3.18)

Thus, the pressure in a finite mass of light fluid is constant.


3.3.2 Fluids on gravitational field of Earth
For a Cartesian reference as in figure below, with the Oz in the sense of the altitude rising (natural
r
in study of the Earth atmosphere), the scalar components of the specific mass force fm acting on a
fluid are the following:
fmx = fmy = 0
= g dz = g z + ct .
fmz = g

In the case of liquids, fluids with constant density = ct . , the potential of the
pressure forces is:
dp

dp =

+ ct .

(3.19)

Fig. 3.2

Thus, the Equation for incompressible fluids at rest can be written as:
p

where:

+ g z = ct p + g z = ct p + z = ct.

(3.20)

is the specific gravity of the fluid;

This is fundamental Equation for hydrostatics. It gives the pressure


distribution in a liquid, which is a linear relationship between depth
(measured from the free surface) and pressure for liquids.
The integral constant ct . must be computed knowing the boundary
conditions. In the case of a liquid in an open tank (see Figure 3.3), at
the level of the free surface, z = h , the pressure is the atmospheric
one p0 . For this case, the Equation (3.20) gives the integral
constant:
z = h p0 + g h = ct p0 + h = ct

(3.21)

Substituting the constant into Equation (3.20) we found that:


p p0 = ( h z )

(3.22)
27

Fig. 3.3

where: p p0

is the gauge pressure; in fluid mechanics we are dealing generally with


relative (gauge) pressure, except for the thermodynamic processes of
gasses. In the following the gauge pressure will be denoted simply as p .

hz

is the depth, measured from the free surface.

For the shown case, pressure has the maximum magnitude on the bottom of the tank:
pmax = h .

(3.23)

For the previous case, if the pressure at the surface is pm , as shown in Figure 3.4 (liquid in a
closed tank, under pressure) in a reference with the Oz in the sense of the depth rising (natural in
study of the liquids), then:
pmax = pm + h .

(3.24)

Equation (3.23) expresses the Pascals principle, which


state that any externally applied pressure is transmitted
undiminished in an enclosed static fluid. This is making
possible a large multiplication of pressure force and has
many practical applications, such as hydraulic pressure,
automatic transmissions used in automobile, brakes,
power steering a.o.
The graphical representation of the pressure variation on
the walls of the tank is known as the pressure diagram.
The plane where the (gauge) pressure is zero is called
the manometric plane and its level, from the free surface,
is defined by manometric high:

Fig. 3.4
hm =

pm

(3.25)

Thus, Equation (3.24) can be written as:


pmax = (hm + h ) .

(3.26)

3.3.3 Relative equilibrium of liquids


In the previous section, the equilibrium of liquids was treated in absolute sense. Two other cases of
static equilibrium of liquids, having technical applications, are presented in the following:

liquids in tanks on accelerating translation, applicable in transport of liquids in large


reservoirs;

liquids in (cylindrical) tanks on rotational motion, applicable in centrifugal casting, design


of impellers a.o.
28

Both cases are concerning the relative equilibrium of liquids, in accelerating systems, attached on
carrying tanks.
3.3.3.1 Liquids in tanks on accelerating translation
For this case, lets consider a liquid, having the specific gravity , in an open tank by length l , up
to the level h (see Figure 3.5), which start to move with the acceleration a = ct .

Fig. 3.5

When the tank accelerates, the specific mass force fm will have two components on considered
reference system:

along Ox , inertial acceleration fmx = a ;

along Oz , local gravitational acceleration fmz = g .

The free surface of liquids will change its orientation under the action of the mass forces. It comes
to the equilibrium at an angle with the horizontal:

= arc tg

a
g

(3.27)

The potential of the mass forces is, from Equation (3.16):

= ( fmx dx + fmz dz ) = ( a dx + g dz ) = a x + g z + ct

(3.28)

Also, the potential of the pressure forces for liquid ( = ct ) is:


dp

dp =

+ ct

(3.29)

Thus, the Equation for the relative equilibrium of liquids in accelerating translations is:
dp

+ U = ct

+ a x + g z = ct

The integral constant ct . can be computed from the following condition:


l

2 p = p0 , (atmospheric pressure).
z = h
x=

29

(3.30)

In this way:
ct = a

l
+gh
2

(3.31)

Replacing (3.31) into (3.30), the Equation for the relative equilibrium of liquids in accelerating
translations becomes (using the gauge pressure):
l

= a x + g ( h z )

(3.32)

The maximum magnitude of pressure is obtained for x = z = 0 :


l

pmax = a + g h .
2

(3.33)

Figure 3.6 shows the pressure diagram on the walls of the tank.

Fig. 3.6
3.3.3.2 Liquids in cylindrical tanks on rotational motion
For this case, lets consider a liquid, having the specific gravity , in an open cylindrical tank by
radius R , up to the level h (see Figure 3.7), which start to rotate
with the angular speed = ct .
In this case the specific mass force fm will have components on
each direction of the reference system:

component of the centrifugal acceleration on Ox and Oy :


fmx = x 2 and fmy = y 2 ;

along Oz , local gravitational acceleration fmz = g .

From Equation (3.16), the potential of the mass forces is,

= ( fmx dx + fmz dz ) =

(3.34)

= x 2 dx + y 2 dy + g dz =
=

x2 + y 2
r2
+ g z + ct =
+ g z + ct
2
2

Fig. 3.7

30

Also, the potential of the pressure forces for liquid ( = ct ) is:


dp

+ ct

(3.35)

Thus, the Equation for the relative equilibrium of liquids in on rotational motion is:
p

r2 2
+ g z = ct
2

(3.36)

The integral constant ct . can be computed knowing that the (gauge) pressure is zero on the free
surface, which is a parabolic type. Thus, from the condition Vinitial = Vfinal
1
2

R 2 h = R 2 hmax R 2 ( hmax hmin ) 2 h = hmax + hmin

(3.37)

Also p = 0 for:
r =R
R2 2
+ g hmax = ct

z = hmax
2

(3.38)

r =0

g hmin = ct
z = hmax

(3.39)

From the system of Equations (3.37), (3.38) and (3.39) we found that:
ct =

R2 2
+gh
4

(3.40)

Replacing (3.40) into (3.36), the Equation for the relative equilibrium of liquids on rotational motion
becomes:
p = g (h z)

2
4

(R2 2 r 2 )

(3.41)

The maximum magnitude of pressure is obtained for r = R and z = 0 :


p= gh+

2R2
4

Figure 3.8 shows the pressure diagram on the walls of the tank.

Fig. 3.8

31

(3.42)

3.3.4 Variable-density pressure fields


If the density is not a constant, in order to compute the potential of the pressure forces, the
relationship between pressure and density must be known. In the following it is presented the
solution for an isothermal process, T = ct :
p

= ct =

p0

p0

(3.43)

Thus:
dp

0 dp
p0

0
p0

ln p + ct

(3.44)

For fluids on gravitational field of Earth:

= g z + ct

(3.45)

ln p + g z = ct

(3.46)

Finally:

0
p0

The solutions for adiabatic or polytropic process can be found similarly.

32

4. FLUID FORCES ON SUBMERGED SURFACES


Engineers interested in the storage of fluids must be concerned with the strength of the container
whether it be a dam, a tank, or a pressure vessel. It is not the purpose of this text to consider the
design of such structures or the gates and valves associated with them, but some knowledge
about the resultant force that the fluid exerts is essential to the designer. Its determination is the
subject of this section.
For purposes of stress analysis, a state of zero stress exists when a fluid is subject only to the
uniform pressure of the atmosphere. It is therefore desirable to use gage pressures whenever
possible in this section.
Consider a surface of arbitrary shape subject to fluid pressure (see Figure 4.1). The surface might
be the side of a tank, the face of a dam, the hull of a ship, or even an imaginary surface cut from
within a body of fluid. At any point on this surface we select an element of area dA having the
r
normal unitary vector n (drawn outwardly from the surface, and thus into the fluid). Acting upon
dA is a differential force that is given by:
r
r
dF = p n dA ,

(4.1)

The negative sign appears because the force exerted by the fluid is directed into the surface, and
r
opposite in sense to n .

Fig. 4.1 Pressure force on elementary area dA


r
The resultant of all the differential forces on the surface can be represented by a force F passing
r
through a conveniently chosen reference point, and a couple M .

The force is obtained by integrating the differential forces over the surface area A .
r
r
F = p n dA ,
A

(4.2)

r
r
r
Each differential force has a moment about the reference point, O , given by dM o = r dF . The
r
moment of the couple is obtained by the integration of dM :

33

r
r
r
r
v
M o = r dF = r p n dA ,
A

(4.3)

r
where r is the position vector locating dA with respect to the moment center.

In evaluating the above integrals, it must be kept in mind that we are summing vector quantities.
The operation of integration must therefore be carried out on a set of scalar components of the
vectors (the pressure distribution in fluid must be known).

4.1 FLUID FORCES ON PLANE SURFACES


r
In the cases of the plane surfaces the normal unitary vector is constant, n = ct , and the Equations

(4.2), (4.3) become:


r
r
r
F = p n dA = n p dA ,

(4.4)

r
r
r
v
v
M o = r p n dA = n r p dA ,

(4.5)

The point where resultant force will act is calling the centre of pressure and has as symbol C (or
r
CP ). The position vector of the centre of pressure, rCP , can be obtained from Varignons
Theorem: the sum of the moments of each elementary force is equals with the moment of the
resultant force:
r

r
r
r
r
r
v
v
v r
n r p dA = rCP F n r p dA = n rCP p dA rCP =
A

p dA

p dA

(4.6)

4.1.1 Resultant force and center of pressure on a submerged plane surface in a gas
For gasses, having a finite volume, the value of pressure within can be considered constant:
p = ct . Equations (4.4) and (4.6) become:
r
r
r
r
F = n p dS = n p dA = n p A F = p A ;
A

r
rCP =

r
p r dA

r
rG A r
=
=
= rCG ;
A
p dA
p dA

where:

p dA

(4.7)

(4.8)

r
rCG is the position vector of the fluid centroid.

These results are also applicable for small volumes of liquids, if the pressure variation is negligible
insight them.

34

4.1.2 Resultant force and center of pressure on a submerged plane surface in a liquid
For this case, it is considered that the plane surface is totally submerged in a liquid of density
and inclined at an angle of to the horizontal (see Figure 4.2). Taking pressure as zero at the
level of free surface and measuring down from the surface, the (gauge) pressure on an element
dA , submerged at distance h , is given by the fundamental Equation of hydrostatics:
p= h

(4.9)

Fig. 4.2 Pressure force on a submerged plane surface in a liquid


By replacing of p in Equations (4.4) and (4.6) the followings are obtained:
r
r
r
r
r
F = n p dA = n h dA = n z sin( ) dA = n sin( ) z dA ,
A

where:

z dA

(4.10)

is the 1st moment of inertia for area about the axis oy :

z dA = zCG

A,

(4.11)

where:

A is the area of the submerged surface;


zCG is the position for centroid of the surface along Oz . Hence:
F = A zCG sin( ) = hCG A ,

where:

(4.12)

hCG is depth to the centroid of the submerged surface, from Figure 4.2.
r
r
r
r
r p dA r h dA r z dA r z dA

r
,
rCP = A
= A
= A
= A
zCG A
p
dA
h
dA
z
dA

(4.13)

Thus, the coordinates of pressure centre in the yoz plane are:


y CP =

where:

y z dA

zCG A

I yz
zCG A

; zCP =

dA

zCG A

Iy
zCG A

I yz

is the product of inertia about the axes Oy and Oz ;

Iy

is the 2nd moment of inertia about the axis Oz .

Note that:

xoy is the manometric plane (where the gauge pressure is null, p = 0 ) ;

35

(4.14)

If yOz is a symmetry plane, than I yz = 0 and y CP = 0 .

In a reference system having the origin in the centroid CG of the surface, the Equations
(4.14) can be write in the following form (using the Steiners theorem):
y CP =

where:

I yz
zCG A

I' yz + y CG zCG A
zCG A

; zCP =

Iy
zCG A

2
I' y + zCG
A

zCG A

I' yz

is the product of inertia about the axes CGy and CGz ;

I' y

is the 2nd moment of inertia about the axis CGz .

(4.15)

Note that in the Equation (4.15) the 2nd moment is always positive so that the center of pressure
CP always falls below the centroid CG .

4.2 FLUID FORCES ON CURVED SURFACES


As stated above, if the surface is curved, the forces on each element of the surface will not be
parallel and must be combined using some vectorial method. It is most convenient to calculate the
horizontal and vertical components and combine these to obtain the resultant force and its
direction. Thus, the components of the resultant pressure forces are given by the following
Equations:
Fp x =

p dS yOz

(4.16)

p dS xOz

(4.17)

p dS xOy

(4.18)

SyOz

Fp y =

S xOz

Fp z =

S xOy

where:

S xOy , S yOz , S xOz are the algebraic projections of the curved surface on the planes

of the reference system, xOy , yOz and xOz ;


The coordinate of the pressure center on Ox axis is:
r
r p dS yOz
r
SyOz
rC Fx =
p dS yOz

(4.19)

SyOz

The other two coordinates are computed similarly.


4.2.1 Pressure forces on a submerged open curved surface in light fluids
For gasses, having a finite volume, the value of pressure within can be considered constant:
p = ct . Equations (4.16) (4.19) become:
Fp x = p dS yOz ; Fp y = p dS xOz ; Fp z = p dS xOy

36

(4.20)

dS yOz
SyOz
r
r
r
r
r
r
rC Fp x =
= rG SyOz ; rC Fp y = rG SxOz ; rC Fp z = rG SxOy
S yOz

(4.21)

4.2.2 Pressure forces on a submerged open curved surface in liquids


The pressure insight a liquid is given by p = z . In a reference system having the xOy as
manometric plane and the axis Oz in the sense of depth rising, the Equations (4.16) (4.19)
become:
Fp x =

z dS yOz

SyOz

= zG SyOz S yOz ; Fp y = zG SxOz S xOz ; Fp z =


r

r
rC Fp x =

where:

zGyoz

z dS yOz

S yOz

zG SyOz S yOz

r
; rC Fp y =

z dS xOz

S xOz

zG SxOz S xOz

z dS xOy

= V

SxOy

r
r
; rC Fz = rG V

(4.22)

(4.23)

is the depth of centroid for S yoz about the manometric plane xoy (where
the gauge pressure is null, p = 0 ) ;

zGxoz

is the depth of centroid for S xoz about the manometric plane;

is the volume of the liquid between wetted curved surface and its projection
on the manometric plane xoy .

These results can easily understand considering a liquid at rest on top of a curved surface AC
(two-dimensional case), as shown in the diagram below.

Fig. 4.3 Pressure force on a submerged curved surface in a liquid


The element of fluid ABC is at equilibrium (as the fluid is at rest). As a consequence the horizontal
force on AC plane must be equal and in the opposite direction to the resultant force Fy on the
curved surface AB . As AC is the projection of the curved surface AB onto the vertical plane
xOz , it can be generalises this to:

The resultant horizontal force of a fluid above a curved surface is equal with the resultant
force on the projection of the curved surface onto a vertical plane.

We know that the force on a vertical plane must act horizontally (as it acts normal to the plane) and
that Fy must act through the same point. Generalising:

37

The resultant horizontal force of a fluid above a curved surface acts horizontally through the
centre of pressure of the projection of the curved surface onto a vertical plane

Concerning the vertical force, there is no shear force on the vertical edges, as consequence the
vertical component Fz can only be due to the weight of the fluid. Generalising:

The resultant vertical force of a fluid above a curved surface is equal with the weight of fluid
directly above the curved surface. It will act vertically downward through the centre of gravity
of the mass of fluid.

The overall resultant pressure force is found by combining vectorialy its components:
F = Fx2 + Fy2 + Fz2

(4.24)

The angle the resultant force makes to the horizontal is:


Fy
Fz

= arctg

(4.25)

4.2.3 Resultant force on a closed curved surface for a constant pressure fluid
For this case, the resultant force on each direction will be zero: Fx = Fy = Fz = 0 (the algebraic
projections of the curved surface on the planes of the reference system are nulls). But the action of
the fluid has as result stresses on the wall of tanks. Knowing the magnitude of these tensions is
useful on computation of the thickness of wall.
For a circular duct, see Figure 4.4, the following Equation is found for computation of the thickness
of wall, , from the balance of the forces of pressure and strength on Oz direction:
pDL = 2 L

=
where:

pD
2 a

is the pressure of fluid;

is the inner diameter of duct;

is the allowable stress characteristic for the material of duct.

Fig. 4.4 Pressure force on a closed curved surface for constant pressure fluid
38

(4.26)
(4.27)

4.3 FLOATING BODIES


4.3.1 Buoyancy of floating bodies. Archimedes principle
The vertical component of the fluid static force on a submerged or floating body is called the
buoyant force (see Figure 4.5).

Fig. 4.5 Buoyancy of floating bodies


About 250 B.C. Archimedes discovered that the buoyant force on a solid body is equal in
magnitude to the weight of the fluid displaced by body.
Thus, the equation of the equilibrium of the forces on a body of specific weight b floating on a
fluid having specific weight f is:
FW = FA b Vb = f Vfd

where:

FW

is the weight of the solid;

FA

is the buoyant force (Archimedes force);

Vb

is the volume of the solid;

Vb

is the volume of the displaced fluid.

(4.28)

The necessary condition of the floating is:

b < f

(4.29)

4.3.2 Stability of floating bodies


The concept of stability relates to the question of whether or not a floating body returns to its initial
state when its equilibrium is disturbed. This is a very important concept in the design of the ships.
The stability of the floating bodies is described by the following equations:

Stable equilibrium:
Ix
e >0
Vdf

(4.30)

Ix
e =0
Vdf

(4.31)

Neutral equilibrium:

39

Unstable equilibrium:
Ix
e <0
Vdf

where:

(4.32)

is the distance (eccentricity) between reference pressure center (at


equilibrium) and the centroid.

Ix

is the 2nd moment of inertia of floating area A .


