Sei sulla pagina 1di 97

HYDROMECHANICS

First Edition

Vladimr Hork
University of Defence,
Brno

Vladimir V. Kulish
Nanyang Technological University,
Singapore

University of Defence, Brno 2010

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Tento studijn text je spolufinancovn Evropskm socilnm fondem


a sttnm rozpotem esk republiky.

Copyright 2010 Vladimr Hork and Vladimir V. Kulish


ISBN: 978-80-7231-725-7

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Contents
Preface..................................................................................................................................... 5
Nomenclature .......................................................................................................................... 6
1 Basic Terms and Quantities of Hydromechanics ............................................................... 8
1.1 Basic State Variables of Fluid ........................................................................................... 8
1.2 Physical Properties of Fluids............................................................................................. 9
1.3 Friction Forces in the Fluid Flow.................................................................................... 11
1.4 Surface Tension of Liquids ............................................................................................. 14
2 Hydrostatics ......................................................................................................................... 15
2.1 Hydrostatic Balance in the Gravitational Field ............................................................... 15
2.2 Hydrostatic Pressure ....................................................................................................... 17
2.3 Pressure Measurements ................................................................................................... 18
2.4 Pascal's Law .................................................................................................................... 19
2.5 Hydrostatic Forces on Submerged Surfaces ................................................................... 20
2.6 Liquid in Linearly Accelerated System .......................................................................... 25
2.7 Liquid in Rotating System .............................................................................................. 27
2.8 Archimedes Law, Buoyancy and Stability .................................................................... 29
3 Flow of Ideal Fluid .............................................................................................................. 31
3.1 Basic Phenomena and Concepts of Fluid Flow .............................................................. 31
3.2 Conservation of Mass Continuity Equation ................................................................. 32
3.3 Conservation of Momentum ........................................................................................... 33
3.4 Basic Equations of an Ideal Fluid Flow .......................................................................... 35
3.5 Practical Applications of an Ideal Fluid Flow ................................................................ 36
3.5.1 Discharge through a small outlet ......................................................................................... 36
3.5.2 Discharge through a large opening ...................................................................................... 37
3.5.3 Discharge through a submerged opening............................................................................. 39
3.5.4 Venturi tube ......................................................................................................................... 40
3.5.5 Pitot and Prandtl tubes ......................................................................................................... 41

4 Flow Similarity Criteria...................................................................................................... 43


4.1 Laminar Flow .................................................................................................................. 43
4.2 Turbulent Flow................................................................................................................ 43
4.3 Theory of Similarity ........................................................................................................ 44
4.3.1 Geometric similarity ............................................................................................................ 44
4.3.2 Kinematic similarity ............................................................................................................ 44
4.3.3 Dynamic similarity .............................................................................................................. 45
4.3.4 Overall and partial similarity ............................................................................................... 48

5 Fluid Flow Losses ................................................................................................................ 50


5.1 Basic Laws of Flow with Friction Losses ....................................................................... 50
5.1.1 Conservation of mass: continuity equation.......................................................................... 50
5.1.2 Conservation of momentum ................................................................................................ 50
5.1.3 Energy conservation ............................................................................................................ 52

5.2 Specific Energy Loss Due to Friction ............................................................................. 52


5.2.1 Straight pipe friction loss ..................................................................................................... 52
5.2.2 The Moody diagram ............................................................................................................ 53
5.2.3 Local losses.......................................................................................................................... 57
5.2.4 Local loss coefficients ......................................................................................................... 58
5.2.5 Resulting specific energy loss ............................................................................................. 61
5.2.6 Pressure loss, head loss, power loss..................................................................................... 61

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

6 Flow with External Energy Supply.................................................................................... 66


6.1 Basic Laws of Flow with External Energy Supply ......................................................... 66
6.1.1 Conservation of mass continuity equation ........................................................................ 66
6.1.2 Conservation of momentum................................................................................................ 66

6.2 Pump Specific Energy..................................................................................................... 68


6.3 Pump Characteristics ...................................................................................................... 68
6.4 Pump Operation at Various Speeds ................................................................................ 71
6.5 Integrating the Pump into Piping System ....................................................................... 71
6.6 Piping Systems in Series and Parallel ............................................................................. 75
6.7 Suction Head, Cavitation ................................................................................................ 75
7 Fluid Flow through Orifices, Mouthpieces, and Nozzles ................................................ 78
7.1 Coefficients of Contraction, Velocity, and Discharge .................................................... 78
7.2 Orifices, Mouthpieces, Nozzles ...................................................................................... 79
7.3 Orifice Plates, Nozzles, Venturi Tubes ........................................................................... 81
8 Momentum and Forces in Fluid Flow ............................................................................... 85
8.1 Control Surface and Control Volume ............................................................................. 85
8.2 Linear Momentum Principle ........................................................................................... 85
8.3 Applications of the Momentum Principle ....................................................................... 87
8.3.1 Jet engine thrust ................................................................................................................... 87
8.3.2 Rocket engine thrust ............................................................................................................ 88
8.3.3 Force of the jet acting upon a flat plate at rest ..................................................................... 89
8.3.4 Force of the jet acting upon the curved vane at rest ............................................................ 89
8.3.5 Force of the jet acting upon the moving plate ..................................................................... 89
8.3.6 Power of the water turbine wheel ........................................................................................ 90
8.3.7 Force of the fluid flow through a reducing bend ................................................................. 91

English-Czech glossary of basic terms used in hydromechanics .......................................... 93


Literature ............................................................................................................................... 96

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Preface
These lecture notes are the basic study material for courses in Hydromechanics, and are
intended for the use by undergraduate students in mechanical engineering of the Military
Technologies study program. Students are expected to have basic knowledge in mathematics
and physics from previous courses. The Czech version of these lecture notes [7] is available
for Czech speaking students.
Hydromechanics is a branch of the general mechanics. In a way similar to the mechanics
of rigid bodies, hydromechanics deals with the balance of forces acting upon fluids in motion
and at rest. Analyzing this motion, the knowledge of inherent laws of rigid-body mechanics is
used. Fluid mechanics describes the matter by macroscopic quantities. Fluid is viewed as a
continuum. The microscopic motion (the motion of molecules, atoms, ions) is neglected. The
continuum mechanics assumes the properties of a continual homogenous environment in any
arbitrarily small elemental volumes remain the same.
A fundamental difference between fluids and rigid bodies lies in the molecular
movability of liquids and gases. In comparison with rigid bodies, the fluid is irreversibly
deformed, does not have its own shape and its particles are put in motion by small forces.
Otherwise, the rigid body is moved as a solid assembly of mass points (if neglect
deformations).
The lecture notes give an integrated view of hydromechanics. The interpretation of
problems comes from the traditional fluid mechanics using the laws of balance of forces and
moments from applied loads, conservation principles of mass and energy, and the momentum
theorem. Analytical solutions are difficult to obtain in many cases. Hence, it is necessary to
take advantage of empirical methods. The theory of similarity is frequently used to generalize
experimental results.
Illustrative problems are often followed by explanation notes, practical remarks, and
examples. Notes and examples are printed in smaller size fonts.
Important passages are emphasized by italics, whereas the fundamental terms are printed
in bold italics. Italic characters are also used for variables within the text. The text in italics is
also used in pictures and tables captions.
The text contains the tables of physical quantities and properties of fluids to enable
students to easily find necessary data for exercises and individual work.
The nomenclature and the units used in the present work follow the International System
of Units (SI) and actual normative technical standards [21-24].
Special thanks go to reviewers: Prof Vclav Tesa, Prof Jaroslav Salga, and Assoc Prof
Ji Maxa, who have spent their time evaluating this text.
Authors

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Nomenclature
A
a
a
B, b
C
Cc
Cd

Area, cross-sectional area


Speed of sound [Chapters 1 and 4]
Acceleration [Chapter 2]
Width of channel
Flow rate coefficient
Coefficient of contraction
Coefficient of discharge

Cv

Coefficient of velocity

c
cF
cM
cp
cp , cV
cd
D, d
Eu
Dxy
e
ez
F
Fm
Fex
Fs
Fp
Ffr
Fb
Fr
f
g
H, h
hM
hz
Ix
K
k
L, l
M
Ma
m
mc
P
Pef
Pmom
Pz
p

Specific heat capacity


Force coefficient
Moment coefficient
Pressure coefficient [Chapter 4]
Const.-pressure, const.-volume specific heat [Chapter 1]
Drag coefficient
Diameter
Euler number
Product moment of area about the x and y axes
Specific energy
Specific energy loss
Force
Mass force
External mechanical force
Force of inertia
Pressure force
Friction force
Buoyancy force
Froude number
Frequency
Acceleration due to gravity
Head, height
Metacentric height
Head loss
Second moment of area about x-axis
Bulk modulus
Hydraulic roughness
Length, characteristic length (dimension)
Moment
Mach number
Mass
Area ratio
Power
Effective power
Rate of momentum
Loss power
Pressure
6

m2
m s-1
m s-2
m

J kg-1K-1

J kg-1K-1
m
m4
J kg-1
J kg-1
N
N
N
N
N
N
N
s-1
m s-2
m
m
m
m4
Pa
m
m
Nm
kg
W
W
kg m s-2
W
Pa

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

pa
pp,, pv

Atmospheric (barometric) pressure


Positive gauge pressure, negative gauge pressure
p z
Pressure loss
p0
Stagnation pressure
p
Pressure of saturated steam
Pr
Prandtl number
Q
Volumetric flow rate, capacity
Qm
Mass flow rate
R
Fluid flow force, reaction force [Chapter 8]
R, r
Radius
RPM, RPS Revolutions per minute, revolutions per second
Re
Reynolds number
s
Displacement
Sh
Strouhal number
Sx
First moment of area about x-axis
T
Thermodynamic (absolute) temperature
t
Time
u
Tangential velocity
Volume, specific volume
V, v
v
Velocity
w
Specific mechanical energy
x
Coordinate, distance
Y
Pump specific energy
Yp
Required specific energy
y
Coordinate, distance
z
Coordinate, potential head
Angle [Chapters 2, 3, 5, 8]

Coefficient of linear thermal expansion [Chapter 1]

Coefficient of volumetric thermal expansion [Chapter 1]

Diameter ratio [Chapter 7]

Coefficients of isochoric thermal expansion

Coefficient of volume compressibility (isothermal)


Relative roughness [Chapter 5]

Turbulent (eddy) viscosity [Chapter 1]

ex
Coefficient of expansion
Efficiency [Chapter 6]

Dynamic viscosity [Chapters 1, 4-7]

Angle
Specific heat ratio (adiabatic exponent)

Friction factor [Chapter 5]

Length ratio [Chapter 4]

Kinematic viscosity

Local loss coefficient (factor)


Density (specific mass)

Surface tension
Shear stress

Angular velocity
7

Pa
Pa
Pa
Pa
Pa
m3 s-1
kg s-1
N
m
min-1, s-1
m
m3
K
s
m s-1
m3, m3 kg-1
m s-1
J kg-1
m
J kg-1
J kg-1
m
m
rad, deg
K-1
K-1
K-1
Pa-1

Pa s
deg

m2 s-1
kg m-3
N m-1
Pa
rad s-1

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

1 Basic Terms and Quantities of Hydromechanics


Fluid mechanics is the study of equilibrium and macroscopic motion of fluids and their
interaction with bodies.
The classical fluid mechanics uses:
Laws of the forces and momentum balance
Laws of the mass and energy conservation
Theorem of the rate of change of the linear momentum.
Fluids are substances that continually deform under an applied shear stress, because their
particles are very moveable. Movement of a fluid is called flow (flux, creep). Fluids do not
have their own shape, but take the shape of the container in which they are placed. Fluids
include gases and liquids.
Gases (or vapors) are substances in the gaseous state. Gases fill up the whole space
available to them and do not create level. They are characterized by a high compressibility
and expansibility. The behavior of gases and vapors, mainly in relation to thermal processes,
is studied by thermodynamics, which is the topic of the second part of this course. The air
motion is studied by aerodynamics.
Liquids (e.g. water, oil, gasoline, mercury, etc.) take the shape of the container in which
they are placed, filling up the lower part of that container and form a level surface. Small
volumes of liquid, bounded completely or almost completely by free surfaces, form drops or
droplets. Liquids are practically incompressible, i.e. the increase of pressure causes a
negligible change of volume. The study of liquids in motion and at rest is called
hydromechanics.
The actual liquid is a real fluid, which is characterized by the viscosity (internal friction).
Other properties of the real liquids are compressibility, volume expansion, surface tension,
evaporation, solubility of gases in liquids, etc. The impact of these properties is often
marginal and practically negligible.
Ideal liquid is a theoretically perfect fluid, by which it is assumed that the internal friction
or shear stress is zero (the fluid is non-viscous), the compressibility and solubility of gases is
zero as well, no evaporation is assumed.
1.1 Basic State Variables of Fluid
The state of a fluid is explicitly determined by the set of the state
variables. They include: pressure p, density , and temperature T.
Pressure is the state variable characterizing the mechanical state
of matter. Pressure is defined as the perpendicular component of the
force dF acting upon the area element dA (Fig. 1.1). The SI unit for
pressure is the Pascal
Fig. 1.1 Definition of pressure
dF
.
(1.1)
p=
dA
Depending on the way the pressure is measured
we can distinguish the following quantities (Fig. 1.2):
The barometric (atmospheric) pressure pa
The gauge (manometric) pressure pp
The vacuum pressure pv
The absolute pressure p is the state variable.
The absolute pressure could be expressed by relations:
For the gauge pressure p > pa: p = pa+ pp.
Fig. 1.2 Absolute, gauge,
For the vacuum pressure p < pa: p = pa pv.
and vacuum pressure
8

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The fluid density (specific mass) is defined by the ratio of substance mass element dm to
its volume dV as
dm 1
(1.2)
=
= .
dV v
In thermodynamics, instead of the density the inversely proportional quantity the specific
volume v is usually used.
The fluid mass remains constant during any process, that is m = V = const. By
differentiating we obtain the formula dV + V d = 0 , from where the relation between the
change in volume and change of density follows in the form
dV
d
.
(1.3)
=
V

Temperature characterizes the thermal state of matter. The temperature scale used in SI is
the thermodynamic or the absolute temperature T, also called the Kelvin temperature, which
is related to the Celsius temperature t by
T (K) = t (C) + 273.15 .

(1.4)

1.2 Physical Properties of Fluids


The state of a fluid is changing under acting external forces, or by adding some external
energy by the thermal or mechanical way. The changes of the fluid state are characterized by
thermodynamic coefficients.
The thermal isobaric expansion of
expansion coefficient as
1
=
V

fluids is characterized by the volumetric thermal

1
=

,
T p

(1.5)

which is defined by the proportional volume change when the temperature is varied at a
constant pressure.
Note: This property can also be taken as a result of the linear expansion expressed by the coefficient of
linear thermal expansion
1 L
.
(1.6)
=
L T
There is valid the following relation: 3 .

The isochoric thermal expansion coefficient

(1.7)
T V
is defined as the proportional pressure change when the temperature is varied at a constant
volume.

1
p

The fluid compressibility is characterized by the coefficient of (isothermal) volume


compressibility as
1 V
1
=
,
=
(1.8)
V p T p T
which is defined as the proportional volume change when the pressure is varied at a constant
temperature.
9

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

A mutual relation exists between the above-mentioned coefficients

= p .
The inversely proportional quantity to
the bulk modulus, given by
1
K = = V

(1.9)

the coefficient of volume compressibility is called

p
p

=
.
V T
T

(1.10)

Fluid compressibility is also related to the speed of sound a as the propagation speed of
small pressure waves in the matter. The following definition is valid for fluids

a=

(1.11)

Typical values of the forenamed physical properties for some fluids are given in Table 1.1.
Note: With respect to the derivatives of an ideal-gas state equation p V = m r T , the following
relations for the fluid physical properties are valid:

= =

1
,
T

1
,
p

a=

= rT ,

(1.12)

where is the adiabatic (isentropic) exponent and r is the specific gas constant. These quantities are
given by the specific heat capacities at constant pressure and volume as
cp
and
(1.13)
r = c p cv .
=
cv
Example 1.1
The petrol storage tank of a cylindrical shape with the diameter of 3 m and the length of 8 m is
completely filled with gasoline at temperature 10 C. Determine the amount of gasoline that discharge
from the tank when the temperature increases to 45 C. ( = 0.0012 K-1)
The tank volume:
V=

d2

l=

32

8 = 56.55 m 3

4
4
The change of the fluid volume represents the amount of gasoline that discharge from the tank:

1
V

1 V

p V T

V = V T = 0.0012 56.55 35 = 2.375 m 3

Example 1.2
The hydraulic test of a pipeline with the diameter of 400 mm and the length of 2 km was carried out at
the initial pressure p1 = 7.5 MPa. After some time the pressure drops to p2 = 7 MPa.
Determine the amount of water leakage if the pipe deformation due to pressure changes is neglected.
( = 4.9210-10 Pa-1)
The pipes volume:

V =

d2

0.4 2

2000 = 251.33 m 3
4
4
The change of the fluid volume represents the amount of water leakage from the pipeline:

1
V

l=

V
1 V


V p
p T

10
6
3
V = V p = 4.92 10 251.33 ( 0.5 10 ) = 0.0618 m

10

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Tab. 1.1 Basic physical properties of some fluids


Quantities, symbols, and units

Liquids

Gases

Water

Oil

Benzene

Air

Oxygen Nitrogen

Pressure

Pa

98066

98066

98066

101325

101325

101325

Density

kg m-3

998

866

879

1.29

1.43

1.25

Temperature

20

20

20

Coefficient of volume
expansion

10-6
K-1

180

690

1150

3671

3674

3671

Coefficient of volume
compressibility

10-9
Pa-1

0.468

0.887

0.973

9868

9871

9867

Coefficient of
thermal expansion

Bulk modulus

Speed of sound

Specific heat at
constant volume

10-3
K-1
10-6
Pa

3922

7831

12054

3.672

3.674

3.672

2136.5

1127

1028

0.1419

0.1418

0.1489

m.s-1

1463

1141

1081

331.6

314.9

345.2

cV

J kg-1K-1

4187

1892

1737

720.0

651.9

744.6

Specific heat at
constant pressure

cp

J kg-1K-1

4187

1892

1737

1009

912.7

1042.5

Adiabatic exponent

1.40

1.40

1.40

Specific gas constant

288.3

260.8

297.9

Dynamic viscosity

100.4

3031

66.89

1.7286

1.9448

1.6875

Kinetic viscosity

J kg-1K-1
10-5
Pa s
10-6
m2 s-1

1.006

35.0

0.761

13.4

13.6

13.5

1.3 Friction Forces in the Fluid Flow


If the real fluid is moving, the shear friction occurs on the boundary between concurrent
neighboring fluid layers, which flow at different speeds. The lower speed layer is accelerated
and the faster one, on the contrary, is restrained. This phenomenon is caused by the friction
(shear) force, which is caused by the internal resistance of the fluid against the shear
deformation, the so-called fluid viscosity. The velocity distribution for a plane fluid flow is
shown in Fig. 1.3. The shear (tangential) stress on the boundary between two elemental
layers of the fluid can be defined as
dF fr
,
(1.14)
=
dA fr
where dFfr is the friction force and dAfr is the
corresponding friction surface. The definition of
the friction surface, the direction and physical
meaning of the friction forces, and the fluid
velocity gradient are evident from Fig. 1.3.
11

Fig. 1.3 Shear stress definition

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The value of the shear stress can be determined for the case of laminar flow (see Chap.
4.1) by using Newton's law of viscosity in the form
dv
L =
,
(1.15)
dy
where is the dynamic viscosity and dv dy is the velocity gradient.
The dynamic viscosity is a fluid transport coefficient, the value of which depends on the
pressure and temperature. The dynamic viscosity of liquids decreases with increasing
temperature (see Tab. 1.2 and Tab. 1.3) and is practically independent of pressure. On the
other hand, the dynamic viscosity of gases increases with increasing temperature and
increasing pressure.
Note: Liquids with a constant dynamic viscosity ( = constant) are called Newtonian fluids. If the
dynamic viscosity of liquids is a function of the shear stress, then we have the so-called nonNewtonian fluids. To them belong some emulsions, mixtures of solid and liquid substances, pastes, etc

The kinematic viscosity is an artificially introduced physical quantity defined by the


ratio of the fluid dynamic viscosity and density as
v=

(1.19)

and is frequently used in hydromechanics.


Particular values of the dynamic and kinematic viscosities for some fluids are given in
Tab. 1.1, Tab. 1.2, and Tab. 1.3, as well as in Fig. 1.4 and Fig. 1.5.
The shear stress for the fully developed turbulent flow (see Chap. 4.2) can be determined
by the formula
dv
T = ,
(1.17)
dy
where is called the turbulent (eddy) viscosity. This quantity is not a physical property of the
fluid, because its value depends on both the pressure and the temperature as well as on the socalled mixing length, which is a function of the location within the flow field. The very values
of the mixing length, characterizing the motion of macroscopic particles within the fluid, are
determined experimentally for various cases of the fluid flow.
For the laminar-to-turbulent transition flow, the combined effect of both phenomena is
used, that is
= L + T .
(1.18)
Whence it is evident that = L , if T = 0 (the mixing length is zero), or = T for the
condition of L << T .
Tab. 1.2 The effect of temperature on the density, the dynamic viscosity, and the kinematic
viscosity for water at pressure 98 066 Pa
Temperature
C
0
5
10
15
20
25
30

Density
-3

kg m
999.8
999.9
999.7
999.0
998.2
997.6
995.6

Dynamic viscosity

Kinematic viscosity

-4

10-6 m2 s-1
1.7891
1.5160
1.3010
1.1407
1.0064
0.8968
0.8054

10 Pa
17.887
15.155
13.061
11.406
10.046
8.941
8.019
12

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Tab. 1.3 The effect of temperature on the density, the dynamic viscosity, and the kinematic
viscosity for the engine oil at pressure 98 066 Pa
Temperature
C
20
40
60
80
100
120

Density
-3

kg m
871
858
845
832
820
807

Dynamic viscosity

Kinematic viscosity

-4

10-6 m2 s-1
14.985
7.9312
4.9349
3.4063
2.4390
1.9071

10 Pa
130.52
68.05
41.77
28.34
20.00
15.39

Fig. 1.4 Density and


kinematic viscosity
of water as a function
of temperature at
pressure 98066 Pa

Fig. 1.5 Dynamic viscosity


of air as a function of
temperature at pressure
101 325 Pa

13

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

1.4 Surface Tension of Liquids


A number of physical phenomena are associated with the surface tension of liquids. These
are e.g. the formation of droplets and bubbles, contraction of the liquid jet and its separation,
spraying of liquids, capillary phenomena, wetting the surface of solids, etc. The liquid located
at the boundary plane with another substance (liquid-gas,
liquid-solid) seems to be very thin film. The surface
tension is caused by the molecular forces of the liquid. In
the bulk of the liquid, each molecule is surrounded by
other molecules and the sum of the intermolecular force
on that molecule is zero (see Fig. 1.6). However, the
molecules on the surface of the liquid are not surrounded
by other molecules on all sides and so the intermolecular
Fig. 1.6 The surface tension
force upon that molecule acts towards the liquid.
The surface tension caused by cohesion on the liquid surface can be expressed by the
ratio of the molecular energy Eim on the liquid surface to the area of that boundary Ar. Thus

Eim
.
Ar

(1.19)

Another definition of the surface tension is possible, namely as the ratio of the elementary
tangential component of the attractive intermolecular forces dFt to the length of the boundary
dl, that is
dF
= t .
(1.20)
dl
The surface tension of liquids decreases with increasing temperature and is independent of
pressure. The values of the surface tension for some cases of interaction between two
substances are given in Table 1.4.
Tab. 1.4 The surface tension for some cases of interaction between two substances

Substances:

N m-1

water-air

mercury-air

mercury-water

oil-water

oil-air

0.073

0.461

0.427

0.02

0.03

Capillarity is the ability of a liquid to climb up or fall down (see Fig. 1.7) inside a narrow
tube, called the capillary (the diameter d is less than 5 mm). For instance, due to the surface
tension, the water level in the capillary is above the surrounding liquid, and vice versa, the
level of mercury in the capillary is below the surrounding liquid. Cases of the elevation, i.e.
an increase of the water level with a convex inner
surface in the capillary, and the depression, i.e. a
decrease of the mercury level of the concave inner
surface, are shown in Fig. 1.7.

