You are on page 1of 132

Prof.univ.Dr.ing.

DUMITRU DINU
S.L.drd.ing. STAN LIVIU
HYDRAULICS
AND
HYDRAULIC MACHINES
1
CONTENTS
PART ONE
HYDRAULICS
1. BASIC MATHEMATICS 11
2. FLUID PROPRIETIES 17
2.1 Compressibility 18
2.2 Thermal dilatation 20
2.3 Mobility 22
2.4 Viscosity 22
3. EQUATIONS OF IDEAL FLUID MOTION
29
3.1 Eulers equation 29
3.2 Equation of continuity 32
3.3 The equation of state 34
3.4 Bernoullis equation 35
3.5 Plotting and energetic interpretation of
Bernoullis equation for liquids 39
3.6 Bernoullis equations for the relative
movement of ideal non-compressible fluid
40
4. FLUID STATICS 43
4.1 The fundamental equation of
hydrostatics 43
4.2 Geometric and physical interpretation
2
of the fundamental equation of
hydrostatics 45
4.3 Pascals principle 46
4.4 The principle of communicating
vessels 47
4.5 Hydrostatic forces 48
4.6 Archimedes principle 50
4.7 The floating of bodies 51
5. POTENTIAL (IRROTATIONAL) MOTION
57
5.1 Plane potential motion 59
5.2 Rectilinear and uniform motion 63
5.3 The source 66
5.4 The whirl 69
5.5 The flow with and without circulation
around a circular cylinder71
5.6 Kutta Jukovskis theorem 75
6. IMPULSE AND MOMENT IMPULSE
THEOREM 77
7. MOTION EQUATION OF THE REAL
FLUID81
7.1 Motion regimes of fluids 81
7.2 Navier Stokes equation 83
7.3 Bernoullis equation under the
permanent regime of a thread of real fluid87
7.4 Laminar motion of fluids 90
7.4.1 Velocities distribution between
two plane parallel boards of infinit length
90
7.4.2 Velocity distribution in circular
conduits 93
3
7.5 Turbulent motion of fluids 97
7.5.1 Coefficient in turbulent motion
99
7.5.2 Nikuradzes diagram 102
8. FLOW THROUGH CIRCULAR
CONDUITS 105
9. HYDRODYNAMIC PROFILES
113
9.1 Geometric characteristics of
hydrodynamic profiles 113
9.2 The flow of fluids around wings116
9.3 Forces on the hydrodynamic profiles
119
9.4 Induced resistances in the case of
finite span profiles 123
9.5 Networks profiles 125
4
PART ONE
Hydraulics
5
1. Basic mathematics
The scalar product of two vectors
k a j a i a a
z y x
+ + and k b j b i b b
y x 2
+ + is a
scalar.
Its value is:
z z y y x x
b a b a b a b a + +
. (1.1)
a b
a b
( )

b a cos
. (1.2)
The scalar product is commutative:
a b b a
. (1.3)
The vectorial product of two vectors
a
and
b
is a
vector perpendicular on the plane determined by
those vectors, directed in such a manner that the
trihedral
a
,
b
and
b a
should be rectangular.
z y x
z y x
b b b
a a a
k j i
b a
. (1.4)
The modulus of the vectorial product is given
by the relation:
6
( )

b a b a b a sin
. (1.5)
The vectorial product is non-commutative:
a b b a
(1.6)
The mixed product of three vectors
a
,
b
and
c
is
a scalar.
( )
z y x
z y x
z y x
c c c
b b b
a a a
c b a
. (1.7)
The double vectorial product of three vectors
a
,
b

and
c
is a vector situated in the plane ( ) c b, .
The formula of the double vectorial product:
( ) ( ) ( ) c b a c a b c b a . (1.8)
The operator is defined by:
z
k
y
j
x
i


. (1.9)
applied to a scalar is called gradient.
. grad
k
z
j
y
i
x

. (1.10)
scalary applied to a vector is called
divarication. . a div a
7
z
a
y
a
x
a
a
z
y
x

. (1.11)
vectorially applied to a vector is called
rotor.
. a rot a
z y x
a a a
z y x
k j i
a


. (1.12)
Operations with :
( ) + + . (1.13)
( ) b a b a + + . (1.14)
( ) b a b a + + . (1.15)
When acts upon a product:
- in the first place has differential and only
then vectorial proprieties;
- all the vectors or the scalars upon which it
doesnt act must, in the end, be placed in
front of the operator;
- it mustnt be placed alone at the end.
( ) ( ) ( ) + +
c c
. (1.16)
( ) ( ) ( ) + + a a a a a c
c
. (1.17)
( ) ( ) ( ) + a a a a a c
c
. (1.18)
8
( ) ( ) ( ) c c b a b a b a + , (1.19)
( ) ( ) ( ) b a b a b a c c , (1.20)
( ) ( ) b a b rot a b ac + , (1.21)
( ) ( ) a b a rot b b a c + , (1.22)
( ) ( ) ( ) a b a rot b b a b rot a b a + + + . (1.23)
c

- the scalar

considered constant,
c

- the scalar

considered constant,
c a
- the vector
a
considered constant,
c b
- the vector
b
considered constant.
If:
, v b a (1.24)
then:
( ) v rot v v v
v
+

,
_

2
2
. (1.25)
The streamline is a curve tangent in each of
its points to the velocity vector of the
corresponding point ( ) k v j v i v v
z y x
+ + .
The equation of the streamline is obtained by
writing that the tangent to streamline is parallel to
the vector velocity in its corresponding point:
9
z y x
v
dz
v
dy
v
dx

. (1.26)
The whirl line is a curve tangent in each of its
points to the whirl vector of the corresponding
point ( ) k j i
z y x
+ + .
v rot
2
1
. (1.27)
The equation of the whirl line is obtained by
writing that the tangent to whirl line is parallel with
the vector whirl in its corresponding point:
z y x
dz dy dx


. (1.28)
Gauss-Ostrogradskis relation:


d a d n a


, (1.29)
where

- volume delimited by surface

.
The circulation of velocity on a curve (C) is
defined by:

, r d v
C

(1.30)
in which
ds r d (1.31)
represents the orientated element of the
curve (

- the versor of the tangent to the curve


(C )).
10
Fig.1.1
( )

+ +
C
z y x
dz v dy v dx v
(1.32)
The sense of circulation depends on the
admitted sense in covering the curve.
ABMA AMBA
. (1.33)
Also:
BA AMB AMBA
+ . (1.34)
Stokes relation:
( )



C
d n v rot r d v
(1.35)
in which
n
represents the versor of the normal
to the arbitrary surface

bordered by the curve


(C).
11
2. FLUID PROPRIETIES
As it is known, matter and therefore fluid
bodies as well, has a discrete and discontinuous
structure, being made up of micro-particles
(molecules, atoms, etc) that are in reciprocal
interaction.
The mechanics of fluids studies phenomena
that take place at a macroscopic scale, the scale at
which fluids behave as if matter were continuously
distributed.
At the same time, fluids dont have their own
shape so are easily deformed.
A continuous medium is homogenous if at a
constant temperature and pressure, its density has
only one value in all its points.
Lastly, a continuous homogenous medium is
isotropic as well if it has the same proprieties in
any direction around a certain point of its mass.
In what follows we shall consider the fluid as a
continuous, deforming, homogeneous and isotropic
medium.
We shall analyse some of basic physical
proprieties of the fluids.
12
2.1. Compressibility
Compressibility represents the property of
fluids to modify their volume under the action of a
variation of pressure. To evaluate quantitatively
this property we use a physical value, called
isothermal compressibility coefficient,

, that is
defined by the relation:
,
1
2
1
]
1


N
m
dp
dV
V

(2.1)
in which dV represents the elementary variation of
the initial volume, under the action of pressure
variation dp.
The coefficient

is intrinsic positive; the


minus sign that appears in relation (2.1) takes into
consideration the fact that the volume and the
pressure have reverse variations, namely dv/ dp <
0.
The reverse of the isothermal compressibility
coefficient is called the elasticity modulus K and is
given by the relation:
.
1
2 1
]
1


m
N
dV
dp
V K

(2.2)
Writing the relation (2.2) in the form:
,
K
dp
V
dV
(2.3)
we can underline its analogy with Hooks law:
13
.
E l
dl
(2.4)
a) The compressibility of liquids
In the case of liquids, it has been
experimentally ascertained that the elasticity
modulus K, and implicitly, the coefficient

, vary
very little with respect to temperature (with
approximately 10% in the interval C
0
60 0 ) and they
are constant for variations of pressure within
enough wide limits. In table (2.1) there are shown
the values of these coefficients for various liquids
at C
0
0 and pressure
200 p
bars.
Table 2.1.
Liquid
[ ] N m /
2

[ ]
2
/ m N
K
Water
10
10 12 , 5

9
10 95 , 1
Petrol
10
10 66 , 8

9
10 15 , 1
Glycerine
10
10 55 , 2

9
10 92 , 3
Mercury
10
10 296 , 0

9
10 7 , 33
Therefore, in the case of liquids, coefficient


may be considered constant.
Consequently, we can integrate the
differential equation (2.2) from an initial state,
characterised by volume
0
V
, pressure
0
p
and density
0