I x = y 2 dA

(4.33)

4.4 STATIC PRESSURE MEASUREMENT BY MANOMETERS


A difference in pressure, if is not too large, can be readily measured by balancing it against a
column of liquid in static equilibrium. An instrument for doing this is known as a manometer (also
know as a liquid gauge). The commonly used are the following types:
4.4.1 The piezometer tube manometer
The simplest manometer is a tube, open at the top, which is attached
to a vessel containing liquid under pressure (higher than atmospheric).
An example can be seen in the Figure 4.6. This simple device is
known as a piezometer tube. As the tube is open to the atmosphere
the pressure measured is relative to atmospheric so is a gauge
pressure.
This method can only be used for liquids (not for gases) and only
when the liquid height is convenient to measure. It must not be too
small or too large and pressure changes must be detectable

Fig. 4.6
The piezometer tube

Pressure at A is the pressure due column of liquid above A:

p A = f g h A [m column of fluid] .

(4.44)

The pressure can be measured in multiple points simultaneously if several piezometer tubes are
connected as in Figure bellow,

Fig. 4.7 Multiple piezometer tubes

40

4.4.2 The U-tube manometer

Using a U-tube, the pressure of both liquids and gases can be measured with the same instrument.
The U-tube is connected as in the figure 4.8 and filled with a fluid called the manometric fluid (or
manometric liquid). The fluid whose pressure is measured should have a mass density less than
that of the manometric fluid and the two fluids should not be miscible. For the manometer shown in
Figure 4.8 the following Equation can be written:

Fig. 4.8 - The U-tube manometer

pressure in a continuous static fluid is constant at any horizontal level:


pB = pC ;

for the left hand arm:


pB = p A + f g h1 ;

for the right hand arm:


pC = lp g h2

Thus the (gauge) pressure in point A is:


p A = lp g h2 f g h1 .

(4.45)

Note that if the fluid being measured is a gas, the density will probably be very low in comparison
to the density of the manometric fluid, i.e. lp >> f . In this case the term f g h1 can be
neglected (because h1 << 1 ), and the gauge pressure can be computed accurately enough as:
p A = lp g h2 .

(4.46)

4.4.3 Measurement of pressure difference using a U-tube manometer.


If the U-tube manometer is connected to a pressurised vessel at
two points the pressure difference between these two points can
be measured. If the manometer is arranged as in the figure below,
then:

pC = pD ;
pC = p A + f g h1 ;
pD = pB + f g ( h2 h ) + lp g h ;
p A + f g h1 = pB + f g ( h2 h ) + lp g h ;

giving the pressure difference:

Fig. 4.9
41

p A pB = f g ( h2 h1 ) + ( lp f )g h .

(4.47)

Also, if the fluid whose pressure difference is being measured is a gas and lp >> f , then the
terms involving f can be neglected:
p A pB = lp g h .

(4.48)

4.4.4 Advanced U-tube manometer.


The U-tube manometer has the disadvantage that the change in height of the liquid in both sides
must be read. This can be avoided by making the diameter of one side very large compared to the
other. In this case the side with the large area moves very little when the small area side move
considerably more.

Fig. 4.10 - Advanced U-tube manometer


Assume that the manometer is arranged as above, to measure the pressure difference of a gas
(with negligible density) and that pressure difference is p1 p2 ( p1 > p2 ). The datum line indicates
the level of the manometric liquid when the pressure difference is zero. The volume of liquid
transferred from the left side to the right is:
V = h1

D2
4

= h2

d2
4

The fall in level of the left side is:


2

d
h1 = h2 .
D

The height different in the two columns gives the pressure difference:
2

d 2
d
p1 p2 = lp g h2 + h2 = lp g h2 1 + .
D

(4.49)

If D is much larger than d ( D >> d ) then ( d / D ) 2 is very small ( d / D ) 2 << 1 , so:


p1 p2 = lp g h2 .

(4.50)

If the pressure to be measured is very small then tilting the arm provides a convenient way of
obtaining a larger (more easily read) movement of the manometer liquid. A manometer with a tilted
arm (inclined manometer) is shown in the figure below.

42

Fig. 4.11 - Manometer with a tilted arm


The pressure difference is given by the height change of the manometric fluid but by placing the
scale along the line of the tilted arm and taking this reading large movements will be observed. The
pressure difference is then given by:
p1 p2 = g h = g l sin .

(4.51)

The sensitivity to pressure change can be increased by a greater inclination of the manometer arm.
Alternatively, the density of the manometric fluid may be changed.
4.4.5 Choice Of Manometer
Some disadvantages of manometers:

Slow response they are useful for very slowly variations of pressure.

For the U-tube manometer two measurements must be performed simultaneously to get the
difference in high value. This may be avoided by using a tube with a much larger crosssectional area on one side of the manometer than the other one.
It is often difficult to measure small variations in pressure; alternatively an inclined

manometer may be used.


Some advantages of manometers:

They are very simple.

No calibration is required - the pressure can be calculated from first principles.

43

4.5 FLUID STATICS EXAMPLES


WORKED EXAMPLES
Exercise 1
Determine the pressure intensity at A ( Pa and mm H 2O ) if the pressure at B is pB = 0.1 at ,
h1 = 0.2 m , h2 = 0.1 m , h3 = 0.4 m , 0 = water = 1000 kg/m 3 and 2 = oil = 800 kg/m 3 .

Fig. 4.12
SOLUTION:
Step 1:
Convert all the quantities in the International System of Units (if is necessary)
pB = 0.1 at = 0.1 9.81 10 4 Pa = 9.81 10 3 Pa ( = 10 3 mmH 2O )
1 mmH2O = 1000 9.81 10 3 = 9.81 Pa

Step 2:

p A 0 g (h1 + h2 + h3 ) = pB 0 g h2 1 g h3
p A = pB + 0 g ( h1 + h3 ) 1 g h3
p A = 9.81 10 3 + 1000 9.81( 0.1 + 0.2 ) 800 9.81 0.4 = 12556.8 Pa
p A = 1280 mmH 2O

44

Exercise 2
A closed tank, having the shape as in Figure below, is holding water under the pressure pm
(manometric pressure).

Fig. 4.13

The following are known:

H = 1 .5 m

depth of water in tank

R = 0 .5 m

the curvature of the thank

L = 1 .0 m

width of the tank

= 1000 kg/m 3

mass density of the water

pm = 0.1 at

(technical atmospheres)

Draw the pressure diagram on the wetted walls of the tank.

Find the magnitude of the force on the vertical wall ( AB )

Find the magnitude of the resultant force on the curved wall ( BC ) .

SOLUTION:
Step 1:
Convert all the quantities in the International System of Units (if is necessary)
pm = 0.1 at = 0.1 9.81 10 4 Pa

Step 2:
Find the level hm of the (horizontal) manometric plane above the free surface of the liquid.
At the level of the manometric datum the absolute pressure is the atmospheric one, respectively
the gauge pressure is zero, p = 0 , thus:
pm = g hm

45

hm =

pm 0.1 9.81 10 4
=
=1 m
g
1000 9.81

Chose the reference system xOyz that:


xoy

is the manometric plane;

oz

in the sense of the depth increasing.

Fig. 4.14
Step 3:
Compute the (gauge) pressure at the level of each characteristic point:
p A = g hA = g hm = pm = 9.81 10 3 Pa (pressure in gasses is constant)
pB = g hB = g [hm + ( H R )] = pm + g ( H R )
pB = 9.81 10 3 + 1000 9.81 (1.5 - 0.5) = 19.62 10 3 Pa
pC = g hC = g (hm + H ) = pm + g H
pC = 9.81 10 3 + 1000 9.81 1.5 = 24.53 10 3 Pa

Draw the pressure diagram on the walls of thank, as shown in Figure 4.15.

46

Fig. 4.15 - Pressure diagram on the walls of thank


Step 4:
Compute the magnitude of the force on the (plane) wall ( AB ) , F( AB ) . Denote H R = h (see
Figure 4.16).
4.1 Using the integral Equation:
F( AB ) = p dA = ( pm + g z) dA = pm dA + g z dA = pm dA + g z dA
A

F( AB ) = pm A(AB)

Lh

z
+ g z dz dx = g hm h L + g
0 2
00

F( AB ) = g hm h L + g

dx = g hm h L + g 2 dx
0
0

h2
h

L = hm + h L
2
2

4.2 Using Equation (4.12):


h

F( AB ) = hCG (AB) A(AB) = hm + h L = 9810 (1 + 0.5 )1 1 = 14715 N


2

Note that the (pressure) force is simply the volume of the pressure distribution. It acts through the
centroid of the pressure diagram.

47

Step 5:
Compute the magnitude of the force on the (curved) wall ( BC ) : F( BC ) . Spit the resultant force into
its components, horizontal Fy ( BC ) and vertical Fz( BC ) (see Figure 4.16). According with Equations
(4.16):
R

Fy = hGxoz S xoz = hm + H R L = 9810 (1 + 1.5 - 0.25) 0.5 1 = 11036.25 N


2

R2

0.5 2

+ hm + h L = 9810
+ 1 + 1 1 = 21546.19 N
Fz = V =
4

where:

S xoz

is the projection of the curved area ( BC ) on the datum plane xoz ;


S xoz = R L

is the volume of the liquid between wetted curved surface and its projection
on the manometric plane xoy (where the gauge pressure is null).
R2

+ ( hm + h ) R L .
V =
4

Hence:

F = Fy2 + Fz2 = 24208.20 N .

Fig. 4.16

48

Exercise 3
The density of liquids can be determined experimentally using a hydrometer as shown below. It
consists on a weighted bulb and an attached tube whit a
calibrated cross section. Compute the density of a fluid f
if the stem is submerged to a depth of h = 50 mm
relative to the equilibrium position in water.
The followings are known:
G = 20 gf :

weight of hydrometer;

d = 8 mm :

diameter of the calibrated section;

w = 1000 kg/m 3 :

density of the water

Solution:
Step 1

Fig. 4.17

Convert all the quantities in the International System of


Units (if is necessary)
G = 20 gf = 9.81

m
s2

0.02 kg = 0.1962 N ;

d = 8 mm = 8 10 -3 m .

Step 2
The buoyancy principle is the basis of the hydrometer. The Equations of the equilibrium for the
floating situations are:
G = FAw G = w g Vdw

in water:
in other liquid:

d2
G = FAl G = l g Vdl = l g Vdw +
h

Thus:

l =

d
g Vdw +
h

d
g
h
+
g
4
w

l = w

l = 1000

= w

4G
4 G + w g d 2 h

4G
4 G + w g d 2 h

4 0.1962
4 0.1962 + 1000 9.81 3.1415 0.008 2 0.05

l = 888.334

49

kg
m3

SELF-ASSESSMENT EXERCISES
Exercise 4
An inclined tube manometer consists of a vertical cylinder of 35 mm diameter. At the bottom of this
is connected a tube 5 mm in diameter inclined upward at an angle of 15 to the horizontal. The top
of the vertical tube is connected to an air duct. The inclined tube is open to the air and the
manometric fluid has relative (to the water) mass density 0.785. Determine the pressure in the air
duct if the manometric fluid moved 50 mm along the inclined tube. What is the error if the
movement of the fluid in the vertical cylinder is ignored? Make a sketch.
Exercise 5
A U tube-manometer (see Figure 4.18) is used to measure the acceleration of a motorcar.
Compute the magnitude of the acceleration if the deflection of the fluid is h = 30 mm . Draw the
calibration curve of the accelerometer.

Fig. 4.18
Exercise 6
A hydraulic tachometer consists on a U tube-manometer (see Figure 4.19) which is used to
measure the speed (revolutions per minute). Compute the speed if the deflection of the fluid is

h = 60 mm . Draw the calibration curve of the tachometer.

Fig. 4.19

50

Exercise 7
A closed tank (see Figure 4.20), filled with gasoline, is carried by a truck. Compute the pressure
force Fp on the back wall of the tank if the truck has acceleration a = 2 m/s 2 . Draw the pressure
diagram on the walls of the tank. The followings are known: h = 1.8 m , b = 1.6 m , l = 6 m ,

gasoline = 0.78 gram/cm 3 .

Fig. 4.20
Exercise 8
Find the magnitude of the resultant force on the vertical wall of a tank that has oil, of relative
density 0.8, floating on water as shown in Figure 4.21. Draw the pressure diagram. The width of
the tank is L = 1.0 m .

Fig. 4.21
Exercise 9
A cylinder, 100 mm in diameter, 250 mm long and mass density = 800 kg/m 3 , floats in water
with vertical axis. Determine the stability of the cylinder.

51

5. IDEAL FLUID DYNAMICS


The topics of this lecture are concerning with the analysis of the ideal fluids in motion - fluid
dynamics. The majority of the engineering problems involving fluids occur in flow processes:
y

the air flow over a road vehicle;

the lift of an airplane on the flow of air over its wings;

the trust of a rocket engine depends on the flow of gasses through a nozzle;

the cooling of the electronic devices depends on the air flow over their components, a.o.

5.1 DESCRIPTIONS OF THE FLOW FIELDS


The motion of fluids can be predicted in the same way as the motion of solids are predicted,
respectively using the fundamental laws of physics together with the physical properties of the
fluid. But, when a fluid flow, the elements of fluid do not maintain their relative position. A complete
description of their motion is therefore inherently more complex than a description of the rigid
bodies motion. It is not difficult to envisage a very complex fluid flow as:
y

wake behind a car,

waves on beaches,

hurricanes and tornadoes, or any other atmospheric phenomenon.

All are example of highly complex fluid flows. Fortunately, we are not generally concerned with the
motion of the individual particles (the Lagrangian approach). It usually suffices to describe the
velocity associated with a point in a specified area of flow field. When the velocity is specified for
each point in a region, a velocity field is defined. Mathematically it can be represent by the
Equation:
r
r
r
r
v =ui +vj +wk

where:

(5.1)

r r r
i , j, k

are the unit vectors in the x , y and z directions;

u, v, w

are the scalar components of the velocity ( v 2 = u 2 + v 2 + w 2 ); the


latter can by expressed by the field equation:
u = u (x, y, z, t)

v = v (x, y, z, t) v = v ( x, y, z, t)
w = w (x, y, z, t)

(5.2)

The use of velocity fields to describe fluid motion was introduced by Euler (XVIII century), who is
generally credited with the founding of the fluid dynamics. The Eulerian method may seem rather
complex, yet the alternative Lagrangian method of describing the motion of each fluid element as a

52

function of time is much more complex. The simplification achieved by the eulerian method is most
apparent in problems of steady flow.
A steady flow is one in which the conditions (velocity, pressure and cross-section) may differ from
point to point but do not change with time. (in practice, there is always a slight variations in velocity
and pressure, but if the average values of these are constant, the flow is considered steady).
In contrast with this type of flow is the unsteady flow, when the conditions are changing with time at
any point in the fluid.

5.1.1 Streamlines
Graphical representations of physical phenomena are generally helpful in gaming insight into
problems and in communicating with others. This is particularly true when the mathematical
description of a phenomenon is complex. A very useful graphical representation of fluid flow is
based on the concept of a velocity field. Lines are drawn in such a way that the tangent at any
point on a line indicates the direction of the velocity associated with that point. These lines are
called streamlines, and a family of them constitutes a streamline pattern (Fig. 5.1).

Fig. 5.1 Flow pattern around an airfoil


A streamline pattern indicates the direction of the velocity at every point in the field at a given
instant. In steady flow, the streamline pattern is stationary with the frame of reference. It is also
stationary in unsteady flow if only the magnitude of the velocity is changing with time. Furthermore,
the steady flow streamlines represent the path lines for moving particles. Generally, in unsteady
motion, a fluid particle will not remain on the same streamline, and hence the streamlines and path
lines do not coincide.
The differential equations for a streamline are obtained by noting that:
u=

dx
dy
dz
; v=
; w=
.
dt
dt
dt

(5.3)

Hence:
dx
dy
dz
=
=
.
u( x , y , z , t ) v( x , y , z , t ) w( x , y , z , t )

(5.4)

r
r
Alternatively, since the velocity vector is tangent to a streamline, then v ( u,v, w ) || dr ( dx ,dy ,dz ) , or:
r
r
(5.5)
v dr = 0

53

Practically, the injection of dye or smoke into a steady flow can make the streamline pattern visible,
as shown in Figure 5.2.

Fig. 5.2 Flow visualisation around an airfoil


But no matter how carefully the injection is made, a streamline pattern is not always revealed.
When none is found, measurements with sensitive instruments indicate that the flow contains
irregular, high frequency fluctuations and is therefore not truly steady. This type of flow is known as
turbulent flow. With the use of the mean velocity, stationary streamline patterns can be drawn for
turbulent flows.
Some things to know about streamlines:
y

Because the fluid is moving in the same direction as the streamlines, fluid can not cross a
streamline.

Streamlines can not cross each other. If they were to cross this would indicate two different
velocities at the same point. This is not physically possible.

The above point implies that any particles of fluid starting on one streamline will stay on that
same streamline throughout the fluid.

5.1.2 Stream-tube control volume


A useful technique in fluid flow analysis is to consider only a part of the total fluid in isolation from
the rest. This can be done by imagining a tubular surface formed by streamlines along which the
fluid flows. This tubular surface is known as a stream-tube (see Figure 5.3).

Fig. 5.3 A 3D stream-tube


The "walls" of a stream tube are made of streamlines. As we have seen above, fluid cannot flow
across a streamline, so fluid cannot cross a stream tube wall. The streamlines are boundaries in
the same sense that real conduit walls are boundaries. Conversely, the boundaries of the real
conduit or any immersed solid are streamlines. Hence, a stream tube can often be viewed as a
solid walled pipe.
54

5.2 ACCELERATION IN A FLOW FIELD


Before we can relate forces and motion within a flow field, we must relate the acceleration of a fluid
element to the velocity field. Referring to Equations (5.1) and (5.2) and taking the x component as
typical, we express the difference by:
du =

u
u
u
u
dt +
dx +
dy +
dz
t
x
y
z

(5.6)

The x component of the acceleration can be obtain dividing the velocity by dt :


ax =

u
u
du u dt u dx u dy u dz u u
=
+
+
+
=
+
u+
v+
w
t dt x dt y dt z dt t x
y
z
dt

(5.7)

v
w
v
w
v v
w w
u+
v+
w , az =
u+
v+
w.
+
+
z
z
y
y
x
t x
t

(5.7)

r
r
r
r
r vr vr
r
v
v
u+
v+
w.
a = a x i + a y j + az k =
+
z
y
t x

(5.8)

Similarly:
ay =

Hence:

r
The first terms, ( v t ) , represent the local acceleration in a transient motion; it is zero for steady

flows. The sum of the three remaining terms express the acceleration associated with the motion of
a fluid element in a non uniform velocity field, called also carrying acceleration or convective
acceleration. The Equation (5.8) can be write also (using the differential operators):
r
r
r
r dv v
r
r

r v
=
+ u +
a=
+ (v ) v
v+
w v =
t x
y
z
dt
t
r
r
r v
r r v
r r
v2
v2
a=
+ grad
+ rot v v
+
+ v v =
t
2
2
t

(5.9)
(5.9)

In the Equation (3.6) are emphasized the potential component, grad v 2 2 and the rotational one,
r r
rot v v , of the convective acceleration. Also, in this form the acceleration is much easy to
integrate.