The relationship for the height h of the


elevation or depression is obtained from the
condition of balance between the gravity force and
the capillary forces , that is
4
.
(1.21)
h=
gd
14

Fig. 1.7 Cases of capillary phenomena

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

2 Hydrostatics
Hydrostatics deals with the problems of the balance of forces acting upon a fluid at rest. A
hydrostatic equilibrium is achieved when the fluid particles do not move relative to each
other. It means that the shape of the fluid volume does not change. If a fluid is at rest there are
no shearing forces and, hence, all derived equations are valid for both ideal and real fluids.
The forces that act upon the fluid can be divided in two groups: surface forces and mass
forces. The surface forces act upon the surface of the fluid control volume and, therefore,
their magnitude depends on the surface area. An example of the surface force is the
hydrostatic pressure in a fluid. The mass forces are proportional to the weight. They are e.g.
gravity force, force of inertia.
We consider primarily the problems of hydrostatic equilibrium in the field of gravity
related to the absolute coordinate system connected with the Earth.
A study of hydrostatics also includes the cases of relative rest, when liquid is at rest in a
container, but the entire system (container-fluid) is accelerated. Generally, the acceleration
may be linear or rotational (angular). Usually it is a linearly accelerated motion or a uniform
rotation. The force balance of fluid is carried out with respect to the relative coordinate
system connected with the container and the fluid is affected by inertial forces due to the
relative motion between both coordinate systems.
The fluid is considered to be incompressible, the density of which can be regarded as a
constant, that is
(2.1)
= constant.
According to the mass conservation principle, the mass of fluid must remain constant
m = constant.

(2.2)

Because there is no energy added or removed, the fluid temperature is constant as well, in
accord with the energy conservation principle. That is

T = constant.

(2.3)

2.1 Hydrostatic Balance in the Gravitational Field


The basic case of hydrostatic balance is related to the coordinate system connected with
the Earth (Fig. 2.1). The fluid is located in the field of gravity and there is no energy addition
or removal.

Fig. 2.1 The hydrostatic balance in the gravitational field


15

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The balance of forces on the fluid element (see Fig. 2.1) of mass dm is given by the sum
of the elemental pressure force dFp and the mass force dFm
dF p + dFm = 0 ,

(2.4)

acting upon the mass element dm in the direction of the z coordinate axis.
The pressure force results from the pressure distribution on the fluid element, as shown in
Fig. 2.1. This force is given by the sum of two surface pressure forces acting perpendicularly
upon both sides of the fluid element, that is in the direction of the z-axis, where the pressure is
varied by dp on the distance dz. It is expressed by the formula
dF p = p dA ( p + dp ) dA = dp dA .

(2.5)

The mass force, acting upon the fluid element of mass dm, is given by the force of gravity
as
dFm = dm g = g dV = g dA dz .

(2.6)

The negative sign shows that the gravitational force acts against the direction of the z-axis.
Upon substituting these forces (2.5) and (2.6) into the balance of forces on the fluid
element, the equation (2.4) takes the form
dp dA g dA dz = 0 .

(2.7)

If we divide the equation (2.7) by the acceleration due to gravity g and the surface of the
fluid element dA, we will obtain the so-called Euler's formula of hydrostatics in the form
dp

+ g dz = 0 .

(2.8)

The general solution of the differential equation (2.8) for a constant density (2.1)
represents the expression of the conservation of energy principle for the fluid at rest, that is
e=

+ g z = constant ,

(2.9)

where e is the integration constant, which represents the overall specific energy of the fluid as
the sum of the pressure energy and the potential energy.
The amount of the fluid specific energy e can be evaluated by using the initial conditions.
For the equipotential surface (i.e. surfaces on which pressure is constant) on the liquid level,
it follows
z = ze

p = pe .

and

(2.10)

Upon substituting the initial conditions into equation (2.9), we can get the value of the
integration constant in the form
e=

pe

+ g ze

(2.11)

and, by comparing the right sides of equations (2.9) and (2.11), we get the final form
p

+g z=

16

pe

+ g ze .

(2.12)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

2.2 Hydrostatic Pressure


The absolute pressure at the given point in the incompressible fluid (Fig. 2.2) being at
rest in the gravitational field can be determined by rearranging equation (2.13) into the form
p = pe + g ( z e z ) ,

(2.13)

where the difference of heights in the z-axis direction represents the distance of the given
point from the equipotential surface (Fig. 2.2), that is the depth
h = ze z .

(2.14)

Fig. 2.2 Hydrostatic pressure

The basic formula of hydrostatics (2.13) can be then rewritten into the commonly used
form
p = pe + g h ,

(2.15)

where the absolute pressure p in the given depth h under the liquid level is equal to the sum of
the pressure pe on this equipotential surface and the hydrostatic pressure g h (the pressure of
the weight of the fluid column).
If the pressure above the liquid level is the atmospheric pressure, p e = p a , the equation
(2.15) becomes
p = pa + g h ,

(2.16)

where the hydrostatic pressure represents the hydrostatic gauge pressure pp = g h.


The pressure differential can be obtained upon differentiating the pressure equation (2.15),
that is
(2.17)
dp = g dh .
The use of the initial conditions for the level of liquid located in the gravity field yields
the equation for the equipotential surface, or the constant-pressure surfaces, as follows
for dp = 0

is

dh = 0 , and so

h = constant.

(2.18)

These constant-pressure surfaces are horizontal straight lines which are perpendicular to
the direction of gravity, i.e. perpendicular to the z-axis.

17

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

2.3 Pressure Measurements


A differential manometer, also known as the two-fluid U-tube manometer, is shown in
Fig. 2.3a. The pressure difference (p2 p1) > 0 causes the level difference hm of the U-tube
pressure. The manometer is filled with two liquids
of different densities m and .
Small reservoirs, placed above the manomater,
have a considerably larger cross-section than the
U-tube. The purpose of this instrument is to ensure
the nearly invariant position of levels inside
reservoirs and to provide a better accuracy of
readings.
The differential manometer is used to measure
the pressure difference p = p 2 p1 by means
of the reading hm as the difference in liquid
column heights within the manometer. For the px
constant-pressure surface of U-tube (see Fig.
2.3a), the pressure balance can be expressed by
from equation (2.15) as
p x = p1 + g h + g m hm ,

and

Fig. 2.3a Schematic of differential


manometer
p x = p 2 + g ( h + hm ) ,

(2.19)

whence the pressure difference can be expressed as


p = p 2 p1 = g ( m ) hm .

(2.20)

If the density of the gauge fluid m is


significantly larger than the density of the
measuring fluid m >> (e.g. measuring gas
pressure as shown in Fig. 2.3b), we can simplify
equation (2.20) by neglecting the gas density to.
Then the equation becomes
p = p 2 p1 = g m hm .

(2.21)

On the contrary, if we want to achieve a


sufficient accuracy of measurement for small
pressure differences p = p2 p1, it is
advantageous to use the gauge fluid, the density of
which is not very different from the measuring
fluid density.

Fig. 2.3b Schematic of gas pressure


measurement

Note: It is obvious, that the gauge fluid and the


measuring fluid must not be mutually miscible.

Example 2.1
Determine the pressure difference measured by the differential manometer (shown in Fig. 2.3a) for the
manometer reading hm = 500 mm. The gauge fluids are mercury m = 13 590 kg m-3 and water
= 1000 kg m-3.
The pressure difference:

p = p 2 p1 = g ( m ) hm = 9.81 (13 590 1000) 0.5 = 61 754 Pa


18

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

A mercury barometer is used to measure atmospheric


(barometric) pressure pa. The barometric glass tube,
closed at one end, is filled with mercury of density
Hg and the open end is placed in a small mercury
reservoir (see Fig. 2.4). If we fill this tube and
reservoir with mercury and then invert, the mercury
column will drop and the closed end of the tube
contains a vacuum.

If we apply equation (2.16) for the atmospheric


pressure pa above the mercury level in the reservoir
and for the vacuum above the mercury column,
where pe = 0, we obtain the following relation for the
atmospheric pressure
p a = g Hg ha .

(2.22)

The atmospheric pressure pa is thus proportional


to the height of the mercury column ha.

Fig. 2.4 Schematic of simple


mercury barometer

Example 2.2
Determine the atmospheric pressure if the height of the mercury column in the barometer is ha = 750
mm. The density of mercury is Hg = 13 590 kg m-3.
The atmospheric pressure:

p a = g Hg ha = 9.81 13 590 0.750 = 99 988 Pa

2.4 Pascal's Law


If the fluid is contained in a closed system in the gravitational field, there are some
practical applications of hydrostatics, where it is possible to neglect the change of the fluid
pressure due to the change of the depth h. A typical example of such a case is the hydraulic
press (see Fig. 2.5).
If the inequality p e g h is valid, we can, according to Pascal's law, neglect the
influence of the depth change in equation (2.15) for the hydrostatic pressure. In this case the
pressure throughout a closed system remains constant
p p e = constant .

(2.23)

Note: In practice, we can apply Pascal's law to engineering devices, where the value of the static
pressure is larger than 1 MPa.

In the case of a hydraulic press (Fig. 2.5), the force F1 is applied to a small piston of the
area A1 and the pressure is transmitted through a hydraulic fluid to a larger piston of the area
A2, the relationship for the force F2 follows from Pascal's law (2.23)

p=

F1 F2
=
A1 A2

F2 = F1

The force F2 is larger than F1 by multiplying factor A2/A1.

19

A2
.
A1

(2.24)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

From the mass conservation principle (2.2)


applied to the hydraulic press shown in Fig. 2.5

m = A1 s1 = A2 s 2 ,

(2.25)

the relationship for the displacement of the


pistons is found
A
s 2 = s1 2 .
(2.26)
A1

Fig. 2.5 Schematic of hydraulic press

A hydraulic pressure booster (Fig. 2.6) is a


device that is used to increase the pressure in
liquids. Using Pascal's law (2.24), it follows the
force balance in the form

p1 A1 = p 2 A2 ,

(2.27)

whence the increased pressure is

p 2 = p1

A1
.
A2

Fig. 2.6 Schematic of


hydraulic booster

(2.28)

Note: Hydraulic brakes, boosters, car lifts, hydraulic jacks, and forklifts make use of these principles.

Example 2.3
Determine the resulting pressure p2 increased by the hydraulic pressure booster as shown in Fig. 2.6.
The lower fluid pressure is p1 = 2.105 Pa, the piston areas are A1 = 10-2 m2 and A2 = 10-4 m2,
respectively.
The increased fluid pressure:

p 2 = p1

A1
10 2
= 2 10 5 4 = 2 10 7 Pa
A2
10

2.5 Hydrostatic Forces on Submerged Surfaces


A surface, which is immersed in the liquid, is loaded with the pressure force. According to
the general definition of pressure (1.1), this force is found by the integration of the pressure
over the surface area

dF = p dA F = p dA = ( p a + g h ) dA .
A

(2.29)

If the system is surrounded by the same ambient pressure pa, than its force effects cancel
out and the resulting force on the submerged surface is caused only by the hydrostatic
pressure. That is
F = g h dA ,
(2.30)
A

where the depth h is, in general, varying.


The hydrostatic force acting upon the flat horizontal bottom of a container, which has a
constant depth h, can be easily determined by the integral equation (2.30) as
F = g h A,

where A is the flat bottom area.


20

(2.31)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The hydrostatic force acting upon the inclined plane surface submerged in a static liquid
of density , as shown in Fig. 2.7, is determined by the solution of the integral equation (2.30)
in terms of
F = g sin y dA ,

(2.32)

where the relation between the depth h


and the y coordinate is given with respect
to to angle as
(2.33)
h = y sin .
The integral in equation (2.32)
represents the first moment of the area
about the x-axis, which is equal to
S x = y dA = y C A ,

(2.34)

where yC is the y-coordinate of the center


of gravity (centroid) C of the surface
area A. Thus the hydrostatic force (2.32)
becomes

Fig. 2.7 Hydrostatic force on inclined


plane surface

F = g sin y C A = g hC A , (2.35)
where hC is the depth of the center of gravity.

The point of application of the hydrostatic force F is not at the center of gravity C, but at a
lower point, called the center of pressure P with coordinates xP, yP (Fig. 2.7). To determine
the application point of the hydrostatic force acting upon the inclined plane surface, we use
the theorem of moment invariants, which states that the resultant moment is equal to the sum
of the moments of the differential forces over the given area. That is
y P F = y dF ,

(2.36)

which can be used to determine the yP coordinate, from equation (2.32), as


yP =

y dF
F

dF

g sin y 2 dA
A

g sin y dA

(2.37)

where we can apply the second moment of the area about the x-axis
I x = y 2 dA

(2.38)

and the first moment of the area about the same axis Sx by equation (2.34). This yields
I
yP = x .
(2.39)
Sx
We can obtain the relation between the coordinates yP and yC by using Steiners theorem
for the second moment of the area about the xC axis that passes through the center of gravity,
as
I x = I xC + A yC2
(2.40)
and then we can rearrange equation (2.39) to the final form
21

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

y P = yC +

I xC
Sx

(2.41)

In a similar way we can get the relationship for the xP coordinate of the center of pressure as
D xy

xP =

Sx

(2.42)

where Dxy is the product moment of the area about the axes x and y given by
D xy = x y dA .

(2.43)

Note: If the area A is symmetrical about the axis yC, then this axis runs through the center of gravity C
as well as through the center of pressure P and then xP = xC.

Example 2.4
Determine the depth hP of the center of pressure and value of the hydrostatic force F acting upon the
inclined plane surface with the angle = 45 submerged by water of the density = 1000 kg m-3. The
surface has the circular shape with the diameter d = 0.1 m. The depth of the surface gravity center (the
center of the circle) is hC = 0.5 m.
The surface area:

A=

d2
4

3.14 0.12
= 7.854 10 3 m 2
4

The coordinate of the center of gravity:

yC =

hC
0.5
=
= 0.707 m
sin 0.707011

The first moment of the area about the x-axis:

S x = y C A = 0.707 7.856 10 3 = 5.554 10 3 m 3


The second moment of the area about the xC axis:

I xC =

d4
64

3.14 0.14
= 4.9087 10 6 m 4
64

The second moment of the area about the x-axis:

I x = I xC + A y C2 = 4.9087 10 6 + 7.854 10 3 0.707 2 = 3.932 10 3 m 4


The resultant hydrostatic force:

F = g hC A = 10 3 9.81 0.5 7.854 10 3 = 38.524 N


The coordinate of the center of pressure:

I x 3.932 10 3
yP =
=
= 0.708 m
S x 5.554 10 3
The depth of the center of pressure:

hP = y P sin = 0.708 0.707011 = 0.50063 m

22

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 2.5
Determine the point of application and the value of the
hydrostatic force F acting upon the vertical rectangular
surface (Fig. 2.8) submerged by water of the density =
1000 kg m-3. The depth of water is h = 1 m and the
width of the surface is b = 2 m.
The surface area:

A = h b = 1 2 = 2 m2
The coordinate of the center of gravity:

h
hC
h
yC =
= 2 =
sin sin 90 2
The first and second moments of the area about the x-axis:

S x = yC A =

h
h2
hb=
b,
2
2

Ix =

Fig. 2.8 Hydrostatic force on


vertical plane surface

h3
b
3

The coordinate of the center of pressure:

h3
b
Ix
2h
3
yP =
= 2 =
= 0.666 m
Sx
3
h
b
2
The resultant hydrostatic force:

F = g hC A = 10 3 9.81 0.5 2 = 9.81 10 3 N


The depth of the center of pressure:

hP = y P sin = 0.666 sin 90 = 0.666 m


Note: The hydrostatic force acting upon the vertical rectangular surface can also be obtained by direct
integration of equation (2.30) as
h

F = g h dA = g h b dh = g b
A

h2
= 10 3 9.81 2 0.5 = 9.81 10 3 N
2

The hydrostatic force acting upon the curved surfaces submerged in a static liquid of the
density and surrounded by the ambient air of the barometric pressure pa, as shown in Fig.
2.9, can be determined as the resultant hydrostatic force F, separated into horizontal and
vertical components Fx and Fz in the direction of the coordinate axes x and z. This curved
surface of area A has the general shape and may be a segment of the submerged wall.
The horizontal and vertical components of the differential hydrostatic force are given by
the following equations:
d Fx = d F cos = g h dA cos = g h dAx ,
(2.44)
and
d Fz = d F sin = g h dA sin = g h dAz .
(2.45)
The horizontal component of the hydrostatic force is found by the integration of equation
(2.44) in the direction of x-axis as
23

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Fx = g S x .

(2.46)

The first moment of the area about the x-axis is


S x = h dAx = hCx Ax .

(2.47)

where Ax is the projected area of the curved surface


on a vertical plane that is perpendicular to the x-axis.
The vertical component of the hydrostatic force
acting upon the area Az in the direction of the z-axis
is given by the weight of the liquid volume V as

Fz = g V ,

(2.48)

V = h dAz .

(2.49)

where
Fig. 2.9 Hydrostatic force on
curved surface

In the case of three-dimensional problems, the horizontal component of the hydrostatic


force acting in the direction of y-axis can be found analogically to equations (2.46) and (2.47)
as solved for the x-axis direction. This yields
Fy = g S y

and

S y = h dAy = hCy Ay .

(2.50)

The total hydrostatic force is then the sum of the individual components given by
F = Fx2 + Fy2 + Fz2 .

(2.51)

Note: Analytical solutions can be found for the simple surfaces such as spheres, cylinders, and cones
only.

Example 2.6
Determine the hydrostatic force of water ( = 1000 kg m-3) acting upon the curved surface of the onefourth of the cylinder as shown in Fig. 2.10. The cylinder radius is r = 2 m and the length is L = 4 m.
The depth of centroid, the surface area, and the first moment of the area:

hCx =

r 2
= =1m
2 2

Ax = r L = 2 4 = 8 m 2

S x = hCx Ax = 1 8 = 8 m 3

The components of the hydrostatic force:

Fx = g S x = 9.81 10 3 8 = 78 480 N
1
1
V = r 2 L = 3.14 2 2 8 = 25.133 m 3
4
4

Fz = g V = 9.81 103 25.133 = 24 6552 N


The resultant hydrostatic force and its direction:

F = Fx2 + Fz2 = 78 480 2 + 246 552 2 = 25 8741 N

= arc tg

Fz
246 552
= arc tg
= 72 o 20 35
Fx
78 480

Fig. 2.10 Hydrostatic force on surface


of one-fourth of the cylinder
24

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

2.6 Liquid in Linearly Accelerated System


One example of the relative rest in hydrostatics is the problem of the hydrostatic balance
of liquid in a linearly accelerated container as shown in Fig. 2.11. The container is moving in
the direction of the x-axis in the field of gravity. No energy is added to the system. The
assumptions of constant density, mass and temperature, stated in equations (2.1), (2.2), and
(2.3), are valid.

Fig. 2.11 Hydrostatic balance of liquid in the linearly accelerated container


The balance of forces acting upon the fluid element of mass dm in the direction s is
given by the sum of the differential pressure force dFps, differential mass force dFms, and
differential force of inertia dFss for the linearly accelerated container (Fig. 2.11) as
dF ps + dFms + dFss = 0 .

(2.52)

The pressure force follows from the pressure distribution on the fluid element dm as
shown in Fig. 2.11. This force is given by the sum of the two pressure forces acting upon the
opposite sides of the fluid element in the s direction perpendicular to the equipotential
surface, where the pressure changes by dp on the distance ds. It is expressed by
dF ps = p dA ( p + dp ) dA = dp dA .

(2.53)

The component of the mass force acting due to gravity upon the fluid element of mass dm
in the direction s is given by
dz
(2.54)
dFms = dm g cos = g dA ds
= g dA dz ,
ds
where the negative sign emphasizes that the direction of gravitation is against the orientation
of the z-axis.
The component of the force of inertia due to acceleration a of the fluid element in the s
direction is
dx
(2.55)
dFss = dFs sin = dm a sin = a dA ds
= a dA dx ,
ds
where the negative sign emphasizes that the force of inertia acts against the orientation of the
x-axis.
By substituting the above-mentioned differential forces from equations (2.53), (2.54), and
(2.55) into the balance (2.52) we get
dp dA g dA dz a dA dx = 0 ,
25

(2.56)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

from where we can obtain the differential pressure equation in the form

dp = ( g dz + a dx ) .

(2.57)

The general integral of the differential pressure equation (2.57)

p = (g z + a x ) + c ,

(2.58)

expresses the pressure distribution within the liquid for the given case of hydrostatic
equilibrium.
From the boundary conditions (see Fig. 2.11)
L
and z = H
is
p = pe ,
x=
2
the constant of integration in equation (2.57) takes the form

(2.59)

(2.60)
c = pe + g H + a ,
2

then, by substituting the constant of integration (2.60) into equation (2.58) we obtain the
pressure distribution in the form

a L

p = p e + g (H z ) + x = p e + g h .
g 2

(2.61)

Of course, the basic equation (2.16) for the hydrostatic pressure is valid as well, thus h
represents the depth under the liquid level in the considered point.
The differential equation of the equipotential surface comes out from the differential
pressure equation (2.57) that under the boundary condition dp = 0 takes the form
a
dz = dx .
(2.62)
g
The general solution of this differential equation is
a
z = x+c,
g

(2.63)

where for the boundary conditions (2.59), the constant of integration in equation (2.63) takes
the form
a L
c=H+
,
(2.64)
g 2
then by substituting the constant of integration (2.64) into equation (2.63) we obtain the
equation of the equipotential (level) surface in the form

z=H+

a L

x ,
g 2

(2.65)

The mass conservation principle (2.2) is used for to solve practical problems of the
relative rest in hydrostatics, where the mass of the liquid within the container must be the
same for both the fluid at rest and the moving fluid. For the case of the linearly accelerated
rectangular container shown in Fig. 2.11, the mass of the liquid inside the container is

H1 + H 2
L b,
(2.66)
2
where b is the width of the container. Naturally, this equation is valid for the container high
enough only, so that no liquid runs over the upper edge by the level shift.
m=V = H Lb=

26

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 2.7
Determine the heights of level H1 and H2, the angle of the liquid surface , and the maximum
hydrostatic pressure pmax for a container in a linearly accelerated system of a = 2.5 m s-2. The length of
container is L = 1 m. The container at rest is filled to the depth of H = 0.5 m with liquid of density
= 800 kg m-3. The pressure over the liquid surface is barometric pa = 105 Pa.
The level heights:

H1 = H +

a L
2,5 1
= 0 .5 +
= 0.627 m
g 2
9.81 2

H2 = H

a L
2 .5 1
= 0 .5
= 0.373 m
g 2
9.81 2

The angle of the liquid surface:

= arctg

H1 H 2
0.627 0.373
= arctg
= 14 o17
L
1

The maximum hydrostatic pressure within the container:


L
1
p max = p a + ( g H + a ) = 10 5 + 800 (9.81 0.5 + 2.5 ) = 104920 Pa
2
2

p max = p a + g H 1 = 10 5 + 800 9.81 0.627 = 104920 Pa

2.7 Liquid in Rotating System


Another case of the relative rest in hydrostatics is rotation with the angular velocity
while the rotation axis is parallel to the direction in which the gravity force is acting as shown
in Fig. 2.12. No energy is added to the system. The assumptions of constant density, mass and
temperature, stated in equations (2.1), (2.2), and (2.3) are valid.