, to a certain final state, where the state


parameters will have the value p V ,
1
and


respectively; we shall successively get:
14


V
V
p
p
dp
V
dV
0 0
,
(2.5)
or
( )
.
0
0
p p
e V V



(2.6)
b) The compressibility of gases
For gases the isothermal compressibility
coefficient depends very much on pressure. In the
case of a perfect gas, the following relation
describes the isothermal compressibility:
pV = cons.,
which, by subtraction, will be:
.
V
dV
p
dp

(2.8)
By comparing this relation to (2.3) we may
write:
.
1
p K

(2.9)
It follows that, in the case of a perfect gas,
the elasticity modulus is equal to pressure.
2.2 Thermal dilatation
15
Thermal dilatation represents the fluid
property to modify its volume under the action of a
variation of temperature. Qualitatively, this
property is characterised by the volumetric
coefficient of isobaric dilatation, defined by the
relation:
,
1
dT
dV
V
(2.10)
where dV represents the elementary variation of
the initial volume V under the action of variation of
temperature dT. Coefficient

is positive for all


fluids, except for water, which registers maximum
density (minimum specific volume) at C
0
4 ;
therefore, for water that has C t
0
4 we shall have
. 0 <
Generally,

varies very little with respect to


temperature, therefore it can be considered
constant. Under these circumstances, integrating
the equation (2.10) between the limits
0
V
and V,
and respectively
0
T
and T, we get:
( ), ln
0
0
T T
V
V

(2.11)
or else
( )
.
0
0
T T
e V V


(2.12)
By dividing the relation (2.12) to the mass of
the fluid
,
0 0
V V m
we get the function of state
for an incompressible fluid:
( )
,
0
0
T T
e


(2.13)
16
In the case of a perfect gas the value of the
coefficient is obtained by subtracting the equation
of isobaric transformation
,
_

. cons
T
V
; we get:
, . dT
T
V
dT cons dV (2.14)
which, replaced into (2.10) enables us to write:
.
1
T
(2.15)
Thus, for the perfect gas, coefficient

is the
reverse of the thermodynamic temperature.
2.3. Mobility
In the case of fluids, the molecular cohesion
forces have very low values, but they arent
rigorously nil.
At a macroscopic scale, this propriety can be
rendered by the fact that two particles of fluid that
are in contact, can be separated under the action of
some very small external forces. At the same time,
fluid particles can slide one near the other and
have to overcome some relatively small tangent
efforts.
As a result, from a practical point of view,
fluids can develop only compression efforts.
In the case of a deformation at a constant
volume, the compression efforts are rigorously nil
and, as a result, the change in shape of the fluid
requires the overcoming of the tangent efforts,
which are very small. Therefore the mechanical
17
work consumed from the exterior will be very small,
in fact negligible.
We say that fluids have a high mobility,
meaning that they have the property to take the
shape of the containers in which they are.
Consequently we should stress that gases, because
they dont have their own volume, have a higher
mobility than liquids (a gas inserted in a container
takes both the shape and the volume of that
container).
2.4 Viscosity
Viscosity is the property of the fluid to oppose
to the relative movement of its particles.
As it has been shown, overcoming some small
tangent efforts that arent yet rigorously nil makes
this movement.
To qualitatively stress these efforts, we
consider the unidimenssional flow of a liquid, which
takes place in superposed layers, along a board
situated in xOy plane (fig.2.1).
Fig.2.1.
Experimental measurements have shown that
velocity increases as we move away from the board
in the direction of axis Oy, and it is nil in the near
vicinity of the board. Graphically, the dependent
18
( ) y f v is represented by the curve . This simple
experiment stresses on two aspects, namely:
- the fluid adheres on the surface of the solid
body with which it comes into contact;
- inside the fluid and at its contact with the
solid surfaces, tangent efforts generate
which determine variation in velocity. Thus,
considering two layers of fluid, parallel to
the plane xOy and that are at an
elementary distance dy one from the other,
we shall register a variation in velocity
dy
dy
dv
, due to the frictions that arise between
the two layers.
To determine the friction stress, Newton used
the relation:
dy
dv

, (2.16)
that today bears his name. This relation that has
been experimentally verified by Coulomb,
Poisseuille and Petrov shows that the friction
stress

is proportional to the gradient of velocity.


The proportionality factor

is called dynamic
viscosity.
If we represent graphically the dependent
( ) dy dv f / we shall get the line 1 (fig.2.2) where
ty
.
The fluids that observe the
friction law (2.16) are called
Newtonian fluids (water, air,
etc). The dependent of the
19
tangent effort to the gradient
of velocity is not a straight line
(for example curve (2) in fig.
2.2), for a series of other
fluids, generally those of
organic nature. These fluids
are globally called non-
Newtonian fluids.
Fig.2.2
The measures for the dynamic viscosity are:
- in the international standard (SI):
[ ]
s m
Kg
m
s N

(2.17)
- in the CGS system:
[ ]
s cm
g
cm
s dyn

. (2.18)
The measure of dynamic viscosity in CGS
system is called poise, and has the symbol P. We
can notice the existence of relation:
P
s m
Kg
10 1

. (2.19)
We can determine the dynamic viscosity of
liquids with the help of Hpplers viscometer,
whose working principle is based on the
proportionality of dynamic viscosity to the time in
which a ball falls inside a slanting tube that
contains the analysed liquid.
The kinematic viscosity of a fluid is the ratio
of dynamic viscosity and its density:
20


. (2.20)
The measures for kinematic viscosity are:
- in the international system:
[ ]
s
m
2
. (2.21)
- in CGS system:
[ ]
s
cm
2
. (2.22)
the latter bearing the name stokes (symbol ST):
s
m
s
cm
ST
2
4
2
10 1 1

. (2.23)
Irrespective of the type of viscometer used
(Ubbelohde, Vogel-Ossag, etc) we can determine
the kinematic viscosity by multiplying the time
(expressed in seconds) in which a fixed volume of
liquid flows through a calibrated capillary tube,
under normal conditions, constant for that device.
In actual practice, the conventional viscosity
of a fluid is often used; this value is determined by
measuring the time in which a certain volume of
fluids flows through a special device, the conditions
being conventionally chosen. The magnitude of this
value thus determined is expressed in conventional
units. There are several conventional viscosities
(i.e. Engler, Saybolt, Redwood etc) which differ
21
from one another both in the measurement
conditions and in the measure units.
Thus, Engler conventional viscosity, expressed
in Engler degrees [ ] E
0
is the ratio between the flow
time of 200 cubic cm of the analysed liquid at a
given temperature and the flow time of a same
volume of distilled water at a temperature of C
0
20 ,
through an Engler viscometer under standard
conditions.
The viscosity of a fluid depends to a great
extent on its temperature. Generally, viscosity of
liquids diminishes with the increase in
temperature, while for gas the reverse
phenomenon takes place.
The dependence of liquids viscosity on
temperature can be determined by using Gutman
and Simons relation:
0
0
T
B
T C
B
e

+

. (2.24)
where the constants B and C depend on the nature
of the analysed liquid (for water we have B= 511,6
K and C= -149,4 K).
For gases we can use Sutherlands formula
T S
T S
T
T
+
+

,
_

0
2 / 3
0
0
. (2.25)
where S depends on the nature of the gas (for air
S=123,6 K).
In relations (2.24) and (2.25),

and
0

are the
dynamic viscosities of the fluid at the absolute
22
temperature T, and at temperature ) 0 ( 15 , 273
0
0
C K T
respectively.
In table 2.2 there are shown the dynamic and
kinematic viscosities of air and water at different
temperatures and under normal atmospheric
pressures.
Table 2.2
Temperature
[ ] C
0
-10 0 10 20 40 60 80 10
0
1
]
1

s m
Kg
9
10

Air 0,16
2
0,1
72
0,17
5
0,18
1
0,1
91
0,2
0
0,2
89
0,2
18
Wat
er
- 1,7
9
1,31 1,01 0,6
58
0,4
78
0,3
66
0,2
95
1
]
1

s
m
2
6
10

Air 1,26 13,


3
14,1 15,1 16,
9
18,
9
20,
9
23,
1
Wat
er
- 1,7
9
1,31 1,01 0,6
58
0,4
78
0,3
66
0,2
95
We must underline the fact that viscosity is a
property that becomes manifest only during the
movement of liquids.
A fluid for which viscosity is rigorously nil is
called a perfect or ideal fluid.
Fluids may be compressible ( ) [ ] p or
incompressible (

is constant with respect to


pressure).
We should emphasise that the ideal
compressible fluid is analogous to the ideal (or
perfect) gas, as defined in thermodynamics.
23
The movement of fluids may be uniform
(velocity is constant), permanent v = v (x,y,z) or
varied v = v (x,y,z,t).
24
3. EQUATIONS OF IDEAL FLUID
MOTION
3.1 Eulers equation
We shall further study, for the most general
case, the movement state of a fluid through a
volume

that is situated in the fluid stream; we


shall not take into consideration the interior
frictions(i.e.viscosity), so we shall analyse the case
of perfect (ideal) fluids that are on varied
movement.
The volume

is situated in an accelerated
system of axes, joint with this system. The
equations, which describe the movement of the
fluid, will be obtained by applying dAlemberts
principle for the fluid that is moving through the
volume