5.3 EQUATIONS OF THE FLUID MOTION


5.3.1 Continuity and conservation of matter
From the definition of a streamline, it is apparent that fluid elements in nonturbulent flow cannot
cross a stream-tube surface. If the density of fluid is independent of time, mass does not
accumulate and:
the mass flow rate at every section of a stream tube must be the same.
This is a specialized statement of the continuity principle the law of mass conservation applied
to a flow field.

55

The continuity principle establishes a condition that any description of the flow field must satisfy,
and it thereby provides a relationship between certain properties within the flow field.
Imagine a stream tube having a cross-sectional area dA at an arbitrary position along the tube. As
dA becomes infinitesimal, the stream tube approaches a streamline. With v representing the

velocity of the fluid at this section, the fluid face coincident with dA moves a distance v dt normal
to dA in time dt . The volume of fluid passing the section is thus given by the expression:
dV = v dA dt .

(5.10)

dm = v dA dt ,

(5.11)

The mass is given by:


& = (dm dt ) is:
and the differential mass flow rate Qm = dm
& = v dA .
dm

(5.12)

& at any section of a stream tube of finite size is obtained by


The instantaneous mass flow rate m

the integration of Equation (3.9) over the cross section.


& = v dA .
m
A

(5.13)

The volumetric flow rate Q (more commonly known as discharge) is often used in preference to
the mass flow rate, especially when the density of the fluid is constant:
Q = v dA .
A

(5.14)

If the density is uniform over the section (as is usually the case), the volumetric flow rate can be
obtained by dividing the mass flow rate by the density:
Q=

Qm

(5.15)

The average velocity v , at any section of a stream tube, is defined by the equation:
v =

Q
.
A

(5.16)

If the properties of a fluid are uniform over a cross section, the flow is said to be one-dimensional.
Many flows may be assumed to be one-dimensional without introducing serious error. When this is
done, the velocity that is used is the average velocity, but the bar notation is generally dropped.
The mass flow rate for one-dimensional flow can thus be expressed as:
Qm = v A .

(5.17)

When this is combined with the continuity principle, we obtain that:


Qm = ( v A )1 = ( v A )2 = ... = ( v A )n = cons tan t .

(5.18)

The subscripts are referring to two different sections along a stream tube. Whenever the density
can be assumed uniform throughout the stream tube, the volumetric flow rate will be the same at
every section and:
Q = v1 A1 = v 2 A2 = ... = v n An = cons tan t .

56

(5.19)

5.3.2 Eulers equation of motion. Conservation of energy.


The Equation of (ideal) fluid motion can be founded from Newtons second law of motion applied to
an element of fluid with mass m and volume V , bounded by a surface having area A .
r
r
r
r
m a = Fext =Fm + Fp .
(5.20)
r
where:
Fext is the sum of the exterior forces, which are acting on the mass of fluid,
r
r
respectively the mass forces Fm and the pressure forces Fp .
For an elementary fluid element:
r
r
r
r
r
dv
dv
a dm = a dV =
dV m a =
dV
dt
dt
V

(5.21)

r
r
r
r
r
dFm = fm dm = fm dV Fm = fm dV

(5.22)

r
r
r
r
dFp = p n dA Fp = p n dA = p dV

(5.23)

r
fm is the unit mass force (it has dimension of acceleration) and it is expressed as:
r
r
r
r
fm = fm x i + fm y j +fm z k .

where:

Generally:

fm x =

r
U
U
U
; fm y =
; fm z =
fm = grad U
y
x
z

(5.24)

U is the potential of the mass forces. It represents the mass potential energy of the fluid. When
fm x , fm y and fm z are known:

U ( x, y, z ) = fm x dx + fm y dy + fm z dz

Substituting the expressions (3.18, 3.19, 3.20) into Equation (3.17) it is obtained that:
r
r
dv
dt dV = fm dV p dV
V
V
V

(5.25)

(5.26)

Note:
In the above Equation, there is no any differential operator before integrals, which can
affect the integration. For an arbitrary volume V 0 , the Equation becomes:
r
r
1
dv r
dv r

= fm p
= fm p
dt
dt

(5.27)

Equation (5.27) is the Eulers Equation of ideal fluid motion, in vectorial form: a fluid in motion is in
r
r
equilibrium under the actions of inertia forces (- dv dt ) , mass forces fm and pressure forces
( p) / .
Taking into consideration the expression (5.9) of the acceleration, the above equation becomes (in
H. Helmholtzs formulation):
r
r r r
v2
1
v
+ grad
+ rot v v = fm grad p
2
t

57

(5.28)

In the case of fluids, which for:


r
fm = grad U ,

the mass forces are deriving from a potential

the mass density is constant or known function of pressure

grad p = grad

the Equation (5.28) can be rewrite in the following form:


r
r r
v2
dp
v
+ grad
+ rot v v = grad U grad

t
r

v2
r r
dp
v
+ grad
+
+ U + rot v v = 0

dp

(5.29)

This is the Equation of the ideal fluid motion formulated by I. S. Gromeka and H. Lamb.

3.3.3 Bernoullis Equation


Integration of the fluid motion Equation (5.29) depends on the stipulated conditions in specific
problems. In this sense:
y

if the flow is steady the transient term is zero

if the motion is not rotational, or along a stream line

r
v
=0,
t
r r
rot v v ,

then:
v2

dp
grad
+
+U = 0
2

(5.30)

The terms between brackets have dimensions of energy for unit mass. Their sum is denoted with
e . It shows that the total energy of the unit fluid mass represents the sum between kinetic energy
and potential energy of the pressure and mass forces. The expression:
v2
dp
+
+U = e
2

is called the Bernoullis function.


r
If the Equation (5.30) is multiply with elementary displacement dr , hence:
v2
r
v2

dp
dp
grad
+
+ U dr = 0 d
+
+U = 0
2

v2
dp
+
+ U = ct .
2

(5.31)

r
r
r
For fluids with constant mass density = ct . , in gravitational field, fm x = fm y = 0 , fm z = g , then:
U = fm z dz = - g dz = g z + ct ,

and The Equation (3.27) becomes:

58

v2 p
+ + g z = ct .
2

(5.32)

This is known as Bernoulli's Equation. For two points along a streamline it can be write:
v12 p1
v 22 p2
+
+ g z1 =
+
+ g z2 .
2
2

(5.32)

In this form all the terms are representing energies per unit mass:
y

Kinetic energy

Pressure energy

Potential energy

v12
.
2
p

gz

Bernoulli's Equation can be expressed in another two forms. One of the most used is obtained
when Equation (5.32) is divided by g :
v12
p
v2
p
+ 1 + z1 = 2 + 2 + z2 = H [m ]
2g g
2g g

The Bernoulli constant H is known as the total head, for it is the sum of:
y

the potential head

z,

the pressure head

p
p
= ,
g

the velocity head

v2
.
2g

All the terms in the above Equation have the dimensions of length (height).
The sum of the potential head and pressure head is known as the piezometric head.

Fig. 5.4 Transformation of energy along a pipeline


59

(5.33)

If an ideal fluid is in motion and the velocity varies along a streamline, the piezometric head does
not remain constant but varies in such a way that the total head remains constant (see Figure 5.4).
A third form of Bernoullis Equation can be write when Equation (5.32) is multiply by :

v12
2

+ p1 + g z1 =

v 22
2

N
+ p2 + g z 2

m2

(5.34)

In this form all the terms are representing pressures:


v12
y Dynamic pressure
pd =
.
2
y Static pressure
ps = p
The sum between dynamic pressure and static pressure represents total pressure pt .

3.3.4 General Momentum Theorem


In General Mechanics, the momentum of a material point having the weight m that moves with
r

velocity is defined as the product m . For a system of material points, the total momentum has
the following expression:
r
r
M = mi i

(5.35)

The Momentum Theorem (which is derived from Newtons second law):


d
dt

mi

r
r
i = Fext

(5.36)
state that the time rate of change of the

momentum for a system of material points equals the sum of all the external forces which are
acting upon the system.
In order to transpose this theorem in the field
of Fluid Mechanics, an incompressible fluid of
density

is considered in steady flow

through a stream tube control volume (see


Figure 5.4).
Note that many of fluids transport cases of
technical interest take place in such stream
Fig. 5.4 General stream tube control volume

tubes, with or without bifurcation (ducts).


At an arbitrary time, the fluid occupies the

control volume bounded by a surface S ABCD (control surface). The lateral surfaces S1 , S2 are
perpendicular on the direction of the flow. At two successive moments of time t1 and t 2 , the mass
of fluid will occupy the positions ABCD , respectively A' B' C' D' . The time rate of change of the

60

momentum can be expressed by the difference of the momentum of the fluid at the two times t1
r

and t 2 : dM = M 2 M1 .
Because we considered a flow developed in steady state conditions, the mass momentum of the
fluid between sections A' B' and CD remains constant in time. Therefore, the time rate of change
of the momentum is given by the difference between the momentum of the fluid contained into the
surface S ABB' A' and the momentum of the fluid contained in the surface SCDD' C' :
r
r
r
r
r
r
r
r
r
dM = M 2 M1 = m2 2 m1 1 = V2 2 V1 1 = S2 2 dt 2 S1 1 dt 1

r
r
r
dM
= Q ( 2 1 )
dt

(5.37)

Thus relation (5.36) becomes:


r
r
Q ( 2 1 ) =

where: Q
1,2

Fext

(5.38)

the flow rate;


average velocities of the fluid in the cross-sections S1 and S2 .

Generally, the sum of all external forces which are acting upon fluid contained in the control
volume is:
r

Fext
r

where: G
r r
Fp1 , Fp2

r r
r
r
r
= G + Fp1 + Fp2 + Fpsl + Ffsl

(5.39)

the gravitational force exerted on the mass of fluid from the control volume;
the pressure forces exerted by the fluid from the stream tube, besides the
considered volume, which is acting on the fluid from control volume, through
the inflow surface S1 , respectively the outflow surface S2 (normal on this
surfaces and pointing inward over the surfaces;

r
Fpls

the force with which the lateral wall of the stream tube belonging to the
control surface acts upon the inside fluid;

r
Ffls

the viscous friction force which is created between the fluid and lateral
surface of the stream tube.

Replacing relation (5.39) in (5.38) we obtain:


r r
r
r
r
r
r
Q ( 2 1 ) = G + Fp1 + Fp2 + Fpsl + Ffsl

(5.40)

Obsevations 1 In order to apply the Momentum theorem it is sufficient to know flow parameters
on the control surface and not to know what happens inside it. More specifically, it is
to be known the pressures and the velocities on this surface.
2 For practical applications of (5.40), the system has to be studied in a fixed
reference frame, conveniently chosen.

61

6. APPLICATIONS OF THE BERNOULLIS EQUATION


6.1 FLOW THROUGH SMALL ORIFICES
Considering the flow from a tank through a hole in the side close to the base. The general
arrangement and a close up of the hole and streamlines are shown in the figure below.

Fig. 6.1 - Tank and streamlines of flow out of the sharp edged orifice
The shape of the hole edges are as they are (sharp) to minimise frictional losses by minimising the
contact between the hole and the liquid.
Looking at the streamlines it can be seen how they contract after the orifice to a minimum value
when they all become parallel. At this point, the velocity and pressure are uniform across the jet.
This convergence is called the vena contracta (from the Latin contracted vein). In order to
calculate the flow, it is necessary to know the amount of contraction.
The velocity through the orifice can be predicted using the Bernoulli equation, by applying it along
the streamline joining point 1 on the surface to point 2 at the centre of the orifice.
v12
p
v2
p
+ 1 + z1 = 2 + 2 + z2
2g g
2g g

(6.1)

At the free surface velocity is negligible v1 0 due to the large area of flow section comparatively
with the area of the orifice, and the pressure is the atmospheric one, p1 = 0 (gauge pressure). At
the orifice, the jet is open to the air so again the pressure is atmospheric, p2 = 0 . If we take the
datum line through the orifice, then z2 = 0 and z1 = h . Hence:
h=

v 22
2g

v2 = 2 g h

(6.2)

This is the theoretical value of velocity. Unfortunately it will be an over estimate of the real velocity
because friction losses have not been taken into account. To incorporate friction we use the

62

coefficient of velocity cv to correct the theoretical velocity. Each orifice has its own coefficient of
velocity, which usually falls in the range (0.97 - 0.99).
To calculate the discharge through the orifice we multiply the area of the jet by the velocity. The
actual area of the jet is the area of the vena contracta not the area of the orifice. We obtain this
area by using a coefficient of contraction for the orifice c c .
Aactual = c c Aoriffice .

(6.3)

So the discharge through the orifice is given by:


Q = cc Aoriffice cv v 2 = cD Aoriffice 2 g h ,

(6.4)

where c D = cv c c is the coefficient of discharge.


If the pressure p1 0 :
p

v 2 = 2 g 1 + h

(6.5)

6.2 TIME FOR A TANK DISCHARGE


As shown previously, the expression for the discharge of a tank is based on the height of water
above the orifice. Often, it is useful to know the time for the tank to empty.
As the tank empties, so the level of water will fall. We can get an expression for the time it takes to
fall by integrating the expression of flow between the initial and final levels.
Lets consider the tank shown in Figure 6.1, with the cross sectional area of A . In a time dt the
level falls by dh or the flow out of the tank is:
Q = A v = A

dh
dt

(6.6)

The minus sign is due to dh , which is falling (is negative). Rearranging and substituting the
expression (6.4) for Q through the orifice gives:
dt =

dh

Aoriffice c D 2 g

(6.7)

This can be integrated between the initial level h1 , and final level h2 , to give an expression for the
time t it takes to fall this distance:
t2

t = dt =
t1

t=

A
Aoriffice c D

A
Aoriffice c D

h2

1
2g

2g

h1

dh
h

h1 h2

63

(6.8)

6.3 FLOW OVER NOTCHES AND WEIRS


A notch is an opening in the side of a tank or reservoir which extends above the surface of the
liquid. It is usually a device for control of stream flow, seldom for measuring of discharge. A weir is
a notch on a larger scale, usually found in rivers. It may be sharp crested but also may have a
substantial width in the direction of flow. It is used as device to control the water level ant its flow
rate.
For the following calculus it will be assumed that the velocity of the fluid approaching the weir is
small so that kinetic energy can be neglected and also that the velocity through any notch depends
only on the depth below the free surface. These are conveniently assumptions for tanks with
notches or reservoirs with weirs, not for flows where the velocities are significant e.g. for a fast
river. For theses cases the kinetic energy must be taken into account.
To determine an expression for the theoretical flow through a notch it is considered a horizontal
strip of width b and high dh at a depth h below the free surface, as shown in the figure 6.2.

Fig. 6.2 - Flow over a notch


From equation (6.2) the velocity of fluid through the strip is:
v = 2gh

(6.9)

dQ = b dh v = b dh 2 g h

(6.10)

The flow rate through the strip is:

Integration from the free surface, h = 0 , to the weir crest, h = H , gives the expression for the
theoretical discharge through the notch:
H

Q = 2 g b

1
h2

dh

(6.11)

This will be different for each differently shaped weir or notch. To solve this equation there is need
an expression relating the width of flow across the weir to the depth below the free surface.

6.3.1 Flow over rectangular notches


For a rectangular notch (see Figure 6.3) the width does not change with depth, so b = B = ct .
Substituting this into the equation (6.11) gives:

64

Fig. 6.3 - Rectangular notch

Q =B 2g

1
h2

2
dh = B 2 g H 2
3

(6.12)

Taking into accounts the losses at the edges of the notch and contractions in the area of flow, the
discharge can be expressed as:
H

Q =B 2g

1
h2

2
dh = c D B 2 g H 2
3

(6.13)

where: c D is the coefficient of discharge through the notch.

6.3.2 Flow over triangular notches


For a triangular notch (see Figure 6.4) having the angle , the relationship between width b and
depth h is:

Fig. 6.4 - Triangular notch

b = 2 (H - h) tg

(6.14)

Thus:
Q = 2 2 g tg

(H - h) h 2 dh =

8
15

2 g tg

H2

(6.15)

Taking into accounts the losses at the edges of the notch and contractions in the area of flow, the
discharge is:
8
Q=
cD
15

2 g tg

5
H2

(6.16)

6.4 MIXING OF FLUIDS. THE EJECTORS


The ejectors are devices used to entrain fluids (secondary fluids) using the other fluids (primary
fluids) with high kinetic energy. They are jet pumps and have the advantage of the simplicity, with
no mechanical part in motion. The general set-up for an ejector is shown in Figure 6.5, together
with variation of the parameters, velocity and pressure.
65

Fig. 6.5
The basic principle underling the operating of an ejector is an increasing in the velocity of the
primary fluid, accompanied by a decrease in pressure at the level of mixing chamber, where the
secondary fluid is entrains by absorption, the result being a mixture of fluids. Three regions are
distinctive:
convergent region, between points 1 and 2; the velocities of both fluids, primary fluid " pf "
and secondary fluid " sf " are increasing; their pressures are decreasing to the same
magnitude;
region of homogenization for the mixture of fluids " mf " , between points 2 and 3;
divergent region, between points 3 and 4; the kinetic energy of the mixture is converted into
potential energy of pressure.
The equations of the mixing of fluids are the following:
mass balance equation:
( Qm )pf + ( Qm )sf = ( Qm )mf

(6.17)

where: Qm is the mass flow rate


power balance equation; for incompressible fluids is:

2
Q v + p + g
m
2

v2 p

z + Qm
+ +g
2

pf

v2 p

z = Qm
+ +g
2

sf

mf

(6.18)

Dividing with ( Qm )pf :


v2 p

+ +g
2

v2 p
z + u
+ +g

pf

v2 p
z = (1 + u )
+ +g

sf

mf

(6.19)

where: u is the coefficient of mixing for the ejector;


u=

( Qm )sf
.
( Qm )pf

66

(6.20)

6.5 FLUID METERING


6. 5.1 The stagnation parameters. The Pitt tube. The Pitt-Prandtl tube
If a stream of uniform velocity flows into a blunt body, the streamlines take a pattern shown in
Figure 6.6.