Obr. 2.12 Hydrostatic balance of angular acceleration system


The balance of forces acting upon the fluid element in the direction s is given by the
sum of differential pressure force dFps, differential mass force dFms, and differential force of
inertia dFss for the linearly accelerated container (Fig. 2.11) as
dF ps + dFms + dFss = 0 .
27

(2.67)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Likewise in the previous case, the pressure force follows from the pressure distribution.
This force is given by the sum of two pressure forces acting upon the opposite sides of the
fluid element in the s direction perpendicular to the equipotential surface, where the
pressure changes by dp on the distance ds. It is expressed by
dF ps = p dA ( p + dp ) dA = dp dA .

(2.68)

Similarly, the component of the mass force acting due to gravity on the fluid element of
mass dm in the direction s is found by
dz
(2.69)
dFms = dm g cos = g dA ds
= g dA dz .
ds
The component of the force of inertia acting upon the fluid element in the s direction
due to rotation with the angular velocity and the radius of rotation r is given by
dr
(2.70)
dFss = dFs sin = dm r 2 sin = r 2 dA ds
= r 2 dA dr .
ds
By substituting the above-mentioned differential forces from equations (2.68), (2.69), and
(2.70) into the balance (2.67) we get

dp dA g dA dz + r 2 dA dr = 0 ,

(2.71)

from where we can obtain the differential pressure equation in the form

dp = g dz + 2 r dr .

(2.72)

The pressure distribution is given by the integral

p = g

dz +

r 2 2

r
dr
=
p
+
g
H

z
+
e
0
1
2g

H1

(2.73)

where the integration limits for the point with the coordinates r and z are obvious from Fig.
2.12. The same follows from the basic equation (2.16) for the hydrostatic pressure as
p

p = dp = p e + g h

(2.74)

pe

where h represents the depth under the liquid level in the considered point.
The differential equation of the equipotential surface follows from (2.72) that under the
boundary condition dp = 0 takes the form

dz =

2
g

r dr .

(2.75)

The equipotential (level) surface is given by the integration of equation (2.75) as


z

dz =
H1

2
g

r dr

r 2 2
z = H1 +
,
2g

(2.76)

which describes the parabolic shape of the surface, were p = pe.


The application of the mass conservation principle (2.2) for the case of the rotating
circular container, as shown in Fig. 2.12, is given by

H1 + H 2
,
2
where the term on the right hand side is volume of the paraboloid.
m = V = R2 H = R2

28

(2.77)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

2.8 Archimedes Law, Buoyancy and Stability


If a body of volume V is immersed in a liquid of density , it is buoyed by the pressure
forces acting upon this immersed body. The resulting upward force Fb acting upon the body
immersed in a liquid can be determined as the
difference of the two pressure forces acting upon
its upper and lower surfaces (Fig. 2.13). That is

Fb = F2 F1 .

(2.80)

If the system is in equilibrium, we can express


these two pressure forces acting in the vertical
direction upon the given surfaces as the mass
forces by

F1 = g V1

and

F2 = g V2 ,

(2.81)

and then the resulting buoyancy force becomes:

Fb = g (V2 V1 ) = g V .

(2.82)

Fig. 2.13 The buoyancy

Equation (2.82) expresses the well-known


Archimedes law, which states that a body immersed in a liquid is buoyed up by a force equal
to the weight of the liquid displaced by the body.
Note: Archimedes law does not consider the surface tension acting upon the body.
If the weight force Fm of the immersed body is larger than the hydrostatic buoyancy force,
Fm > Fb, the body will sink. When these two forces are equal, Fm = Fb, the body floats. If the
weight of body, however, is less than the hydrostatic buoyancy force, Fm < Fb, the body floats
upwards. When the body floats partially immersed in the liquid, the weight of the body equals
the weight of the liquid displaced by its submerged portion.
Because the fluid is homogeneous, the point of application of the buoyancy force Fb is in
the center of buoyancy Cb, which is at the center of gravity of the volume of the displaced
liquid (Fig. 2.14). If the body is immersed partially, the buoyancy force is produced by its
submerged portion. The weight force acts in the center of gravity C of the floating body. The
floating body is in equilibrium when the line joining the center of buoyancy Cb and the center
of gravity C, called centerline, is normal to the water surface.
When the body is heeled, the center of buoyancy moves laterally. These two forces make
a couple of moment

M = Fb hM sin = Fm hM sin ,

(2.83)

where is the angle of the heel and hM sin is the horizontal distance between the center of
buoyancy and the center of gravity referred to as the righting arm.
If the moment M tends to reduce the heel angle , a floating body will be stable and the
heeled body turns back to the stable position when = 0.
When the moment M tends to increase the heel angle , a floating body will be unstable
and will tend to capsize.

29

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Fig. 2.14 Stability of floating body


The point where the direction of the weight force Fm intersects the centerline of the cross
section is referred to as the metacenter CM. The position of the metacenter can be determined
using the equation
Cb C M =

IC
,
V

(2.84)

where Cb is the center of buoyancy, IC is the second area moment of the waterplane about its
centerline, and V is the volume of displacement.
The distance hM = CC M between the center of gravity and the metacenter is the
metacentric height, which depends on the shape of the floating body.
The metacentric height is used to determine the stability of a ship. If the metacenter CM
lies above the center of gravity C of the ship, the ship is stable. Otherwise, if CM is situated
below, the ship is unstable and tends to capsize.
Example 2.8
Determine the buoyancy of the ball buoy, which is immersed in water of density = 103 kg m-3 up to
the half of its diameter d = 1 m. Determine the mass of the buoy.
The displaced volume of water:

V=

1 d 3 1 3.14 13
=
= 0.2618 m 3
2 6
2
6

Buoyancy:

Fb = g V = 9.81 10 3 0.2618 = 2568 N .


The buoy mass:

Fb = Fm = m g

m=

Fb 2568
=
= 261.8 kg
g
9.81

30

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

3 Flow of Ideal Fluid


Explorations of fluid motion are more difficult in comparison with the motion of rigid
bodies, because fluid particles change their mutual location easily. Fluid mechanics includes
kinematics and dynamics.
Kinematics of fluids is concerned with description of track, velocity and acceleration of
fluid particles in relation to time.
Dynamics of fluids studies causes of the fluid particles movement, which is implied by
operation of external forces on these fluid particles, eventual incoming or outgoing of energy
to the fluid particles from the environment.

3.1 Basic Phenomena and Concepts of Fluid Flow


With respect to the kinematic characteristics of fluid particles, the flow can be steady-state
and non-steady-state flow. The fluid state quantities in the given location of the flow field
r
vector of velocity v , pressure p, density , and temperature T may be time-dependent or
time-independent, respectively.
The motion of fluid particles is either
rotational or irrotational (potential). In the
case of a irrotational flow, the fluid particles
perform only translational straight-lined or
curvilinear motion during dilatation and
deformation. If the flow is rotational, the
fluid particles are furthermore rotating round
their own axes. Individual particles are
moving along the tracks called trajectories.
Streamlines are the curves created by certain
fluid particles, to which the velocity vectors
are tangential (see Fig. 3.1). At a non-steadystate flow, the trajectories and streamlines
are different curves. In the case of a steadyFig. 3.1 Trajectory and streamline
state flow, these lines coincide. Only one
streamline can come through the given point.
Streamlines cannot intersect or contract,
because, in this case, the fluid particle would
have its velocity vector pointing at two
different directions.
Density of streamlines presents their
number passing through the unit area,
perpendicular to these lines. Streamlines thus
give us an idea of the velocity distribution in
the fluid at a given time moment.
Stream fiber of fluid is bounded by the
stream tube (see Fig. 3.2), which is a set of
particular streamlines. The fluid flowing
through the stream tube cannot cross the
boundary surface of this tube, because the
velocity vectors of the fluid particles are
Fig. 3.2 Stream fiber
always tangential to the surface of a given
stream tube.
31

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

As regards the geometric classification of fluid flow, we can distinguish threedimensional (spatial), two-dimensional (plane) and one-dimensional flow.
r
In the case of the one-dimensional flow, we consider that the vector of velocity v and also
the pressure p, density and temperature T are function of the one variable only the
coordinate of displacement s, see Fig. 3.3. We deal with the average values of the flux state
variables in the given cross sectional area A. The mean value of the velocity vector v mean is
given from the condition of coincidence of the flow volume for the real and rectangular
distribution of velocity in the given normal cross sectional area of stream tube.

Fig. 3.3 One-dimensional fluid flow


The simplest case of flow is a one-dimensional steady-state flow of an ideal fluid in the
stream tube with a gradual change of the normal cross sectional area. The flow takes place in
the field of gravity. There is no adding of energy to the fluid from the environment. The
influence of the shear forces is neglected.

An ideal liquid is non-viscous and incompressible fluid, the density of which can be
considered as constant, that is
= constant.
(3.1)
3.2 Conservation of Mass Continuity Equation
During the fluid flow through a general stream tube (see Fig. 3.4), the mass of fluid
bounded by the control surface must remain constant while moving along that stream tube

dm = constant.

(3.2)

The time change of mass of the fluid element


along the stream tube must be zero, that is
d
(dm ) = 0 .
(3.3)
dt
The mass of the fluid element in the volume
limited by the control surface can be determined
by means of density as
dm A ds .

(3.4)

The size of the stream tube element ds can be


expressed through the fluid velocity v and the
differential of time dt, by relation
ds = v dt .

(3.5)
32

Fig. 3.4 Conservation of mass

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

For a steady-state fluid flow, the flux variables depend on the displacement of the mass
particle within the stream tube only. Generally, the fluid density , normal cross-sectional
area A, and the length of the displacement element ds are changing.
The law of conservation of mass given by equation (3.3) and taking into account equations
(3.4) and (3.5) can be written in the differential form

d
(dm ) = d ( A ds ) = d ( A v dt ) = d ( v A) = 0 .
dt
dt
dt

(3.6)

The continuity equation can be obtained by integrating equation (3.6)

Qm = v A = constant ,

(3.7)

where Qm is the mass flow rate of the fluid.


The integral form of the continuity equation represents the physical fact that the same
mass of the fluid flows in and out of the given control volume per the same time interval. In
other words, the change of the fluid mass flux is zero.
Applying the continuity equation (3.7) to two
different cross-sectional areas of the stream tube as
shown in Fig. 3.5 we obtain the relation valid for
the mass flow rate of the fluid as

Qm = 1 v1 A1 = 2 v 2 A2 .

(3.8)

For the flow of an incompressible fluid, the


density of which is constant (3.1), the continuity
equation (3.7) can be modified to the form valid for
liquids as
Q
(3.9)
Q = m = v A = constant ,

Fig. 3.5 Stream tube

where Q is volumetric flow rate of the fluid.


Between two cross-sections of the stream tube (Fig. 3.5), a relation similar to equation
(3.8) can be applied, to express the volumetric flow rate of the liquid, that is

Q=

Qm

= v1 A1 = v 2 A2 .

(3.10)

3.3 Conservation of Momentum


Applying Newton's second law of motion to the mass element dm bounded by the control
surface, as shown in Fig. 3.4, gives
d
(3.11)
( dm v s ) = dFs ,
dt
which expresses that the rate of change of momentum of a given mass dm in the direction of
the flow s equals to the component of the differential resultant external force dFs in the
same direction.
The rate of change of momentum of the mass element dm in the volume limited by the
control surface of the moving fluid in the direction s is given by

d
(dm v s ) = d ( A ds v ) = d ( A v dt v ) = v A dv ,
dt
dt
dt
33

(3.12)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

where the continuity equation (3.7) was used and the component of velocity v s in the given
direction s is approximately equal to the average fluid velocity v .
The component of differential external force in the s direction, dFs, influences the
mentioned mass element dm bounded by the control surface as a sum
(3.13)
dFs = dF ps + dFms .
The differential pressure force dFps acts upon the
control surface in the s direction of the flow (see
Fig. 3.6). This can be expressed by a sum of the
pressure forces affecting the fluid element as

dp

dF ps = pA ( p + dp )( A + dA) + p +
dA , (3.14)
2

where the first term represents the pressure force


acting upon area A, the second term is the pressure
force acting upon area A+dA, and the third term is
the component of the pressure force acting on the
remaining part of the control surface in the direction
Fig. 3.6 Pressure force
s. This third component of the pressure force is
determined as a product of the average pressure on
surface in question and the given area projection normal to the direction s.
By expending the terms in equation (3.14) and neglecting the terms containing the
product of two differentials the value of which is negligible with respect to the terms having
one differential only we obtain the final relation for the differential pressure force in the
form
(3.15)
dF ps = A dp .
The component of the differential mass force dFms in the estimated flow direction s is
given by projecting the gravity force onto the axis of the stream tube (see Fig. 3.7) and can be
expressed by
dz
(3.16)
dFms = dFm cos = g dm cos = g A ds
= g A dz .
ds
In the above equation, the cos was expressed as a
function of the angle for the triangle shown in
Fig. 3.7.
Note: The negative signs of the estimated forces indicate
that the direction of these forces is opposite to the
specified direction s.

Newton's second law (3.11) for the given mass


element, applied in the direction of s, when the
pressure and mass forces are considered, takes the
form
d
(3.17)
( dm v s ) = dF ps + dFms .
dt
By replacing the rate of change of momentum
Fig. 3.7 Determination of mass force
from equation (3.12) and using equations (3.15)
and (3.16) for the components of the pressure and mass forces, we obtain

v A dv = A dp g A dz ,
34

(3.18)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

which expresses the balance of forces acting upon the mass element of the fluid in the given
direction.
If we divide all terms in equation (3.18) by the element of mass dm, using relation (3.4),
we obtain the so-called Euler's differential equation in the form

dv 1 dp
dz
+
+g
= 0.
ds ds
ds

(3.19)

Multiplying the Euler equation (3.19) by the displacement ds, we obtain,

v dv +

dp

+ g dz = 0 .

(3.20)

The vector equation was reduced to the scalar equation by the last transformation. In this
new equation, each component is written in its scalar form. This means, all terms in equation
(3.20) are quantities that do not depend on the direction and orientation, in contrast to vector
values.
For the case of flow of an ideal liquid in the field of gravity, it is possible to easily
integrate the differential equation (3.20), considering the fluid density invariant (=constant).
We obtain the form called Bernoulli's equation

v2 p
+ + g z = constant .
(3.21)
2
Bernoulli's equation can be interpreted as a sum of specific kinetic energy v 2 2 , specific
pressure energy p and specific potential energy g z of the fluid, which remains constant
along the stream line and is equal to the overall specific energy of the fluid e. It means that
the mutual conversion of the mentioned energy types (kinetic, pressure and potential) is
possible only, while the internal energy u and temperature T of the fluid remain constant.

e=

3.4 Basic Equations of an Ideal Fluid Flow


For practical applications, two basic equations are used for the analysis of the onedimensional flow of incompressible ideal fluid (=constant) in the field of gravity within the
flow systems of variable cross-sectional area (see
Fig. 3.8):

The continuity equation

Q=

Qm

= v A = constant

(3.22)

Bernoulli's equation

e=

v2 p
+ + g z = constant
2

(3.23)

Note: For the accurate application of Bernoulli's equation


it is necessary to appropriately choose the coordinate
origin for the evaluation of the potential heads z (see Fig.
3.8).

35

Fig. 3.8 Two basic equations for


an incompressible ideal
fluid flow

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

3.5 Practical Applications of an Ideal Fluid Flow


3.5.1 Discharge through a small outlet
The vessel of the cross-sectional area A1 is
filled with the liquid with the density up to the
height h above the outlet with the cross-sectional
area A2, which is located in the side wall (see Fig.
3.9). The pressure above the level is denoted pN.
The fluid is discharging from the vessel to the
environment with the barometric pressure pa.
The task is to determine the fluid flow velocity
v 2 and the volumetric flow rate Q in the vessel
outlet.
The problem is solved by using the continuity
equation (3.22) and the Bernoulli equation (3.23)
in points of the stream line on the level and in the
vessel outlet as shown in Fig. 3.9. Thus

Q = v1 A1 = v 2 A2 ,

(3.24)

Fig. 3.9 Discharge of fluid from the


vessel through the small outlet

v12 p1
v2 p
+
+ g z1 = 2 + 2 + g z 2 . (3.25)
2

The boundary conditions determine pressures on the stream line (Fig. 3.9) as
p1 = p N and
p2 = pa .

(3.26)

From solving two equations (3.24) and (3.25) simultaneously, we can evaluate the two
variables v1 and v 2 . Substituting the velocity v1 from the continuity equation (3.24) into the
Bernoulli equation (3.25), we obtain
2

A2
v2 p
p
+ N + g z1 = 2 + a + g z 2 ,

2
A1
from where results the relation for the velocity in the outlet cross-section in the form

v 22
2

v2 =

2 g h +
g

,
2
A2
1
A1

(3.27)

(3.28)

where the following notation of the variables was used h = z1 z 2 and p = p N p a .


In the case of the outflow through a small outlet we can assume A2 << A1 and v1 << v 2 .
This allows us to simplify equation (3.28) to the form

p
.
v 2 = 2 g h +
g

(3.29)

The volumetric flow rate of the discharging fluid can be determined from the continuity
equation (3.24) as
Q = v 2 A2 .
(3.30)
36

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 3.1
Determine values for the outflow of a liquid with density = 103 kg m-3 from the round vessel
(Fig.3.9). The level height above the outlet is h = 1 m, the diameter of the outlet opening is d = 0.1 m,
the diameter of the vessel is D = 1 m, the pressure above the level is pN = 0.12 MPa and the barometric
pressure is pa = 105 Pa.
We now determine velocity, volumetric and mass flow rates of the discharged liquid:
2

A
For the given proportions: 2 = d = 0.1 = 10 4 << 1 , the simplified solution may be used.
A
D
1.0
1
The velocity of the discharged liquid:

p pa

v 2 = 2 g h + N
g

(1.2 1.0) 10 5
= 2 9,81 1 +
9.81 10 3

= 7.72 m s -1

The volumetric flow rate:


d2
3.14 0.12
Q = v2
= 7.72
= 6.06 10 2 m 3 s 1
4
4
The mass flow rate:
Q m = Q = 10 3 6.06 10 2 = 60.61 kg s 1

3.5.2 Discharge through a large opening


Fluid of density discharges from a wide basin with a large opening in the side wall of the
cross-sectional area A2 (Fig.
3.10). The pressure of the
surrounded air is barometric pa.
We determine the velocity
v 2 and the pressure p2
distribution at the cross-sectional
area A2 of the basin opening and
the flow rate Q of the
discharging liquid. The solution
to the problem follows from the
application
of
Bernoulli's
equation (3.23) at the marked
points on the liquid level and in
the opening (Fig. 3.10)

Fig. 3.10 Discharge through a large opening

v12 p1
v 22 p 2
+
+ g z1 =
+
+ g z2 .
2

(3.31)

If we consider the large basin, from where the liquid discharges, then the velocity at the
level is practically zero, hence, v1 = 0 . Furthermore, we can assume the same barometric
pressure over the whole opening from the boundary conditions (Fig. 3.10), hence, p1 = p a
and p 2 = p a . Upon substituting these conditions into equation (3.31) and rearranging the
terms, we obtain a simple formula

v 2 = 2 g ( z1 z 2 ) = 2 g h .

(3.32)

From equation (3.32), it follows that the discharge velocity v 2 depends on the depth h of
the estimated point under the level. The velocity v 2 in the cross-sectional area of the jet
37

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

increases with the increase of the depth h. This has to be taken into account when determining
the discharging flow rate.
The volumetric flow rate of the liquid in the elemental cross-sectional area of the opening
dA is given by the formulas
dQ = v 2 dA = 2 g h dA = 2 g h b dh ,
(3.33)
where the elemental cross-sectional area dA is expressed by means of the width b of the
opening (see Fig. 3.10). The overall volumetric flow rate of the liquid through a large opening
is given by the integration of (3.33) and taking into account the relation of the velocity v 2 on
the depth h (3.32). That is

Q = 2g

H2

h b dh ,

(3.34)

H1

where the width b of the cross-sectional area is a function of the depth under the level h,
which can be written as b = f (h ) .
In the case of the discharge through a rectangular opening of the width B, as shown in Fig.
3.10, b = B = constant. The volumetric flow rate is then given by the integration of the
general expression (3.34) in the form
H2
3
32

Q = 2 g h B dh =
2 g B H 2 H 12
(3.35)
3
H1

If the liquid overflows through the rectangular weir (see Fig. 3.11), it is possible to apply
equation (3.35) to determine the volumetric flow rate Q. For the boundary conditions H 1 = 0
and H 2 = H , it becomes
2
(3.36)
BH 2gH .
3
The integration of the general expression (3.34) for the
liquid discharging through the large opening can be easily
done for the triangular weir, shown in Fig. 3.12. The
dependency of the width on the depth is given by

Q=

b = f (h ) = 2 (H h ) tg

Fig. 3.11 Rectangular weir

(3.37)
2
and the integration limits are from h = 0 to h = H . That is
H2
8
Q = 2 g h b dh = tg H 2 2 g H . (3.38)
15 2
H1
This can be simplified for the angle = 90 as
8
Q = H2 2gH
15

(3.39)

Fig. 3.12 Triangular weir

Note: The shapes of the weirs considered here are frequently used for the measurement of flow rates in
laboratory conditions or on water streams.

Example 3.2
Determine the volumetric flow rate of a liquid of the density = 103 kg m-3 through the rectangular
opening as shown in Fig. 3.10. The heights of level above the upper edge of the opening is H1 = 0.2 m
and above the lower edge is H2 = 0.4, width is B = 0.6 m.