.
The three categories of forces that act upon
the fluid that is moving through the volume

,bordered by the surface

(fig.3.1), are:
Fig.3.1
- the mass forces
m
F ;
25
- the inertia forces
i
F ;
- the pressure forces
p
F (with an equivalent
effect; these forces replace the action of
the negligible fluid outside volume

).
According to dAlemberts principle, we shall
get:
0 + +
p i m
F F F . (3.1)
Equation (3.1) represents in fact the general
vectorial form of Eulers equations.
Lets establish the mathematical expressions
of those three categories of forces.
If
F
is the mass unitary force (acceleration)
that acts upon the fluid in the volume

, the mass
elementary force that acts upon the mass
d
, will
be:
d F F d
m
, (3.2)
hence:

d F F
m
. (3.3)
As the fluid velocity through the volume

is a
vectorial function with respect to point and time:
( ) t r v v , , upon the mass
d
that is moving with
velocity
v
the elementary inertia will act:
d
dt
v d
F d
i
. (3.4)
26
So, the inertia will be:

d
dt
v d
F
i
. (3.5)
If d is a surface element upon which the
pressure p acts, and
n
- the versor of the exterior
normal (Fig.3.1), the elementary force of pressure
is:
d n p F d
p
. (3.6)
Having in mind Gauss-Ostrogradskis theorem,
the resultant of pressure forces will be:


d p d n p F
p


. (3.7)
By replacing equations (3.3), (3.5) and (3.7) in
the equation (3.1), we shall get:
0

,
_

d
dt
v d
p F
, (3.8)
Hence:
dt
v d
p F

1
, (3.9)
Or
( ) v v
t
v
p F +

1
, (3.10)
27
The equation (3.10) Eulers equation in a
vectorial form for the ideal fluid in a non-
permanent movement.
Projecting this equation on the three axes, we
shall obtain:
z
x
y
x
x
x x
x
v
z
v
v
y
v
v
x
v
t
v
x
p
F

1
;
z
y
y
y
x
y y
y
v
z
v
v
y
v
v
x
v
t
v
y
p
F

1
; (3.11)
z
z
y
z
x
z z
z
v
z
v
v
y
v
v
x
v
t
v
z
p
F

1
.
3.2 Equation of continuity
This equation can be obtained by writing in
two ways the variation in the unity of time for the
mass of fluid that is in the control volume

,
bordered by the surface

(fig.3.1). By splitting
from the volume

one element d , and taking into


consideration that the density is a scalar function
of point and time, ( ) t r, , we can write the total
mass of the volume

d m
. (3.12)
The variation of the total mass in the unity of
time will be:
28

d
t t
m
. (3.13)
The second form of writing the variation of
mass is obtained by examining the flow of the mass
through surface

that borders volume

.
Denoting by
n
the versor of the exterior
normal to the area element d , and by
v
the vector
of the fluid velocity, the elementary mass of fluid
that passes in the unity of time through the area
element d is:
d v dM
n

. (3.14)
In the unity of time through the whole surface

will pass, the mass:

d v M
n
(3.15)
that is the sum of the inlet and outlet mass in
volume

, by crossing surface

.
By equalling equations (3.13) and (3.15), it
will result:

+

0 d v
t
n . (3.16)
According to Gauss-Ostrogradskis theorem:
( )



d v d v
n
. (3.17)
Taking into consideration (3.17), the equation
(3.16) will take the form:
29
( ) 0
1
]
1

d v
t
, (3.18)
hence, successively:
( ) , 0 +

v
t

(3.19)
0 + +

v v
t

, (3.20)
0 + v
dt
d

. (3.21)
The equation (3.21) represents the equation
of continuity for compressible fluids.
In the case of non-compressible fluids (
. cons
, 0
dt
d
), the equation of continuity takes the form:
0 v
, (3.22)
or
0

z
v
y
v
x
v
z
y
x
. (3.23)
It follows that the inlet volume of non-
compressible liquid is equal to the outlet one in and
from the volume

.
3.3. The equation of state
30
From a thermodynamically point of view, the
state of a system can be determined by the direct
measurement of some characteristic physical
values, that make up the group of state parameters
(e.g. pressure, volume, temperature, density etc.).
Among the state parameters of a
thermodynamically system generally there are link
relationships, explained by the laws of physics.
In the case of homogenous systems, there is
only one implicit relationship, which carries out the
link among the three state parameters, in the form
of:
( ) 0 , , T p F . (3.24)
Adding to vectorial equations (3.10) and
(3.21) the equation of state, we get three
equations with three unknowns: ( ) ( ) ( ) t r p t r t r v , , , , , ,
that enable us solve the problems of motion and
repose for the ideal fluids.
3.4. Bernoulli s equation
Bernoullis equation is obtained by integrating
Eulers equation written under a different form
(Euler Lamb), that stresses the rotational or non-
rotational nature of the ideal fluid (see the relation
(1.25)).
Euler Lambs equation:
31
v rot v
v
t
v
p F

,
_


2
1
2

. (3.25)
Considering the case when the mass force
derives from a potential U, thus being a
conservative force (the mechanical energy-kinetic
and potential-will be constant):
U F
. (3.26)
In the case of compressible fluids, when
( ) p , we insert the function:
( )

p
dp
P

. (3.27)
Thus:
( )
p
p
P

1
. (3.28)
The equation (3.25) takes the form:
v rot v
t
v v
P U

,
_

+ +
2
2
. (3.29)
The equation (3.29) can be easily integrated
in certain particular cases.
In the case of permanent motion 0

t
v
, and:
- along a stream line:
z y x
v
dz
v
dy
v
dx

, (3.30)
32
- along a whirl line:
z y x
dz dy dx


, (3.31)
- in the case of potential motion 0 v rot :
0
z y x

, (3.32)
-in the case of helicoid motion (the velocity
vector
v
is parallel to the whirl vector):
z
z
y
y
x
x
v
v
v


. (3.33)
Multiplying by
r d
the equation (3.29), we shall
get under the conditions of permanent motion (
0

t
v
):
( ) v rot v r d
v
P U d

,
_

+ +
2
2
. (3.34)
Since 2 v rot , we shall get:
z y x
z y x
v v v
dz dy dx
v
P U d

2
2
2

,
_

+ +
. (3.35)
The determined is zero for one of the four
cases above. By integrating in these cases we shall
get Bernoullis equation:
C
v
P U + +
2
2
. (3.36)
33
If the fluid is a non-compressible one, then

p
P
.
If the axis Oz of the system is vertical,
upwards directed, the potential U is:
C gz U +
. (3.37)
It results the well known Bernoullis equation
as the load equation:
C z
p
g
v
+ +
2
2
. (3.38)
The kinetic load
g
v
2
2
represents the height at
which it would rise in vacuum a material point,
vertically and upwards thrown, with an initial
velocity v, equal to the velocity of the particle of
liquid considered.
The piezometric load

p
is the height of the
column of liquid corresponding to the pressure p of
the particle of liquid.
The position load z represents the height at
which the particle is with respect to an arbitrary
chosen reference plane.
Bernoullis equation, as an equation of loads,
may be expressed as follows: in the permanent
regime of an ideal fluid, non-compressible,
subjected to the action of some conservative
forces, the sum of the kinetic, piezometric and
position loads remains constant along a streamline.
34
Multiplying (3.38) by

we get the equation


of pressures:
C z p
v
+ +
2
2
, (3.39)
where:
2
2
v
dynamic pressure;
p
piezometric (static) pressure;
z
position pressure.
Multiplying (3.38) by the weight of the fluid G,
we get the equation of energies:
C z G
p
G
g
v
G + +
2
2
, (3.40)
where:
g
v
G
2
2
- kinetic energy;

p
G
- pressure energy;
Gz - position energy.
3.5. Plotting and energetic
interpretation of Bernoullis equation for
liquids
Going back to the relation (3.38) and
considering C = H (fig.3.2):
35
H z
p
g
v
+ +
2
2
. (3.41)
Fig.3.2
The sum of all the terms of Bernoullis
equation represents the total energy (potential and
kinetic) with respect to the unit of weight of the
moving liquid.
This energy measured to a horizontal
reference plane N-N, arbitrarily chosen is called
specific energy and it remains constant during the
permanent movement of the ideal non-
compressible fluid that is under the action of
gravitational and pressure forces.
3.6. Bernoullis equation for the
relative movement of ideal non-
compressible fluid
Lets consider the flow of an ideal non-
compressible fluid through the channel between
two concentric pipes that revolve around an axis Oz
with angular velocity

(fig.3.3.).
36
Fig.3.3
In the equation (3.38) v is replaced by w,
which represents the relative velocity of the liquid
to the channel that is revolving with the velocity
r u
.
Upon the liquid besides the gravitational
acceleration g, the acceleration r
2
acts as well.
The unitary mass forces decomposed on the
three axes will be:
.
;
;
2
2
g F
y F
x F
z
y
x