Figure 6.6 - Streamlines around a blunt body


Except one, all the streamlines are moving around the blunt body. The streamline in the center
goes to the tip of the blunt body and stops. Along this streamline the velocity is decreased to zero
in the impact point, and transform into pressure. The contact point is known as the stagnation
point, SP .
The pressure at this point can be computed applying Bernoullis equations along the central
streamline from a point upstream ( 0 ) , where the stream is not disturbed, to the stagnation point of
the blunt body, where the velocity is zero, v 2 = v * = 0 . For z1 = z2 = 0 , then:

v 2
+ p = p* = ptot
2

(6.21)

Terms having the subscript " " represent the parameters of the stream unperturbed by the blunt
body, theoretically in upstream and downstream at infinite. Terms having the superscript *
represent the parameters of the stream in the stagnation point. The total pressure is known, also,
as the stagnation pressure.
The blunt body stopping the stream does not have to be a solid. I could be a static column of liquid.
In this way the total pressure can be measured using a L tube with the open end facing a stream
as shown in Figure 6.7. This is the Pitt tube, named after the French physicist Henry Pitt, who
measured in 1732 the velocity of the Seine.

Figure 6.7 - A static tube and a Pitt tube


67

Two piezometers, one as normal and one as a Pitt tube within the pipe can be used in an
arrangement shown previously to measure velocity of flow. From equation (6.21):
v = 2 g ( htot hst )

(6.22)

The Pitt tube may be combined with the static tube as shown in Figure 6.8. In this design, first
used by Ludwig Prandtl, the static pressure tube jackets the Pitt tube, resulting in a compact
mater. The static holes in the Pitt-Prandtl tube are so located that the slight pressure decrease
caused in the stream by the tube is compensated by the pressure increase due to stagnation on
the stem. Figure 6.8 show also the relative dimensions of the Pitt-Prandtl tube suitable for
incompressible flows.

Figure 6.8 The Pitt-Prandtl tube

lp
v = 2 g
1 h

(6.23)

6. 5.2 The orifice meters


The orifice meters are devices used to measuring the flow of fluid through closed conduits. The
basic principle underlying the operation of the orifice meter is an increase in the velocity of flow,
accompanied by a decrease in the pressure of the fluid under consideration. Thus, the quantity of
fluid can be computed from the measured pressure drop across the orifice.
Figure 6.9 shows the principle of this method of flow metering, considering an orifice plate provided
with a sharp-edged circular orifice, fitted into a conduit between two flanges. Upstream from the
plate, in the cress section S1 , the velocity is v1 and the pressure p1 . Downstream from the orifice
a vena contracta is formed and the cross section area decreases to a minimum value S2 ; the
68

parameters of the fluid in this section are v 2 and p2 . Further downstream the stream gradually
expands to the normal flow.

Figure 6.9 The orifice meter


From the continuity equation, the discharge (volumetric flow rate) is:
Q = 1 S1 = 2 S2 ,

where: S1 =

D2
4

S2 = c c

d2
4

(6.24)

area of the first cross section;


area of the vena contracta cross section;

diameter of conduit;

diameter of orifice;

cc

coefficient of contraction.

Neglecting the pressure losses, the velocity 2 can be determined from Bernoullis equation:
12
p
2
p
2
+ 1 = 2 + 2 + 2
2g g 2 g g 2 g

(6.25)

From equation of continuity we can write:


d
1 = c c 2
D

(6.26)

Replacing velocity v1 in equation (6.25) and rearranging the terms it is obtained the following
expression for the velocity in the vena contracta:
v2 =

1
1 (c c m)

69

p1 p2

(6.27)

d
where: m = .
D

Hence the discharge becomes:


Q = A2 v 2 = c c

d 2
4

1 ( c c m)

(p1 p2 )

(6.28)

From practical considerations, because there is measured the pressure drop pI pII , it is
convenient to express the discharge in the following form:
Q = cd

where: c d
E

d 2
4

(pI

pII ) ,

(6.29)

is the discharge coefficient of the orifice meter;


is the approach factor:
E=

1
d
1
D

1
1 m2

(6.30)

The discharge coefficient is experimentally obtained. Thus the errors due to the omission of the
contraction coefficient under the square root and losses between sections 1 and 2 are
automatically adjusted. For a specified design, the discharge coefficient is not constant; it is a
function both of ratio m and an important parameter of flow known as Reynolds number (see the
real fluids flow). For higher Reynolds numbers than 10 5 the influence of Reynolds on c d
becomes negligible.
In order to minimise the loss upstream and downstream the orifice, either an orifice nozzle (see
Figure 6.10) or a Venturi meter (see figure 6.11) can be used.

Figure 6.10 The orifice nozzles


For a well designed orifice nozzle the minimum stream area is located at the exit of the nozzle, so
it is equal with the nozzle area. For the conditions generally encountered in metering of lowviscosity fluids, such air or water Reynolds number is on order of 10 5 - 10 6 , thus using of this type
of flow meter is more advantageous. Best results are obtained with a quarter ellipse profile as
shown above.

70

The Venturi meter consist on converging, parallel and diverging sections as shown in Figure below.
The converging section is short and has a taper angle of 15 20 , while the diverging section is
longer, having a taper angle of 5 7 .

Figure 6.11 The Venturi meter


SELF-ASSESSMENT EXERCISES
Exercise 1
A vertical cylindrical tank 2m diameter has, at the bottom, a 0.05m diameter sharp edged orifice for
which the discharge coefficient is 0.6.
a) If water enters the tank at a constant rate of 0.0095 m3/s find the depth of water above the
orifice when the level in the tank becomes stable.
b) Find the time for the level to fall from 3m to 1m above the orifice when the inflow is turned off.
c) If water now runs into the tank at 0.02 m3/s, the orifice remaining open, find the rate of rise in
water level when the level has reached a depth of 1.7m above the orifice
Exercise 2
A reservoir with vertical sides has a plan area of 56000m2. Discharge from the reservoir takes
place over a rectangular weir, the flow characteristic of which is Q=1.77BH3/2 m3/s. At times of
maximum rainfall, water flows into the reservoir at the rate of 9m3/s. Find:
a) the length of weir required to discharge this quantity if head must not exceed 0.6m;
b) the time necessary for the head to drop from 60 cm to 30 cm if the inflow suddenly stops.
Exercise 3
A Venturi meter with an entrance diameter of 0.3m and a throat diameter of 0.2m is used to
measure the volume of gas flowing through a pipe. The discharge coefficient of the meter is 0.96.
Assuming the specific weight of the gas to be constant at 19.62 N/m3, calculate the volume flowing
when the pressure difference between the entrance and the throat is measured as 0.06 m on a
water U-tube manometer.
71

7. APPLICATIONS OF THE MOMENTUM EQUATION


7.1 HYDRODYNAMIC FORCES ON A FLAT PLATE
One of the applications of the momentum theorem refers to the theoretical computation of the
hydrodynamic (impact) forces. For example lets consider the case of a jet of fluid with circular
section, which hits under the angle a flat plate. On the assumption, the diameter of the jet is
significantly smaller that the diameter of the plate.

Fig. 7.1 - Hydrodynamic force on large flat plate


As shown in figure 7.1, near the wall, the jet is radially deflected around the impact point, as the
exit section is cylindrical.
The control volume is chosen so that the first section is find upstream, where the jet is not
perturbed by the plate, and the second one after the contact point, where the velocities of the fluid
particles become parallel with the plate. Because the jet freely develops in the environment, the
pressure forces distributed on the control surface are balancing themselves. The equation of the
momentum theorem for the control volume is:
r r
r
r
r
Q ( 2 1 ) = G Fh + Ffsl

(7.1)

If the friction forces and the weight of the fluid inside the control volume are neglected (they are
smaller in comparison with the pressure forces) the velocities in the characteristic section are
equals: 1 = 2 = . Projecting relation (7.1) on ox axis, the equation for the hydrodynamic force
is:
Fh = Q sin ,

(7.2)

where the flow rate is:


Q=

d2
4

Thus:
72

(7.3)

Fh =

d 2 2 sin , or Fh =

Q2
d2

sin

(7.4)

If the dimension of the plate is comparable with that of the jet, as shown in figure 7.2, the
theoretical hydrodynamic force is:
Fh =

d 2 2 (1 cos ) ,

(7.5)

where is the angle under the jet is deviated.

Fig. 7.2 - Hydrodynamic force on small flat plate


The equation (7.5) is also valid for other types of surfaces: concave, convex, and others.

7.2 HYDRODYNAMIC FORCE ON A PIPE NOZZLE


Considering the flow through a nozzle, as shown in Figure 7.3. Because the fluid is contracted in
the converging section, forces are induced in the nozzle. Anything holding the nozzle (e.g. a
fireman) must be strong enough to withstand these forces. In the following will be computed this
induced force in the nozzle by the fluid, Fh .
y
o

S2 , v2
S1 , v1
Fig. 7.3 - Hydrodynamic force on a pipe nozzle

The control volume coordinates axis system is chosen as in figure above. Neglecting the friction
forces and the weight of the fluid inside the control volume, the equation of the momentum theorem
for the control volume is, on ox :
Q ( 2 1 ) = Fp1 Fh

(7.6)

1
1

Fh = Fp1 Q ( 2 1 ) = p1 S1 Q 2
S2 S1

(7.7)

Hence, using the continuity:

The (ideal) Bernoulli equation is used to calculate the (gauge) pressure in the section 1 :
73

v12
p
v2
+ 1 = 2 .
2g

2g

(7.8)

Then:
p2 =

(v

2
2

Q 2 1

1
Q2
S12

v12 =

S2
2

1
.
S12

(7.9)

1
1

S2 S1

(7.10)

Finally:
Fh =

Q 2 S1 1
2

S2
2

7.3 ENERGY FROM WIND. The Axial Momentum Theory Betzs Theory
Wind is simply air in motion. It is the movement of the air mass, caused by uneven heating of the
atmosphere by the sun. That is the source of wind energy. The wind contains energy by the virtue
of its motion, and this is called kinetic energy.
A windmill extracts power from the wind by slowing down the wind. At stand still, the rotor
obviously produces no power, and at very high rotational speeds the air is more or less blocked by
the rotor, and again no power is produced.
From the air stream tube of Figure 7.4, conservation of mass dictates that:

v1

vax

v1

S1
S

S1

Figure 7.4 - Axial Momentum illustration for Betzs theory


Qm = ct = air S1 v1 = air S v ax = air S2 v 2 ,

where:

air

is the density of air;

S1

is the area of the air stream far before the rotor;

v1

is the wind velocity of the air stream far before the rotor;

is the rotor swept area;

v ax

is the wind velocity at the plane of the rotor;


74

(7.11)

S2

is the area of the air stream far behind the rotor;

v2

is the wind velocity of the air stream far behind the rotor.

The thrust force F on the rotor is given by the change in momentum:


F = (air S1 v1 ) v1 (air S2 v 2 ) v 2 .

(7.12)

With equation (7.11), the thrust force reduces to equation (7.13):


F = air S v ax (v1 v 2 ) .

(7.13)

The difference in power before and after the rotor is the power extracted by the windmill, which
equals the product of the thrust force, F and the velocity, v ax , given by equation (7.14):
Pkin =

1
1
air v12 S1 v1 air v 22 S2 v 2 = air S v ax (v1 v 2 ) v ax ,
2
2

(7.14)

where:
Pkin

is the kinetic power extracted by the rotor [W].

Solving for v ax :
v ax =

( v1 + v 2 )
.
2

(7.15)

Velocity of the wind in the rotor plane is, therefore, the average of the upstream and low stream
wind speed. Substituting the value of v ax from equation (7.15) into equation (7.14), the power
becomes:
Pkin =

1
1
air Sv ax (v12 v 22 ) = air S (v13 v 23 v1 v 22 + v12 v 2 ) .
2
4

(7.16)

To find the maximum power extracted by the rotor, differentiate equation (7.16) with respect to v 2
and equate it to zero.
dPkin 1
= air S ( 3 v 22 2 v1 v 2 + v12 ) = 0 .
dv 2
4

(7.17)

Since the area of the rotor S and the density of the air air cannot be zero, the expression in the
bracket of equation (7.17) has to be zero. Hence, using the formula for the solution of a quadratic
equation, equation (7.17) yields:
v2 =

1
v1 .
3

(7.18)

Substitution of equation (7.18) into equation (7.16) results in equation (7.19):


Pkin =

16 1
air S v13 .
27 2

(7.19)

The theoretical maximum fraction of the power in the wind that could be extracted by an ideal
windmill is therefore 16/27 = 59.3% . This fraction is called the Betz Coefficient. Because of
75

aerodynamic imperfections in any practical machine and of mechanical loses, the power extracted
is less than that calculated above (above 60 70% from Phin computed with equation (7.19).
The shaft power of a wind rotor is given by equation (2.10):
Pmech = CP

1
3
air S v air
,
2

(7.20)

where:
Pmech

is the shaft power;

CP

is the rotor coefficient (ratio of shaft power of the windmill to the power in the
wind in the cross-sectional area of the rotor);

v air

is the velocity of the air (wind).

Equation (7.20) clearly shows that:

The power is proportional to the density of the air, air which varies slightly with
altitude and temperature.

The power is proportional to the area S swept by the blades and thus to the square of
the radius of the rotor.

3
The power varies with the cube of the wind speed v air
. This means that the power

increases eight times if the wind speed is doubled.


The maximum axial force acting in the plane of rotation occurs when the rotor extracts the
maximum power from the wind, i.e. when the low stream wind speed is 1/3 of the upstream wind
speed. By substituting this value into equation (7.13), the maximum axial wind thrust is obtained as
shown by equation (7.21).
Fmax =

4
air S v12 .
9

(7.21)

SELF-ASSESSMENT EXERCISES
Exercise 1
The force exerted by a 25 mm diameter jet against a flat plate normal to the axis of the jet is
650 N . What is the flow in m 3 /s and l/s ?

Exercise 2
A 75 mm diameter jet of water having a velocity of 25 m/s strikes a flat plate, the normal of which
is inclined at 30 to the jet. Find the force normal to the surface of the plate.
Exercise 3
A 90 reducing bend, 0.6 m diameter upstream, 0.3 m diameter downstream, has water flowing
through it at the rate of 0.45 m 3 /s under a pressure of 1.45 bar . Neglecting any loss is head for
76

friction, calculate the force exerted by the water on the bend, and its direction of application.
Exercise 4
Using the general Momentum Theorem calculate the force Fx in the x direction which is acting on
a Pelton turbine blade, as shown in the Figure bellow (known parameters: flow rate Q , diameter of
jet d , angle ). What is the value of for a maximum force Fx ?

Exercise 5
A jet aircraft, flying at a speed corresponding to a Mach number Ma = 0.85 , takes in air a rate of
Qm = 70 kg/s . If the gasses leave the nozzle with a velocity of v out = 400 m/s and the fuel-to-air

ratio is 1 : 40 , compute the thrust on the aircraft due to chance of momentum. The magnitude of
velocity of sound is c = 298.5 m/s .

air

gases

77

8. REAL (VISCOUS) FLUIDS FLOW


Viscosity is fundamentally due to cohesion and interaction between fluid molecules and any action
aimed at upsetting an existing balance is resisted by the fluid.
Pressure and temperature changes affect viscosity, but whilst the pressure changes affect
viscosity to a negligible extent only, temperature changes cause considerable variations. It is noted
that viscosity of liquids decreases with increasing temperature, whilst the viscosity of gases
increases. Because the effects of viscosity are of primary interest in fluid flow problems, all other
effects, such as compressibility and vaporization will be ignored for the time being, and the
behaviour of fluids, affected only by viscosity, will be analysed.
Viscosity opposes fluid motion. Objects moving in viscous fluids soon come to rest unless power is
expended to maintain the motion. Viscosity causes fluids moving over solid bodies to adhere to the
surface and allows motion only by a sliding action between adjacent fluid layers. The sliding action
between the layers introduces shear stresses and continuous power is required to maintain the
flow. Dissipated by viscous action, the power invested ultimately reappears in the form of heat.
None of these phenomena is manifested in perfect fluids.

8.1 LAMINAR AND TURBULENT FLOWS. EXPERIMENTS OF REYNOLDS


Reynolds' experimental investigations first published in 1883 disclosed that fluids flowing inside
pipes behave differently if the flow velocity is varied.
Reynolds' apparatus consisted essentially of a constant head tank filled with water, a small tank
containing dye, a horizontal glass pipe provided with a rounded entry and a regulating valve. The
layout is shown in Figure 8.1. The water could flow from the tank through the glass pipe into the
atmosphere and the flow velocity could be changing by adjusting the regulating valve. The dye was
introduced into the flow at the bell-mouth through a small diameter tube.

Fig. 8.1 - Reynolds' apparatus


The experiments disclosed that the dye remained in the form of a straight and stable filament when
the velocity was low (see Figure 8.2). With increasing velocity a critical state was reached, the dye
78

showed irregularities, began to waver and finally, at high velocities, rapidly diffused over the entire
cross-section (see Figure 8.3).