Q=

3
3
3
2
3

2
2 g B H 22 H 12 =
2 9.81 0.6 0.4 2 0.2 2 = 289.8 10 3 m 3s 1
3

38

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

3.5.3 Discharge through a submerged opening


A liquid of the density flows from a wide basin to another one, as shown in Fig. 3.13.
The level difference between both basins is H. The pressure of the surrounding air is
barometric, pa. We determine the velocity v 2 and the pressure p2 distribution at the crosssectional area A2 of the submerged opening and
the flow rate Q of the passing liquid.
The solution to the problem follows from the
application of Bernoulli's equation (3.23) at the
marked points on the level of the liquid and in
the opening (Fig. 3.13)

v12 p1
v 22 p 2
+
+ g z1 =
+
+ g z2 .
2

(3.40)

For the point number one on the level, we


can assume that the velocity is negligibly small,
that is, v1 = 0 , and from the boundary conditions
follows that p1 = pb . For the second point,
located in the cross-section of the submerged
opening, we can assume the hydrostatic pressure
given by the depth under the level of the lower
basin
p 2 = p a + g (h H ) .
(3.41)

Fig. 3.13 Discharge of liquid through


the submerged opening

Upon substituting these conditions into Bernoulli's equation (3.40) together with the
relation between the potential heads and the depth of the given points

h = z1 z 2 ,

(3.42)

we obtain the formula for the fluid flow velocity in the opening between the basins in the
form
v2 = 2 g H .
(3.43)
As follow from equation (3.43), the velocity v 2 of the liquid flow is constant in the whole
cross-sectional area A2. The velocity v 2 depends only on the difference of the levels H of the
mentioned basins and is independent of the depth h.
The volumetric flow rate of the liquid through a submerged opening between the given
basins can be expressed using the continuity equation (3.24) as

Q = v 2 A2 .

(3.44)

Example 3.3
Water of density = 103 kg m-3 is discharging through the submerged opening with the cross-sectional
area A2 = 2 m2 (see Fig. 3.13). The difference of the fluids levels is H = 1 m and the barometric
pressure is pa = 105 Pa.
Determine velocity of the liquid flow in the cross-sectional area and the volumetric flow rate.
The flow velocity:

v 2 = 2 g H = 2 9.81 1.0 = 4.43 m s 1


The volumetric flow rate:
Q = v 2 A2 = 4.43 2 = 8.86 m 3 s 1
39

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

3.5.4 Venturi tube


Fluid of a specific mass flows through a Venturi tube (see Fig. 3.14) characterized by
the varying cross-sectional areas A1 and A2. The attached differential U-tube manometer
indicates the height hm of the liquid column of the density m as a consequence of the change
of the cross-sectional area. This simple apparatus enables us to determine the flow velocity
and the volumetric flow rate in a pipeline.
The solution for the velocity v2 and the
volumetric flow rate Q as a function of the
pressure difference p follows from the continuity
equation (3.24) and Bernoulli's equation (3.23) at
the given points of the stream tube. Hence,

Q = v1 A1 = v 2 A2

(3.45)

v12 p1 v 22 p 2
+
=
+
,
2

(3.46)

where the modified form of Bernoulli's equation


accounts for the fact that the both potential heads
Fig. 3.14 Venturi tube
are equal, z1 = z2. Solving equations (3.45) and
(3.46) simultaneously, we can determine the two unknown variables velocities of flow v1
and v2 . Upon substituting v1 from the continuity equation (3.45) into the Bernoulli equation
(3.46), we have

v 22
2

A2

A1

p
v2 p
+ 1 = 2 + 2 ,

(3.47)

from which the velocity v2 in the cross-section A2 as a function of pressure difference


p = p1 p 2 or for the manometer column height p = g ( m ) hm can be determined in
the form

p p2
2 1
2g m
hm

v2 =
=
.
(3.48)
2
2
A2
A2
1
1
A
1
A1
The volumetric flow rate measured by the Venturi tube is then determined from the
continuity equation as

2
Q = v 2 A2 = A2

p1 p 2

A
1 2
A1

2
2

p1 p 2

d
1 2
d1

2
2

2g

m
hm

d
1 2
d1

(3.49)

where the circular cross-sectional area of the Venturi tube was estimated as:

A1 =

d 12
4

, A2 =

d 22
4

A2 d 2
=
A1 d 1

and

A
, 2

A1

d
= 2

d1

(3.50)

Note: The classical Venturi tubes, together with orifices and nozzles are frequently used to determine
flow rates through pipes. For exact geometry and specifications for these apparatuses, see [25].
40

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 3.4
Determine the volumetric and mass flow rates of the fluid of the density = 103 kg m-3 measured by
the Venturi tube (Fig. 3.14) with diameters d1 = 0.1 m and d2 = 0.05 m. The manometer mercury
column height is hm = 0.3 m and the density of mercury is m = 13590 kg m-3.
The volumetric flow rate:

Q=

d 22
4

2g

m
hm

d
1 2
d1

3.14 0.052
=
4

13590 1000
2 9.81
0.3
1000
= 5.52 10 2 m 3s -1
4
0.05
1

0.1

The mass flow rate:


Q m = Q = 10 3 5.52 10 2 = 55 .2 kg s -1

3.5.5 Pitot and Prandtl tubes


The Pitot probe is frequently used to measure local flow velocity. A probe tube is
perpendicular to the flow direction (Fig. 3.15). The
moving fluid inside the tube is brought to rest
(stagnates), the fluid velocity is zero and the pressure
inside reaches the stagnation pressure p0. The
pressure in a fluid stream of the density is called
the static pressure p. The difference between the
stagnation pressure and the static pressure p is
measured by manometer.
One can determine the fluid velocity v as a
function of the dynamic pressure pd or of the
manometer height hm from Bernoulli's equation
(3.23) at the points on the tube axis (Fig. 3.15).

Fig. 3.15 Pitot probe

At a stagnation point the fluid velocity is zero v0 = 0 and the total kinetic energy has been
converted into the pressure energy. This can be expressed by Bernoulli's equation (3.23) as

e=

v 2 p p0
+ =
,
2

(3.51)

from where the stagnation pressure p0 can be written in the form

p0 = p +

1
v 2 = p + pd ,
2

(3.52)

where the dynamic pressure pd is equal to the difference between the stagnation pressure p0
and the static pressure p.
Equation (3.52) can also be used for to determine the fluid velocity. This yields

v= 2

p0 p

= 2

41

pd

(3.53)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Eventually, it is possible to use the measured difference between the stagnation pressure
and the static pressure by the U-tube manometer p = p 0 p = g ( m ) hm to determine
the flow velocity v as

v= 2

= 2g

m
hm .

(3.54)

The Prandtl probe (also called Pitotstatic tube) is a pressure-sensitive instrument


that is most often used to determine the free
stream velocity or aircraft speed. The
schematic drawing of a Prandtl probe is
shown in Fig. 3.16. Equations (3.53) and
(3.54) to determine the flow velocity v by
the Prandtl probe are the same as for the Pitot
probe.
Note: The Pitot tube is called the probe for
withdrawing of the stagnation pressure. The static
tube has port for the static pressure only.

Fig. 3.16 Prandtl probe

The static pressure is obtained through a static


port. The schematic diagram of the static port is
shown in Fig. 3.17. The static port is a flushmounted hole, perpendicular to the direction of the
flow. From the flow boundary condition, the
pressure in the cavity is equal to the pressure within
the flow.
The probe for reading the stagnation pressure
represents a center tube, which is always pointed in
the direction of the flow (see Fig. 3.18). The
stagnation pressure is the ram fluid pressure at the
stagnation point, where the flow velocity is zero.
The measured stagnation pressure is sometimes
called the Pitot pressure.

Fig. 3.17 Static port

Fig. 3.18 Stagnation pressure probe

Example 3.5
Determine the velocity of the fluid with the density = 1000 kg m-3 measured by the Pitot probe,
shown in Fig. 3.15. The U-tube gauge mercury column height is hm = 0.7 m and the density of
mercury is m = 13590 kg m-3.
The flow velocity:

v= 2g

m
13590 1000
hm = 2 9.81
0.7 = 13.15 m s 1
1000

Example 3.6
Determine the flow velocity of air with the density = 1.25 kg m-3 measured by the Pitot-static system
(see Fig. 3.16). The pressure difference measured by manometer is 5880 Pa.
The airflow velocity:

v= 2

= 2

5880
= 97 m s 1
1.25
42

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

4 Flow Similarity Criteria


The motion of fluid is called flow. In 1883, during his experiments with the flow within a
circular pipe, Osborne Reynolds observed some phenomena resulting from the action of the
friction forces within the viscous fluid flow. If a fluid flows through a clear circular pipe with
low velocity (see Fig. 4.1), the fluid particles move following equidistant tracks in the axial
direction, thus the thin colored stream maintains its diameter and axial direction. Reynolds
called such a flow regime laminar. As the flow velocity increases, the fluid particles change
their relative positions in various directions.
Hence, local eddies are formed and the colored
stream dispersed gradually (see Fig. 4.1). This
flow regime is called turbulent.
In addition, Reynolds found out that the
transition between the laminar and turbulent
flows occurs when the dimensionless Reynolds
number is
vd
(4.1)
Re =
2300 ,
v
Fig. 4.1 Reynolds experiment
where v is the mean flow velocity, d is the pipe
diameter and is the kinematic viscosity.
Reynolds experiments have opened the era of the investigation of turbulence and the
development of the theory of physical similarity in fluid mechanics.

4.1 Laminar Flow


In the laminar flow regime, the fluid moves as if in laminas and there is no exchange of
fluid particles between these laminas within the flow (see Fig. 4.2).
The velocity distribution within the laminar
flow is parabolic. Therefore, the velocity profile
is in the shape of a paraboloid of revolution for
a circular pipe.
The mean velocity (cross-sectional velocity)
of the laminar flow, from the condition of the
flow rate constancy, is

v=

v max
.
2

(4.2)

Fig. 4.2 Laminar flow

4.2 Turbulent Flow


If fluid particles have, besides the translation velocity, a transverse (fluctuation/turbulent)
velocity component (see Fig. 4.3), such a
flow regime is called turbulent. This regime
is characterized by chaotic, stochastic
changes.
Turbulent velocity profiles are much
flatter compared to the laminar profiles (see
Fig. 4.3) and the mean velocity is about

v = (0.82 0.87 ) v max .

Fig. 4.3 Turbulent flow

(4.3)

43

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

4.3 Theory of Similarity


Methods of the similarity theory are used in many applied sciences, mainly for the
processing of experimental data and their generalizations within the theoretical background.
The main goal of the similarity theory is to ensure a fast and accurate processing of
experimental results and matching the results between the model and the prototype.
In fluid mechanics, we deal with the similarity of mechanical phenomena, and, therefore,
we speak here about mechanical similarity. Mechanical similarity includes geometric,
kinematic, and dynamic similarity.

4.3.1 Geometric similarity


Geometric similarity between the prototype and the model is achieved if the basic
conversion module of lengths (body sizes) corresponds to the ratio

l 2 b2
=
= ..... constant.
l1 b1

(4.4)

Thus, each corresponding dimension of the prototype and the model are proportional (see
Fig. 4.4). From this requirement it follows that the corresponding angles of similar objects are
to be equal.

Fig. 4.4 Formulation of mechanical similarity


In practice, if we want to satisfy the requirements of geometric similarity, it means that the
model is scaled down in a given ratio, including the reduction of the surface roughness, which
poses substantial demands on the surface finish of models.

4.3.2 Kinematic similarity


Kinematic similarity of the fluid flow around the prototype and the model is achieved, if
the basic conversion module of time is
t
= 2 = ..... constant,
(4.5)
t1
in other words, the characteristic times of the processes for both objects are proportional.
To describe the flow around objects, the velocity is used, that is the ratio of the
characteristic length l and time t scales:
l
(4.6)
v= .
t
The basic criterion for the similarity in time between the prototype and the model for
kinematic similarity follows from the relationships (4.4) and (4.5) where the expression for
time is substituted from equation (4.6). This criterion is the dimensionless Strouhal number,
Sh, that describes oscillating flow mechanisms, and is defined as
44

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

l2
t 2 v2
l2
l
l
=

= 1
Sh =
.
(4.7)
l
t1
v 2 t 2 v1 t1
vt
1
v1
The kinematics of the fluid flows around the prototype and the model are similar if the
Strouhal numbers are equal. The values of the Strouhal number are obtained from the
corresponding velocities, times and lengths.
In the case of periodic processes (e.g. flow around the propeller), the formulas that relate
time t, frequency f, angular frequency , and revolutions per second RPS are to be used:

t=

1 2
1
=
=
.
f
RPS

(4.8)

4.3.3 Dynamic similarity


To achieve dynamic similarity, the corresponding forces (see Fig. 4.4) acting upon the
prototype and the model are to be proportional. Hence, the ratios of these forces external
forces, inertial forces, friction forces, mass forces, pressure forces are to be equal
F1 Fs 2 F fr 2 Fm 2 Fp 2
=
=
=
=
= ..... constant.
(4.9)
F2 Fs1 F fr1 Fm1 Fp1
Then, the dynamic effects the fluid flow acting upon the prototype and the model become
similar.
The Newton number, Ne, and the force coefficient, cF, are used as the basic criteria of
dynamic similarity. They follow from the ratio of the resultant external force F to the force of
inertia Fs as
F
(4.10)
= constant.
Fs
The external force F is measured on the model and the results are applied on the prototype.
The force of inertia Fs can be determined by relations
v
v
Fs = m a l 3 l 3 v 2 l 2 .
(4.11)
l
t
v
This expression is based on the fact that mass m is proportional to the product of the density
and the cube of the characteristic length l. The acceleration a is proportional to the ratio of
the velocity v and the characteristic time t. The characteristic time t, in turn, is proportional to
the ratio of the characteristic length l and velocity v . Thus, all these quantities are
conventional and they are determined by means of the characteristic properties of the given
physical effect.
From equation (4.11), it is evident that we do not operate with the value of particular
force of inertia, but with the value proportional to the conventional force of inertia, which is
determined through the derived characteristic quantities for the physical process in question.
Upon substituting (4.11) into the definition formula (4.10), we obtain the dimensionless
Newton number in the form
F
F
(4.12)
=
= Ne .
Fs v 2 l 2
45

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

In practice, the force effects are often determined though the criteria of dynamic
similarity in terms of the force coefficient cF and the moment coefficient cM, defined by the
formulas
F
M
and
,
(4.13)
cF =
cM =
1
1
2
2
v A
v Al
2
2
where 1 2 v 2 is the dynamic pressure, A is the characteristic area and l is the characteristic
length.

Another significant criterion of dynamic similarity is the Reynolds number, Re, which is
given by the ratio of the force of inertia Fs and the friction force Ffr
Fs
= constant.
(4.14)
F fr
If we express the friction force Ffr by the shear stress (1.14), using Newtons law of
viscosity (1.15), we obtain
dv
v
F fr =
A l2 vl ,
(4.15)
dy
l
where the friction force Ffr is proportional to the product of the dynamic viscosity , velocity
v , and the characteristic length l . By substituting the inertia and friction forces (4.11) and
(4.15) into the basic forces ratio (4.14), we obtain the dimensionless Reynolds number Re in
the form
Fs
v2 l 2 vl v l

=
= Re ,
(4.16)
F fr
vl

v
where relation (1.16) between dynamic viscosity and kinematic viscosity has been used.
Note: The Reynolds number enables us to characterize fluid flows in channels and flows around solid
bodies. The Reynolds number is conveniently used as a basic parameter in processing experimental
results and their further generalization. As an example, we can show in Fig. 4.5 the dependence of the
sphere drag coefficient cd on the Reynolds number measured in a wind tunnel. The characteristic
length is the sphere diameter d and the similarity criteria are as follows
vd
Fd
and
.
Re =
cd =
2
v

d
1
v2
2
4

Fig. 4.5 Drag coefficient for a smooth sphere versus Reynolds number
46

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 4.1
Determine the resistance force Fd of the sphere of diameter d = 0.01 m moving through a fluid with
the density = 10 3 kg m 3 and the kinematic viscosity v = 10 6 m 2 s 1 . The free stream velocity is
v = 2 m s 1 .
The Reynolds number:

v d 2 0.01
=
= 2 10 4
6
v
10
The value of the drag coefficient of the sphere from the chart in Fig. 4.5:
Re =

c d = 0.45
The resistance force:

Fd = c d

d2
1
1
3.14 0.012
v2
= 0.45 10 3 2 2
= 7.069 10 2 N
2
4
2
4

The Froude number Fr is the criterion of the dynamic similarity, which is used to
indicate the influence of gravity on the fluid motion. This follows from the ratio of the inertia
force Fs and the mass force Fm
Fs
= constant .
(4.17)
Fm
The mass force is the force due to gravity, which can be expressed by the characteristic
quantities as
(4.18)
Fm = m g g l 3 .
Substituting the inertia and mass forces (4.13) and (4.18) into the basic ratio of forces
(4.17), we obtain the dimensionless Froude number Fr in the form
Fs
v2 l 2 v2

= Fr .
Fm
gl
g l3

(4.19)

The dimensionless Euler number Eu is proportional to the ratio of the pressure forces
and the inertia forces and is defined as
Fp
p

= Eu .
(4.20)
Fs v 2
The relative pressures throughout a flow field describes the pressure coefficient which is
a very useful parameter for studying the flow of incompressible fluids

cp =

p p
,
1
2
v
2

(4.21)

and is defined by the difference between the local pressure p and the free-stream pressure p
related to the dynamic pressure.
The dimensionless Mach number Ma is defined by the ratio of the flow velocity v , or
the speed of an object moving through a fluid, to the speed of sound, a, in that fluid. That is
Ma =

47

v
.
a

(4.22)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

4.3.4 Overall and partial similarity


If we want to solve flow problems by applying the mechanical similarity, firstly, we have
to fulfill the geometric similarity conditions for the prototype and the model. The model has
to be manufactured in scale, including the surface roughness.
The kinematic similarity of a given case of fluid flow around the prototype and the model
is provided by the equality of the Strouhal numbers

Sh2 = Sh1 .

(4.23)

The fluid flow around solid bodies is characterized by the periodic separation of vortices.
The similarity condition of the same frequency of the vortices' separation on the prototype and
the model is satisfied by (4.23).
In order to achieve the full dynamic similarity of mechanical processes for the flow
around the prototype and the model, it is necessary to have

c F 2 = cF1 ,

Fr2 = Fr1 ,

Re2 = Re1 ,

Eu 2 = Eu1 ,

etc.

(4.24)

This means that the main criterion of the dynamic similarity the coefficient of the
observed external force, as well as the criteria for dynamic similarity: Froude number,
Reynolds number, Euler number and others are to be equal for the prototype and the model.
The fulfillment of all the criteria of the dynamic similarity leads to the identity of the
prototype and the model.
The partial flow similarity requires the fulfillment of the dynamic similarity of the main
forces acting upon the body under consideration. Usually, the satisfaction of the main
dynamic criterion the force coefficient and additional one or two other similarity criteria are
required.
In order to maintain the similarity of inertia and gravity forces, the equality of the two
criteria is required
c F 2 = c F 1 and Fr2 = Fr1 ,
(4.25)
which is usually given by an empirical dependence

c F = f (Fr ) .

(4.26)

As an example, the dependence of the resistance coefficient of a floating body on the


Froude number can be mentioned. This is because the creation of waves, which play the main
role in the resistance against the motion of the body on the water level, is mostly influenced
by the gravity forces.
To maintain the similarity of inertia and friction forces, we have to achieve the equality
of the following criteria
c F 2 = c F 1 and Re2 = Re1 ,
(4.27)
which is usually given by an empirical dependence

c F = f (Re) .

(4.28)

Another example is how the drag coefficient of a sphere depends on the Reynolds
number (see Fig. 4.5), since the friction force of fluid flow around the body is the main
resistance force.
For the concurrent fulfillment of the similarity of inertia, gravity and friction forces, it is
necessary to get the equalities of all the three criteria

cF 2 = cF1 ,

Fr2 = Fr1 ,
48

and Re2 = Re1 .

(4.29)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

This is usually given by an empirical relation

c F = f (Fr , Re) .

(4.30)

Yet another example is how the resistance coefficient of the fast floating body on the
water surface depends on the Froude number and the Reynolds number.
Note: The characteristic length (dimension) l is a conventional parameter representing the geometric
size of the flow through or around a solid body. If the fluid happens in noncircular pipes or channels,
the so called hydraulic diameter is commonly used, which is defined as the diameter of an equivalent
circular cross section that substitutes the given cross section of an arbitrary shape as
4A
,
(4.31)
lD=

where A is the flow cross-sectional area of the channel and o is the wetted perimeter.

Example 4.2
Determine the drag force Fd of the cylinder of the diameter d = 0.125 m moving through a fluid with
the density = 1.29 kg m 3 and the kinematic viscosity v = 13.4 10 6 m 2 s 1 . The free stream
velocity is v = 1 m s 1 . Determine the frequency f of the vortices generated in the wake of the cylinder
on one side per second.
The Reynolds number:
v d 1.0 0.125
Re =
=
= 9328
v
13 .4 10 6
The values of the drag coefficient of the cylinder and the Strouhal number from the chart in Fig. 4.6:

c d = 1.0 and Sh = 0.2


The drag force per unit length of the cylinder:

1
1
v 2 d = 1.0 1.29 12 0.125 = 0.0806 N m 1
2
2
The frequency of vortices generated in the wake:
v
1
f = Sh = 0.2
= 1.63 s 1
d
0.125
Fd = c d

Fig. 4.6 Drag coefficient and Strouhal number for a smooth cylinder
versus Reynolds number
49

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

5 Fluid Flow Losses


The description of the viscous fluid flow, in which friction forces act, is based on the
following basic assumptions:
The flow is one-dimensional and steady-state
The change in the flow cross-section occurs
Flow takes place in the gravitational field and there is no added energy by the
external source (mechanical, thermal, etc.)
The effect of internal friction in the fluid is expressed by the coefficient of friction
force, the value of which is determined from the similarity theory
The liquid is considered as an incompressible fluid of constant density

= constant .

(5.1)

5.1 Basic Laws of Flow with Friction Losses


The set of equations describing the fluid flow with friction losses follows from the three
basic laws of conservation, namely mass, momentum and energy.

5.1.1 Conservation of mass: continuity equation


In the case of incompressible fluids (constant density) the continuity equation assumes the
form
Q
(5.2)
Q = m = v A = constant,

where Q is the volumetric flow rate and Qm denotes the mass flow rate of the liquid.
The derivation of this equation is based on the same assumptions as were used for the
ideal liquid and was presented in detail in Chap. 3.2.

5.1.2 Conservation of momentum


Newton's second law, applied to the mass element dm bounded by the control surface in
the direction of flow s (see Fig. 5.1), has the form
d
(dm v s ) = dFs = dFps + dFms + dF frs .
dt

(5.3)

The change of momentum with time is equal


to the component of the resultant external force
in the direction s. This component of the
resultant external force is given by the sum of
the components of the differential forces, namely
the pressure force dF ps , the mass (gravitational)
force dFms , and the friction force dF frs .
The change of momentum with time of the
mass element, the component of pressure force
and the mass force acting in the direction s
were described in detail in Chap. 3.3, where
equations (3.12), (3.15), and (3.16) provide the
following relationships:
50

Fig. 5.1 Formulation of the law


of momentum conservation

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

d
(dm v s ) = v A dv ,
dt

dF ps = A dp ,

dFms = g A dz .