(3.42)
In this case, the potential U will be:
C
r
gz U +
2
2 2

. (3.43)
By adding (3.43) to Bernoullis equation, we
get:
37
C z
p
g
r
g
w
+ +

2 2
2 2 2
, (3.44)
or
C z
p
g
u w
+ +

2
2 2
. (3.45)
In the theory of hydraulic machines we use
the following denotations:
v absolute velocity;
w relative velocity;
u peripheral velocity.
The equation (3.45) written for two particles
on the same streamline is:

2
2
2
2
2
2
1
1
2
1
2
1
2 2
z
p
g
u w
z
p u w
+ +

+ +


(3.46)
38
4. FLUID STATICS
The fluid statics hydrostatics is that part of
the mechanics of fluid which studies the repose
conditions of the fluid as well as their action,
during the repose state, on solid bodies with whom
they come into contact.
Hydrostatics is identical for real and ideal
fluids, as viscosity becomes manifest only during
motion. In hydrostatics the notion of time does no
longer exist.
4.1 The fundamental equation of
hydrostatics
If in Eulers equation (3.9) we assume that
0 v
, we get:
0
1
p F

. (4.1)
We multiply everywhere by
r d
:
0
1
r d p r d F

. (4.2)
39
or

dp
dz F dy F dx F
z y x
+ +
. (4.3)
If the axis Oz of the system
xOyx
is vertical,
upwards directed, then:
0
y x
F F
, , g F
z

and equation (4.3) becomes:
0 +

dp
gdz
. (4.4)
In the case of liquids (

= cons.), by
integrating equation (4.4) we get:
. const
p
gz +

(4.5)
or
. const
p
z +

(4.6)
or
. const z p +
(4.7)
Equation (4.7) is called the fundamental
equation of hydrostatics.
40
If
0
p
is the pressure at the surface of water (in
open tank the atmospheric pressure), pressure p,
situated at a distance h from the surface, will be
(fig.4.1):
Fig.4.1
1 0 2
z p z p + +
, (4.8)
h p p +
0
. (4.9)
p is called the absolute pressure in the point
2, and
h
is the relative pressure.
4.2 Geometrical and physical
interpretation of the fundamental equation
of hydrostatics (fig.4.2)
Fig.4.2
41
According to (4.6) we can write:
2
2
2
1
1
1
z
p
z
p
+ +

. (4.10)
In fig.4.2 we have:

p
- piezometric height corresponding to the
absolute hydrostatic pressure;
2 , 1
z
- the quotes to an arbitrary plane (position
heights).
4.3 Pascals principle
We rewrite the fundamental equation of
hydrostatics between two points 1 and 2.
2 2 1 1
z p z p + + . (4.11)
Supposing that in point 1, the pressure
registers a variation
p
, it becomes p p +
1
. In order
that the equilibrium state shouldnt be altered, for
point 2 the same variation of pressure has to be
registered.
2 2 2 1 1 1
z p p z p p + + + + . (4.12)
Hence:
2 1
p p . (4.13)
42
Pascals principle:
Any pressure variation created in a certain
point in a non-compressible liquid in equilibrium, is
transmitted with the same intensity to each point
in the mass of this liquid.
4.4 The principle of communicating
vessels
Let us consider two communicating vessels
(fig.4.3) that contain two non-miscible liquids,
which have specific weights
1
and
2
, respectively.
Writing the equality of pressure in the points 1 and
2, situated in the same horizontal plane N N that
also contains the separation surface between the
two liquids, we get:
2 2 0 1 1 0
h p h p + +
, (4.14)
or else
1
2
2
1

h
h
, (4.15)
where
1
h and
2
h are the heights of the two liquid
columns that, according to this relation, are in
reverse proportion to the specific weights of the
two liquids.
43
Fig.4.3
If ,
2 1

then
2 1
h h .
In two or more communicating vessels, that
contain the same liquid (homogenous and non-
compressible), their free surfaces are on the same
horizontal plane.
4.5 Hydrostatic forces
The pressure force that acts upon a solid wall
is determined by means of the relation:

A
dA n p F
, (4.16)
where dA is a surface element having the versor
n
,
and p is the relative pressure of the fluid.
Let A be a vertical plane surface that limits a
non-compressible fluid, with specific weight


(fig.4.4).
44
Fig.4.4
45
Then the hydrostatic pressure force will be:


A
y
M A z zdA F
0
, (4.17)
where:
0
z
- the quote of the specific weight for
surface A;
y
M
- the static moment of the surface A with
respect to the axis Oy.
The application point of the pressure force F
is called pressure centre. It has the following co-
ordinates:
y
y
A
M
I
zdA
dA z
F
zdF


2
, (4.18)
y
yz
A
M
I
zdA
yzdA
F
ydF


.
y
I
- the inertia moment of surface A with
respect to the axis Oy;
yz
I
- the centrifugal moment of surface A with
respect to axes Oy and Oz.
The hydrostatic pressure force that acts
upon the bottom of a container does not depend on
the quantity of liquid, but on the height of the
liquid and on the section of the bottom of this
container.
46
The above statement represents the
hydrostatic paradox and is illustrated in fig.4.5. The
force that presses on the bottom of the three
different shaped containers, is the same because
the level of the liquid in the container is the same,
and the surface of the bottom is the same.
Fig. 4.5
4.6 Archimedes principle
Lets consider a solid body and further to
simplify a cylinder, submerged in a liquid; we
intend to compute the resultant of the pressure
forces that act upon it (fig.4.6).
Fig.4.6
The resultant of the horizontal forces
'
x
F and
' '
x
F is obviously nil:
47
.
,
0
' '
0
'
x x
x x
A z F
A z F

(4.19)
The vertical forces will have the value:
.
;
2
' '
1
'
z z
z z
A z F
A z F

(4.20)
Thus their resultant will be:
( ) V h A z z A F F F
z z z z z
+
1 2
' ' '
. (4.21)
This demonstration may easily be extended
for a body of any shape.
An object submerged in a liquid is up thrust
with an equal force with the weight of the
displaced liquid.
4.7. The floating of bodies
A free body, partially submerged in a liquid is
called a floating body.
The submerged part is called immerse part or
hull.
The weight centre of the hulls volume is
called the hull centre.
The free surface of the liquid is called floating
plane.
The crossing between the floating plane and
the floating body is called the floating surface.
48
Its weight centre is called floating centre, and
its outline is called floating line or water line.
In order that the floating body be in
equilibrium, it is necessary that the sum of the
forces that act upon it as well as the resultant
moment should be nil.
Upon a floating body there can act two forces:
the archimedean force and the weight force also
called displacement (D = mg) (fig.4.7)
Fig.4.7
As a result, a first condition to achieve the
equilibrium is:
V mg D
, (4.22)
where m is the mass of the floating body, V is the
volume of the hull, and

is the specific weight of


the liquid.
Furthermore, in order that the moment of the
resultant should be nil these two forces must have
the same straight line as support or, in other
words, that the weight centre G should be found on
the same vertical with the centre hull.
Equation (4.22) is called the equation of
flotability.
49
Stability is the ability of the floating body to
return on the initial floating of equilibrium after the
action of perturbatory forces that drew it out of
that position has ceased.
With respect to a Cartesian system of axes
Oxyz, having the plane xOy in the floating plane
and axis Oz upwards directed (fig.4.8), the floating
body has six degrees of freedom: three translations
and three rotations. The rotation around Ox and Oy
is most important.
These slantings are due to the actions of
waves or wind.
By definition, the rotation of the floating body
thus produced as the volume of the hull to remain
unchanged as a value but which can vary in shape
is called isohull slanting.
Let
0 0
L L
be the plane of the initial floating.
After the slanting of the isohull around a certain
axis, the floating body will be on a floating
1 1
L L .
If initially the centre of hull were situated in
the point
0
C
after the isohull slating with an angle

, the centre of hull would move


50
further, in the sense of slanting, to a point
1
C .
This movement takes place due to the
alteration of the shape of the hull volume.
The locus of the successive positions of the
centre of the hull for different isohull slantings
around the same axis is called the curve of the
centre of hull (trajectory C).
The curvature centre of the curve of the hull
centres is called metacentre and its curvature
radius is called metacentric radius.
For transversal slantings around the
longitudinal axis Ox we shall talk about a
transversal metacentre M and about a transversal
metacentric radius r (fig.4.8 a).
Fig.4.8 a, b
For longitudinal slantings around the
transversal axis Oy the longitudinal metacentre
will be denoted by

, and the corresponding


metacentric radius will be R (fig. 4.8 b).
51
Causing a transversal slanting to the floating
body, isohull, with a small angle,