Fig. 8.2 Laminar flow

Fig. 8.3 Turbulent flow

Having repeated these experiments Reynolds discovered that the value of the critical velocity was
governed by the relationship between the inertia (kinetic) and viscous forces. At low velocities, he
reasoned, the viscous forces were predominant and the flow was largely viscous in character,
adjacent fluid layers flowing in 'lamina' form. At greater velocities however, the inertia forces
predominated over the viscous forces and the layers no longer remained in 'lamina' form, but
diffused through fluctuations.
Reynolds related the inertia to viscous forces and arrived at a non-dimensional parameter,
Reynolds number, Re :
Re =

vd

vd =

(8.1)

In his experiments Reynolds invariably found that critical flow occurred whenever his number
approached Re = 2320 and concluded that depending on the value of this parameter two types of
flow were in existence:
Laminar:

Re < 2300 ,

Turbulent:

Re > 2300 .

For Reynolds numbers lying between 2320 and 5000 the flow is of transition character and is
called transition flow.

8.2 VELOCITY DISTRIBUTION ON LAMINAR AND TURBULENT FLOWS


It appears that the existence of laminar and turbulent flow depends primarily on the value of the
Reynolds number. Investigations show further fundamental differences. Omitting details at this
stage the differences appear in:
-

velocity distribution

shear stress

resistance to flow.

The velocity distribution of a laminar flow in pipes follows a parabolic (see Figure 8.4) law given by
the expression:
79

r 2
v = v max 1 ,
R

(8.2)

where the maximum velocity v max attained at the centre, is exactly twice the value of the average
flow velocity v mean .

Fig. 8.4 - Velocity distribution on laminar flow


The velocity profile of a turbulent flow is much flatter than the corresponding laminar-flow parabola
for the same average velocity and becomes even flatter with increasing Reynolds number (see
Figure 8.5).

Fig. 8.5 - Velocity distribution on turbulent flow


(Flow pattern changes with increasing Reynolds' number)
Roughly, the maximum velocity is about 1.23 times the value of the average flow velocity. In Figure
8.5, a comparison between three velocity profiles at different values of Re is shown.
The velocity distribution for turbulent flow approximately follows a logarithmic law given by the
expression:
v max v

= 5.75 log

y
,
R

(8.3)

where v is the 'temporal mean' velocity and 0 the boundary shear stress.
Furthermore, in turbulent flow, random speed fluctuations are superimposed on the temporal mean
velocity so that the actual velocity and the velocity profile changes from instant to instant. The
characteristic velocity profile shown in Fig. 8.5 is based on the temporal mean velocity.

80

Experimental recording of the actual movement presents an extremely difficult problem, because
these random fluctuations are three dimensional.

Fig. 8.6 Fluctuations of the velocity profile


The intensity of the velocity fluctuations, v 'x , v 'y , v 'z , in a turbulent flow is given by the parameter
called turbulence, T :
T =

2
2
2
100 v' x + v' y + v' z
v' 2
= 100
[%].
v
3
v

(8.4)

Fig. 8.7 - Fluctuations of the velocity in time

8.3 STEADY FLOW IN PIPELINES


HEAD LOSS AND BERNOULLIS AUGMENTED EQUATION
Limiting our consideration to constant-density flow, it is clear that with shear forces included in the
equations of motion, at least one additional term must be added to each of Bernoulli's equations.
The presence of additional terms clearly indicates that the total head cannot remain constant along
a streamline.
Generally, shear forces oppose the flow and the total head decreases. All terms due to shear
stress are therefore lumped together in a single term, denoted by H L , the head loss. Bernoulli's
augmented equation is therefore written as:
v12
p
v2
p
+ 1 + z1 = 2 + 2 + z2 + H L12 [m ]
2g g
2g g

(8.5)

Following numerous studies and research works it was convened that the energy losses can be
related to the kinetic term, computed with the average velocity in the cross-section in the following
general form:

81

HL =

where:

[-]

2
[m of fluid column]
2g

(8.6)

the coefficient of energy loss (called also hydraulic loss coefficient, head loss
coefficient or hydraulic resistance coefficient).

These losses consist of linear hlin (major losses) and the local ones hloc (minor losses), in order
to facilitate their computation.
H L12 = hlin + 1nhloc [m of fluid column], (for a duct with constant cross-section)

(8.7)

8.3.1 Linear (distributed) losses of head


These represent the part of hydraulic energy of the fluid that is lost by this during flowing, its value
being proportional with the length of the completed track. They are the expression of the frictions
with the interior walls of the pipe.
According to H.P.G. Darcy, the computational formula for these losses for pipes of constant
circular cross-section is:
hlin = f

where:

l 2

[m of fluid column]
d 2g

f [-]

linear head loss (skin friction) factor - Darcy-Weissbach coefficient;

l [m]

length of the completed track between sections 1 and 2 ;

d [m]

inner diameter of the pipe.

(8.8)

For laminar flow, the skin friction factor can be deduced analytically. The answer is:
f =

64
[-]
Re

(8.9)

For turbulent flow f can be determined from experimental curve fits. One such fit is provided by
Colebrook:
e D
2.51
= 2 log
+
3
.
7
f
Re f

[-],

(8.10)

where: e is roughness of the pipe;


D is the pipe diameter.

The ratio e D is the relative roughness. For commercial pipes this is usually a very small number.
Note that perfectly smooth pipes would have a roughness of zero.
Note: For pipes of non-circular cross-section, formula (8) is applied as:
hd = f

where:

Rh

l
2
[m of fluid column]

4 Rh 2 g

(8.11)

hydraulic radius, defined as the ratio of the wetted pipe cross-section area to
the wetted perimeter of the cross-section.
82

The solutions of the equation (8.10) plotted versus Re make up the popular Moody Chart for pipe
flow.
8.3.2 Local head loss
They occur along a short portion of the flowing (called singularities) and are due to the variation of
the average velocity vector (in modulus and/or direction): in bends, ramifications, fittings, changes
of cross-section a.o. Their computational formula is:
hloc =

where:

[-]

2
[m of fluid column]
2g

(8.12)

the local loss factor, depending mainly on the geometrical characteristics of


the singularity, but also on the flow regime (Reynolds criterion, Re ).

8.3.3 Pipeline problems


Pipeline problems generally fall into one of three classifications:
1. The pipe, flow rate, and fluid properties are specified, and the head loss is to be found.
2. The pipe, head loss, and fluid properties are specified, and the flow rate is to be
determined.
3. The type of pipe, flow rate, head loss, and fluid properties are specified, and the pipe
size is to be determined.
In the first category, all the information necessary to obtain the friction factor from the
Moody diagram is available. With the friction factor known, the Darcy equation provides a direct
solution.
In the second and third categories, the friction factor cannot be directly determined and an iterative
procedure must be employed. Fortunately, convergence is quite rapid.
For problems in the second category, it is our inability to compute Reynolds number that thwarts a
direct solution. However, in the most commonly encountered engineering applications, Reynolds
number tends to be quite high, and a good starting point is to assume that the flow corresponds to
the rough-pipe zone on the Moody diagram. Since the relative roughness is known, a first
approximation of the friction factor is obtained and the corresponding velocity is computed from the
Darcy equation. Reynolds number is then determined, and the assumed friction factor checked. If
necessary, the procedure is repeated until a satisfactory check is achieved.
Problems in the third category are the most difficult, for neither Reynolds number nor relative
roughness can be determined directly. With experience, a good estimate of pipe size and Reynolds
number can be made, and thus a good first approximation of the friction factor is attainable. In any
event, a friction factor must be assumed and an iterative procedure followed until a satisfactory
check is achieved.

83

8.4 FRICTION FACTOR COMPUTATION


The relationship between friction factor, Reynolds number, and relative roughness was
established experimentally by Nikuradse. One of the chief experimental difficulties was the
measurement and control of pipe roughness. Nikuradse solved this problem by coating the inside
of smooth pipes with finely graded sand. This provided a uniform roughness that could be characterized by the diameter of the sand grains.
For commercial pipes, having non-uniform roughness, Moody summarized the Nikuradses work,
plotting the variation of the friction factor in the form of a Stanton diagram (see Figure 1). A
logarithmic plot of friction factor versus Reynolds number, with relative roughness as a parameter,
is generally referred to as a Stanton diagram.
Several features of these results are particularly worthy of note:
1. There is an abrupt change in the friction factor between Reynolds numbers of 2000 and 4000.
At higher Reynolds numbers the flow is observed to be turbulent; at lower Reynolds numbers it
is non-turbulent, or laminar. Actually, laminar flow has been observed in the laboratory at
Reynolds numbers as high as 50000, but above 4000 laminar flow is very unstable and a small
disturbance causes an immediate transition to turbulent flow. Below the lower critical Reynolds
number of 2000, laminar flow is stable and any initial turbulence is damped out.
2. When the flow is laminar, the friction factor is a function of Reynolds number only. The single
straight line on the logarithmic plot has the equation:
f =

64
[-]
Re

(8.13)

This is known as the Stokes friction factor for smooth pipe.


3. At high values of Reynolds number the friction factor becomes a function of the relative
roughness only. The region of the Stanton diagram in which this occurs is known as the roughpipe zone. The Reynolds number corresponding to the beginning of this zone depends on the
relative roughness: the greater the relative roughness, the smaller the value of this Reynolds
number.
4. Except for very large values of relative roughness, there is a range of Reynolds numbers within
the turbulent-flow regime in which the friction factor is independent of the roughness. The
extent of this range depends on the relative roughness; the smaller the relative roughness, the
greater the range of Reynolds numbers in which the pipe is hydraulically smooth. For Reynolds
numbers between 3000 and 100000 the hydraulically smooth pipe curve can be approximated
by a straight line on the Stanton diagram, and its equation is:
f =

0.3164
Re 0.25

[-]

This is known as the Blasius friction factor for smooth pipe.


84

(8.14)

5. For complete turbulence Re > 4000 , rough pipe ( e D ) = 0.00008 - 0.0125 , the friction factor
can be computed using the Equation (Idelcik):
e 68
f = 0.11 +

D Re

0.25

[-]

(8.15)

6. Between the smooth-pipe curve and the rough-pipe zone is the transition zone. In this zone the
friction factor is a function of both relative roughness and Reynolds number and can be
computed with the Moodys Equation:
f =

260 D
[-].
Re e

(8.16)

Fig. 8.9 Moody Chart

WORKED EXAMPLES
Example 1
The below figure shows a pump delivering water to a tank through a pipe with the following
characteristics: D = 30 mm (inner diameter), e = 0.2 mm (roughens), L = 30 m (length) and

= 0.6 (the sum of the minor losses coefficients: in elbow and bore of tank). The kinematic

viscosity of the water is = 10 6 m 2 / s . Find the pressure at point (1) if the flow rate is
1.4 dm 3 /s .

85

SOLUTION
Convert all the quantity in International System of Units
D = 30 mm = 30 10 3 m
e = 0.2 mm = 0.2 10 3 m
Q = 1.4 dm 3 /s = 1.4 10 3 m 3 /s
L = 30 m

Also

= 1000 kg / m 3

(the density of water).

Apply Bernoulli between points (1) and (2): at the free surface of the water in tank.
v12

v 22
p1
p
+
+ z1 =
+ 2 + z2 + H L12 [m ] .
2g g
2g g

Notice:
Choose the level of the point (1) the datum level =>

z1 = 0 and z2 = 25 m ;

The pressure on the free surface is zero gauge pressure =>

p2 = 0 N / m 2 ;

The velocity at the level of point (2) is very small and can be neglected => v 2 = 0 m / s .
In this conditions the Bernoulli s Equation can be rewrite:
v12
p
+ 1 = z2 + H L12 [m ] .
2g g

The head loss is:


H L12 = f

v2
v2 L
L v12

+ 1 = 1 f + [m ] .
D2g
2 g 2 g D

86

Thus:
v2 L
v12
p

+ 1 = z2 + 1 f +
2g g
2 g D

p1 = g z2 +

v12 L

f + 1 [m ]
2 D

The velocity in point (1) (mean velocity in pipe) can be computed from the Equation of flow rate
(discharge):
Q = ct. = v1

D2
4

v1 =

4Q

D2

4 1.4 10 3

30 10 3

= 1.98 m / s .

In order to obtain the friction factor, is necessary to establish the flow regime (to compute the
Reynolds number):
Re =

v D 1.98 30 10 3
=
= 5.94 10 4 .

10

Because Re > 4000 , the flow is complete turbulent. In these conditions is necessary to establish if
the flow is affected or not by the roughness (to compute the relative roughness):
e =

e 0.15 10 3
=
= 0.005
D
30 10 3

The condition ( e D ) = 0.00008 - 0.0125 is fulfilling, so the flow is rough, also.


e 68
f = 0.11 +

D Re

0.25

68
= 0.11 0.005 +
5.94 10 4

0.25

= 0.031 .

Finally:
p1 = g z2 +
= 1000 9.81 25 + 1000

v12
2

f + 1 =
D

1.98 2
30
N

+ 0.6 1 = 3.052 10 5 2
0.031

3
2
30 10
m

Example 2
The Figure below shows a tank that is drained by a horizontal pipe with the following
characteristics: D = 50 mm (inner diameter), e = 0.2 mm (roughens), 10 m length.

87

Calculate the pressure head at point (2) when the valve is partly closed so that the flow rate is
reduced to half. The energy loss due to section change is equal to S = 0.5 . The kinematic
viscosity of the water is = 10 6 m 2 / s .
SOLUTION
The flow rate (or velocity through pipe) is unknown and the friction factor cannot be directly determined. An
iterative procedure must followed as below.
Notice:
In the first stage the velocity what is corresponding to maximum flow rate can be computed from the
Bernoulli s Equation between points (1) and (3). The valve is complete open so that its (minor)

loss coefficient is zero.


v 32
p
p1
+
+ z1 =
+ 3 + z3 + H L13 [m ] .
2g g
2g g
v12

As previously:
- Choose the level of the point (3) the datum level =>

z3 = 0 and z1 = 15 m ;

- The (gauge) pressure on the free surface is zero, =>

p1 = p3 = 0 N / m 2

same at the level of the exit.


v12 = 0 m / s

- The velocity at the level of point (1) is very small =>


and can be neglected.
- The head loss is:
H L13 = f

v2
v2 L
L v 32

+ S 3 = 3 f + S [m ] .
D2g
2 g 2 g D

Thus, for this case the Bernoulli s Equation can be write:


v 32

v 32 L
2 g z1

.
z1 =
+
f + S v 3 = v max =
L
2 g 2 g D

1 + f + S
D

A value f = 0.01 0.1 must be assigned for friction factor. So, f1 = 0.03 .
v max1 =

2 g z1
2 9.81 15
=
= 6.264 m/s .
L
10
1 + f1 + S
1 + 0.03
+ 0.5
D
0.05

Having this value for velocity, the friction factor must be checked.
Re1 =

v max1 D

6.264 0.05
10

= 3.131 10 5 .

Because Re > 4000 , the flow is complete turbulent. In these conditions is necessary to establish if
the flow is affected or not by the roughness (to compute the relative roughness):
88

e =

e 0 .2
=
= 0.004 .
D 50

The condition ( e D ) = 0.00008 - 0.0125 is fulfilling, so the flow is rough, also.


f1 checked

e 68
= 0.11 +

D Re

0.25

68

= 0.11 0.004 +

5
3.131 10

0.25

= 0.028 .

The relative error is:

1 =

f1 f1 checked
100 = 6.67 % .
f1

In order to decrease the error, the cycle of friction factor computation must be reloaded. So,
f2 = 0.028 , thus:
v max 2 =

2 g z1
2 9.81 15
=
= 6.438 m/s .
L
10
1 + f2 + S
1 + 0.028
+ 0.5
D
0.05

Re2 =

v max 2 D

e
68
f2 checked = 0.11 +
D Re2

2 =

0.25

6.438 0.05
10

= 3.219 10 5 .

68
= 0.11 0.004 +
3.219 10 5

0.25

= 0.028 .

f2 f2 checked
100 = 0.0 % .
f2

Finally, the maximum velocity through pipe and the corresponding flow rate are:
v max = 6.438 m/s ;
Qmax = v max

A half flow rate is

D2
4

= 6.438

0.05 2
4

= 12.641 10 3 m 3 /s .

Qmax
= 6.321 10 3 m 3 /s and corresponding velocity, also in point (2),
2

v 2 = 3.219 m/s .

The pressure head at point (2) h2 =

p2
for this case can be computed from the Bernoullis
g

Equation between points (1) and (2):


v 22

2
v 22
p2
L v2
z1 =
+
+f
+ S

2g g
D2g
2g

v 22
p2
L

= z1
1 + f + S
g
2 g
D

For this case friction factor must be recomputed:


v D 3.219 0.05
Re = 2 =
= 1.610 10 5
6

10

89

e 68
f = 0.11 +

D Re

0.25

68
= 0.11 0.004 +
1.610 10 5

0.25

= 0.028( 37 )

And finally:
v2
p2
10
L
3.219 2

+ 0.5 = 11.25 m .
= z1 2 1 + f + S = 15
1 + 0.028
2 9.81
0.05
g
2 g
D

SELF-ASSESSMENT EXERCISES
Exercise 3
The figure below shows a horizontal nozzle discharging into the atmosphere. The inlet has a bore
area of 600 mm 2 and the exit has a bore area of 200 mm 2 . Calculate the flow rate when the
inlet pressure is 400 Pa . The energy loss coefficient in nozzle is N = 0.8 .

Exercise 4
Determine the friction coefficient for a pipe 100 mm bore with a mean surface roughness of 0.06
mm when a fluid flows through it with a Reynolds number of 20 000 .
Exercise 5
Oil flows in a pipe 80 mm bore with a mean velocity of 4 m/s . The mean surface roughness is
0.02 mm and the length is 60 m . The dynamic viscosity is 0.005 N s/m 2

and the density is

900 kg/m 3 . Determine the pressure loss.

Exercise 6
A tank of water empties by gravity through a siphon (see figure below). The difference in levels is
3 m and the highest point of the siphon is 2 m above the top surface level and the length of pipe

from inlet to the highest point is 2.5 m .

90

The pipe has a bore of 25 mm and length 6 m . The friction coefficient for the pipe is 0.03 . The
inlet (in pipe) loss coefficient is 0.7 . Calculate the volume flow rate and the pressure at the highest
point in the pipe.