(5.4)

The differential friction force dF frs acting upon the mass element dm within the control
surface in the direction of flow s can be expressed by using the force coefficient from the
theory of similarity (4.13) in the form
1
(5.5)
dF frs = c fr v 2 dA fr ,
2
where c fr is the coefficient of friction force and
1
v 2 is the dynamic pressure.
2
The differential friction area dA fr can be
expressed by means of the perimeter o and the length
ds of the stream tube as
(5.6)
dA fr = o ds .
It is obvious from Fig. 5.2, that the friction area
is equivalent to the so-called wetted area, which is
the area where the stream tube walls are in contact
with the fluid.

Fig. 5.2 Definition of


the friction force

The required motion equation can be found by substituting equations (5.4), (5.5), and (5.6)
into the basic relationship (5.3) as
1
(5.7)
v A dv = A dp g A dz c fr v 2 o ds .
2
If we divide all terms in equation (5.7) by dm expressed by (3.4), we will obtain Euler's
differential equation for the frictional fluid flow in the form

dv
1 dp
dz
v2 o
=
g c fr
.
ds
ds
ds
2 A

If we use the concept of the hydraulic diameter (4.31)


A
D=4
o
and if we define the friction factor as
= 4 c fr ,

(5.8)

(5.9)

(5.10)

it becomes possible to rewrite Eulers equation (5.8) in the final form

dv 1 dp
dz
v2 1
+
+ g +
= 0.
ds ds
ds
2 D

(5.11)

The differential form of Bernoulli's equation, which expresses the work of all the
differential forces acting along the streamline is then obtained by multiplying the equation
(5.11) by the displacement ds . That is

v dv +

dp

+ g dz +

51

v 2 ds
= 0.
2 D

(5.12)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Integration of equation (5.12), assuming the fluid density to be constant, =constant (5.1),
leads to the modified Bernoulli's equation in the extended form

e=

v2 p
+ + g z + e z = constant ,
2

(5.13)

where the specific energy loss due to friction is defined by the integral
s

ez =
0

v 2 ds
.
2 D

(5.14)

From the physical point of view, it is possible to interpret the Bernoulli equation (5.13) for
the viscous fluid flow in the gravity field as the sum of the specific kinetic energy v 2 2 , the
specific pressure energy p , the specific potential energy g z , and the specific energy loss
ez of the fluid, which remains constant along the stream tube and is equal to the overall
specific energy of the fluid e.

5.1.3 Energy conservation


In the case of a frictional flow, the work of the friction forces dissipates into heat, which
increases the internal energy of the fluid. The increase of the internal energy leads to the fluid
temperature rise by dT as

de z = du = c dT ,

(5.15)

where c is the specific heat of the fluid.


Then the temperature change dT of the fluid, flowing along the stream tube, due to friction
is given by
dT =

de z v 2 ds
=
.
c
c 2 D

(5.16)

Usually, because values of the specific heat of fluids are large, the temperature increase is
marginal and can be neglected in most cases. However, for closed hydraulic systems, it is
necessary to consider the temperature rise, because the temperature of the hydraulic fluid
circulating through the system can be substantially increased.
Note: Friction in fluids is a typical irreversible process, which is, according to the second law of
thermodynamics, accompanied by an increase of the specific entropy. This problem will be discussed
in more detail in the course of thermodynamics.

5.2 Specific Energy Loss Due to Friction


Friction forces become significant in fluid flows within the pipelines or hydraulic
systems. Friction occurs in straight pipes, all fittings, valves, and hydraulic devices. The
calculation of the total losses due to friction is the sum of friction losses in straight pipes and
ducts of constant cross-section and local losses, which result from the flow in the other parts
or elements of the piping systems.

5.2.1 Straight pipe friction loss


Straight circular tubes of constant diameter are basic parts of pipelines and hydraulic
systems.

52

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Specific energy loss due to friction has been defined by (5.14), which can be applied for
the straight tube of constant cross-section (as shown in Fig. 5.3) within the given limits, from
the tube beginning s1 to its end s2, as
S2

ez =
S1

v 2 ds
.
2 D

(5.17)

For a circular pipe of constant diameter


d, it follows from the equation (4.31) that the
hydraulic diameter D is equal to the diameter
of the pipe d, i.e. d = D . From the
continuity equation (5.2), it follows that the
fluid velocity is constant v = v1 = v 2
throughout the pipe. Therefore, it can be
Fig. 5.3 Circular pipe of constant diameter
concluded that the diameter d of the pipe and
the velocity v are independent of the variable s.
Upon integrating (5.17) for constant friction factor , flow velocity v , and pipe diameter
d, the resulting specific energy loss due to friction in the straight circular pipe of constant
cross-section is
ez =

s 2 s1 v 2
l v2
=
.
d
2
d 2

(5.18)

The friction factor is determined from


measured values using similarity as an empirical
function of the Reynolds number
vd
k

,
(5.19)
, Re =
v
d

where the ratio k / d is called the relative


roughness of the pipe and k is the hydraulic
roughness, which represents an average height of
the irregularities on the pipe surfaces.

= f Re,

Fig. 5.4 Pipe roughness

The definition of these quantities is shown in


Fig. 5.4. Typical values of the hydraulic wall
roughness of the common pipe materials are given
in Tab. 5.1.
The relationship among the friction factor,
Reynolds number and relative roughness, which is
plotted on a chart is called Moody (Nikuradze)
diagram. The scheme of how to determine the
friction factor from this diagram is shown in Fig.
5.5. The value of is also independent of the
coordinate s.

Fig. 5.5 Schematic of Moody diagram

5.2.2 The Moody diagram


As already stated, the Moody diagram expresses the dependence of the friction factor
on the Reynolds number for various relative roughness of the straight circular pipe of constant
diameter.
53

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The characteristic dimension for the Reynolds number, defined by equation (4.16), is the
pipe diameter d. Then the Reynolds number takes the form
vd vd
Re =
=
.
(5.20)

v
The relative pipe roughness k / d is a parameter that characterizes the geometric
similarity. Sometimes the geometric similarity is referred to the relative pipe smoothness
defined by the ratio r / k , where r is the pipe radius (see Fig. 5.4).
The Moody diagram, which is shown in Fig. 5.6, is based on experiments carried out
within pipes with artificial roughness produced by uniform sand grains stuck to the walls. The
coordinates of the friction factor and the Reynolds number Re are plotted in the logarithmic
scale.

Fig. 5.6 The Moody diagram for a pipe with artificial roughness
From the Moody chart in Fig. 5.6, it is possible to observe according to the value of the
Reynolds number three regions of flow: laminar, transitional, and turbulent.

For Re < 2300 , there is the laminar flow, where the friction factor changes linearly. Its
value is given only by the Reynolds number and it is not affected by the value of the
relative roughness of the pipe walls k / d .

For the values of the Reynolds number of 2300 < Re < ca. 4000 the transition between
laminar and turbulent flow occurs.

For Re > ca. 4000 , the turbulent flow is already fully developed and a substantial
influence of the relative roughness k / d is observed.

In the turbulent flow region, the line of k / d = 0 represents the losses in the hydraulically
smooth regime, where the value of the friction factor depends only on the Reynolds number.
From this line, for a hydraulically smooth pipe, the curves for various relative roughnesses
k / d are successively branched out, where the friction factors reach constant values for the
higher Reynolds numbers.
54

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The explanation of the physical principle of this phenomenon can be made on the basis of
the so-called laminar sub-layer. Fluid particles are stuck on the pipe surface and therefore do
not move. A near-wall laminar sub-layer is formed. The thickness of this sub-layer decreases
with an increase of the Reynolds number. Therefore, for the lower Reynolds numbers, the
laminar sub-layer covers the surface irregularities and the turbulent flow is not affected by the
roughness. The friction factor is the same as for a hydraulically smooth pipe in this case.
As the Reynolds number increases, the thickness of the laminar sub-layer decreases and
can become smaller than the height of the surface irregularities. In such a case the laminar
sub-layer is destroyed and the turbulent motion is affected by the surface roughness.
To use computers, it is necessary to express the dependence of the friction factor on the
Reynolds number and the relative roughness (5.18) analytically.
For the laminar flow, when the Reynolds number is Re < 2300 , the Hagen-Poiseuille
equation is frequently used in the form
64
.
(5.21)
=
Re
For the turbulent flow region, when the Reynolds number is Re > 2300 , the Colebrook
equation can be used
2.51
k
= 2 log
+ 0.27 .
d

Re

(5.22)

The Moody diagram, expressed as a graphical solution of the Colebrook equation, is shown in
Fig. 5.7. The dashed line in the diagram represents the borderline, above which the region
where the friction factor can be considered constant for the given relative roughness is
located.

Fig. 5.7 Dependence of the friction factor on the Reynolds number Re


and the relative roughness = k / d
55

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Tab. 5.1 Hydraulic roughness for various materials


Hydraulic roughness k [mm] (average height of the irregularities)
Pipe material
Seamless steel tubes
Drawn steel pipes
Welded steel tubes
Galvanized steel pipes
Cast iron pipes
Drawn brass, copper, or aluminum pipes
Glass tubes, plastic pipes
Rubber and plastic hoses
Concrete pipes

Initial state

Corroded state

0.04 0.1
0.03 0.12
0.05 0.1
0.15 0.5
0.1 0.3
0.0015 0.003
0.0015 0.01
0.01 0.03
0.3 6.0

0.1 0.9
0.12 0.9
0.15 0.9
0.5 3.5
1 4.5
0.003 0.1

Example 5.1
Determine the pressure loss due to friction in the
water pipeline as shown in Fig. 5.8. Where the
height is h = 10 m, the pipe diameter is d = 0.01
m, the hydraulic roughness is k = 0.1 mm and the
pipeline length is l = 500 m. Water flow velocity is
v = 1 m s-1 and temperature is t = 20 C.
Properties of water are given in Tab. 1.1 and Tab.
1.2, from where the density is = 998 kg.m-3, the
kinematic viscosity is = 1.00610-6 m2 s-1, and the
specific heat is c = 4187 J kg-1 K-1.
The volumetric flow rate:

d2

3.14 0.012
= 7.854 10 5 m 3 s 1
Fig. 5.8
4
4
The Reynolds number:
vd
1.0 0.01
Re =
=
= 9.940 10 3
v
1.006 10 6
Relative roughness:
k 0 .1
=
= 0.01
d 10
The friction factor is = 0.044, read off from the Moody diagram in Fig. 5.7.
Specific energy loss due to friction:

Q=v

= 1 .0

l v2
500 1.0 2
= 0.044
= 1100 J kg -1
d 2
0.01 2
Pressure loss due to friction from the Bernoulli equation:
ez =

p
p
v2
v2
+ 1 + g z1 + 0 =
+ 2 + g z 2 + e z , h = z 2 z1
2

p = p 2 p1 = g h e z = 9.81 998 10 998 1100 = 1.196 10 6 Pa


e=

Rise of the water temperature due to friction:

T =

e z 1100
=
= 0.2627 K
c 4187
56

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

5.2.3 Local losses


The piping and hydraulic systems consist of various components such as manifolds,
flanges, fittings, valves, filters, the pressure containing parts of other components such as
expansion joints, etc. The fluid flow through such components causes hydraulic losses. Such
elements are referred to as local hydraulic resistances.
The specific energy loss due to local losses is defined by the formula

v2
,
(5.23)
2
where is the local loss coefficient and v is the characteristic flow velocity related to the
chosen cross-section, before or behind the local resistance (see Fig. 5.9).
ez =

Fig. 5.9 Definition of the specific energy loss and local loss coefficient
The local loss factor is usually expressed as a function of the configuration of the given
local hydraulic resistance and the Reynolds number, which can be written as

= f (configuration, Re) .

(5.24)

Typical dependence of the local loss factor versus the Reynolds number is shown in Fig. 5.9.
Note: The proper application of the basic equation (5.23) to determine the specific energy loss due to
local losses requires the correct identification of the reference cross-section of the piping, where the
characteristic dimension is defined. The fluid velocity determined in this cross-section is then used for
to accurately determine the Reynolds number. Determination wrong estimate of the reference crosssection can lead to significant computational errors. These errors can be expressed in terms of the
relationship (5.26) between the local loss factor
and flow velocity v in different cross-sections
A. The specific energy loss of the local hydraulic
resistance for two cross-sections, as shown in
Fig. 5.10, can be expressed by using (5.23) as

e z = 1

v12
v2
= 2 2 ,
2
2

(5.25)

from where, using the continuity equation (5.2),


the relationship for the ratio of local loss factors,
velocities in given cross-sections, and piping
diameters follows

Fig. 5.10 Conversion of local loss coefficients


2

2 v1 A2 d 2
= .
= =
1 v 2 A1 d 1
57

(5.26)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

5.2.4 Local loss coefficients


The values of local loss coefficients depend on their configuration and the Reynolds
number as was described in the previous
section by the general function (5.24). Such
dependencies for various components of
piping and hydraulic systems are
determined experimentally. The results of
such measurements are available in
literature, e.g. [4], [17], and [18], in the
form of tables and charts. Some typical
cases of local hydraulic resistances are
described in the following passage.
Elbows and branches are the basic
components of piping systems. Losses due
to changes in flow direction occur in such
elements. Values of local loss factors for
elbow depending on the bending angle are
plotted in Fig. 5.11.

Fig. 5.11 Local loss coefficient for circular


cross-section segment elbow

Sudden contraction or expansion of the tube cross-section causes the loss due to the
change of the flow velocity. Basic formulas for determining the local loss coefficients are
given in Fig. 5.12. These expressions are independent of the Reynolds number, and therefore
their use has a wide range of validity.

Fig. 5.12 Local loss coefficients for the sudden contraction and expansion
The values of the inlet and outlet loss coefficients for the intake to the pipe from the
reservoir and for the outlet from the pipe into the reservoir given in Fig. 5.13 represent the
limiting cases of the above relationships for the local losses due to sudden contraction and
expansion.

Fig. 5.13 Local loss coefficients for the pipe inlet and outlet

58

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Gradual expansion is called diffuser (see Fig. 4.14). Depending on the cross-sectional
area ratio and the angle of the diffuser , there may or may not be separation losses. The local
loss coefficient is given by
2

A
= k 1 2 ,
(5.27)
A
1

where values of the diffuser


coefficient k depending on the cone
angle are plotted in Fig. 5.14.
The lowest value is for an angle
of 6. The worst case is at an angle of
about 60. Thus a sudden expansion
(90) may be better than an attempt at
a short diffuser.

Fig. 5.14 Diffuser coefficient as a function of cone angle

Orifice plates inserted into the piping system cause local losses due to the throttling
effect. Orifices are used for metering
flows as part of a system for
measuring the fluid flow rate (see
Chap. 7.3). Restriction orifices are
used as local flow restrictors in the
various branches of piping systems or
as devices for reducing fluid pressure.
A plot of the local loss coefficient
as a function of the Reynolds
number Re = v D / v for various
contractions mc = (d /D ) is shown in
Fig. 5.15, where d is the orifice
diameter and D is the pipe diameter.
Here, the characteristic dimension is
the pipe diameter and the reference
velocity is the fluid velocity in the
pipe before the contraction.
2

Fig. 5.15 Local loss coefficient for the orifice

Fittings valves, gate valves, ball valve, and clack valves are used to open and close
pipes or to control the flow and pressure level. Schematics of the basic valve configurations
are shown in Fig. 5.16, where d is the characteristic dimension, h is the valve lift, and is the
angle of opening of the cock or valve.

Fig. 5.16 Schematics of: a) gate valve, b) valve, c) ball valve, d) clack valves
59

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Hydraulic losses in fittings include effects of friction and mainly turbulence losses. The
flow cross-section is considerably reduced in valves and cocks. Local loss coefficients are
given by the fittings design and configuration. Values of the local loss coefficient are usually
determined experimentally for varied opening. These values are available from fitting
manufacturers and suppliers in the form of tables and charts.
An example of the measured dependence of the local loss coefficient on the Reynolds
number for a varied size of the flat valve opening d / h is shown in Fig. 5.17. The control
parameter here is the ratio of the
valve opening h to the inlet
diameter d. It is obvious from Fig.
5.17 that values of the local loss
coefficient are constant for the
higher Reynolds numbers and
depend only on the value of
control parameter.
These constant values of local
loss coefficients for basic valve
configurations as shown in Fig.
5.16 are given as the dependence
on the control parameter in Tab.
5.2.
Even when fully open, valves
show a loss. Values of the local
loss coefficients of some types of
fittings for the full opening are
listed in Tab. 5.3

Fig. 5.17 Local loss coefficient of flat valve

Tab. 5.2 Values of the local loss coefficient in dependence of the control parameter
for basic valve configurations as shown in Fig. 5.16

a)
b)
c)
d)

d/h

d/h

1.14
0.07
0.5
2.235
85
0.05
60
3.2

1.33
0.26
1
1.94
80
0.29
50
6.6

1.6
0.81
2
1.56
70
1.56
45
9.5

2
2.06
4
1.64
60
5.17
40
14

2.67
5.52
6
2.84
50
17.3
30
30

4
17
8
5.16
45
31.2
20
62

8
98
10
8.6
40
52.6

32
160
20
22
30
206

Tab. 5.3 Values of the local loss coefficient for some valves at full opening
Type of fitting
Straight shut-off
valve
Angled shut-off
valve
Return valve

Nominal inside diameter d [mm]


50

100

150

200

250

300

350

400

450

500

3.3

4.1

4.7

5.3

5.8

6.2

6.4

6.6

6.8

6.9

5.4

5.8

6.2

6.6

7.0

7.4

7.7

8.0

8.3

1.3

1.5

1.7

1.9

2.0

2.1

2.2

2.3

2.3

2.5

60

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

5.2.5 Resulting specific energy loss


Particular branches of piping systems or hydraulic circuits can be considered as sets
consisting of sections of straight tubes of constant diameter and inserted local resistances. The
overall specific energy loss represents the sum of individual losses of particular elements of
the considered branch. The sum of friction losses in n straight sections of piping and losses in
k local resistances can be written as
n
k
v 2j
l v2
e z = i i i + j
,
(5.28)
d i 2 j =1
2
i =1
where the reference velocities v i or v j are found from the continuity equation for individual
elements of the system through the pipe diameters di or characteristic dimensions dj.
Note: When analyzing the overall specific energy loss in long piping systems, the local losses are
usually negligible in comparison with the friction losses in straight pipe sections. This is advantageous
due to some uncertainty in determining the coefficients of local losses.

5.2.6 Pressure loss, head loss, power loss


When quantifying the friction losses during fluid flow, besides the specific energy loss,
there are frequently used terms: pressure loss, head loss, and power loss.
The pressure loss p z can be defined as a pressure drop of the flowing fluid as the result
of friction losses. The specific energy loss and the pressure loss are related by

p z = e z .

(5.29)

From the physical point of view, the pressure loss represents the energy loss per unit volume.
The head loss hz is defined as the measure of the reduction in the total head of the liquid
as it flows through a system. The head loss is linked with the specific energy loss by

hz =

e z p z
=
.
g
g

(5.30)

Note: When fluid flows, the work of friction forces is irreversibly dissipated into heat energy, which
increases the internal energy of the fluid and leads to a temperature rise (see Chap. 5.1.3).

The power loss Pz then represents the amount of energy loss per unit time. From (3.9)
between the mass flow rate Qm and the volumetric flow rate Q, there arise the relations of the
power loss to the specific energy loss ez, the pressure loss p z , and the head loss hz as
Pz = Qm e z ,

Pz = Q p z , and

Pz = Q g hz .

(5.31)

The power loss also represents the heat energy rate supplied to the fluid due to friction,
which may adversely affect the function of closed hydraulic systems by the substantial
increase of the hydraulic fluid temperature.
If there is no heat exchange with the environment, the energy loss addition causes the
increase of the fluid temperature (see 5.15). This can be expressed through the power loss Pz
with regard to the fluid mass m of specific heat c in the given time interval by the differential
relationship
de
P
dT
dT
Pz = m z = m c

= z ,
(5.32)
dt
dt
dt m c
where the rate of increase in the fluid temperature dT / dt is expressed.
Note: If the increase of the fluid temperature in the hydraulic system is too high, it is necessary to
ensure adequate cooling.
61

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 5.2
Determine the pressure loss and the power loss due to the sudden
tube expansion (Fig. 5.18). The dimensions are given as follows:
d1 = 10 mm a d2 = 20 mm. The flowing fluid is the engine oil of
temperature 100 C. Mass flow rate of oil is Qm = 0.1 kg s-1.
Physical properties of engine oil are given in Tab. 1.3, from
where the density is = 820 kg m-3 and the kinematic viscosity
is = 2.43910-6 m2 s-1.
The flow velocity from the mass flow rate:
4 Qm
4 0.1
v1 =
=
= 1.553 m s 1
2
d 1 820 3.14 0.012

Fig. 5.18

The Reynolds number:


v d
1.553 0.01
Re1 = 1 1 =
= 6366
v
2.439 10 6
The local loss coefficient (Fig. 5.12):
2

2
d 2
0.01 2

A1
1 = 1 = 1 1 = 1
= 0.5625
A2
d 2

0.02
Specific energy loss:
v2
1.553 2
e z = 1 1 = 0.5625
= 0.678 J kg 1
2
2
Pressure loss and power loss:

p z = e z = 820 0.678 = 556 Pa

and

Pz = Q m e z = 0.1 0.678 = 6.78 10 2 W

Example 5.3
Determine the pressure loss on the orifice inserted into the pipe (Fig. 5.19)
for the air flow at temperature 0 C. The mass flow rate is Qm = 10-3 kg s-1.
The pipe diameter is D = 10 mm and the orifice diameter d = 8 mm.
Physical properties of air are given in Tab. 1.1, from where the density is
= 1.29 kg m-3 and the dynamic viscosity is = 1.728610-5 Pa s.

Fig. 5.19

The flow velocity from the mass flow rate:


4 Qm
4 10 3
v=
=
= 9.87 m s 1
2
2
D 1.29 3.14 0.01
The Reynolds number:

Re =

v d 1.29 9.87 0.01


=
= 7366

1.7286 10 5

Contraction of the flow cross-section and local loss coefficient:


2

d
0.008
mc = =
= 0.64 and = 1.0 from diagram in Fig. 5.14 for given Re
D
0.01
Pressure loss and power loss:
v2
9.87 2
p z = e z =
= 1.29 1.0
= 62.8 Pa
2
2
Q
10 3
Pz = m p z =
62.8 = 4.87 10 2 W

1.29
62

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 5.4
Determine the parameters of
flow of oil at temperature 80 C

in the connecting pipe


between vessels as shown in
Fig. 5.20. The diameter of the
connecting pipe is d = 4 mm,
the pipe length is l = 0.2 m and
the hydraulic roughness is k =
0.02 mm.
The level heights are: h = 0.1 m,
h1 = 0.05 m, and h2 = 0.03 m.

Fig. 5.20

The pressures above the levels are pN1 = 1.2105 Pa and pN2 = 105 Pa.
Physical properties of the engine oil are given in Tab. 1.3, from where the density is = 832 kg m-3
and the kinematic viscosity is = 3.406310-6 m2 s-1
Relative roughness:

k 0.02
=
= 0.005
d
4

From the Moody diagram in


Fig. 5.7, the friction factor is
= 0.031 for Re > Rem = 2105.
The scheme of the friction
factor determination from the
Moody diagram is shown in
Fig. 5.21.
The local loss coefficient for
the pipe inlet is in = 0.5
from Fig. 5.13.