, the centre of
hull will move to point
1
C (fig.4.8 a). In this case,
the force of flotability
V
, normal on the slanting
flotability
1 1
L L , having as application point the
point
1
C wont have the same support as the weight
(displacement) of the floating body.
As a result, the two forces will make up a
couple whose moment,
r
M , will be given by the
relation:
sin h D M
r
, (4.23)
where
a r h . (4.24)
is called metacentric height, and a is the distance
on the vertical between the weight centre and
centre of hull; denoting by
G
z
and
C
z
the quotes of
these points to a horizontal reference plane, we
shall have:
C G
z z a
. (4.25)
The metacentric height, expressed by the
relation (4.24) may be positive, negative or nil. We
shall in turn analyse each of these cases.
a) if h > 0 the metacentre will be above the weight
centre, and the moment
r
M , given by the
relation (4.24) will also be positive. From
fig.4.8.it can be noticed that, in this case, the
moment
r
M will tend to return the floating body
52
to the initial floating
0
L
; for this reason it is
called restoring moment. In this case the floating
of the body will be stable.
b) if h < 0, the metacentre is below the centre of
weight (fig.4.9 a). It can be noticed that, in this
case, the moment
r
M will be negative and will
slant the floating body even further. As a result,
it will be called moment of force tending to
capsize, the floating of the body being unstable.
c) If h = 0, the metacentre and the centre of hull
will superpose (fig.4.9 b). Consequently, the
restoring moment will be nil, and the body will
float in equilibrium on the slanting floating.
Fig.4.9 a, b
In this case the floating is also unstable. Thus,
the stability conditions of the floating are: the
metacentre should be placed above the weight
centre, namely
. 0 > a r h (4.26)
53
According to (4.24) and (4.23), we may write:
( )
g f r
M M a D r D a r D M + sin sin sin
, (4.27)
where:

sin r D M
f

, (4.28)
is called stability moment of form, and:
sin a D M
g

, (4.29)
is called stability moment of weight.
As a result, on the basis of (4.27) we can
consider the restoring moment as an algebraic sum
of these two moments.
In the case of small longitudinal slantings, the
above stated considerations are also valid, the
restoring moment being in this case:
( ) sin sin a R D H D M
r
, (4.30)
where
a R H . (4.31)
represents the longitudinal metacentric height, and
R is the longitudinal metacentric radius.
54
5. POTENTIAL (IRROTATIONAL)
MOTION
The potential motion is characterised by the
fact that the whirl vector is nil.
0
2
1
v rot , (5.1)
hence its name: irrotational.
If

is nil, its components on the three axes


will also be nil:
. 0
2
1
, 0
2
1
, 0
2
1

,
_


,
_

,
_

y
v
x
v
x
v
z
v
z
v
y
v
x
y
z
z x
y
y
z
x

(5.2)
55
or:
.
,
,
y
v
x
v
x
v
z
v
z
v
y
v
x
y
z x
y
z

(5.3)
Relations (5.3) are satisfied only if velocity v
derives from a function

:
. , ,
z
v
y
v
x
v
z y x


(5.4)
or vectorially:
v . (5.5)
Indeed:
( ) 0 grad rot v rot . (5.6)
Function ( ) t z y x , , , is called the potential of
velocities.
If we apply the equation of continuity for
liquids,
0
2
2
2
2
2
2

z y x z
v
y
v
x
v
z
y
x

, (5.7)
we shall notice that function

verifies equation of
Laplace:
0
, (5.8)
thus being a harmonic function.
56
5.1 Plane potential motion
The motion of the fluid is called plane or
bidimensional if all the particles that are found on
the same perpendicular at an immobile plane,
called director plane, move parallel with this plane,
with equal velocities.
If the director plane coincides with xOy, then
0
z
v .
A plane motion becomes unidimensional if
components
x
v
and
y
v
of the velocity of the fluid
depend only on a spatial co-ordinate.
For plane motion, the equation of the
streamline will be:
y x
v
dy
v
dx

, (5.9)
or else:
0 dx v dy v
y x
, (5.10)
and the equation of continuity:
0

y
v
x
v
y
x
. (5.11)
The left term of the equation (5.10) is an
exact total differential of function

, called the
stream function:
57
x
v
y
v
y x


,
, (5.12)
0 dx v dy v d
y x

. (5.13)
Function

verifies the equation of continuity


(5.11):
0
2 2

x y y x y
v
x
v
y
x

. (5.14)
Function

is a harmonic one as well:


0
2
1
2
1
2
2
2
2

,
_

,
_

y x y
v
x
v
x
y
z

, (5.15)
0
. (5.16)
The total of the points, in which the potential
function

is constant, define the equipotential


surfaces.
In the case of a potential plane motion:

- constant, equipotential lines of velocity;

- constant, stream lines.


Computing the circulation of velocity along a
certain outline, in the mass of fluid, between points
A and B (fig.5.1), we get:


B
A
B
A
A B
B
A
d r d r d v . (5.17)
58
Thus, the circulation of velocity doesnt
depend on the shape of the curve AB, but only on
the values of the function

in A and B. The
circulation of velocity is nil along an equipotential
line of velocity ( . const
B A
).
If we compute the flow of liquid through the
curve AB in the plane motion (in fact through the
cylindrical surface with an outline AB and unitary
breadth), we get (fig.5.1):
Fig.5.1
( )


B
A
B
A
A B y x
d dx v dy v Q 1 1 . (5.18)
Thus, the flow that crosses a curve does not
depend on its shape, but only on the values of
function

in the extreme points. The flow through


a streamline is nil ( ) . const
B A
.
A streamline crosses orthogonal on an
equipotential line of velocity. To demonstrate this
propriety we shall take into consideration that the
gradient of a scalar function F is normal on the
level surface F = cons. As a result, vectors

and

are normal on the streamlines and on the
equipotential lines of velocity.
Computing their scalar product, we get:
59
0 +


y x y x
v v v v
y y x x


. (5.19)
Since their scalar product is nil, it follows that
they are perpendicular, therefore their streamlines
are perpendicular on the lines of velocity.
Going back to the expressions of
x
v
and
y
v
:
.
;
x y
v
y x
v
y
x



(5.20)
Relations (5.20) represent the Cauchy-
Riemanns monogenic conditions for a function of
complex variable.
Any potential plane motion may always be
plotted by means of an analytic function of complex
variable,
( )
i
re z iy x z + .
The analytic function;
( ) ( ) ( ) y x i y x z W , , + , (5.21)
is called the complex potential of the plane
potential motion.
Deriving (5.21) we get the complex velocity:
60
y x
v i v
y
i
y x
i
x dz
dW


. (5.22)
Fig.5.2
( )


i
e v i v
dz
dW

sin cos . (5.23)
Having found the complex potential, lets
establish a few types of plane potential motions.
5.2 Rectilinear and uniform motion
Lets consider the complex potential:
( ) z a z W , (5.24)
where a is a complex constant in the form of:
K
v i v a
0
, (5.25)
with
0
v
and
K
v real and constant positive.
61
Relation (5.24) can be written in the form:
( ) ( ) ( )i x v y v y v x v i z W
K K
+ + +
0 0

, (5.26)
where from we can get the expressions of functions

and

:
( )
( ) . ,
, ,
0
0
x v y v y x
y v x v y x
K
K

+

(5.27)
By equalling these relations with constants we
obtain the equations of equipotential lines and of
streamlines.
.
.
2 0
1 0
cons C x v y v
cons C y v x v
K
K

+
(5.28)
From these equations we notice that the
streamlines and equipotential lines are straight,
having constant slopes (fig.5.3).
Fig.5.3
62
. 0
, 0
0
2
0
1
>
<
v
v
tg
v
v
tg
K
K

(5.29)
We can easily check the orthogonality of the
stream and equipotential lines by writing:
1
2 1
tg tg . (5.30)
Deriving the complex potential we get the
complex velocity:
K
v i v a
dz
dW

0
, (5.31)
that enables us to determine the components of
velocity in a certain point:
. 0
, 0
0
>
>
K y
x
v v
v v
(5.32)
The vector velocity will have the modulus:
2 2
0 K
v v v + , (5.33)
and will have with axis Ox, the angle
2
, given by
the relation (5.29).
We can conclude that the potential vector
(5.25) is a rectilinear and uniform flow on a
direction of angle
2
with the abscissa axis.
The components of velocity can be also
obtained from relations (5.20):
63
.
,
0
K y
x
v
x y
v
v
y x
v



(5.34)
If we particularise (5.25), by assuming
0
k
v
,
the potential (5.24) will take the form:
( ) z v z W
0

, (5.35)
that represents a rectilinear and uniform motion on
the direction of the axis Ox.
Analogically, assuming in (5.25)
0
0
v
, we get:
( ) z v i z W
K
, (5.36)
that is the potential vector of a rectilinear and
uniform flow, of velocity
K
v , on the direction of the
axis Oy.
The motion described above will have a
reverse sense if the corresponding expressions of
the potential vector are taken with a reverse sign.
5.3 The source
Lets consider the complex potential:
( ) z
Q
z W ln
2
, (5.37)
64
where Q is a real and positive constant.
Writing the variable
i
e r z , this complex
potential becomes:
( ) ( )

i r
Q
i z W + + ln
2
, (5.38)
where from we get function

and

:
.
2
, ln
2

Q
r
Q

(5.39)
which, equalled with constants, give us the
equations of equipotential and stream lines, in the
form:
. .
, .
cons
cons r