8.5 TRANSIENT FLOW IN PIPELINES. WATER HAMMER


If a velocity of a liquid in a pipeline is abruptly decreased by a valve movement, the phenomenon
which arises is called water hammer. The terminology of water hammer is a little bit misleading,
because this phenomenon occurs in any kind of fluid, both liquids and gases.
The changing of the velocity has as results a significant increase of the pressure and formation of
some pressure waves which are stresses the pipeline. In the consequence is a very dangerous
phenomenon, especially in liquids. A very important application of the water hammer phenomenon
is encountered in hydraulic plants, where the flow of water must be rapidly varied in order to
accommodate the load changes of the turbine.
The wave velocity a of pressure propagation is given by Allievis equation:
a=

c
1+

where:

d
E

is the inner diameter of the pipe;

is the wall thickness of the pipe;

is the Youngs modulus of the wall material;

c=

(8.17)

is the sonic velocity (velocity of sound in fluid);

is the elasticity modulus of the fluid;

is the density of the fluid.

For practical purposes, the pressure change (for a partial closure of the valve) can be computed
with the equation:

p = p2 p1 = c ( v1 v 2 )

91

(8.18)

9. REAL FLUIDS FLOW


BOUNDARY LAYER THEORY
9.1 GENERAL DESCRIPTION OF THE BOUNDARY LAYER
It is observed experimentally that when a fluid flows over a solid surface (e.g. the bed of a river, or
the wall of a pipe, or body of car) a thin layer is formed next to the boundary. Inside this layer the
velocity grows from zero at the surface to a maximum value (of the mainstream velocity, or free
stream, v 0 ) at height (see Figure 9.1). In theory, the value of is infinity, but in practice it is
taken as the height needed to obtain 99 % of the free stream velocity. This layer is called the
boundary layer and is the boundary layer thickness. It is a very important concept (which was
developed by the German engineer-mathematician L. Prandtl in a series of publication starting in
1904) and is discussed in below work. The inverse gradient of the boundary layer is du dy and
this is the rate of shear strain.

dv

boundary layer
thickness

dy

distance from
no slip plane

0.99 v

velocity v

Fig. 9.1 Thickness of boundary layer


It was also observed experimentally that boundary layers formed on solid surfaces exposed to a
free fluid stream become thicker in going downstream. Further, the flow inside the boundary layer
may be of laminar or turbulent character. The stages of the formation of the boundary layer are
shown in the figure below, for a flat plate.
Boundary layer is starting to form at the leading edge of the plate. In this point the thickness of this
is zero. Experiments disclose two types of boundary layer flow. Near the leading edge, the
boundary layer flow is laminar. Downstream the laminar flow changes at a point called the
transition point, and after going through a transition zone of finite (short) length, the boundary layer
flow assumes a fully developed turbulent character. Within the turbulent boundary layer, a thin,

92

laminar sub-layer with thickness lt (a few hundredths of a mm) is formed next to the boundary.
Outside the boundary layer, the main fluid flow may be either laminar or turbulent.

vx

Transition point

Leading
edge (LE)

laminar
boundary layer

tranzition
zone

turbulent
boundary layer

lt

Fig. 9.2 Laminar and turbulent boundary layers (on a flat plate)
Despite its thinness, the laminar sub-layer can play a vital role in the friction characteristics of the
surface. This is particularly relevant when defining pipe friction. In turbulent flow if the height of the
roughness of a pipe is greater than the thickness of the laminar sub-layer then this increases the
amount of turbulence and energy losses in the flow. If the height of roughness is less than the
thickness of the laminar sub-layer the pipe is said to be smooth and it has little effect on the
boundary layer. In laminar flow the height of roughness has very little effect
Boundary layer

Fully developed laminar


(parabolic) velocity profile

(a)

Entry length L l(aminar)

Laminar
boundary layer

Re < Re critic

Turbulent
boundary layer

(b)
Fully developed turbulent
velocity profile

Entry length L t(urbulent)

Re > Re critic

Fig. 9.3 Boundary layer at pipe inlet

93

The establishment of fully developed flow conditions inside a tube (pipe) takes place similarly.
Downstream from the entrance to the tube, the thickness of the annular boundary layer increases
until at a certain point it is equal to the radius of the tube, as shown in Figure 9.3.
Between the entrance to the tube and this point the flow in the boundary layer may be either
laminar or turbulent. As the layer thickens, the main fluid body accelerates. Beyond a distance L
(entrance length) the flow becomes fully developed, and is laminar if the Reynolds' number (based
on the tube diameter and the mean velocity is less than Re < Recritic (figure 9.3 (a)). Fully
developed turbulent flow may be observed beyond the distance Lt , provided the Reynolds' number
is greater than Re > 4000 . In the case of turbulent flow with turbulent boundary layer there exists,
nevertheless, a laminar boundary layer near the entry to the tube, as shown in Figure 9.3(b).
Boundary layer investigations are concerned with the determination of:
-

the thickness of the layer,

the velocity distribution in the layer, and

the shear stresses set up in the layer and the associated frictional drag forces experienced
over the surface.

The thickness of the turbulent boundary layer, which is dependent on the distance x from its
origin, can be computed with the following Equation:

t ( x ) = 0.377

x
Re 0x.2

(9.1)

Notice: Equation (1) is an approximation for a 2D turbulent boundary layer at constant pressure.
For the same conditions, the velocity in the boundary layer may be assumed to be proportional to
the seventh root of the distance y from the wall, accordingly with:
y
v x = v
t

7
.

(9.2)

Because the thickness cannot be established with precision since the point separating the
boundary layer from the zone of negligible viscous influence is not a sharp one, in some
computations concerning the boundary layer, another two parameters are used, namely the
displacement thickness * and momentum thickness . The displacement thickness can be
determined with much better precision than the overall thickness (see Figure 9.4). It is defined
by:
v * =

(v

h
v
v x ) dy or * = 1 x
v
0

94

dy .

(9.3)

Equation (9.3) shows that * is the thickness of an imaginary layer of fluid of velocity v and
mass flux rate equal to the amount of the mass flow rate deficit due to the boundary layer
(compared to the mass flow rate that would pass through same zone in the absence of the
boundary layer).
The flow retardation within the also causes a reduction in the momentum flux. Momentum
thickness is defined as the thickness of an imaginary layer of fluid of velocity v for which the
momentum flux rate equals the reduction caused by the velocity profile. It is defined by:
v 2 =

(v v x

v x2 dy or =

vx
v
1 x
v
0

dy .

(9.4)

0.99 v

Area = (v v x ) dy
0

Fig. 9.4 Definition diagram for displacement thickness *


For a 2D turbulent boundary layer at constant pressure * and can be computed with:
* =

7
1
; =
.
72
8

(9.5)

Notice: For Reynolds number greater than Re > 3 10 6 the velocity distribution deviates from the
1 / 7 power low.

It has be found that the exponent decreasing with increasing Reynolds number. The investigations
performed in order to find valid laws in the whole range of Reynolds number (especially valid for
large Re ) were led to the replacement of the seventh power law by the logarithmic velocity
distribution law. The main results, concerning turbulent flow over a flat plate, are summarising
below:
1

= 0.38 x (c D ) 2 .
1


v x = v 1 4.15 (c D ) 2 lg .
y

Where the drag coefficient c D is given by the (Schoenherr s) formula:


95

(9.6)

(9.7)

0.242
cD

= lg (Re x c D ) .

(9.8)

Thus, with the aid of the drag coefficient, both the layer thickness and the velocity distribution may
be calculated.

9.2 BOUNDARY LAYER SEPARATION


If flow over a boundary occurs when there is a pressure decrease in the direction of flow, the fluid
will accelerate and the boundary layer will become thinner (as for convergent flows). The
accelerating fluid maintains the fluid close to the wall in motion. Hence the flow remains stable and
turbulence reduces.
When the pressure increases in the direction of flow the situation is very different. Fluid outside the
boundary layer has enough momentum to overcome this pressure which is trying to push it
backwards. The fluid within the boundary layer has so little momentum that it will very quickly be
brought to rest, and possibly reversed in direction. If this reversal occurs it lifts the boundary layer
away from the surface as shown below:

Edge of
boundary layer

D
E

A
v x
>0
y

p
<0
x

B
vx
=0
Stagnation y
p
point
=0
x

v x
<0
y

p
>0
x

Separation
zone

Fig. 5 Boundary layer separation


This phenomenon is known as boundary layer separation. At the edge of the separated boundary
layer (line B E ), where the velocities change direction, a line of vortices occurs (known as a
vortex sheet). This happens because fluid to either side is moving in the opposite direction. The
vortices result in very large energy losses in the flow. These separating (divergent) flows are
inherently unstable and far more energy is lost than in parallel or convergent flow
9.2.1 Examples of boundary layer separation and control of separation
The boundary layer separation can occur in divergent ducts or diffusers, in T-junctions, Y-junctions,
bends, flows over solid surfaces as airfoils (wing cross-sections) a.o.
Normal flow over an airfoil is shown in the figure below with the boundary layers greatly
exaggerated.
96

Boundary layer

Wake

Fig. 9.6 Normal flow over an airfoil


If the angle of the wing becomes too great boundary layer separation occurs on the top of the
airfoil the pressure pattern will change dramatically. This phenomenon is known as stalling.

Fig. 9.7 Boundary layer separation on an airfoil


When stalling occurs, all, or most, of the 'suction' pressure is lost, and the plane will suddenly drop
from the sky! The only solution to this is to put the plane into a dive to regain the boundary layer. A
transverse lift force is then exerted on the wing, which gives the pilot some control and allows the
plane to be pulled out of the dive.
Fortunately there are some mechanisms for preventing stalling. They all rely on preventing the
boundary layer from separating in the first place.
-

Arranging the engine intakes so that they draw slow air from the boundary layer at the rear of
the wing though small holes helps to keep the boundary layer close to the wing. Greater
pressure gradients can be maintained before separations take place.

Slower moving air on the upper surface can be increased in speed by bringing air from the
high pressure area on the bottom of the wing through slots. Pressure will decrease on the top
so the adverse pressure gradient that would cause the boundary layer separation reduces.

Slot

Fig. 9.8 Control of boundary layer separation


-

Putting a flap on the end of the wing and tilting it before separation occurs increases the
velocity over the top of the wing, again reducing the pressure and chance of separation
occurring.
97

Fig. 9.9 Control of boundary layer separation

9.3 NAVIER-STOKES EQUATIONS


Due to inside the boundary layer the friction forces are significant, this is also called friction layer.
General theory concerning the friction between layers of fluid state that the expression of these
friction forces is:
r r 1
r
f = v +
( v ) ,

where:

r
fv

unitary friction forces;

dynamic viscosity;

fluid density;

r
v

second-order differential operator (Laplaces operator);


velocity of fluid particles.

Tacking into account the friction forces, the equation of motions for real (viscous) fluids can be
expressed as, in vectorial form (see also the Eulers equation of fluid motion):
r
r
dv r
1
= fm p + f
dt

r
where:
dv r
= a acceleration of fluid particles;
dt
p
pressure insight fluid;
r
unitary mass forces.
fm

(9.9)

Replacing the equation (9.8) in (9.9), the following system of equation is obtained in the Cartesian
form:
2u 2u 2u u w
1 p
u
u
u
u

+u
+
+w
= f mx
+ 2 + 2 + 2 +
+
+
x
t
x
x
y
z
y
z 3 x x y y

2 2 2
1 p

+u
+
+w
= f my
+ 2 + 2 + 2
x
t
y
x
y
z
y
z

(9.91)

u w
+

+
+
3 y x y y

(9.92)

2 w 2 w 2 w u w
w
w
w
w
1 p

+u
+
+w
= f mz
+ 2 + 2 + 2 +
+
+
x
t
x
y
z
z
y
z 3 z x y y

(9.93)

where:

u , , w

components of velocity, u = v x , v = v y , w = v z .
98

kinematic viscosity;

These are the Navier-Stokes equations for the motion of viscous fluids. For incompressible
r
fluids, ( = ct . and v = 0 ,) the equation (9.9) become:
du
1 p
= fmx
+ u
dt
x

(9.101)

d
1 p
= fmy
+
dt
y

(9.102)

dw
1 p
= fmz
+ w
dt
z

(9.103)

In vectorial form the Navier-Stokes equations can be write as:


r
r
dv r
1
= fm p + v
dt

(9.11)

Solutions of the NavierStokes equations


Due to the presence of the nonlinear terms, the number of the exact solutions of the
Navier-Stokes equations is limited. In general, the solutions of these equations can be found
using CFD (Computational Fluid Dynamics) techniques, as shown in figure 9.10.

Fig. 9.10 Pressure distribution on body of car, together with wake behind of this

Exact solutions are possible when:


- there are no inertial forces

r
dv
=0,
dt

(9.12)

r
- density is constant and flow is steady, consequently v from continuity equation.

Such of example is the fully developed flow between two (infinite) parallel plates.

99

Steady flow between parallel plates


Consider the fully developed stage of the steady flow between two infinite parallel
plates, as shown in figure bellow. The fluid density and viscosity are known, = ct . and
= ct . , and the distance between the plates is h , each plate having a different velocity, v1

respectively v 2 .

Fig. 9.11 Flow between parallel plates


In the Cartesian coordinates as in figure above, the only component of the velocity is
along Ox axis, and v x = v x (z ) , thus
v y v z
r
v
+
=0
v =0 x +
y
z
x
v y v z
v
v
x =
=
= 0 i = 0.
x
y
z
xi
r
Also, the unitary mass force f m action along Oz , having the components

f mx = 0, f my = 0, f mz = g .

(9.13)

(9.14)

In this case the Navier-Stokes equations give the following system


1

d2 v x
p
=
x
dz 2
p
=0
y
p
= g.
z

(9.15 1)
(9.15 2)
(9.15 3)

The equation (9.15 3) represents the hydrostatic low along Oz : p + gz = ct. In order to
solve the flow equation.
Integrating the x-momentum twice, we obtain
1 p
v x =
dz dz =
x

1 p

1 p

x z + C1 dz = 2 x z

+ C1 z + C2 .

The constants C1 and C 2 can be found from boundary conditions:

100

(9.16)

C = v 2
z = 0 v x = v 2 2
1 p
v + v2

h.
z = h v x = v1 C1 = 1
h
2 x

(9.17)

Thus, the velocity profile of flow is


vx =

v + v2
1 p
z ( z h) + 1
z v2 ,
2 x
h

(9.18)

Figure 9.12 shows it with continuous line for a negative pressure gradient, (p / x ) < 0 ,
and with dashed line for (p / x ) > 0 .
Using the equation (9.18), flow rate, average velocity, and shear stress can be
computed for the unit length along Oy :
h

y =1 m

= v x dz ,

(9.19)

v med

y =1 m

zx =

y =1 m

dv x
.
dz

(9.20)
(9.21)

Fig. 9.12 Velocity profile of steady flow between parallel plates


Figure 9.13(a) shows a particular solution, when the pressure gradient is (p / x ) = 0 .
vx =

v1 + v 2
z v2 .
h

(9.22)

Fig. 9.13 (a) Velocity profile of steady flow between parallel plates for (p / x ) = 0 ;
(b) Couette flow, if v 2 = 0 .

101

Plane Couette flow


If v1 = v , v 2 = 0 and (p / x ) = 0 , solution gives the plane Couette flow, fig. 9.13(b), which
represent the Newtons assumption used to define the shear stress:
=

dv
v
= = ct . ,
dz
h

y =1 m

= v x dz =
0

v med

(9.23)
1

h z dz = 2 vh ,

(9.24)

y =1 m

y =1 m

1
v.
2

(9.25)

Plane Poiseuille flow


If v1 = v 2 = 0 , solution gives the plane Poiseuille flow, fig.9.14.

Fig. 9.14 Plane Poiseuille flow


In this case, boundary conditions are
C = 0
z = 0 vx = 0 2
1 p

h,
z = h v x = 0 C1 =
2 x

(9.26)

and the velocity is


vx =

1 p
z ( z h) ,
2 x

(9.27)

Similarly
h

y =1 m

= v x dz =
0

1 p

1 p

2 x z( z h) dz = 12 x h

(9.28)

v med

y =1 m

y =1 m

1 p 2
h ,
12 x

dv
1 p
=
(2 z h ) .
dz 2 x

102

(9.29)

(9.30)

The sign - from the above equation is due to the negative pressure gradient along
Ox axis, (p / x ) < 0 , which is also the condition for a flow with physically relevance. The

maximum velocity is for z = (h / 2)


vmax =

h 2 dp
.
8 dx

(9.31)

9.4 THE NOTION OF RESISTANCE, DRAG, AND LIFT


The notion of resistance to motion or drag of a fluid on an immersed body is an intuitive concept
easy to grasp. A fluid moving relative to a rigid boundary exerts a dynamic force on the boundary
that is caused by two factors.
-

First, shear stresses due to viscosity and velocity gradients at the boundary surface cause
forces tangential to the surface.

Second, pressure intensities, which vary along the surface due to dynamic effects, result in
forces normal to the boundary.

For an immersed body, the vector sum of the normal and tangential surface forces integrated over
the complete surface gives a resultant force vector and a resultant moment vector, as illustrated in
Fig. 9.15, for a road vehicle.
Fz

Fx

My
Fy

Mx

x
y

Mz
z

Fig. 9.15 - Components of the resultant force

The component of this resultant force in the direction of the relative velocity v past the body is
the drag D ( Fx ) . The component normal to the relative velocity is a lift L ( ( F z) , or lateral (side)
force S ( Fy ) . The components of the resultant moment are rolling moment RM ( M x ) , pitching
moment PM ( M y ) and yawing moment YM ( M z ) .
Each component of the resultant force includes frictional and pressures parts, e.g. for the total drag
we can write:
D = Df + D p

103

(9.32)

with the components:

Df

frictional drag;

Dp

pressure drag.

The frictional drag is also known as surface resistance or skin-friction drag. The pressure drag
depends largely on the shape or form of the body and is known as form (shape) drag. Bodies like
hydrofoils, and slim ships have large and sometimes completely dominant surface resistance. Bluff
objects like spheres, bridge piers, and automobiles have large form drag relative to surface
resistance.
Dimensionless force coefficients are useful parameters for expressing the steady state dynamic
force components. The following relations define them:
D

Drag coefficient:

cD =

Lift coefficient:

cL =

Pitching moment coefficient:

c PM =

Where

(9.34)

q A
PM
q Ac l c

Ac

is a suitably chosen reference (characteristic) area;

lc

is a suitably chosen reference length;

is the reference dynamic pressure.


q =

where:

(9.33)

q A

1
v2 [N/m 2 ]
2

(9.35)

(9.36)

is the mass density of the free stream.