Fig. 5.21

Application of the Bernoulli equation (5.13) between points N1 and 2:

e=

( )

v N2 1 p N1
v2 p
+
+ g z N1 = 2 + 2 + g z 2 + e zi
2

e=0+

p N1

+ g (h2 + h ) =

N1, 2

v 2 p N 2 + g h2
v2
l v2
+
+ 0 + in
+

2
2
d 2

from where the flow velocity is:

v=

p pN 2
2 g h + N 1
g

l
in + + 1
d

(1.2 1.0)10 5
2 9.81 0,1 +
9.81 832

0,2
0.5 + 0.031
+1
0.004

= 4.073 m s 1

The Reynolds number:

Re =

v d 4.073 0.004
=
= 4.783 10 3 < Re m = 2 10 5
6
v
3.4063 10

the above-mentioned assumption does not comply, therefore, the next evaluation of the friction
factor for the newly determined Re is necessary (see Fig. 5.7), which is = 0.044.
63

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The flow velocity is then:

p pN 2
2 g h + N 1
g

l
in + + 1
d

v=

(1.2 1.0)10 5

2 9.81 0.1 +
9.81 832

0,2
0.5 + 0.044
+1
0.004

= 3.677 m s 1

The Reynolds number:

Re =

v d 3.677 0.004
=
= 4.318 10 3
v
3.4063 10 6

Further evaluation of the friction factor from the Moody diagram for the currently determined Re gives
the same value = 0.044, and hence, the iterative calculation is terminated.
The mass flow rate:

Qm = v

d2

3.14 0.004 2
= 832 3.677
= 3.845 10 2 kg s 1
4

From the pressure boundary condition, p2 is equal to the hydrostatic pressure:

p 2 = p N 2 + g h2 = 10 5 + 9.81 832 0.03 = 1.00245 10 5 Pa


Application of the Bernoulli equation (5.13) between points N1 and 1:

e=

( )

v N2 1 p N1
v2 p
+
+ g z N1 = 1 + 1 + g z1 + e zi
2

e=0+

p N1

+ g (h2 + h ) + 0 =

N1,1

v 2 p1
v2
+
+ g (h2 + h h1 ) + in

2
2

From where pressure p1:

p1 = p N1 + g h1 (1 + in )

v2
3.677 2
= 1.2 10 5 + 9.81 832 0.05 (1 + 0.5) 832
= 111972 Pa
2
2

Specific loss energy for flow between points 1 and 2:

(e z )1, 2 = (e z

1, 2

2
2
l
0.2

3.677
= in + + 1
= 0.5 + 0.044
+ 1
= 25.019 J kg 1
d
0.004 2

The pressure loss:

p z = (e z )1, 2 = 832 25.019 = 20816 Pa


The power loss:

Pz = Qm (e z )1, 2 = 3.845 10 2 25.019 = 0.96 W


Note: If we apply the Bernoulli equation between points N1 and N2:

e=

( )

v N2 1 p N1
v2
p
+
+ g z N1 = N 2 + N 2 + g z N 2 + e zi
2

e=0+

p N1

+ g (h2 + h ) = 0 +

pN2

+ g h2 + in

N1, N2

v2
l v2
v2
+
+ out
2
d 2
2

Then, by comparison with the application of the Bernoulli equation between points N1 and 2, it is
evident, that the value of the local loss coefficient for the pipe outlet is always out = 1.
64

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 5.5
Determine the parameters of water
flow at temperature 10 C in the
discharge pipe as shown in Fig.
5.22. The pipe diameter is d = 52
mm and the length is l = 50 m.
The pipe surface is hydraulically
smooth. Height is h1 = 1 m. The
pipe inlet has rounded edges with
the local loss coefficient in = 0.1.
Water volumetric flow rate is Q =
310-3 m3s-1. Atmospheric pressure
is pa = 105 Pa.

Fig. 5.22

Physical properties of water are given in Tab. 1.2, from where the density is = 1000 kg m-3 and
kinematic viscosity is = 1.30110-6 m2 s-1.
Flow velocity from the volumetric flow rate:

4Q
4 3 10 3
v=
=
= 1.413 m s 1
2
2
d
3.14 0.052
The Reynolds number:

v d 1.413 0.052
=
= 5.646 10 4
6
v
1.301 10
The friction factor is = 0.019 for the smooth pipe = k d = 0 from the Moody diagram in Fig. 5.7
Re =

Application of the Bernoulli equation (5.13) between points 0 and 2:

e=

( )

v 02 p 0
v2 p
+
+ g z 0 = 2 + 2 + g z 2 + e zi
2

e=0+

pa

+gh=

0, 2

v 2 pa
v2
l v2
+
+ 0 + in
+

2
2
d 2

From where the elevation of the level:


2
2
l
50

1.413
h = in + + 1
= 0.1 + 0.019
+ 1
= 1.3 m
d
0.052 29.81

2g
Application of the Bernoulli equation (5.13) between points 0 and 1:

e=

( )

v 02 p 0
v2 p
+
+ g z 0 = 1 + 1 + g z1 + e zi
2

e=0+

pa

+ g h+0=

0, 1

v 2 p1
v2
,
+
+ g (h h1 ) + in

2
2

From where pressure p1:

p1 = p a + g h1 (1 + in )

v2
1.413 2
= 10 5 + 9.81 10 3 1 (1 + 0.1) 10 3
= 1.087 10 5 Pa
2
2

Head loss:

hz = h = 1.3 m
Power loss:

Pz = g Q hz = 9.81 10 3 3 10 3 1.3 = 38.3 W


65

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

6 Flow with External Energy Supply


Sources of external mechanical energy supplied to (or removed from) the fluid are
hydraulic machines, which are mainly divided into pumps and turbines.
The specific energy of the fluid is increased by the pump operation as a result of supply of
mechanical energy. On the contrary, in turbines, the mechanical energy is produced and the
specific energy of the fluid is decreased.
In deriving the fundamental equations of the fluid flow with external mechanical energy
supply and their application to particular engineering problems, we focus on pumps as
hydraulic machines, which are most commonly used in practice. These fundamental equations
(basic laws) are valid generally, however. The basic theory for the hydraulic machines has a
common basis and a number of conclusions can be used for turbines.
The description of fluid flow with external mechanical energy supply is based on the
following assumptions:
The flow is assumed to be one-dimensional and steady-state
The change in the flow cross-section occurs
The mechanical energy supplied (or produced) by the external source in a certain
part of the piping system
Flow takes place in the gravity field
The effect of hydraulic friction is taken into account through the losses
The fluid is considered to be incompressible the density of which can be regarded as
constant, so that
= constant .
(6.1)

6.1 Basic Laws of Flow with External Energy Supply


The set of equations, describing the fluid flow with an external energy supply, is based on
the application of three basic laws of conservation: mass, momentum, and energy.

6.1.1 Conservation of mass continuity equation


For the incompressible fluid flow, the density of which is constant, it is possible to write
the continuity equation in the known form valid for liquids
Q
(6.2)
Q = m = v A = constant ,

where Q is volumetric and Qm is mass flow rate.

6.1.2 Conservation of momentum


Newton's second law, applied to the mass
element dm bounded by the control surface in the
direction of flow s (see Fig. 6.1), takes the form
d
(dm v s ) = dFs .
(6.3)
dt
This component of the differential resultant force
in the s direction, acting on the mass element dm,
can be for the given case of flow expressed by
dFs = dF ps + dFms + dFexs .

(6.4)
66

Fig. 6.1

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

This is the sum of the components of differential forces, namely: the pressure force dF ps ,
the mass force (gravitational) dFms , and the external force from a hydraulic mechanism dFexs
acting on the fluid in the control surface in the direction s. This force supplied by the
hydraulic machine can be viewed, for example, as a pressure force, which is delivered to the
fluid by an elemental part of the pump blade.
The change in momentum with time of the mass element, the component of pressure force
and mass force acting in the direction of s were described in detail in Chap. 3.3, where by
equations (3.12), (3.15), and (3.16) the following relationships are given:
d
(dm v s ) = v A dv ,
dt

dF ps = A dp ,

dFms = g A dz .

(6.5)

By substituting equations (6.4) and (6.5) into the basic relationship (6.3) we obtain the
motion equation in the form

v A dv = A d g A dz + dFexs .

(6.6)

By applying the force balance per unit mass, i.e. dividing equation (6.6) by the element of
mass dm = A ds , we will obtain Euler's differential equation in the form
v

dv
dz dFexs
1 dp
.
=
g
+
ds
ds
ds
dm

(6.7)

By multiplying the Euler's equation (6.7) by the displacement ds, we will obtain the work
performed along the flow tube related to unit mass as
de = dw =

dFs ds
dp
= v dv +
+ g dz ,
dm

(6.8)

where the change in the total specific energy of fluid de is equal to the specific work dw done
by a hydraulic mechanism.
This work represents the specific mechanical energy supplied to the fluid from an external
source, which causes the increase of the kinetic energy, the pressure energy, and the change of
potential energy per unit mass of the flowing fluid.
The integration of equation (6.8), considering the fluid density invariant = constant (6.1),
leads to the Bernoulli equation applied to the flow with an external mechanical energy supply
in the form

v2 p
v2 p
v 2 v12 p 2 p1
w = e2 e1 = 2 + 2 + g z 2 1 + 1 + g z1 = 2
+
+ g (z 2 z1 ) , (6.9)
2

which states that the sum of differences of the specific kinetic energy, the pressure energy and
the potential energy of the flowing fluid before and after the hydraulic mechanism in question
is equal to the difference of the total specific energies in the given cross-sections of the stream
tube.
The difference of the total specific energies e2 e1 represents the resultant amount of the
mechanical energy (work) w, per unit mass, supplied to the fluid during its passage through a
hydraulic machine.
Note: The friction losses associated with the fluid flow within the hydraulic mechanism were not
included into the consideration explicitly, because it is assumed that part of the shaft power is spent to
cover the friction losses. This is reflected in the pump efficiency.
67

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

6.2 Pump Specific Energy


Pumps (Fig. 6.2) are hydraulic machines in which the mechanical energy is supplied to
the fluid. The pump specific energy represents the mechanical energy per unit mass, which is
supplied to the fluid as it passes through the pump.
This pump unit energy (work), which is commonly
denoted by Y, can be expressed by the Bernoulli
equation (6.9) for the flow with external
mechanical energy supply applied between points s
and d, located in the suction and in the discharge
sides (see Fig. 6.2) of the pump as

Y =w=

v d2 v s2 p d p s
+
+ g ( z d z s ) . (6.10)
2

Here, we do not deal with the process of the


mechanical energy supply in the pump, but only
with the resulting effect of such energy supply,
which is expressed by the difference of kinetic,
pressure, and potential energies per unit mass of
the flowing fluid before and after the pump.

Fig. 6.2 Schematic of pump


When quantifying the amount of energy
supplied to the fluid, besides the pump specific energy, there are frequently used terms:
differential head and differential pressure of the pump. The concept of the total differential
head is usually connected with the problems of the dynamic pumps, while the term of the total
differential pressure is used for the analysis of the positive displacement pumps.
Conversion formulas between the pump specific energy Y, the total differential head H,
and the total differential pressure p are given by
H=

and

p = .

(6.11)

6.3 Pump Characteristics


Pumps are of two major types: the positive displacement pumps and the dynamic pumps.

Positive displacement pumps operate on the principle of forcing a fixed pump working
volume. According to the design we distinguish: reciprocating pumps, gear pumps, peristaltic
pump, vane pumps, roots blowers, etc. A typical feature of the positive displacement pumps is
the fact that the pump delivery capacity is proportional to the pump size and speed.
Roto-dynamic pumps increase the specific fluid energy which is achieved by the action of
the hydrodynamic forces. Roto-dynamic pumps are further subdivided according to the way in
which fluid flows through them: centrifugal pumps, axial flow pumps, and mixed flow
pumps. The delivery capacity is not explicitly linked to the pump size and speed. The pump
capacity is determined on the basis of the pump characteristic.
The pump characteristic curve is a graphical representation of the relationship between
the pump specific energy Y (or the total head H) and the mass flow rate Qm (pump capacity)
for the actual pump at constant speed of rotation RPM.
Characteristic curves are obtained by testing of the actual pumps or given by
manufacturers. Typical characteristic curves for the positive displacement pump and the
centrifugal pump are shown in Fig. 6.3.
68

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Fig. 6.3 Characteristic curves for: a) positive displacement pump; b) centrifugal pump
The characteristic of the positive displacement pump shown in Fig. 6.3 a) is so-called
hard, i.e. the pump delivers at a fixed speed a nearly constant flow rate at a wide range of
pressures. Hence, the pump total head (or total pressure) can grow almost unlimitedly. The
deviation from the ideal course, which is the perpendicular line Qm = constant, is given by the
internal leakage in the pump, when fluid leaks from the discharge back to the pump suction at
higher pressures.
The characteristic of the centrifugal pump shown in Fig. 6.3 b) is so-called soft, when the
pump maintains at a fixed speed fairly constant specific energy (total head) over a relatively
wide range of flow rates. Dynamic pumps produce energy when operate with fluid but with
no flow rate. This beginning point of the pump characteristic curve is called the shut-off head
point. There usually exists a maximum head point on the centrifugal pump characteristic
curve. For higher flow rates, the specific energy supplied to the fluid decreases due to an
increase in the fluid friction loss at a higher flow velocity in the pump.
Pump characteristics are obtained
experimentally. Pumps are tested on a
special stands in test laboratories. The
schematic of such a pump testing system is
shown in Fig. 6.4
The fluid flow rate Q can be measured
by means of a pressure differential device
or by using a flow-meter. Pumps are tested
at different speeds and flow rates. The
fluid flow is controlled by the valve in the
range from the maximum rate, which the
system allows, to the zero flow rate, when
the control valve is fully closed.
The parameters usually measured are
as follows:
Suction pressure ps at the pump inlet
Discharge pressure pd at the pump
outlet
Fluid flow rate Q
Pump speed of rotation RPM
Pump driving moment (torque) Mt

Fig. 6.4 Schematic of the pump


testing system
69

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The pump characteristics are usually described graphically as the pump performance
curves. The pump characteristic curves describe relations of the specific energy Y, the power
consumption Pef, and the efficiency versus the mass flow rate Qm at the constant pump
speed RPM as shown in Fig. 6.5.
Fluid velocities in the suction and discharge pipes are determined from the measured flow
rate Q through the continuity equation as
4Q
4Q
vs =
and v d =
.
(6.12)
2
ds
d d2
The pump specific energy Y can be obtained
from the above-mentioned equation (6.10) by
introducing the suction and discharge pressures, ps
and pd, measured in the suction and discharge sides
of the pump (Fig. 6.2) of heights zs and zd. Thus,

v d2 v s2 p d p s
Y=
+
+ g (z d z s ) (6.13)
2

If the diameters of the suction and discharge


pipes are the same, and if the difference in heights
zs and zd can be neglected, the equation (6.13) takes
the simplified form
p ps
Y= d
(6.14)

Fig. 6.5 Characteristic curves for


the centrifugal pump

The pump power consumption can be determined from the torque moment Mt on the
pump shaft and the speed of rotation RPM (revolutions per minute) by
RPM
.
(6.15)
Pef = M t 2
60
The hydraulic output power represents the
energy supplied to the fluid per second, which can
be obtained from the specific energy Y and the mass
flow rate Qm by

Ph = Qm Y .

(6.16)

The pump hydraulic efficiency is given by the


ratio of the hydraulic output power Ph and the input
power consumption Pef as

Ph
.
Pef

(6.17)

Fig. 6.6 Characteristic curves for


the positive displacement pump

It is seen from the pump characteristic (Fig. 6.5) that the curve of hydraulic efficiency has
the point of maximum efficiency. This is the optimal operating point, at which the pump
losses are the lowest.
Note: The pump best operating conditions, which are close to the maximum efficiency point, are
reported as nameplate data by the manufacturer. These nameplate data should be chosen as operating
parameters, especially for long-term operation.

The characteristic curves for the positive displacement pumps are usually plotted as the
dependence of the mass flow rate Qm, the power consumption Pef, and the efficiency versus
the discharge pressure at the constant speed, as shown in Fig. 6.6.
70

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

6.4 Pump Operation at Various Speeds


The characteristic curves describe the relations of the pump properties at the constant
speed. If we plot a set of head, power and efficiency curves at varying speeds into one
diagram, we obtain the composite
characteristics as shown for a
centrifugal pump in Fig. 6.7.
The plotted field of efficiencies in
Fig. 6.7 shows that the maximum
efficiency of the pump can be achieved
at certain head, flow rate and speed,
i.e. at operation conditions for which
the pump was designed. The deviation
from
these
optimal
operating
conditions always causes a decrease in
pump efficiency.
Similar composite characteristics
can be plotted for several impeller
diameters of a centrifugal pump.
The change in the flow rate, the
specific energy and the power
consumption due to the change in the
pump speed of rotation can be
determined by conversion formulas
that follow from the theory of
geometric and kinematics similarity as
Q2 RPM 2
,
=
Q1 RPM 1

Fig. 6.7 Composite characteristics for


the centrifugal pump

Y2 H 2 RPM 2
=
=
Y1
H 1 RPM 1

and

Pef 2 RPM 2

=
.
P RPM
ef
1
1

(6.18)

Note: These relations for the conversion of the pump parameters with the change in speed of rotation
(revolutions per minute) are sufficiently accurate only for relatively small changes in speed, because
the assumption of the constant pump efficiency was applied.

6.5 Integrating the Pump into Piping System


The pump inserted into a piping
system (Fig 6.8) provides fluid delivery
between two tanks. Locations of the
fluid levels in the tanks are denoted as 1
and 2. The elevation of the liquid
surface is given by the height h and
there are to distinguish the suction
height hs and the discharge height hd,
depending on the pump position in the
system. The tanks can be pressurized or
evacuated with respect to the
barometric pressure.
The schematic of such a pump
system is shown in Fig 6.8.

Fig. 6.8 Schematic of pump system


71

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The specific energy Yp required for the delivery of fluid between points 1 and 2 of the
piping system can be determined by applying the Bernoulli equations (6.9) and (5.13) to the
fluid flow with an energy supply and considering the hydraulic losses in the form

Y p = e2 e1 =

v 22 v12 p 2 p1
+
+ g h + e zi
2

( )

1, 2

(6.19)

The difference in the kinetic energies between the given points of the hydraulic system is
obviously very small in comparison with other kinds of energy and, therefore, can be
neglected.
The piping characteristic, also known as
the system characteristic, (see Fig. 6.9) is a
graphical representation of the required
specific energy Yp versus the mass flow rate
Qm. This dependence is based on equation
(6.19), which, by neglecting the difference in
the kinetic energies, takes the form
Yp =

p 2 p1

( )

+ g h + e zi

1, 2

(6.19)

If the pressures in points 1 and 2 are


equal, i.e. p1 = p2, then equation (6.20) can be
further simplified to

( )

Y p = g h + e zi

1, 2

(6.21)

Fig. 6.9 Piping characteristic curves

The piping characteristic curve (Fig. 6.9) has a typical parabolic shape due to the
quadratic dependence of the hydraulic losses on the fluid velocity. The course of the dashdotted piping curve in Fig. 6.9 shows changed conditions due to the increase in hydraulic
losses by e.g. reduction of the pipe cross-section (diameter) or by insertion of the substantial
local hydraulic resistance (valves, orifices, etc.).
The characteristic curve origin (see Fig. 6.9) is shifted from the zero axis by Yp0, which
represents the value of the sum of the differences in pressure energy and potential energy
(6.20), or only the potential energy (6.21) in the case of equal pressures. That is because, for
the zero flow rate, the flow velocity is zero and then the hydraulic losses are zero. Thus,
Y p0 =

p 2 p1

+ g h , or Y p 0 = g h . (6.22)

If we combine the piping curve and the


actual pump curve as shown in Fig. 6.10, we
obtain the point P at their intersection. This
point is called the pump operating point, in
which the specific energy Yp required for fluid
delivery is equal to the pump specific energy
Y, thus
(6.23)
Yp = Y .
The operating point defines the fluid flow
rate through the given piping system and for
the actual integrated pump.
72

Fig. 6.10 Pump operating point

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Combined characteristic curves of piping systems and pumps (Fig. 6.10) are often used
for the piping design and for actual pump selection. Furthermore, it is possible to analyze the
possibilities of the control of pumping rates, e.g. by the fluid throttling in the control valve,
using of which the operating point P can be shifted to the dash-dotted curve, as shown in Fig.
6.10.
If the fluid is pumped in the opposite direction in the given piping system, which is shown
in Fig. 6.8, then the fluid flow occurs from point 2 to point 1 and this causes the change of
signs in equation (6.22). That is
Y p0 =

p1 p 2

g h,

or

This brings the change in characteristics as


shown in Fig. 6.11. The operating points P are
then at higher flow rates if compared with the
previous one in Fig. 6.10.
However, if the given piping system is not
provided with the pump, then the fluid is
flowing freely under the force of gravity without
external energy supply. This kind of flow is
called gravitational flow. The operating point P
of such a flow represents the intersection of the
piping curve with the Qm axis. For this operating
point, all the pressure and potential energy is
absorbed in flow losses and so the required
energy is zero
(6.25)
Yp = Y = 0 .

Yp0 = g h .

Fig. 6.11 Piping characteristic curves

Example 6.1
The gravitational flow of water occurs in
the system, as shown in Fig.6.12.
Determine the piping characteristic curves
for connecting pipe diameters 150, 200, 250
and 300 mm. The pipe is hydraulically
smooth of length l = 150 m. The difference
in heights between the water surface levels
is h = 15 m.
Water temperature is 10 C.
Fig. 6.12 Schematic of piping system

Physical properties of water are given in


tab. 1.2, from where the density is = 1000
kg m-3 and the kinematic viscosity is =
1.30110-6 m2 s-1.

The local loss coefficient for the pipe inlet is in = 0.5 (from Fig, 5.14) and for outlet is out = 1.
The friction factor is obtained from the Moody diagram (Fig. 5.7) for = k d = 0 .
Velocity from the mass flow rate:

v=

4 Qm
d 2
73

(6.24)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Reynolds number:

Re =

vd
v

Required specific energy:

( )

Y p = g h + e zi

1, 2

2
l
150 v 2

= g h + in + + out
= 9.81 15 + 0.5 +
+ 1
d
d

The evaluation of these variables is to be carried out for the connecting pipe diameters d = 150, 200,
250, and 300 mm and for the chosen values of the mass flow rate 0, 100, 200, 300, 400, and 500 kg s-1.
The results of calculation for the given variables are presented in Tab. 6.1 for the pipe diameter d =
300 mm. Similar tables for other diameters are not outlined here.
Tab. 6.1 Resulting values of velocity, Reynolds number, friction factor, specific loss energy and
required specific energy for the pipe diameter d = 300 mm
Qm

kg s-1

100

200

300

400

500

m s-1

1.415

2.831

4.246

5.662

7.077

Re

326386

652772

979159

0.014

0.0132

0.0115

0.0118

0.0112

1305545 1631931

(ezi)1,2

J kg-1

8.5

32.5

65.4

118.6

177.8

Yp

J kg-1

-147.15

-138.6

-114.7

-81.8

-28.5

30.7

The graphical representation of the overall results is given in Fig. 6.13 as the pipe characteristic
curves. For the gravitational flow, the intersections of pipe curves with the axis of Yp = 0, as operating
points, determine the mass flow rates Qm for the given pipe diameters.