(5.40)
It can be noticed that the equipotential lines
are concentric circles with the centre in the origin
of the axes, and the streamlines are concurrent
lines in this point (fig.5.4).
Fig.5.4
65
Knowing that:
sin cos r y and r x
, (5.41)
in a point ( ) , r M , the components of velocity
will be:
. 0
1
,
2

r
v
r
Q
r
v
S
r
(5.42)
It can noticed that on the circle of radius r =
cons., the fluid velocity has a constant modulus,
being co-linear with the vector radius of the
considered point.
Such a plane potential motion in which the
flow takes place radially, in such a manner that
along a circle of given radius velocity is constant as
a modulus, is called a plane source.
Constant Q, which appears in the above -
written relations, is called the flow of the source.
The flow of the source through a circular
surface of radius r and unitary breadth will be:
1 2
r
v r Q . (5.43)
Analogically, the complex potential of the
form:
( ) z
Q
z W ln
2
, (5.44)
will represent a suction or a well because, in this
case, the sense of the velocity is reversing, the
66
fluid moving from the exterior to the origin (where
it is being sucked).
If the source isnt placed in the origin of the
axes, but in a point
1
O , of the real axis, of abscissa
a t , then:
( ) ( ) a z
Q
z W t ln
2
. (5.45)
5.4. The whirl
Let the complex potential be:
( ) z
i
z W ln
2

. (5.46)
where is a positive and real constant, equal to
the circulation of velocity along a closed outline,
which surrounds the origin.
Proceeding in the same manner as for the
previous case, we shall get the functions

and

:
, ln
2
,
2
r

(5.47)
from which we can notice that the equipotential
lines, of equation . const are concurrent lines, in
the origin of axes, and the streamlines, having the
equation
. const r
, are concentric circles with their
centre in the origin of the axes (fig.5.5).
67
* The dipole or the duplet is a plane potential motion that consists of two equal
sources of opposite senses, placed at an infinite small distance , so that the product ,
, called the moment of the dipole should be finite and constant. .
Fig.5.5
The components of velocity are:
0
2
1
0 >

r r
v and
r
v
S r


. (5.48)
Thus, on a circle of given radius r, the velocity
is constant as a modulus, has the direction of the
tangent to this circle in the considered point and is
directed in the sense of angle increase.
If the whirl is placed on the real axis, in a
point with abscissa a t , the complex potential of
the motion will be:
( ) ( ) a z
i
z W ln
2

. (5.49)
68
* The dipole or the duplet is a plane potential motion that consists of two equal
sources of opposite senses, placed at an infinite small distance , so that the product ,
, called the moment of the dipole should be finite and constant. .
5.5. The flow with and without
circulation around a circular cylinder
The flow with circulation around a circular cylinder
is a plane potential motion that consists of an axial
stream (directed along axis Ox), a dipole of
moment
*
2 M (with a source at the left of suction)
and a whirl (in direct trigonometric sense).
The complex potential of motion will be:
( ) z
i
z
r
z v z W ln
2
2
0
0

,
_

+
, (5.50)
where we have done the denotation:
0
2
0
1
v
r
. (5.51)
By writing the complex variable
i
e r z , we
shall divide in (5.50) the real part from the
imaginary one, thus obtaining functions

and


2
cos
2
0
0

+

,
_

+
r
r
r v
, (5.52)
r
r
r
r v ln
2
sin
2
0
0

,
_


. (5.53)
69
* The dipole or the duplet is a plane potential motion that consists of two equal
sources of opposite senses, placed at an infinite small distance , so that the product ,
, called the moment of the dipole should be finite and constant. .
The stream and equipotential lines are
obtained by taking in relations (5.52), (5.53),
C C ,
respectively. We notice that if in (5.53)
we assume
0
r r
, function

will become constant;


therefore we can infer that the circle of radius
0
r

with the centre in the origin of the axes is a
streamline (fig.5.8).
Admitting that this streamline is a solid
border, well be able to consider this motion
described by the complex potential (5.50) as being
the flow around a straight circular cylinder of
radius
0
r
, having the breadth normal on the motion
plane, infinite.
If we plot the other
streamlines we shall get
some asymmetric curves
with respect to axis Ox
(fig.5.6). On the inferior
side of the circle of
radius
0
r
, the velocity
due to the axial stream
is summed up with the
velocity due to the whirl.
Fig.5.6
As a result, here we shall obtain smaller
velocities, and the streamlines will be more rare.
In polar co-ordinates, the components of
velocity in a certain point ( ) , r M , will be:
cos 1
2
2
0
0

,
_


r
r
v v
r
, (5.54)
70
If the considered point is placed on the circle
of radius
0
r
, well have:
.
2
sin 2
, 0
0
0
r
v v
v
S
r

(5.55)
The position of stagnant points can be
determined provided that between these points the
velocity of the fluid should be nil.
The flow without circulation around a circular
cylinder is the plane potential motion made up of
an axial stream (directed along axis Ox) and a
dipole of moment 2 M (whose source is at the
left of suction).
Thus, this motion can be obtained
particularising the motion previously described by
cancelling the whirl.
By making 0 , in relations (5.50), (5.52) and
(5.53) we get the complex potential of the motion,
the function potential of velocity and the function
of stream, in the form:
( ) ,
2
0
0

,
_

+
z
r
z v z W
(5.56)
, cos
2
0
0

,
_

+
r
r
r v
(5.57)
. sin
2
0
0

,
_


r
r
r v
(5.58)
71
By writing the equation of streamlines

=
cons. in the form:
.
2 2
2
0
0
const C y
y x
r
y v
+

(5.59)
we notice that the nil streamline (C = 0) is made up
of a part of the real axis (Ox) and the circle of
radius
0
r
(fig.5.7).
The other
streamlines are
symmetric curves with
respect to axis Ox.
Obviously, if we
consider the circle of
radius
0
r
, as a solid
border, the motion
can be seen as a flow
of an axial stream
around an infinitely
long cylinder, normal
on the motion plane.
Fig.5.7
The components of velocity are:
. sin 1
, cos 1
2
2
0
0
2
2
0
0

,
_

,
_


r
r
v v
r
r
v v
S
r
(5.60)
which, on the circle of radius
0
r
, become:
72
. sin 2
, 0
0
v v
v
S
r

(5.61)
The position of stagnant points is obtained by
making
0
S
v v
, which implies 0 sin . Thus the
stagnant points are found on the axis Ox in the
points
( ) ,
0
r A
and
( ) 0 ,
0
r B
.
5.6 Kutta Jukovskis theorem
Let us consider a cylindrical body normal on
the complex plane, the outline C being the crossing
curve between the cylinder and the complex plane.
Around this outline there flows a stream,
potential plane, having the complex potential ( ) z W .
The velocity in infinite of the stream, directed in
the negative sense of the axis Ox, is

v .
In this case the resultant of the pressure
forces will have the components:

. 1
, 0

v R
R
y
x

(5.62)
The forces are given with respect to the unit
of length of the body.
The second relation (5.62) is the mathematic
expression of Kutta-Jukovskis theorem, which will
be only stated below without demonstrating it:
If a fluid of density

is draining around a
body of circulation and velocity in infinite

v , it
73
will act upon the unit of length of the body with a
force equal to the product

v , normal on the
direction of velocity in infinite called lift force
(lift).
The sense of the lift is obtained by rotating
the vector of velocity from infinite with
0
90 in the
reverse sense of circulation.
74
6. IMPULSE AND MOMENT
IMPULSE THEOREM
We take into consideration a volume

of fluid.
This fluid is homogeneous, incompressible, of
density

, bordered by surface

. The elementary
volume d has the speed
v
.
The elementary impulse will be:
d v I d . (6.1)

d v I
. (6.2)

d
dt
v d
dt
I d
. (6.3)
At the same time
i
F
dt
I d
. (6.4)
But: 0 + +
i p m
F F F (dAlembert principle).
(6.5)
Therefore:
e p m
F F F
dt
I d
+ . (6.6)
75
The total derivative, of the impulse with
respect to time, is equal to the resultant
e
F of the
exterior forces, or
i i e e e
v M v M F , (6.7)
where
e i
M M ,
are the mass flows through entrance/
exit surfaces.
Under permanent flow conditions of ideal
fluid, the vectorial sum of the external forces which
act upon the fluid in the volume

, is equal with
the impulse flow through the exit surfaces (from
the volume

), less the impulse flow through the


entrance surfaces (to the volume

) .
r
- the position vector of the centre of volume
with respect to origin of the reference system.
The elementary inertia moment with respect
to point O (the origin) is:
( ) d v r
dt
d
d
dt
v d
r M d
i

,
_


, (6.8)
since
( ) .
dt
v d
r
dt
v d
r v v
dt
v d
r v
dt
r d
v r
dt
d
+ + (6.9)
then
( )



d v r
dt
d
M d M
i i . (6.10)
If:
d v I d the elementary impulse, (6.11)
76
d v r k d the moment of elementary impulse,
(6.12)