For road vehicle Ac is the projected frontal area and l c is the wheelbase (or length of the vehicle).
For a wing with rectangular planform Ac is the product between chord and span, and l c is the
chord.
Another useful parameter is the pressure coefficient c p in a point of a wetted area, defined by:
p p
cp =
,
(9.37)
q
where

is the static pressure in point of surface;

is the static pressure of the free stream;

The diagrams of c p variation over a surface of a body (see figure 9.16) can be built in order to
characterise the interaction between the later and a fluid stream. The pressure force components
can be computed with the aid of these diagrams.
104

Fig. 9.16 - c p distribution on a road vehicle

9.5 APLICATIONS OF BOUNDARY LAYER THEORY


9.5.1 Boundary layer properties
The important boundary conditions that are used in the formulation of boundary layer laws are:
1. The velocity is zero at the wall

v = 0 at y = 0 .

2. The velocity is a maximum at the top of the layer

v = v at y = .

3. The gradient of the boundary layer is zero at the top of the BL:

dv
= 0 at y = .
dy

4. The gradient is constant at the wall

dv
= ct . at y = 0 .
dy

d 2v

5. Following from (4):

dy 2

= 0 at y = 0 .

One of the laws that seem to work for laminar flow is:
y
v = v sin
.
2

(9.38)

Exercise 1

Find the displacement thickness * for a laminar boundary layer modeled by the previous
equation.
Solution:

y
v
y
dy = 1 sin
* = 1
dy = dy sin
dy
v
2
2
0
0
0
0

=y

2
2
y
+
cos
= 0.364
=

2 0

Exercise 2 (self-assessment)

Find the momentum thickness for the laminar boundary layer modeled by the equation (9.38).

105

Exercise 3 (self-assessment)

The velocity profile in a laminar boundary layer on a flat plate is to be modeled by the cubic
expression:
v
= a0 + a1 y + a2 y 2 + a3 y 3 .
v

Evaluate the constants ai in terms of the boundary layer thickness .

9.5.2. Skin friction drag


Skin friction drag is due to the viscous shearing that takes place between the surface and the layer
of fluid immediately above it. This occurs on surfaces of objects that are long in the direction of flow
compared to their height. Such bodies are called streamlined. When a fluid flows over a solid
surface, the layer next to the surface may become attached to it (it wets the surface). This is called
the no slip condition. The layers of fluid above the surface are moving so there must be shearing
taking place between the layers of the fluid. The shear stress acting between the wall and the first
moving layer next to it is called the wall shear stress, .
The skin drag is due to the wall shear stress and this acts on the surface area (wetted area:

Aw ). On a small (elementary) area the drag is:


dDf = dAw .

(9.39)

Df = dAw = Aw .

(9.40)

The frictional drag force is hence:

Using drag coefficient c D =

D
the wall shear stress can be defined as:
q Aw

= c D q .

(9.41)

The dynamic pressure is the pressure resulting from the conversion of the kinetic energy of the
stream into pressure and is defined by the expression:
pdyn q =

v 2
.
2

(9.42)

If the body is not a thin plate and has an area inclined at an angle to the flow direction (see
figure below), the drag force in the direction of flow become:

dDf = dAw cos( ) .

106

(9.43)

Notice:

The friction drag force acting on the entire surface area is found by integrating over
the entire area. Solving this equation requires more advanced studies concerning
the boundary layer theory.
Df = cos( ) dAw .

(9.44)

Exercise 4
Calculate the frictional drag force on each side of a thin smooth plate by 2 m long and 1 m wide
with the length parallel to a flow of fluid moving at v = 30 m/s . The density of the fluid is
= 800 kg/m3 and the dynamic viscosity is = 8 cP .

Solution:
1 cP( oise ) = 10 2 P = 10 2
Re =

g
kg
= 10 3
cm s
ms
v L
800
= v L =
30 2 = 6 10 6 .

0.008

Because Re > 3 10 6 the drag coefficient c D is given by the (Schoenherr s) equation:


0.242
cD

= lg (Re x c D ) .

or, for a smooth surface, by the following equation:


c D = 0.074 (Re )

1
5

= 0.074 6

1
6 5
10

= 0.00326 .

The dynamic pressure is:


pdyn =

2
v
800 30 2
=
= 360 kPa .
2
2

The tangential stress:

= c D pdyn = 0.00326 360 10 3 = 1173.6 Pa .


Finally, the frictional drag force is:

Df = Aw = 1173.6 2 1 = 2347.2 N .
Exercise 5 (self-assessment)
A smooth thin plate by 5 m long and 1 m wide is placed in an air stream moving at v = 3 m/s
with its length parallel with the flow. Calculate the drag force on each side of the plate. The density
of the air is = 1.2 kg/m3 and the kinematic viscosity is = 1.6 10-5 m 2 /s .

107

9.5.3 Total drag


Exercise 6

A cylinder 80 mm diameter and 200 mm long is placed in a stream of fluid flowing at 0.5 m/s . The
axis of the cylinder is normal to the direction of flow. The density of the fluid is 800 kg/m 3 . The
drag force is measured and found to be 30 N . Calculate the drag coefficient. At a point on the
surface the pressure is measured as 96 Pa above the ambient level. Calculate the velocity at this
point.
Solution:
The total drag coefficient is:
cD =

D
D
30
=
=
= 18.75
2
2
pdyn Ac
v
800 (0.5 )
3
3
d L
80 10
200 10
2
2

)(

From the Bernoullis equation:

v 2
2 ( p p )
v2
+ p =
+ p => v = v 2 +
=

2
2

(0.5 )2

2 ( 96 )
= 0.1
800

Exercise 7 (self-assessment)
Calculate the drag force for a cylindrical chimney 0.9 m diameter and 50 m tall in a wind blowing at

30 m/s given that the drag coefficient is 0.8 . The atmospheric conditions are p = 720 mmHg and
t = 20 C .

Exercise 8 (self-assessment)
Using the graph (see figure below) to find the drag coefficient, determine the drag force per metre length
acting on an overhead power line 30 mm diameter when the wind blows at 8 m/s . The atmospheric
conditions are: p = 710 mmHg and t = 5 C .

108

10. ROTODYNAMIC MACHINERY


10.1 INTRODUCTION
The collective name rotodynamic applies to a group of machines which function through rotary
action. Transformation of fluid energy into mechanical work or mechanical work into fluid energy is
obtained by means of rotating impellers. Although the basic theory is substantially the same for all
of these machines, they can be classifying in following main groups:
1.1 Pumps: deliver fluids at the expense of power input, +W (work).
1.2 Turbines: supply power at the expense of fluid energy, W .
2.1 Centrifugal pumps: the fluid pressure develops radially to the axis of rotation. Similarly, in
radial turbines (e.g., Francis) the fluid gives up energy in radial direction to the axis of
rotation.
2.2 Axial flow pumps: the stream remains parallel to the axis. In axial flow turbines (e.g.,
Kaplan) the stream remains parallel to the axis.
If the working fluid is a gas, the pumps are called:
3.1 Fans: if they deliver gases at low pressures
3.2 Compressors (or superchargers): if they deliver gases at higher pressure.
For incompressible fluids work and fluid energy relationship is:
v 2 v12 p2 p1
W1 2 = 2
+
+ z2 z1
2g
y

(10.1)

The positive sign denotes work done on the fluid and the negative sign work done by the fluid.

10.2 CENTRIFUGAL PUMPS


10.2.1 General Description
Centrifugal pumps consist substantially of one or several impellers rotating in a suitably shaped
casing. In single stage pumps (see figure 10.1) a single impeller 1 rotates in a casing 2 of spiral or
volute form, whilst in multistage pumps two or more impellers are fitted to a common shaft. Fluid
enters the impeller axially through the eye, the flow continuing radially and discharging around the
entire circumference into the casing. In passing through the impeller the fluid receives energy from
vanes 9 incorporated in the impeller, resulting in an increase of both pressure and velocity. The
kinetic energy of the fluid leaving the impeller is partially transformed into pressure inside the
casing.

109

1
9

10
4

Fig. 10.1 Single stage, single inlet centrifugal pump


According to requirements, a variety of pump designs exist in practice such as single and double
suction pumps, turbine pumps, mixed flow pumps, and so on. If diffuser vanes 6 surround the
impeller, we speak of turbine pumps probably because its construction is similar to that of turbines
having guide vanes. The function of the diffuser is to guide the fluid, whereas in volute type pumps
the fluid, after leaving the impeller, discharges freely into the casing. In double inlet pumps (see
Figure 10.2), fluid enters from two sides as if two impellers were placed back to back, thus
doubling discharge for the same head.

Fig. 10.2 Single stage,double inlet centrifugal pump


Mixed flow pumps are designed for large discharge and limited lift requirements; the impeller
passages being wide the flow has radial and axial components as well. Again, in multi-stage
pumps the discharge is the same as for one stage but the total head developed is the product of
the lift of one stage times the number of stages. In each stage the fluid discharged from the
impeller is returned to the eye of the impeller of the next stage through suitably shaped passages.
10.2.2 Pump Characteristics. Design point
Pump characteristics (curves) generally relate to the performance of a pump under varying
conditions. Usually, lift (also called pressure head) H , power P (useful and consumed) and
110

efficiency are plotted against discharge Q for a constant speed. A typical set of curves is shown
in Figure 10.3.

H [m]

Hd

Pc [kW]
max

Pu [kW]

Qd

Q [m3 /s]

Fig. 10.3 Pump characteristics


It appears that lift H decreases and power consumption Pc increases with increasing of the
discharge. The overall efficiency increases from zero to a maximum value, then decreases again.
The value of the design discharge Qd and the design lift H d are corresponding for efficiency peak
(design point).
Pump pressure head, represents the change of liquid specific energy difference between the outlet
" o" and inlet " i" section of the pump:
H = eo e i =

o2 i2 p o pi
+
+ (zo zi ) [m]
g
2g

(10.2)

Useful or hydraulic power Pu is computed with the Equation:


Pu = Q H [W].

(10.3)

It represents the part of the consumed power Pc (the power at pump shaft) converted in hydraulic
(useful) power:
Pc = M [W]

where:

angular speed of the impeller;

torque at the pump shaft.

(10.4)

In this way, the overall efficiency can be written as:

Pu
[-]
Pc

Generally, the pump characteristics are determined experimentally.

111

(10.5)

10.2.3 Cavitation in Centrifugal Pumps


In pumps carrying liquids, a phenomenon termed cavitation is frequently encountered. As a result
of the dynamics of motion the pressure varies along the path of the flowing liquid, and, in particular,
low pressure regions are encountered at points where high local velocities are observed. If
conditions are such that the static pressure attains the vapour pressure of the liquid, vaporization
results. In centrifugal pumps these low pressure regions are mostly observed at the inlet to the
impeller where the liquid is locally accelerated over the vane surfaces.
Where vaporization occurs small cavities or bubbles filled with vapour form near the points of
lowest pressure. Subsequently these expand whilst moving forward with the flow and then
suddenly collapse on reaching zones of higher pressures. Experiments show that the rate of
growth and collapse of these bubbles is extremely high and the local pressures, resulting from the
sudden collapse of the bubbles near the surfaces, are of the order of thousands of pounds per
square inch.
The undesirable features of cavitation are:

pitting of the surfaces which is due to the continuous hammering action of the collapsing
cavities;

marked drop in efficiency, owing to vapour formation which increases the volume of the
fluid;

knocking and vibration of the machine resulting in noise, emanating from cavitation
zones.

In order to avoid cavitation it is necessary to ensure that the lowest pressure regions are
maintained above vapour pressure. The critical value of the net suction head H s which mark the
appearance of cavitation is determined experimentally.

Fig. 10.4 Pump characteristics


At critical height the performance of the pump begins to deteriorate and eventually the flow may
break down.
10.2.4 Similitude laws of Centrifugal Pumps
To have complete similitude of two centrifugal pumps, the following conditions must be fulfilled:
112

Q
nD

H
2

n D
P
3

n D

where: Q

= ct .
= ct .

= ct.

Q1
n1 D1
H1
2

n1 D1

P1
3

1 n1 D1

(10.6)

Q2

n2 D2

=
=

(10.7)

H2
2

n2 D2

(10.8)

P2
3

2 n2 D2

discharge;

speed of the pump [rot/min];

characteristic dimension (diameter);

density of working fluid.

These are called the similitude laws of the centrifugal pump. Are valid also for other turbomachines.

10.3 AXIAL FLOW MACHINERY


In an axial flow machine the fluid passes through in a substantially axial direction. The impeller
blades are enclosed by a cylindrical duct of sufficient length to provide uniform flow on both the
upstream and downstream sides. Upstream from the impeller, the duct is usually provided with a
bell-mouthed intake to avoid sudden contraction of the stream, whilst immediately downstream
from the impeller straightener vanes are employed to remove the whirling motion of the fluid. Since
the flow contracts on approaching the impeller a suitably shaped hub is provided in front of the
rotor and downstream from the straightener vanes a tail-fairing aids diffusion.

Fig. 10.5 Schematic of an axial flow machine


10.3.1 Axial fans
The fans are hydraulic generators that are working with gases. In this way, they transform
mechanical energy supplied by an electrical motor in pneumatic energy. Between inlet and outlet,
there will be an increase of pressure. A particular characteristic of axial fans is that they are used
for significant volumetric flow rate of gases at small pressure.
113

Internal characteristic of a fan represent the dependency between total pressure ptot of fan and
mass flow rate Qm (or volumetric flow rate) of this, ptot = f ( Qm ) (or ptot = f ( Q ) ) and describe
the working behaviour of the fan.
The mass flow rate (or volumetric flow rate) is defined as flux of velocity through the inlet section
Ai (or outlet section Ao ) per unit time.

(Qm )i,o = ( A n dA)i,o

(10.9)

where n is velocity of fluid through the control section, that is normal to the flow direction.
Total pressure ptot of the fan is the pressure change of the gas through fan (the difference of
average total pressure between inlet and outlet):

ptot = ( ptot )o ( ptot )i = ( p st + pdin )o ( p st + pdin )i


where:

( pst )i , o

- static pressure into inlet section and outlet section, respectively;

( pdin )i ,o

- average dynamic pressure in the same sections.

(10.10)

Seldom, in practical applications, total pressure can be determined with equation:

ptot = ( p st +
where:

( )r,a

2
2

)o ( p st +

2
2

)i

(10.11)

- average velocities into inlet and exit sections;

Taking into consideration that useful power is defined as real power transferred to the gas, we can
make the following statement: energetically, ptot represent the hydraulic power per flow rate unit.
Pu = Q ptot

A typical characteristic curve of an axial fan is shown in Figure 10.6.

Fig. 10.6 Axial fan characteristic

114

(10.12)

10.4 INTRODUCTION TO HYDRAULIC TURBINES


Hydraulic turbines fall into that group of rotodynamic machinery which converts fluid pressure into
useful power. The fluid medium is invariably water and its supply to the turbine must come from a
certain height above the installation. Since the power is proportional to the product of supply Q
(flow rate) and head H (height of water pressure), it is evident that a given power requirement
specifies the product Q H . For that product to remain constant one needs:
(a) either values of high H and low Q or
(b) low H and high Q .
The first class (a) will be satisfied by the use of action (impulse) turbines and class (b) by reaction
turbines. This classification is rather broad as there is some overlapping of conditions when both
action or reaction turbines may be used.
Impulse occurs when the direction of the fluid is changed with no pressure change. It follows that
the magnitude of the velocity remains unchanged.
Reaction occurs when the water is accelerated or decelerated over the vanes. A force is needed to
do this and the reaction to this force acts on the vanes.
10.4.1 Action (Impulse) Turbine. Pelton Turbine
In impulse turbines the head available is first completely converted into kinetic energy in one or
more nozzles; the emerging jet then successively engages suitably shaped vanes called buckets
spaced at equal pitch, which deflect the stream, and a force is created due to the change in
momentum. The vanes are fitted to the perimeter of a circular disc which maintains steady rotation
resulting from the tangential jet force. It is noted that the jet leaving the nozzle is surrounded by
atmospheric pressure which also prevails both on entering and leaving the vane system.

Fig. 10.7 - Schematic diagram of Pelton wheel with two nozzles


115

The most known impulse turbine is Pelton wheel which are mainly used with high pressure heads
such as in mountain hydroelectric schemes. Pelton turbines are usually arranged with horizontal
shafting supporting single or twin runners. Power may be derived from a single jet or from twin jets
(see Fig. 10.7), although some designs embody as many as four jets. The nozzle (1) incorporating
the needle (2) must be carefully streamlined to produce parallel flow at exit and the surfaces must
be smooth to reduce losses resulting from friction. In order to keep the boundary layer as thin as
possible the velocity should be increased quickly at the nozzle tip and for this reason the nozzle
must be short. The buckets (3) may be cast in one piece with the disc (4) or may be separately
cast, machined and bolted to the disc.