Fig. 6.13 Piping characteristic curves for several pipe diameters,


solution of Example 6.1

74

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

6.6 Piping Systems in Series and Parallel


Branching pipe systems contain several interconnected pipes in series (one after another)
or in parallel (side by side).
In the case of pipes in series (Fig. 6.14),
the piping system is composed by sections of
different diameters and lengths in series. For
the system shown in Fig. 6.14, applying the
continuity principle, the mass flow rates must
remain the same

Qm = Qm1 = Qm 2

(6.26)

and, neglecting the loss from pipe expansions


and contractions, the required specific
energies of the sections are added up
Y p = Y p1 + Y p 2 .

Fig. 6.14 Pipes in series

(6.27)

The branched piping system contains


junctions and pipes in parallel (Fig. 6.15). In
this case, the mass flow rates of branches are
summed up
Qm = Qm1 + Qm 2
(6.28)
and the required specific energy remains
constant
(6.29)
Y p = Y p1 = Y p 2 .
The
principle
of
the
graphical
superposition method, used for combined
characteristic curves of piping systems in
series and in parallel, is indicated in Fig. 6.14
and Fig. 6.15.

Fig. 6.15 Pipes in parallel

Note: Similar principles of the graphical superposition can be used to combine the characteristic
curves for pumps operated in series or in parallel. The mathematical description of the serial and
parallel operation of pumps uses the same equations (6.26) to (6.29) as given above for the piping
systems combination.

6.7 Suction Head, Cavitation


Since the mechanical energy is supplied to the fluid by the pump, the fluid specific energy
is lower in the suction pipe than in the discharge pipe of the pump system (Fig. 6.8). For this
reason, it is necessary to turn our attention to the value of the suction pressure ps in the pump
inlet, so that this pressure would not decrease under the saturated vapor pressure p''.
The value of the saturated vapor pressure is directly proportional to the liquid temperature.
From the physical point of view, this is the pressure at which the liquid boiling point is
reached at a given temperature. The relationship between saturated vapor pressure and
temperature for various liquids is given in Tab. 6.2. If the fluid is boiling, the phase changes
from liquid to gas. This process is followed by an almost three orders of magnitude increase
in the fluid volume. In addition, the reduced pressure causes the release of gases absorbed in
the liquid.
75

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Vapor bubbles are created in a region of the


lowest pressure, where the pressure of the liquid
drops below the saturated vapor pressure. The
supply of mechanical energy causes increase in
the fluid pressure along the pump vane
associated with the backward condensation of
the vapor bubbles. This creation and dissipation
(collapse) of the vapor bubbles is a very
Fig. 6.16 Schematic of bubble collapse
dynamic
process
accompanied
by
a
considerable change in the fluid local volume. The rapid collapse of the bubbles produces the
shock waves. The process of the collapse of the bubbles along the vane surface is shown in
Fig. 6.16. These changes in the fluid phase produce certain micro-explosions and microshocks effected on the cavitating surfaces and causing an erosive damage of internal pump
components. The phenomenon described above is called cavitation.
Cavitation phenomena are varied and adversely affect the operation of hydraulic
machines. They may occur in both water pumps and turbines. Operation of hydraulic
machines in the cavitation regime can cause their gradual mechanical damage, but also can
reduce hydraulic performance and efficiency. Moreover, cavitation causes the increase in
noise and vibration, which can cause other faults in hydraulic machines. Therefore, the
cavitating operating conditions should not be admitted.
Cavitation may also play an important role in pumping and delivery of liquid fuels. It is
because the hydrocarbon fuels have a larger value of the saturated vapor pressure p'' at
relatively low temperatures (Tab. 6.2). Cavitation vapor bubbles can reduce or stop the fuels
pumping process.
Tab. 6.2 Temperature dependence of the saturated vapor pressure for some liquids
Temperature
t (C)

Water
p (k Pa)

0.01
10
20
30
40
50
60
70
80
90
100

0.612
1.228
2.339
4.247
7.381
12.351
19.946
31.200
47.414
70.171
101.410

Petrol
p (k Pa)

Kerosene
p (k Pa)

Acetone
p (k Pa)
8.933

10.666
16.398
19.731
27.464
38.129
51.861
75.459
100.66
129.72

4.666
7.466
11.199
15.598
21.331
27.997
35.996
44.395
54.128

19.838
56.195
115.51
214.81
371.86

Mutual dependence of variables that characterize the fluid flow in the suction pipe line, as
shown in Fig. 6.8, can be expressed by the application of the Bernoulli equation (6.9) between
points 1 and s, i.e. at the liquid surface in the suction tank and in the pump inlet, in the form

( )

v s2 p s
=
+
+ ghs + e zi

p1

76

1, s

(6.30)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

from where it becomes possible to express the suction head hs by

( )

p1 p s v s2 e zi
hs =

g
2g
g

1, s

(6.31)

The minimum value of the suction pressure ps is given by the saturated vapor pressure p''
of the pumped liquid. The difference between the suction head liquids vapor and the suction
head is called the net positive suction head. That is
NPSH =

p1 p
g

v s2

2g

(e )

z i 1, s

hs =

p s p
,
g

(6.32)

which can be expressed in terms of the difference in pressure of a fluid at the inlet pump ps
minus the saturated vapor pressure p'' at the liquid temperature.
Besides the saturated vapor pressure p'' of the pumped liquid, the value of the suction
height is influenced by the pressure p1 in the suction tank and by the hydraulic losses in the
suction pipe line. Therefore, it is necessary to minimize such losses by the shortest length and
sufficient diameter of the suction pipe as hydraulic losses depend on the square of the flow
velocity. This is why the pump inlet diameter is often larger than the diameter of the discharge
port. This is also generally the way how to distinguish the inlet and outlet port of the pump at
the first glance.
The maximum theoretical suction lift for cold water is given by:
p1 = p a 10 5 Pa, p s = p 0, = 103 kg m 3 , v s = 0 , from where

(e )

zi 1, s

= 0.

By substituting these values into equation (6.31), we obtain the maximum theoretical suction
height hs max 10 m. However, since the saturated vapor pressure, the flow velocity, and
hydraulic losses are never zero, the real value of the maximum suction lift for water pumping
is in the range of hs max = 6 8 m .
The maximum suction lift decreases with an increased liquid temperature, as how the
saturated vapor pressure increases exponentially with temperature.
With regard to the fairly high saturated vapor pressures p'' of liquid hydrocarbons at
elevated temperatures (Tab. 6.2), the cavitation phenomena can lead to the creation of vapor
bubbles in the fuel systems of vehicles. Therefore, the fuel pump should be located inside or
near the fuel tank being below the surface of the fuel level.
Example 6.2
Determine the change in the net positive suction head with the change in the temperature of the
pumped water from 20 C to 40 C.
The saturated vapor pressure of water for the given temperatures from (Tab. 6.2): p 20 = 2.339 kPa and

p 40 = 7.381 kPa.
The change in the net positive suction head is given by application of equation (6.31):

NPSH = NPSH 40 NPSH 20 =


NPSH =

p1 p 40
p p 20
1
g
g

p 40

p 20
2339 7381
=
= 0.514 m
g
9.81 1000

The given change in the water temperature causes the suction height decrease by 0.514 m.
77

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

7 Fluid Flow through Orifices, Mouthpieces, and Nozzles


The fluid flow through an orifice represents, from the
hydraulic point of view, the flow of viscous liquid through
the local resistance. The orifice is an opening in the wall of a
tank. The shape of the orifice affects the liquid discharge.
The diameter of the discharging jet is smaller than the
diameter of the orifice due to the convergence of streamlines,
as shown in Fig. 7.1. This phenomenon, called the
contraction, is caused by the surface tension. In addition to
the friction loss in the orifice, the actual fluid velocity and
the actual discharge flow rate are lower than the ideal ones,
which would be achieved without friction.

Fig. 7.1 Contraction

7.1 Coefficients of Contraction, Velocity, and Discharge


The above-mentioned phenomena related to the liquid jet
contraction can be quantified by the empirical coefficients.
The coefficient of contraction of the liquid jet discharged
freely from the sharp edged orifice, as shown in Fig. 7.2, is
defined by the ratio
A
(7.1)
Cc =
Ai
of the actual cross-sectional area A of the jet divided by the
ideal one. The ideal cross-sectional area Ai is the orifice area.

Fig. 7. 2 Discharge from


sharp edged orifice

The coefficient of velocity is analogously defined by the


ratio
v
v
Cv =
=
,
(7.2)
vi
2gh
of the actual liquid velocity v and the ideal velocity vi of
liquid discharging from the vessel (Fig. 7.3) given by (3.32).
The coefficient of discharge is similarly defined by the
ratio of the actual volumetric flow rate of discharge Q and
the ideal volumetric flow rate of discharge Qi as
Cd =

vA
Q
=
= C v Cc ,
Qi v i Ai

(7.3)

Fig. 7.3 Schematic of the


liquid discharge

where we have obtained the mutual relation among the above-mentioned coefficients by using
the continuity equation.
Furthermore, there exists the explicit relation between the coefficient of velocity Cv and
the local loss coefficient for given discharge opening. This relationship can be obtained by
the application of the Bernoulli equation (6.9) for the liquid discharge, as shown in Fig. 7.3,

pa

+gh=

v 2 pa
v2
+
+
,
2
2

78

(7.4)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

from where we obtain the relation for actual discharge flow velocity in the form
v=

1
1+

2gh.

(7.5)

If we compare this relation with the definition equation for the coefficient of velocity (7.2)
in the form
v = C v vi ,
(7.6)
we can easily obtain the demanded relation between the coefficient of velocity Cv and the
local loss coefficient
1
1
= 2 1.
Cv =
,
or
(7.7)
Cv
1+

7.2 Orifices, Mouthpieces, Nozzles


The values of the abovementioned coefficients (7.1), (7.2), and (7.3), which describe
differences between the ideal and actual fluids, are determined experimentally for various
types and shapes of openings. The dependence of the coefficients of contraction Cc , velocity

Cv , and discharge Cd on the Reynolds number for the sharp edged circular orifice (Fig. 7.2)
is shown in the diagram in the Fig. 7.4. The Reynolds number is here defined by
v d
Re = i
,
(7.8)
v
where vi is the ideal liquid discharging velocity and the characteristic dimension is the
orifice diameter d.

Fig. 7.4 Dependence of coefficients of contraction Cc , velocity Cv , and discharge Cd


on Reynolds number for sharp edged circular orifice
It is apparent from this diagram that the values of the given coefficients are constant for
larger values of the Reynolds number. For computations we can use the following values:

Cc = 0.62 , C v = 0.98 , C d = 0.61 , = 0.04.


79

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

In order to eliminate the contraction of the fluid jet, different types of mouthpieces can be
used. The mouthpieces are short pipes attached internally or externally to the opening.
The dependence of the coefficient of discharge on the
Reynolds number for the external cylindrical mouthpieces
(Fig. 7.5) of various length/diameter ratios is shown in Fig.
7.6. The constant values of the abovementioned
coefficients, for the larger values of the Reynolds number
and the recommended length/diameter ratio l/d = 3, are as
follows:

Cc = 1.0 , C v = C d = 0.82 , = 0.49.

Fig. 7.5 External cylindrical


mouthpiece

Fig. 7.6 Dependence of coefficients of contraction on Reynolds number


for Venturi mouthpiece of various length/diameter ratios

The internal cylindrical mouthpieces (Fig. 7.7) of the


recommended length/diameter ratio l/d = 3 have the
following constant values of coefficients:

Cc = 1.0 , C v = C d = 0.81, = 0.52,


Fig. 7.7 Internal cylindrical
mouthpiece

for larger values of the Reynolds number.

The nozzle mouthpiece, shown in Fig. 7.8, is also


called the streamlined mouthpiece. The optimal geometry
of the streamlined shaping ensures the following values of
the coefficients of contraction, velocity, discharge and
local loss:

Cc = 1.0 , C v = C d = 0.99 , = 0.02.


Fig. 7.8 Nozzle mouthpiece
80

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Example 7.1
Determine parameters of water discharge from an external cylindrical mouthpiece (Fig. 7.5) of the
diameter d = 0.1 m and the length l = 0.3 m. The height of the water surface level above the
mouthpiece axis is h = 2.0 m. The barometric pressure is pa = 105 Pa. Properties of water are given by
the density = 1000 kg m-3 and the kinematic viscosity = 10-6 m2 s-1.
The mouthpiece flow cross-sectional area:

Ai =

d2
4

3.14 0.12
= 7.854 10 3 m 2
4

The ideal discharge velocity:

v i = 2 g h = 2 9.81 2.0 = 6.264 m s 1


The volumetric flow rate for discharge of ideal liquid:
Q i = v i Ai = 6 .264 7 .854 10 3 = 4 .92 10 2 m 3 s 1

The Reynolds number:

Re =

v i d 6.264 0.1
=
= 6.264 10 5
6
v
10

Evaluation of the discharge coefficients from diagram in Fig. 7.6 for the given Reynolds number and
length/diameter ratio:
For

l 0 .3
=
=3
d 0. 1

C d = 0.82

The actual discharge volumetric flow rate:


Q = C d Q i = 0 .82 4 .92 10 2 = 4 .034 10 2 m 3 s 1

The actual discharge velocity for C v = C d = 0.81 is:


v = C v v i = 0.82 6.264 = 5 .137 m s 1

7.3 Orifice Plates, Nozzles, Venturi Tubes


The flow throttling devices are frequently used for
measuring the fluid flow rate in piping systems. Among
such flow throttling devices are orifice plates, nozzles,
and Venturi tubes.
Drawings of such throttling meters are shown in Fig.
7.9, Fig. 7.10, and Fig. 7.11.

Fig. 7.9 Orifice plate

The fluid flow rate is determined by measuring the


differential pressure p = p1 p 2 generated by the
constriction of the throttling device inserted into the
piping.
To determine the actual volumetric flow rate Q,
measured by the throttling device, equation (3.50) can be
used for the ideal fluid flow within the Venturi tube in the
form
Fig. 7.10 Nozzle
81

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Q = A2 v 2 = A2

p1 p 2

A
1 2
A1

(7.9)

which is applied to the actual cross-sectional


area A2 at the point of the maximum
convergence of streamlines which occurs
shortly downstream of the throttling hole, as
shown in Fig. 7.12.

Fig. 7.11 Venturi tube

The character of the fluid flow in the throttling device is determined by the value of the
Reynolds number
Re =

v1 D
,
v

(7.10)

where the characteristic dimension is the pipe diameter D and the reference velocity is v1 .

Fig. 7.12 Contraction of the stream


Differences between the ideal and actual fluid flow are quantified by the above-defined
coefficients of contraction (7.1), velocity (7.2) and discharge (7.3). If we apply these
coefficients to the fluid flow through a throttling device, as shown in Fig. 7.12, we obtain the
following formulas

Cc =

A2
,
A2i

Cv =

v2
,
v 2i

Cd =

Q2
,
Q2 i

C d = C v Cc .

(7.11)

The geometry of the throttling devices is given by the diameter ratio , which is the ratio
of the hole diameter d to the pipe diameter D, or by the area ratio mC of the cross-sectional
areas of the hole and the pipe as

d
=
D

or

A
d
mC = = = 2 i .
A1
D
2

(7.12)

By combining these coefficients and ratios with equation (7.9) and rearranging the terms,
we obtain the following relations for the volumetric flow rate
Q = C v C c A2i v 2i = C d A2i v 2i =

82

C d A2i
1 Cc 4
2

p1 p 2

(7.13)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

To evaluate the fluid volumetric


flow rate Q, measured by the throttling
device (Fig. 7.12), the simplified
formula is used
Q=C

d2
4

(7.14)

where C is called the flow rate


coefficient, which is given by
Cd
.
(7.15)
C=
2
1 Cc 4
The flow rate coefficient is often
presented as a function of the Reynolds
number (7.10). This relation of the flow
rate coefficient C to the Reynolds
number Re for the orifice plate of
various area ratios mC is graphically
shown in Fig. 7.13. It is apparent from
this diagram that the values of the flow
rate coefficient are constant for larger
values of the Reynolds number. The
limit of constancy is shown by the
dashed line in Fig. 7.13.
The mass flow rate Qm is then
given by multiplying equation (7.14) by
the fluid density as

Fig. 7.13 Flow rate coefficient for


the orifice plate

2 p .
(7.16)
4
The method of measurement and computation of the fluid flow rate is exactly defined by
the international standard ISO 5167:2003 Measurement of liquid flow by means of pressure
differential devices inserted in circular cross-section conduits running. This standard defines
terms and symbols, establishes the general principles and requirements for methods of
measurement, installation and determination of the uncertainty of the measurement of the
flow rate. It also defines the general specified limits of the pipe size and the Reynolds
number, for which these pressure differential devices are to be used.
Qm = C

The standard ISO 5167 [25] in parts 2, 3 and 4 specifies the geometry and method of use
(installation and operating conditions) of orifice plates, nozzles, and Venturi tubes when they
are inserted in a conduit running full to determine the flow rate of the fluid through the
conduit.
This standard determines the mass flow rate of the fluid by
C

Qm =
ex d 2 2 p 1 ,
4
4
1

(7.17)

where the coefficient of expansion ex for gases is given by the empirical formula

ex = 1 (0.41 + 0.35 4 )

p
p1

and for liquids is

83

ex = 1.

(7.18)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The standard ISO 5167 gives special formulas for determining the flow rate coefficients
of orifice plates, nozzles, and Venturi tubes. These empirical formulas for various types of
throttling devices are often rather complicated. For example, the flow rate coefficient for the
orifice plate with corner pressure tapings is given by the Stolz equation in the form
C = 0.5959 + 0.0312

2 ,1

10 6
0.1840 + 0.0029
Re
8

0.75

(7.19)

Because the value of the Reynolds number in equation (7.19) is unknown, it is necessary
to use an iterative method of solution. Therefore, we have to choose the initial value of the
Reynolds number or to take it as infinity and then the last term in equation (7.19) becomes
zero.
Computation of the fluid mass flow rate Qm is performed by iteration of equations (7.17)
and (7.19). The corrected value of the Reynolds number (7.10) is provided by

Re =

v1 D 4 Qm
,
=
v
D

(7.20)

where is the dynamic viscosity.


The value of the mass flow rate in equation (7.20) is known from the previous step of the
calculation. The method has fast convergence and needs few steps of iteration only.
The standard ISO 5167 also includes data needed to determine the values of the flow rate
coefficients C and the coefficients of expansion ex in the form of tables for orifice plates,
nozzles, and Venturi tubes.
Example 7.2
An approximate computation of the volumetric flow rate of water measured by the orifice plate (see
Fig. 7.9) of dimensions: the diameter of the pipe is D = 0.2 m and the diameter of the orifice hole is d
= 0.1265 m. The measured differential pressure on the orifice plate is p = 104 Pa.
Properties of water are given by the density = 1000 kg m-3 and the kinematic viscosity = 10-6 m2 s-1.
The area ratio of the orifice plate:
2

d 0.1265
mC = =
= 0.4
D 0.2
Evaluation of the constant value of the flow rate coefficient from diagram in Fig. 7.13 for the given
area ratio:
C = 0.66

Re > 2105 (limit of constancy)

for

The volumetric flow rate:

Q=C

d2

2 p

= 0.66

3.14 0.1265 2
4

2 10 4
= 3.71 10 2 m 3 s 1
10 3

Verification of the assumption of the constant value of the flow rate coefficient:
v1 =

4Q
4 3.71 10 2
=
= 1.181 m s 1
D2
3.14 0.2 2

Re > 2105

and

Re =

v1 D 1.181 0.2
=
= 2.3612 10 5
v
10 6

the previous assumption has been fulfilled.

84

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

8 Momentum and Forces in Fluid Flow


In the flows around or through bodies, the effects of the fluid force are determined by the
Newton second law, which also implies the law of conservation of momentum. We
investigate the fluid forces and the rate of change of the linear momentum for the given flow
around or through bodies by using a suitably chosen control volume bounded by the control
surface. The advantage of applying the momentum principle is given by the fact that we can
determine the forces acting on bodies only from the knowledge of the fluid variables (velocity
and pressure) on the control surface. Therefore, it is not necessary to solve for the velocity
and pressure distribution within the control surface.

8.1 Control Surface and Control Volume


The control surface (CS) is the
hypothetical closed area through which the
fluid is flowing. The control surface is
connected to the body under consideration.
The control surface may be simple or
composite. Examples of control surfaces
are shown in Fig. 8.1. The simple control
surface (Fig. 8.1 a) is denoted by CS. The
composite case (Fig. 8.1 b) then includes
external CS1 and internal CS2 control
surfaces.

Fig. 8.1 Control surfaces

The volume of fluid contained within the control surface is called the control volume V of
mass m. External pressure forces which act upon the fluid bounded by the control surface
then represent the surface forces substituting influences from the surroundings. Volume forces
then express the force effects of external fields on the parcel of fluid in the control volume.
The typical volume force is the mass force.

8.2 Linear Momentum Principle


The Newton second law of motion
applied to the mass element dm bounded
by the simple control surface in the stream
tube, as shown in Fig. 8.2, can be written
in the vector differential form
r
d
(dm vr ) = dF .
(8.1)
dt
If we express the mass element of fluid
at the common position s in the stream tube
of the cross-sectional area A (Fig. 8.2) as

dm = A ds = A v dt = Qm dt ,

(8.2)

then the basic vector differential form of


Newtons second law of motion (8.1) can
be modified to the form expressing the rate
of change of the linear momentum by

Fig. 8.2 Determination of rate


of change of linear momentum

85

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

r
d
(dm vr ) = d ( A v dt vr ) = d (Qm vr ) = dPmom .
(8.3)
dt
dt
In writing this equation, we have introduced the rate of linear momentum given by the
definition:
r
r
Pmom = Qm v ,
(8.4)
r
where Qm is the mass flow rate and v is the vector of fluid velocity in any given flow crosssection of the stream tube.
Note: It is necessary to distinguish the physical difference between the rate of linear momentum
r
r
r
Pmom = Qm v and the linear momentum m v .

Integration of equation (8.3) within the limits designated in Fig. 8.2, i.e. between fluid
inlet and outlet cross-section, leads to the equation of fluid motion in the form
r
r
r
r
r
r
d
(
)
v
dm
=
Q
v

Q
v
=
P

P
=
F
,
(8.5)

m
2
2
m
1
1
mom
2
mom
1

dt CV
which expresses the fact that the rate of change of the linear momentum of the considered
mass within the control volume is given by the difference in the rate of the linear momentum
for the fluid flow incoming and outgoing through the control surface. The sum on the right
side of equation (8.5) represents the summation of the external forces acting upon the fluid
within the control surface.
The differential pressure force is determined from the pressure distribution on the
surfaces that are shown in Fig. 8.3, which correspond to the control surface in Fig. 8.2. That is
(8.6)
dF p = p A ( p + dp ) ( A + dA) = d ( p A) ,
where the direction and orientation are those of the vector of
velocity of the mass element in the stream tube. The form of
equation (8.6) follows from the relation A dp + p dA = d ( pA)
and neglecting the product of the two differentials dp dA.
By integrating between limits given by the inlet and outlet
cross-sections of the stream tube (Fig. 8.2), the pressure force
acting on the fluid element takes the form
r
r
r
r
(8.7)
F p = d ( p A) = p 2 A2 + p1 A1 .
The mass force is due to the force of gravity acting on the
fluid within the control surface as
r
r
r
Fm = g dV = m g .