, d v r k
(6.13)
( )
i
M d v r
dt
d
dt
k d


. (6.14)
The derivative of the resultant moment of
impulse with respect to time is equal with the
resultant moment of inertia forces with reversible
sign.
ex p m
M M M
dt
k d
+ , (6.15)
where
m
M - the moment of mass forces,
p
M - the moment pressure forces,
ex
M - the moment of external forces.
oi oe
r r , - the position vector of the centre of
gravity for the exit /entrance surfaces.
( ) ( )
i oi i e oe e ex
v r M v r M M . (6.16)
Under permanent flow conditions of ideal
fluids, the vectorial addition of the moments of
external forces which act upon the fluid in the
volume

, is equal to the moment of the impulse


flow through the exit surfaces less the moment of
the impulse flow through the entrance surfaces.
77
78
7. MOTION EQUATION OF THE
REAL FLUID
7.1 Motion regimes of fluids
The motion of real fluids can be carried out
under two regimes of different quality: laminar and
turbulent.
These motion regimes were first emphasised
by the English physicist in mechanics Osborne
Reynolds in 1882, who made systematic
experimental studies concerning the flow of water
through glass conduits of diameter mm d 25 5 .
The experimental installation, which was then
used, is schematically shown in fig.7.1.
79
Fig.7.1
The transparent conduit 1, with a very
accurate processed inlet, is supplied by tank 2, full
of water, at a constant level.
80
The flow that passes the transparent conduit
can be adjusted by means of tap 3, and measured
with the help of graded pot 6.
In conduit 1, inside the water stream we
insert, by means of a thin tube 4, a coloured liquid
of the same density as water. The flow of coloured
liquid, supplied by tank 5 may be adjusted by
means of tap 7.
But slightly turning on tap 3, through conduit
1 a stream of water will pass at a certain flow and
velocity.
If we turn on tap 7 as well, the coloured liquid
inserted through the thin tube 4, engages itself in
the flow in the shape of a rectilinear thread,
parallel to the walls of conduit, leaving the
impression that a straight line has been drawn
inside the transparent conduit 1.
This regime of motion under which the fluid
flows in threads that dont mix is called a laminar
regime.
By slowly continuing to turn on tap 3, we can
notice that for a certain flow velocity of water, the
thread of liquid begins to undulate, and for higher
velocities it begins to pulsate, which shows that
vector velocity registers variations in time
(pulsations).
For even higher velocities, the pulsations of
the coloured thread of water increase their
amplitude and, at a certain moment, it will tear, the
particles of coloured liquid mixing with the mass of
water that is flowing through conduit 1.
81
The regime of motion in which, due to
pulsations of velocity, the particles of fluid mix is
called a turbulent regime.
The shift from a laminar regime to the
turbulent one, called a transition regime is
characterised by a certain value of Reynolds
number
*
, called critical value (
cr
Re
).
82
* Number , is the number that defines the
similarity criterion Reynolds.
For circular smooth conduits, the critical value
of Reynolds number is
2320 Re
cr
.
For values of Reynolds number inferior to the
critical value (
cr
Re Re <
), the motion of liquid will be
laminar, while for
cr
Re Re >
, the flow regime will be
turbulent.
7.2 Navier Stokes equation
Navier Stokes equation describes the
motion of real (viscous) incompressible fluids in a
laminar regime.
Unlike ideal fluids that are capable to develop
only unitary compression efforts that are
exclusively due to their pressure, real (viscous)
fluids can develop normal or tangent
supplementary viscosity efforts.
The expression of the tangent viscosity effort,
defined by Newton (see chapter 2) is the following:
y
v


. (7.1)
Newtonian liquids are capable to develop,
under a laminar regime, viscosity efforts

and

,
that make-up the so-called tensor of the viscosity
efforts,
v
T
(in fig. 7.2, efforts manifest on an
elementary parallelipipedic volume of fluid with the
sides
dz and dy dx,
):
83
1
1
1
]
1

zz yz xz
zy yy xy
zx yx xx
v
T



. (7.2)
The tensor
v
T
is symmetrical:
yz zy xz zx xy yx
; ;
. (7.3)
Fig.7.2
The elementary force of viscosity that is
exerted upon the elementary volume of fluid in the
direction of axis Ox is:
( ) ( ) ( )
. dz dy dx
z y x
dy dx dz
z
dy dx dy
y
dz dy dx
x
dF
zx
yx
xx
zx
yx
xx
vx

,
_

(7.4)
According to the theory of elasticity:
84
z
x
y
.
;
; 2

,
_

,
_

z
v
x
v
y
v
x
v
x
v
x z
zx
x
y
yx
x
xx



(7.5)
Thus:
.
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
dydz dx
z
v
y
v
x
v
z
v
y
v
x
v
x
z x
v
z
v
y
v
y x
v
x
v
dF
x x x z
y
x
z x x
y
x
vx
1
1
]
1

,
_

,
_

1
1
]
1

,
_

,
_



(7.6)
But
0

z
v
y
v
x
v
z
y
x
, according to the equation
of continuity for liquids.
Then:
dz dy dx v dF
x x

. (7.7)
Similarly:
, dz dy dx v dF
y vy

(7.8)
. dy dy dx v dF
z vz

(7.9)
Hence:
, d v F d
v
(7.10)
85
.

d v Fv
(7.11)
Unlike the ideal fluids, in dAlemberts
principle the viscosity force also appears.
. 0 + + +
i v p m
F F F F (7.12)
Introducing relations (3.3), (3.5), (3.7) and
(7.11) into (7.12), we get:

,
_

0 d
dt
v d
v p F
, (7.13)
or:
dt
v d
v p F +

1
. (7.14)
Relation (7.14) is the vectorial form of Navier-
Stokes equation. The scalar form of this equation
is:
.
1
;
1
;
1
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
z
z
y
z
x
z z z z z
z
z
y
y
y
x
y y y y y
y
z
x
y
x
x
x x x x x
x
v
z
v
v
y
v
v
x
v
t
v
z
v
y
v
x
v
z
p
F
v
z
v
v
y
v
v
x
v
t
v
z
v
y
v
x
v
y
p
F
v
z
v
v
y
v
v
x
v
t
v
z
v
y
v
x
v
x
p
F

,
_

,
_

,
_

(7.15)
86
7.3 Bernoullis equation under the
permanent regime of a thread of real fluid
Unlike the permanent motion of an ideal fluid,
where its specific energy
*
remains constant along
the thread of fluid and where, from one section to
another, there takes place only the conversion of a
part from the potential energy into kinetic energy,
or the other way round, in permanent motion of the
real fluid, its specific energy is no longer constant.
It always decreases in the sense of the movement
of the fluid.
A part of the fluids energy is converted into
thermal energy, is irreversibly spent to overcome
the resistance brought about by its viscosity.
Denoting this specific energy (load) by
f
h
,
Bernoullis equation becomes:
f
h z
p
g
v
z
p
g
v
+ + + + +
2
2
2
2
1
1
2
1
2 2
. (7.16)
In different points of the same section, only
the potential energy remains constant, the kinetic
one is different since the velocity differs in the
section, ( ) z y x v v , , . In this case the term of the
kinetic energy should be corrected by a coefficient

, that considers the distribution of velocities in


the section ( ) 1 , 1 05 , 1 .
f
h z
p
g
v
z
p
g
v
+ + + + +
2
2
2
2 2
1
1
2
1 1
2 2

. (7.17)
87
* the weight unit energy
By reporting the loss of load
f
h
to the length l
of a straight conduit, we get the hydraulic slope
(fig.7.3):
Fig.7.3

l
h
l
z
p
g
v
z
p
g
v
I
f

,
_

+ +

,
_

+ +

2
2
2
2 2
1
1
2
1 1
2 2

. (7.18)
If we refer only to the potential specific
energy, we get the piezometric slope:
l
z
p
z
p
I
p

,
_

,
_

2
2
1
1

. (7.19)
In the case of uniform motion (
ct v
):
l
h
tg I I
f
p
. (7.20)
Experimental researches have revealed that
irrespective of the regime under which the motion
of fluid takes place, the losses of load can be
written in the form:
88
m
f
v b h
, (7.21)
where b is a coefficient that considers the nature of
the fluid, the dimensions of the conduit and the
state of its wall.
1 m for laminar regime;
2 75 , 1 m
for turbulent regime.
If we logarithm (7.21) we get:
v m b h
f
lg lg lg +
. (7.22)
In fig. 7.4 the load variation
f
h
with respect to
velocity is plotted in logarithmic co-ordinates.
Fig.7.4
For the laminar regime
0
45 . The shift to the
turbulent regime is made for a velocity
corresponding to
2320 Re
cr
.
89
7.4 Laminar motion of fluids
7.4.1 Velocities distribution between two
plane parallel boards of infinite length (fig.7.5).
To determine the velocity distribution
between two plane parallel boards of infinite
length, we shall integrate the equation (7.15)
under the following conditions:
Fig.7.5
a) velocity has only the direction of the axis
Ox:
; 0 , 0
z y x
v v v
(7.23)
from the equation of continuity
0 v
, it results:
, 0

x
v
x
(7.24)
therefore velocity does not vary along the axis Ox.
90
b) the movement is identically reproduced in
planes parallel to xOz:
0

y
v
x
. (7.25)
From (7.24) and (7.25) it results that
( ) z v v
x x

.
c) the motion is permanent:
0

t
v
x
. (7.26)
d) we leave out the massic forces (the
horizontal conduit).
e) the fluid is incompressible.
The first equation (7.15) becomes:
0
1
2
2
+

dz
v d
x
p
x

, (7.27)
Integrating twice (7.27):
( )
2 1
2
2
1
C z C z
x
p
z v
x
+ +

. (7.28)
For the case of fixed boards, we have the
conditions at limit:
. 0 ,
; 0 , 0


x
x
v h z
v z
(7.29)
91
Subsequently:
. 0
;
2
1
2
1

C
h
x
p
C

(7.30)
Then the law of velocity distribution will be:
( ) ( ) z h z
x
p
z v
x


2
1
. (7.31)
It is noticed that the velocity distribution is
parabolic, having a maximum for
2
h
z :
x
p h
v
x