Fig. 10.8 Picture of a Pelton wheel with case removed


In designing the buckets care must be taken that no water is allowed to pass through the wheel
without reacting fully on the buckets. To avoid shock losses upon contact with the jet the buckets
must be carefully shaped. Their number and size depends to some extent upon jet diameter
although experiments indicate that a large number of buckets is desirable. The casing (5), housing
the wheel and nozzle, is usually made in two parts and is open at the bottom to allow the water to
escape freely.
10.4.2 Reaction Turbines
In reaction turbines the head available is partially transformed into kinetic energy in a set of
stationary guide vanes, the balance being transformed inside the runner. Whilst in the impulse
turbine the jet engages a few vanes at a time and leaves the rest rotating in the surrounding air, in
the reaction turbine fluid fills the runner passages completely. Two distinct types of reaction
turbines are being produced at the present time: the Francis and the Kaplan.
Francis Turbine
The Francis wheel is an example of a mixed impulse and reaction turbine. They are adaptable to
varying heads and flows and may be run in reverse as a pump such as on a pumped storage
scheme. The principle of operation of the Francis type turbines is similar to a centrifugal pump
working under reversed flow conditions. Whilst in centrifugal pumps a whirl is developed in the
impeller and subsequently transformed into pressure, in reaction turbines one portion of the head
116

available is first transformed into a whirl and is subsequently converted into power inside the
runner. The other portion of the head is employed for the acceleration of the relative velocity of the
water passing through the channels formed by the runner vanes. The vanes themselves are curved
surfaces. The impulse part is due to the guide vanes which are used to produce an initial velocity

which is directed at the rotor. The diagram form figure 10.9 shows the layout of a vertical axis
Francis wheel
Water first enters a spiral casing (2) from which it flows radially through a set of fixed guide vanes
(3) arranged in the form of a ring around a second inner ring of adjustable guide vanes, frequently
called wicket gates (4). The flow turns inside the runner (1) and leaves in the axial direction; it then
enters the diffuser or draft tube 5 and finally discharges into the tail-race.
At the throat of the draft tube the pressure may be well below atmospheric which compensates for the
height at which the turbine runner is set. The draft tube serves a dual purpose: it reduces the velocity
from throat to exit and permits the turbine to be set above the level of the free surface of the water.

Fig. 10.9 Schematic section of a Francis turbine


Kaplan
The Kaplan turbine is a pure reaction turbine. The main point concerning this is that all the flow energy
and pressure is expended over the rotor and not in the supply nozzles. The picture below shows the
rotor of a large Kaplan turbine. They are most suited to low pressure heads and large flow rates such
as on dams and tidal barrage schemes. The principle of operation of the Kaplan type turbine is similar
to an axial flow pump in which the stream passes over blades of hydrofoil cross-section.

The guide vanes (3) for the Kaplan turbine (see figure 10.11) are arranged in the same way as for
the Francis and produce a whirl in the whirl chamber (4). The runner (1) itself is found at the throat
of the ducting and a considerable space exists between the ends of the guide vanes and the
leading edge of the propeller. Within this space the flow direction changes from radial to axial and
simultaneously the flow pattern approaches, but does not fully attain the free vortex state. Tests
117

show that at the inlet to the propeller the whirl velocity is not exactly inversely proportional to the
radius because the flow had not had time to reach the ideal state of radial equilibrium before
entering the wheel blading. The runner blades remove this whirl and transform it into useful power.
A suitably shaped hub (2) is provided in the downstream of the turbine to avoid shock losses in the
draft tube (5).

Fig. 10.10 Picture of the rotor of a large Kaplan turbine

Fig. 10.11 Schematic section of a Kaplan turbine


The advantage of Kaplan turbines over the Francis type lies in the fact that both guide vane and
blade setting angles can be adjusted simultaneously to compensate for changing flow or load
conditions whilst the Frances turbine with fixed blade runners can only change the wicket gate
setting. Thus, off-design-point operation results in shock losses in the Francis turbine with a
subsequent loss in efficiency, whilst the Kaplan maintains optimum operation over a wider range of
loading. It is noted that the speed of these machines is held constant under changes of load.

118

11. DIMENSIONAL ANALYSIS


11.1 INTRODUCTION
In engineering the application of fluid mechanics in designs make much of the use of empirical
results from a lot of experiments. This data is often difficult to present in a readable form. Even
from graphs it may be difficult to interpret. Dimensional analysis provides a strategy for choosing
relevant data and how it should be presented.
This is a useful technique in all experimentally based areas of engineering. If it is possible to
identify the factors involved in a physical situation, dimensional analysis can form a relationship
between them. The resulting expressions may not at first sight appear rigorous but these
qualitative results converted to quantitative forms can be used to obtain any unknown factors from
experimental analysis.

11.2 DIMENSIONS AND UNITS


Any physical situation can be described by certain familiar properties e.g. length, velocity, area,
volume, acceleration etc. These are all known as dimensions. Of course dimensions are of no use
without a magnitude being attached. We must know more than that something has a length. It must
also have a standardised unit, such as a meter, a foot, a yard etc.
In consequence it can said that the dimensions are properties which can be measured, and units
are the standard elements used to quantify these dimensions. In dimensional analysis we are only
concerned with the nature of the dimension i.e. its quality not its quantity. Generally, the following
common abbreviations are used:

length = L ;
mass = M ;
time = T ;
force = F ;

temperature = .

In this chapter we are only concerned with L , M , T and F . We can represent all the physical
properties we are interested in with L , T and one of M or F , because F can be represented by
a combination of LTM . In the following are presented the dimensions of most common physical
quantities:
QUANTITY

SI Unit

DIMENSION

velocity

m s -1

L T -1

acceleration

m s -2

L T -2

force

N (Newton)

119

kg m s -2

pressure

M L T -2

Pa (Pascal) or N/m 2
kg m

-1

M L-1 T -2

-2

kg m -3

density

M L-3

N/m 3

specific weight

kg m

-2

M L-2 T -2

-2

N s/m 2

(dynamic) viscosity

M L-1 T -1

kg m -1 s -1
J (Joule) or N m

energy (or work)

kg m s

M L2 T -2

-2

W ( Watt) or N m/s

power

M L2 T -3

kg m 2 s 3

11. 3 DIMENSIONAL HOMOGENEITY


Any equation describing a physical situation will only be true if both sides have the same
dimensions. That is it must be dimensionally homogenous. This property of dimensional
homogeneity can be useful for:
1. Checking units of equations.
2. Converting between two sets of units.
3. Defining dimensionless relationships.
Example

Show that the equation:


Power = Force velocity ,

is homogeneous in both SI units and basic dimensions.


Solution:
The equation to be checked is:
P =F v

The SI unit of power is the Watt. The Watt is a Joule per second and a Joule is a Newton metre of
energy. Hence a Watt is 1 N m/s .
Also the SI units of force and velocity are Newton and respectively the metre/second. Hence, the
unit of F v is N m/s , and the equation is homogeneous in SI units.
Writing the MLT dimensions of each term we have:
[P] = M L2 T -3

[F] = M L T -2

Substituting into the equation we have:


120

[v] = L T -1

M L2 T -3 = M L T -2 L T -1 = M L 2 T -3 ,

The equation is homogeneous also in basic dimensions.

11.4 Results of dimensional analysis


The result of performing dimensional analysis on a physical problem is a single equation. This
equation relates all of the physical factors involved to one another. This is probably best seen in an
example.
If we want to find the force on a propeller blade we must first decide what might influence this
force. It would be reasonable to assume that the force F depends on the following physical
properties:
d

diameter;

forward velocity of the propeller (velocity of the plane);

fluid density;

speed (revolutions per second);

fluid viscosity.

In this way the following equations can be write:


F = f ( d, u, , n, ) ;

or
0 = f1 (F, d, u, , n, ) ,

where f and f1 are unknown functions.


These can be expanded into an infinite series which can be reduced to:
F = K d a1 u a2 a3 n a4 a5

where: K
a1 , a2 , a3 , a4 , a5

is a constant;
are unknown constant powers.

From dimensional analysis can be obtain these powers and can be arranged the variables into
several dimensionless groups. The value of K or the functions f and f1 must be determined from
experiment. The knowledge of the dimensionless groups often helps in deciding what experimental
measurements should be taken.

11.5 BUCKINGHAM'S THEOREMS


Although there are other methods of performing dimensional analysis, as the indecial method, the
method based on the Buckingham theorems gives a good strategy for obtaining a solution, as
shown below.

121

11.5.1 1st Theorem


A relationship between m variables (physical properties such as velocity, density etc.) can be
expressed as a relationship between m-n non-dimensional groups of variables (called groups),
where n is the number of the fundamental dimensions (such as mass, length and time) required to
express the variables.
In the consequence if a physical problem can be expressed as:
f ( Q1 , Q2 , Q3 ,..., Qm ) = 0 ,

then, according to 1st theorem, this can also be expressed as:


f ( 1 , 2 , 3 ,..., Qm - n ) = 0 .

In fluids, usually n = 3 , corresponding to M , L , T .

11.5.2 2nd Theorem


Each group is a function of n governing or repeating variables plus one of the remaining
variables.

11.5.3 Choice of repeating variables


Repeating variables are those which will appear in all or most of the groups, and have influence
in the problem. Before commencing the analysis of a problem, the repeating variables must
choused. There are some basic rules which should be followed.
1. From the 2nd theorem there can be n = 3 repeating variables.
2. When are combined, these repeating variables variable must contain all of dimensions M , L ,
T (that is not to say that each must contain M , L and T ).

3. A combination of the repeating variables must not form a dimensionless group.


4. The repeating variables must not to appear in all groups.
5. The repeating variables should be chosen to be measurable in an experimental investigation.
They should be of major interest to the designer. For example, a pipe diameter (dimension L )
is more useful and measurable than roughness (also dimension L ).
6. If, when a problem is defined, unimportant variables are introduced, then extra groups will
be formed. They will play a not significant role in the physical behaviour of the problem and
should be identified during experimental work. If an important variable was missed, then a
relevant group would be missing. Experimental analysis based on these results may miss
significant behavioral changes. It is therefore, very important that the initial choice of variables
must carried out very carefully.
Notice: In fluids it is usually possible to take , v and d as the three repeating variables.

122

11.5.4 Operations with the groups


Once the groups are identified, manipulation of these is permitted. These operations do not
change the number of groups involved, but may change their appearance significantly.
For the equation
f ( 1 , 2 , 3 ,..., m - n ) = 0 .

the following operations can be performed:


1. Any number of groups can be combined by multiplication or division to form a new group which
replaces one of the existing. E.g. 1 and 2 may be combined to form 12 = 1 / 2 so the
defining equation becomes:
f ( 12 , 2 , 3 ,..., m - n ) = 0 .

2. The reciprocal of any dimensionless group is valid:


f ( 1 ,

, 3 ,...,

m-n

)=0.

3. Any dimensionless group may be raised to any power:


am n
f ( 1a1 , 2a2 , 3a3 ,..., m
n ) = 0 .

4. Any dimensionless group may be multiplied by a constant:


. f (a1 1 , a1 2 , a3 3 ,..., am - m m - n ) = 0
5. Any group may be expressed as a function of the other groups:

2 = f ( 1 , 3 ,..., m - n ) = 0 .

11.5.5 Example
Taking the example discussed above of force F induced on a propeller blade, we have the
equation:
0 = f (F, d, u, , n, ) ,

Thus m = 6, n = 3, corresponding to M , L , T , and there are m - n = 3 groups. Previous equation


can be write as:
f ( 1, 2 , 3 ) = 0 .

The choice of , u , d as the repeating variables satisfies the criteria presented in 11.5.3. They
are measurable, good design parameters and, in combination, contain all the dimension M , L , T .
Hence, three groups can be formed according to the 2nd theorem:

1 = a1 u b1 d c1 F ;

2 = a2 u b2 d c2 n ; 3 = a3 u b3 d c3 .

As the groups are all dimensionless i.e. they have dimensions M 0 L0 T 0 , the principle of
dimensional homogeneity can be used to solve the dimensions for each group. For the first
group, in terms of SI units, it can be write:
1 = ( kg m -3 )a1 (m s -1 )b1 (m)c1 (kg m s -2 ) .

In terms of dimensions the first group can be write as:


123

M 0 L0 T 0 = ( M L-3 )a1 (L T -1 )b1 (L)c1 (M L T -2 ) .

For each dimension M , L and T the powers must be equal on both sides of the equation, thus:
-

for M :

0 = a1 + 1 a1 = 1.

for L :

0 = - 3 a1 + b1 + c1 + 1 0 = 4 + b1 + c1 .

for T :

0 = -b1 - 2 b1 = 2

From previous equation:


c1 = - 4 - b1 = - 2

In this way, the first 1 can be expressed as:

1 = 1 u - 2 d - 2 F 1 =

u-2 d -2

A similar procedure can be followed for the other groups (self-assessment exercise), giving:

2 = 0 u -1 d 1 n 2 =
3 = 1 u -1 d -1 3 =

nd
;
u

ud

Thus the problem may be described by the following function of the three non-dimensional
groups:
f ( 1 , 2 , 3 ) = 0

F
2

u d

nd

.
= f
,
u ud

11.6 COMMON GROUPS


During dimensional analysis several groups can appear for different problems. These often have
names (in honour of some known physicists). Most common non-dimensional numbers (groups)
are the following:
-

Re =

vd

Reynolds number: inertial, viscous force ratio;

Ma =

v
c

Mach number: local velocity, local velocity of sound ratio.

Eu =

Fr =

We =

v2
v2
:
gd

vd
:

Euler number: pressure, inertial force ratio;

Froude number: inertial, gravitational force ratio;

Weber number: inertial, surface tension force ratio.

124

11.7 SCALE MODELS TESTING


In order to predict the performance of some real things, seldom scale models are tested. In these
cases the results are only valid when the forces acting on the model are in the same ratio to each
other as they are on the real thing. When this occurs the dynamic similarity is fulfil. This must be
also accomplished by the geometric similarity.

11.7.1 Geometric similarity


Geometric similarity exists between model and real thing if the ratio of all corresponding
dimensions in the model and real thing are equal:
Lmod el
L
= m = L ,
Lreal thing
Lr

For lengths:

where: L is the scale factor for length.


Amod el
L2
= m
= 2L .
2
Areal thing
Lr

For areas:

Notice: All corresponding angles are the same.

11.7.2 Kinematic similarity


Kinematic similarity is the similarity of time as well as geometry. It exists between model and
prototype if:
1. the paths of moving particles are geometrically similar
2. the rations of the velocities of particles are similar
Some useful ratios are:
-

Velocity:

v m Lm Tm

=
= L = v ;
vr
Lr Tr
T

Acceleration:

am Lm Tm2

=
= 2L = a ;
2
ar
Lr Tr
T

Discharge:

Qm L3m Tm
3
= 3
= L = Q .
Qr
T
Lr Tr

Notice: This has the consequence that streamline patterns are the same.

11.7.3 Dynamic similarity


Dynamic similarity exists between geometrically and kinematically similar systems if the ratios of all
forces in the model and real thing are the same.
2

Force ratio:


Fm M m am
L3
=
= m 3m 2L = 2L L = 2L v2 .
Fr
M r ar
r Lr T
T

125

This occurs when the controlling dimensionless group on the right side of the defining equation is
the same for model and real thing.

11.7.4 Dynamically similar model examples


Example 1
A model aeroplane is built at 1:10 scale and is to be tested in a wind tunnel operating at a pressure
of 20 times atmospheric. The real aeroplane is design to fly at 500 km/h. At what speed should the
wind tunnel operate to give dynamic similarity between the model and prototype? If the drag
measure on the model is 337.5 N what will be the drag on the plane?
Solution:
The drag D of a body moving through a fluid is dependent on the following physical properties:
-

v (L T -1 )

relative velocity between fluid an solid;

l (L)

characteristic length of the flowing (e.g. mean aerodynamic chord);

(M L-3 )

density of the fluid;

(M L-1 T -1 )

dynamic viscosity of the fluid.

In terms of dimensional analysis the defining equation for this case is:
f (D, v, l, , ) = 0 .

Thus, m = 5, n = 3, so there are 5 - 3 = 2 groups:


For the first group:

1 = a1 v b1 d c1 D ;

2 = a2 v b2 l c2 .

M 0 L0 T 0 = ( M L-3 )a1 (L T -1 )b1 (L)c1 (M L T -2 ) ,

leading to:

1 =

v2 l2

For the second group:


M 0 L0 T 0 = ( M L-3 )a2 (L T -1 )b2 (L)c2 (M L-1 T -1 ) ,

leading to:

2 =

=
= 2a .
v l Re

Notice: 1 / 2 is the Reynolds number, Re . This can be denoted as 2 a .


The defining equation for resistance to motion becomes:
D

= 0 D = v 2 l 2 f (Re) .
f ( 1 , 2 a ) = 0 f
,
Re
2 2

v l

126

This equation applies whatever the size of the body i.e. it is applicable to a to the prototype and a
geometrically similar model. In this way, in order to fulfil the dynamic similarity the two Reynolds
numbers must be equals:
Rem = Rer v m = v r

r dr r
m dm m

The value of does not change much with pressure so m = r .


From the equation of state for an ideal gas:
p = RT ,

if temperature is the same, then the density of the air in the model can be obtained as:
pm 20 pr

=
= m m = 20 r
pr
pr
r

Hence, the model velocity is:


vm = vr

1
1
1
= v r = 250 km/h = 69.44 m/s.
20 1 / 10 2

The ratio of forces is found from:


Dm ( v 2 l 2 )m 20 0.5 2 0.12
1
=
=
Dr =
Dm = 6750 N .
2
2
Dr
1 1
1
0.05
( v l )r

Example 2
A submarine having diameter 2 m and length 10 m is tested in a water tunnel to determine the
forces acting on the real prototype. A 1:10 scale model is to be used. If the maximum allowable
speed of the prototype submarine is 2 m/s, what should be the speed of the water in the tunnel to
achieve dynamic similarity?
Solution:
For dynamic similarity the Reynolds number of the model and prototype must be equal:
vd
vd
=

Rem = Rer
m r

Thus the model velocity should be:


vm = vr

r dr r
.
m dm m

As both the model and prototype are in water then, r = m and r = m , thus:
vm = vr

dr
2
=2
= 40 m / s .
dm
1/10

Observations:
-

This is a high velocity for water. This is one reason why model tests are not always done at
exactly equal Reynolds numbers. Some relaxation of the equivalence requirement is often
acceptable when the Reynolds number is very high.
127

Using a wind tunnel may have been possible in this example. If this were the case then the
appropriate values of the and ratios need to be used in the above equation.

SELF-ASSESSMENT EXERCISES
Exercise 1
The discharge Q through an orifice is a function of the diameter d , the pressure difference p , the
density , and the viscosity . Show that:

Q=

1
12
p

1
2

1
1

d 2 p2
,

where is some unknown function.


Exercise 2
The pressure drop per unit length p due to friction in a pipe depends upon the diameter d , mean
velocity v , density and the dynamic viscosity . Find the relationship between these variables.

128

You might also like