Fig. 8.3 Pressure force


(8.8)

CV

Note: The influence of the mass force can be neglected in most of practical applications.

The reaction force R expresses the force effect of the fluid flow through bodies. In the
case of the simple control surface (Fig. 8.2), it represents the force effect of the fluid flow on
the tube inner walls determined by this control surface. From equation (8.5) it follows that
r
r
r
r
r
Pmom 2 Pmom1 = F p + Fm R .
(8.9)
By the use of equations (8.5), (8.7), (8.8) and upon rearranging the terms, we obtain the final
relation for the reaction force in the form
r
r
r
r
r
r
R = (Qm1 v1 + p1 A1 ) (Qm 2 v 2 + p 2 A2 ) + m g .
(8.10)
86

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

8.3 Applications of the Momentum Principle


The following practical applications illustrate the effect of the fluid flow force in some
cases.

8.3.1 Jet engine thrust


A jet engine represents the system,
the schematic of which is shown in Fig.
8.4. The suction air stream of velocity
v1, pressure p1, and density 1 enters
through the inlet cross-section of area
A1. The exhaust gases are discharged
from the jet engine outlet nozzle of the
cross-sectional area A2 at velocity v2,
pressure p2, and density 2. The
ambient pressure is atmospheric pa.

Fig. 8.4 Schematic of jet engine

There is applied the composite control surface. The external control surface CS1 passes
around the entire engine, enclosing the inlet and outlet (Fig. 8.4). The internal control surface
CS2 excludes the engine surface.
The equation of fluid motion (8.9), written for the s direction in the axis of the stream,
which passes through the jet engine, is written in the form
Pmom 2 s Pmom1s = Fps + Fms R ,

(8.11)

where the incoming and outgoing rates of the linear momentum are

Pmom1s = Qm1 v1

Pmom2 s = Qm 2 v 2 ,

and

(8.12)

furthermore, the incoming and outgoing mass flow rates are given by

Qm1 = 1 v1 A1

Qm 2 = 2 v 2 A2 .

and

(8.13)

The incoming and outgoing mass flow rates are not equal for jet engines with regard to the
mass flow rate Qmf, which is supplied by burning fuel. The mass flow rate of the discharged
exhaust gases is given by the summation of the mass flow rate of incoming air and supplied
fuel as
(8.14)
Qm 2 = Qm1 + Qmf .
The pressure force Fp in the
equation of fluid motion (8.9)
represents the resultant force acting
upon the external control surface CS1.
The pressure distribution on the control
surface of the jet engine (Fig. 8.4) is
schematically shown in Fig. 8.5.
The component of pressure force
Fps, acting in the direction of flow s,
is then given by

Fig. 8.5 Pressure force

F ps = ( p1 p a ) A1 + ( p 2 p a ) A2 .

87

(8.15)

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The component of the mass force Fms, acting in the direction of flow s, is zero

Fms = 0 ,

(8.16)

because the acceleration due to gravity is perpendicular to the given direction.


The reaction force R expresses the force effect of the fluid flow on the whole jet engine
bounded by the internal control surface CS2. Substituting the rates of the linear momentum
(12.12), the pressure force (8.15), and the mass force (8.16) into the equation of fluid motion
(8.11) and rearranging the terms, we obtain

R = Qm 2 v 2 Qm1 v1 + ( p 2 pa ) A2 ( p1 p a ) A1 .

(8.17)

From the boundary condition imposed upon the flow, the pressure in the inlet crosssection is equal to the atmospheric pressure
p1 = pa ,
(8.18)
then the equation for the jet engine thrust (8.17) is simplified to:

R = Qm 2 v 2 Qm1 v1 + ( p2 pa ) A2 .

(8.19)

Example 8.1
Determine the thrust of the jet engine (Fig. 8.4). The suction mass flow rate of air is 20 kg s-1 and the
fuel consumption is 0.1 kg s-1. The incoming air velocity is 400 m s-1 and the velocity of the outgoing
exhaust gases is 800 m s-1. The outlet cross-sectional area is 0.6 m2 and the pressure in this crosssection is 1.1105 Pa. The atmospheric pressure is 105 Pa.
The mass flow rate of the discharged exhaust gases:

Q m 2 = Q m1 + Q mf = 20 + 0 .1 = 20 .1 kg s 1
The jet engine thrust:
R = Q m 2 v 2 Q m1 v1 + ( p 2 p a ) A2 = 20 .1 800 20 400 + (1 .1 1) 10 5 0 .6 = 14 080 N

8.3.2 Rocket engine thrust


The schematic of the rocket engine system is shown in Fig. 8.6. The burned gases are
discharged from the outlet nozzle of the cross-sectional area A at velocity v, pressure p, and
density . The ambient pressure is atmospheric
pa .
The application of the linear momentum
principle is akin to the one for the jet engine
(8.19). In this case, however, no fluid enters
into the control surface. Then, the equation
(8.19) is simplified for the rocket engine thrust
to:
R = Qm v + ( p p a ) A .
(8.20)

Fig. 8.6 Schematic of rocket engine

Example 8.2
Determine the thrust of the rocket engine (Fig. 8.6). The cross-sectional area of the outlet is 0.1 m2.
The discharged mass flow rate is 2.5 kg s-1, the outgoing velocity is 1200 m s-1, and the pressure in the
outlet cross-sectional area is 1.2105 Pa. The atmospheric pressure is 105 Pa.
Rocket engine thrust:
R = Q m v + ( p p a ) A = 2 .5 1200 + (1 .2 1) 10 5 0 .1 = 5 10 3 N
88

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

8.3.3 Force of the jet acting upon a flat plate at rest


The horizontal jet of fluid of density is leaving the nozzle of cross-sectional area A with
velocity v1 and impinging on a fixed vertical flat plate, as shown in Fig. 8.7. This jet causes
the reaction force R. The external control surface CS1 is taken around the fluid jet and the
internal control surface CS2 excludes the whole plate surface.
The rates of the linear momentum in the s direction of the jet are given by

Pmom1s = Qm v1

and Pmom2 s = 0 ,

(8.21)

because the velocity component in the direction


of the jet is reduced to zero.
As we consider the uniform spreading of
fluid along the plate (Fig. 8.7), from the
boundary condition for pressure

p1 = p 2 = p a = p

(8.22)

and if we neglect the influence of the mass force,


the equation of fluid motion (8.11) is reduced to:

R = Qm v1 ,

(8.23)

where the mass flow rate of the fluid leaving the


nozzle is
Qm = v1 A .
(8.24)

Fig. 8.7 Jet force on flat plate

8.3.4 Force of the jet acting upon the curved vane at rest
If the jet of fluid is impinging on a concave surface of the blade shaped in such a way that
the incoming jet is divided into two streams which are turned to the opposite direction of the
angle (Fig. 8.8), than the outgoing rate of momentum in the direction of the jet is not zero
but given by the component of velocity v2 in the s direction, that is

Pmom2 s = Qm v 2 cos .

(8.25)

From the boundary condition (8.22), the equality


of fluid velocities follows from the Bernoulli equation
v1 = v 2 = v

(8.26)

and upon substituting the rates of momentum (8.21)


and (8.25) into the equation of fluid motion (8.11), we
obtain the jet force acting upon the vane:

R = Qm v (1 + cos ) .

(8.27)

Fig. 8.8 Shape of vane

8.3.5 Force of the jet acting upon the moving plate


If the plate is moving with the velocity u in the direction of the jet (Fig. 8.9), then the
outgoing rate of momentum is not zero. The components of incoming and outgoing rates of
the linear momentum in the direction of the jet are

Pmom1s = Qm v

Pmom2 s = Qm u .

and

(8.28)

Substituting the rates of momentum (8.28) into the equation of fluid motion (8.11), we
obtain the jet force acting upon the moving plate:
89

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

R = Q m (v u ) ,

(8.29)

where the difference v u represents the relative


velocity of the fluid jet with respect to the moving
plate surface. The linear velocity u of the rotating
paddle wheel (Fig. 8.9) is given by the angular
velocity , or revolutions per minute RPM, and the
paddle wheel radius r as
RPM
.
(8.30)
u = r = r 2
60

Fig. 8.9 Jet force on moving plate

The relative mass flow rate of the fluid impinging on the moving surface is similarly
given by the relative velocity, that is

Qm = A (v u ) .

(8.31)

The theoretical power output of the single moving vane, by using (8.29) and (8.31), is
given by
2
(8.32)
P = R u = Q m (v u ) u = A (v u ) u .
The maximum available power with regard to the relative velocity u is given by the local
extremum of the function (8.32), when its derivative is equal to zero, that is
P
v

(8.33)
= A (v 2 4 v u + 3 u 2 ) = 3 A (u v ) u = 0 ,
u
3

from where the condition for the optimal velocities to achieve the maximum power of the
single vane is
v
(8.34)
u opt = .
3

8.3.6 Power of the water turbine wheel


In the case of the Pelton wheel (Fig. 8.10), the curved
turbine vanes are designed to obtain a maximum water
directional change, allowing for a very efficient turbine.
If the direction of the jet is deflected through 180, as
shown in Fig. 8.1I, then the incoming and outgoing rates
of the linear momentum in the direction of the jet are as
follows:

Pmom1s = Qm (v u ) and Pmom2 s = Qm (v u ) . (8.35)

Substituting the rates of momentum (8.35) into the


equation of fluid motion (8.11), we obtain the maximum
theoretical jet force acting upon the moving curved
Pelton turbine vane as
(8.36)
R = 2 Q m (v u ) .

Fig. 8.10 Water turbine wheel

The theoretical power output of the Pelton wheel is


P = R u = 2 Q m (v u ) u = 2 A v (v u ) u ,

(8.37)

where, for the range of vanes, the entire discharged water


mass flow rate from the nozzle is utilized Qm = A v .
90

Fig. 8.11 Pelton vane

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The maximum available power output of the Pelton wheel with regard to the relative
velocity u is given by the local extremum of the function (8.37), when its derivative is equal
to zero, that is
P
(8.38)
= 2 A v2 2u v = 0 ,
u

from where the condition for the maximum power of the Pelton wheel, the optimal wheel
speed, is given by
v
(8.39)
u opt = .
2
Then the maximum power output of the Pelton wheel is given by substituting the optimal
wheel speed (8.39) into equation (8.37) as
P = A

1 3
v .
2

(8.40)

Example 8.3
Determine the reaction force of the water jet leaving the nozzle of the cross-sectional area 0.01 m-2
with the velocity 30 m s-1 and the density 1000 kg m-3 acting upon the moving plate. The paddles are
attached on the wheel (Fig. 8.9) of the diameter 1 m and rotating with the angular velocity 30 rad s-1.
Determine the theoretical power output of the paddle wheel and the torque on the runner.
The linear velocity:

u = R = 0.5 30 = 15 m s 1
Revolutions per second:

RPS =

30
=
= 4.775 s 1
2 2

The mass flow rate from the nozzle:

Q m = v A = 1000 30 0.01 = 300 kg s 1


The jet force acting upon the moving plate:

R = Q m (v u ) = 300 (30 15) = 4500 N


The theoretical power output:

P = R u = 4500 15 = 67 500 W
The torque:

Mt =

67 500
= 2250 N m
30

8.3.7 Force of the fluid flow through a reducing bend


If the fluid flows though the reducing bend sketched in Fig. 8.12, then the hydrodynamic
r
force R acts upon the bend due to the change in the flow direction and velocity.
The fluid of density 1 and pressure p1 is passing through the inlet cross-sectional area A1
r
r
with velocity v1 and is discharging from the outlet cross-sectional area A2 with velocity v2 ,
density 2 and pressure p2. The ambient pressure is atmospheric pa.

91

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

The external control surface CS1 passing


around the entire bend and the internal
control surface CS1 in the tube were chosen,
as shown in Fig. 8.12, to determine the fluid
r
flow force R .
The rates of linear momentum for the
fluid flow incoming and outgoing through
the control surface are given by the definition
(8.4), as follows
r
r
r
r
Pmom1 = Qm v1 and Pmom2 = Qm v 2 , (8.41)
where the mass flow rate is given by

Qm = 1 v1 A1 = 2 v 2 A2 .

(8.42)

The pressure forces acting upon the


control surface by the gauge pressure at the
given cross-sectional areas are as follows
r
r
F p1 = ( p1 p a ) A1
and

Fig. 8.12 Fluid flow force on reducing bend

r
r
F p 2 = ( p 2 p a ) A2 .

The equation of fluid motion (8.9) takes the form:


r
r
r
r
r
r
Pmom 2 Pmom1 = F p1 + F p 2 + Fm R .

(8.43)

(8.44)

If we neglect the influence of the mass force (e.g. horizontal bend), then the resultant fluid
r
flow force R acting upon the reducing bend is
r r
r
r
r
R = Pmom1 Pmom 2 + F p1 + F p 2 .
(8.45)
Upon substituting the rates of linear momentum (8.41), the pressure forces (8.43), and the
mass flow rates (8.16) into equation (8.45) we obtain
r
r
r
r
r
R = 1 A1 v1 v1 2 A2 v 2 v 2 + ( p1 p a ) A1 ( p 2 p a ) A2 .
(8.46)
This vector sum is shown in Fig. 8.12.
Example 8.3
Determine the force of water flow through the horizontal 90
degree elbow (Fig. 8.13) of the cross-sectional area 1 m-2. The
flow velocity is 5 m s-1. The pressure of water is 1.15 105 Pa and
the atmospheric pressure is 105 Pa. Water density is 1000 kg m-3.
The mass flow rate:
Q m = v1 A = 1000 5 1 = 5 000 kg s 1

The rates of linear momentum:


Pmom = Pmom1 = Pmom 2 = Q m v 1 = 5 000 5 = 25 000 N
The pressure forces:

F p = F p1 = F p 2 = ( p p a ) A = (1.15 1) 10 5 1 = 15 000 N

Fig. 8.13 Fluid flow force on elbow

The water flow force:

R=

(P

mom

+ F p ) + (Pmom + F p ) = 2 (Pmom + F p ) = 2 (25 + 15 ) 10 3 = 56.57 10 3 N


2

92

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

English-Czech glossary of basic terms used in hydromechanics


acceleration due to gravity
actual fluid
armature
bend
blade
buoyancy
boundary layer
bottom width, bottom area
branch, branch line
cavitation
centrifugal pump
characteristic curve
characteristic length
clack valve
cock
coefficient of contraction
coefficient of dynamic viscosity
coefficient of kinematic viscosity
coefficient of velocity
continuity equation
contraction
control surface, control volume
compressibility, compressibility coefficient
cross-sectional area
density
depth
diameter
discharge, discharge coefficient
drag coefficient
draught
duct
efficiency
energy loss
elbow, elbow-pipe
fitting
float
flow, flow rate, flow rate coefficient
flowmeter, flow indicator
fluid
fluid velocity
flux
force
free surface
friction, friction coefficient (factor)
gravitational flow

thov zrychlen
skuten (reln) tekutina
armatura
ohyb
lopatka
vztlak, vtlak
mezn vrstva
ka koryta u dna, plocha dna
odboka, vtev potrub
kavitace
odstediv erpadlo
charakteristika
charakteristick dlka, charakteristick rozmr
klapka
kohout
souinitel zen, kontrakce
dynamick viskozita
kinematick viskozita
souinitel vtokov rychlosti
rovnice kontinuity
zen
kontroln plocha, kontroln objem
stlaitelnost, souinitel stlaitelnosti
prton plocha (prez)
hustota
hloubka
prmr
vtok, souinitel vtoku
souinitel odporu
ponor
trubice, kanl
innost
ztrtov energie
koleno
armatura
plavn, plovk, lopatka
proudn, prtok, souinitel prtoku
prtokomr
tekutina
rychlost proudn tekutiny
proud
sla
voln hladina
ten, souinitel tec ztrty
gravitan proudn

93

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

head
head loss
head over weir
height
hydraulic diameter, hydraulic radius
hydraulic press,
hydraulic wall roughness
hydrodynamics
hydromechanics
inlet
jet
junction
law of conservation of energy
law of conservation of mass
law of momentum conservation
layer thickness
level
lift
linear thermal expansion coefficient
liquid
leak
linear momentum
local loss, local loss coefficient
loss due to friction
manometer
mass flow
mean velocity
metacentric height
motion
mouthpiece
nameplate data
negative gauge pressure
nozzle
opening
operating point
orifice, orifice plate
outlet
overflow
paddle
pipe, pipeline, piping
plate
potential head
positive displacement pump
positive gauge pressure
power output
pressure: absolute, ambient, barometric, critical,
hydrostatic, partial, saturated, static

vka, spd
ztrtov vka
vka pepadu
vka
hydraulick prmr, hydraulick polomr
hydraulick lis
hydraulick drsnost
hydrodynamika
hydromechanika
vstup, vstupn
proud, trysk, tryskat
vtven, rozvtven
zkon zachovn energie
zkon zachovn hmotnosti
zkon zachovn pohybu
tlouka vrstvy
hladina
vztlak
souinitel teplotn dlkov roztanosti
kapalina
otvor, netsnost
hybnost
mstn ztrta, souinitel mstnch ztrt
ztrta tenm
manometr
hmotnostn prtok
prezov (stedn) rychlost
metacentrick vka
pohyb
ntrubek
ttkov daje
podtlak
dza
otvor
pracovn bod
otvor, clona
vstup, vstupn
pepad
lopatka
trubka, potrub
deska
polohov vka
objemov erpadlo
petlak
vkon
tlak: absolutn, okoln, barometrick, kritick,
hydrostatick, parciln, nasycen, statick
94

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

pressure gauge
pressure head
probe
pump
rate of linear momentum
reaction force
reservoir
revolutions per minute, per second
roughness , relative roughness
safety valve
saturated vapor
side-contraction
similarity: geometric, kinematic, dynamic
shear stress, shear force
shock wave
sleeve
slide-valve, slider
socket
specific energy, specific energy required
specific heat capacity
specific weight
stability moment
standard
stream, streamline, stream tube
submerged
suction, suction head
surface tension
T-joint, tee
tank
temperature: absolute, ambient, barometric,
critical, dew-point, fluid, saturated, surface
throat
thrust (jet engine, rocket)
top width
tube, tubular
unit energy
valve
vane
velocity
Venturi tube, Venturi nozzle
viscosity, viscous fluid
volume, specific volume
volumetric thermal expansion coefficient
water turbine, water wheel, water power plant
wetted perimeter
width
weir

manometr, manometrick
tlakov vka
sonda, midlo
erpadlo
tok hybnosti
reaktivn sla
ndr
otky za minutu, za sekundu
drsnost, relativn drsnost
pojiovac ventil
nasycen pry
zen
podobnost: geometrick, kinematick, dynamick
smykov napt, smykov sla
rzov vlna
trubice, hrdlo, ntrubek
ouptko
hrdlo, ntrubek
mrn energie, mrn potebn energie
mrn tepeln kapacita
mrn hmotnost, hustota
stabilitn moment
norma
proud, proudnice, proudov trubice
ponoen
sn, sac vka
povrchov napt
T-kus
ndr
teplota: absolutn, okol, barometrick, kritick,
rosnho bodu, tekutiny, nasycen, povrchu
krcen
tah (proudovho motoru, rakety)
ka koryta na hladin
trubice, trubka, trubkov
mrn energie
ventil
lopatka
rychlost
Venturiho trubice, Venturiho dza
viskozita, viskzn tekutina
objem, mrn objem
souinitel teplotn objemov roztanosti
vodn turbna, obn kolo, vodn elektrrna
smoen obvod
ka
pepad
95

Hydromechanics
Vladimr Hork and Vladimir V. Kulish

Literature
[1] engel, Y.A. Cimbala, J.M.: Fluid Mechanics: Fundamentals and Applications.
McGraw-Hill, Boston, 2006.
[2] Chase, M.W. (ed.): Nist Janaf Thermochemical Tables. Monograph, No 9.
National Institute of Standard, 1998
[3] Crowe, C.T. Elger, D.F. Roberson, J.A.: Engineering Fluid Mechanics. John Wiley
and Sons, New York, 2001.
[4] Dickenson, T.Ch.: Valves, Piping, and Pipelines Handbook. Elsevier Science Ltd., New
York, 1999.
[5] Glich J.F.: Centrifugal Pumps. Springer-Verlag, Berlin, Heidelberg, 2008.
[6] Hork, V. Kulish, V.V.: Introduction to Engineering Thermodynamics. University of
Defence, Brno, 2009.
[7] Hork, V.: Zklady hydromechaniky. Univerzita obrany, Brno, 2007.
[8] Hwang, H.C. Houghtalen, R. J.: Fundamentals of Hydraulic Engineering Systems.
Prentice Hall, 1996.
[9] Krivchenko, G.I.: Hydraulic Machines: Turbines and Pumps. Lewis Publishers, 1994.
[10] Matovsk, O.: Hydromechanika. SNTL, Praha,1963.
[11] Maxa, J.: Hydromechanika. VAAZ, Brno, 1988.
[12] Mays, L.W. (ed.): Hydraulic Design Handbook. McGraw-Hill Book, New York, 1999.
[13] Murdock, J.W.: Fundamental Fluid Mechanics for the Practicing Engineer. Marcel
Dekker, New York, 1993.
[14] Munson, B.R. Young, D.F. Okiishi, T.H. Huebsch, W.W.: Fundamentals of Fluid
Mechanics: SI Version. John Wiley and Sons, New York, 2009.
[15] Noskievi, J. a kol.: Mechanika tekutin. SNTL, Praha, 1987.
[16] Reid, R.C. Prausnitz, J.M. Poling, B.E: The Properties of Gases and Liquids.
McGraw-Hill Book Co., Singapore, 1988.
[17] Sahu, G.K.: Handbook of Piping Design. New Age International, New Delhi, 1998.
[18] Saleh, J.: Fluid Flow Handbook. McGraw-Hill Handbooks, New York, 2002.
[19] White, F.M.: Fluid Mechanics. McGraw-Hill, New York, 2003.
[20] Young, D.F. Munson, B.R. Okiishi, T.H. Huebsch, W.W.: A Brief Introduction to
Fluid Mechanics. John Wiley and Sons, New York, 2007.
[21] SN 01 1320 Veliiny, znaky a jednotky v hydromechanice. (Quantities, symbols and
units in hydromechanics)
[22] SN ISO 31-4 Veliiny a jednotky. Teplo. (Quantities and Units. Heat)
[23] SN ISO 80000-4 Veliiny a jednotky. Mechanika. (Quantities and Units. Mechanics)
[24] SN ISO 4006 Men prtoku tekutin v uzavench potrubch (Measurement of fluid
flow in closed conduits)
[25] SN ISO 5167 Men prtoku tekutin pomoc snma diferennho tlaku vloench do
zcela zaplnnho potrub kruhovho prezu. (Measurement of liquid flow by means of
pressure differential devices inserted in circular cross-section conduits running)

96

Potrebbero piacerti anche