8
2
max
*
. (7.32)
Computing the mean velocity in the section:
( )


h
x
x
p h
dz z v
h
u
0
2
12
1

, (7.33)
well notice that
max
3
2
v u .
The flow that passes through a section of
breadth b will be:
x
p h b
h b v Q


12
3
. (7.34)
92
* is positive, since (the sense of the flow, the positive
sense of axis Ox, corresponds to a decrease in pressure).
7.4.2 Velocity distribution in circular conduits
Lets consider a circular conduit, of radius
0
r

and length l, through which an incompressible fluid
of density

and kinematic viscosity

(fig.7.6)
passes.
We report the conduit to a system of
cylindrical co-ordinates (
and r x,
), the axis Ox,
being the axis of the conduit. The movement being
carried out on the direction of the axis, the velocity
components will be:
0 , 0

v v v
r x
. (7.35)
The equation of continuity
0 v
, written in
cylindrical co-ordinates:
( ) ( )
0
1

1
]
1


x
r v v
r
v r
r
v
x r

, (7.36)
becomes:
0

x
v
x
, (7.37)
where from we infer that the velocity of the fluid
doesnt vary on the length of the conduit.
On the other hand, taking into consideration
the axial symmetrical character of the motion,
velocity will neither depend on variable .
As a result, for a permanent motion, it will
only depend on variable r, that is ( ) r v v .
93
The distribution of velocities in the section of
flow can be obtained by integrating the Navier-
Stokes equations (7.14).
Noting by
r
i i, and

i the versors of the three


directions of the adopted system of cylindrical co-
ordinates, we can write vector velocity:
( ) i r v v
x
. (7.38)
Bearing in mind that in cylindrical co-
ordinates, operator " " has the expression:


r
i
r
i
x
i
r
. (7.39)
On the basis of (7.38), we can write:
( ) ( ) 0


x x
v i
x
v v v , (7.40)
since, as we have seen, velocity
x
v
only depends on
variable r.
On the other hand, in cylindrical co-ordinates, the
term
v
may be rendered in the form:
.
1

,
_

1
]
1

,
_

+
,
_

+
,
_


r
r
v
r r
i
r
x
v
x r
v
r
r
v
r r
i
v i v
x
x x x
x

(7.41)
Keeping in mind the permanent character of
the motion, relation (7.40) and (7.41) the projection
of equation (7.14) onto the axis Ox may be written
in the form:
94
x
p
r
r
v
r r
x

,
_

1
, (7.42)
since, on the hypothesis of a horizontal conduit,
0
x x
g F
.
Assuming that the gradient of pressure on the
direction of axis Ox is constant (
. / cons x p
), and
integrating the equation (7.42), we shall
successively get:
,
2
1
1
r
C
r
x
p
r
v
x
+

(7.43)
, ln
4
1
2 1
2
C r C r
x
p
v
x
+ +

(7.44)
The integrating constants
1
C and
2
C are
determined using the limit conditions:
- in the axis of conduit, at r = 0, velocity
should be finite, thus constant
1
C should be
nil;
- on the wall of conduit, at
0
r r
, velocity of
fluid should be nil; consequently:
2
0 2
4
1
r
x
p
C

, (7.45)
and relation (7.44) becomes:
( )
2 2
0
4
1
r r
x
p
v
x

. (7.46)
95
From (7.46) we notice that if the motion takes
place in the positive sense of the axis
( ) 0 >
x
v Ox
, then
0 / < x p
; therefore pressure decreases on the
direction of motion if I is the piezometric slope
(equal in this case to the hydraulic slope), we can
write:
I
l
p
x
p

, (7.47)
where
p
is the fall of pressure on the length l of
the conduit. Subsequently, relation (7.41) becomes:
( )
2 2
0
4
r r
I
v
x

. (7.48)
Fig.7.6
It can be noticed that the distribution of
velocities in the section of flow is parabolic (fig.7.6
a), the maximum velocity being registered in the
axis of conduit (r = 0), therefore we get:
2
0 max
4
1
r
I
v
x

. (7.49)
Let us now consider an elementary surface d A
in the shape of a circular crown of radius r and
96
breadth d r (fig.7.6 b). The elementary flow that
crosses surface d A is:
rdr v dA v dQ
x x
2
, (7.50)
and:
( )


0
0
4
0
2 2
0
8 2
r
r
I
dr r r r
I
Q

. (7.51)
The mean velocity has the expression:
2 8
max ,
2
0
x
v
r
I
A
Q
u

. (7.52)
Further on we can write:
g
v
d d g
d v
d g
v
r
v
l
h
I
f
2
1
Re
64
Re
32
32 8
2
2
2
2 2
0

. (7.53)
Relation (7.53) is Hagen-Ppiseuilles law,
which gives us the value of load linear losses in the
conduits for the laminar motion:
g
v
d
l
g
v
d
l
h
f
2 2 Re
64
2 2

, (7.54)
Re
64
is the hydraulic resistance coefficient for
laminar motion.
7.5 Turbulent motion of fluids
97
In a point of the turbulent stream, the fluid
velocity registered rapid variation, in one sense or
the other, with respect to the mean velocity in
section. The field of velocities has a complex
structure, still unknown, being the object of
numerous studies.
The variation of velocity with the time may be
plotted as in fig.7.7.
Fig.7.7
A particular case of turbulent motion is the
quasipermanent motion (stationary on average). In
this case, velocity, although varies in time, remains
a constant means value.
In the turbulent motion we define the
following velocities:
a) instantaneous velocity ( ) t z y x u , , , ;
b) mean velocity
( ) ( )

T
dt t z y x u
T
z y x u
0
, , ,
1
, ,
; (7.55)
c) pulsation velocity
( ) ( ) ( ) z y x u t z y x u t z y x u , , , , , , , ,
'
. (7.56)
98
There are several theories that by simplifying
describe the turbulent motion:
a) Theory of mixing length (Prandtl), which
admits that the impulse is kept constant.
b) Theory of whirl transports (Taylor) where
the rotor of velocity is presumed constant.
c) Karamans theory of turbulence, which
states that, except for the immediate vicinity of a
wall, the mechanism of turbulence is independent
from viscosity.
7.5.1 Coefficient in turbulent motion
Determination of load losses in the turbulent
motion is an important problem in practice.
It had been experimentally established that in
turbulent motion the pressure loss
p
depends on
the following factors: mean velocity on section, v ,
diameter of conduit, d , density

of the fluid and


its kinematic viscosity

, length l of the conduit


and the absolute rugosity
*

of its interior walls;


therefore:
( ) , , , , , l d v f p , (7.57)
or:
d
l v
p
2
2

, (7.58)
99
d
l
g
v p
h
f
2
2

, (7.59)
r
or
d

- relative rugosity
where:

,
_

d
Re, 2
1

. (7.60)
100
*mean height of the conduit prominence ; -relative rogosity.
As it can be seen from relation (7.60), in
turbulent motion the coefficient of load loss may
depend either on Reynolds number or on the
relative rugosity of the conduit walls.
In its turbulent flow through the conduit, the
fluid has a turbulent core, in which the process of
mixing is decisive in report to the influence of
viscosity and a laminar sub-layer, situated near the
wall, in which the viscosity forces have a decisive
role.
If we note by
l

the thickness of the laminar


sub-layer, then we can classify conduits as follows:
- conduits with smooth walls;
l
<
;
- conduits with rugous walls;
l
>
.
From (7.60) we notice that, unlike the laminar
motion in turbulent motion is a complex function
of Re and
d

.
It has been experimentally established that in
the case of hydraulic smooth conduits, coefficient
depends only on Reynolds number. Thus, Blasius,
by processing the existent experimental material
(in 1911), established for the smooth hydraulic
conduits of circular section, the following empirical
formula:
25 , 0
4 / 1
Re
3164 , 0
3164 , 0
,
_

d v
, (7.61)
valid for
5
10 Re 000 , 4 < < .
Using Blasius relation in (7.59) we notice that
under this motion regime the load losses are
proportional to velocity to 1,75
th
power.
101
Also for smooth conduits, but for higher
Reynolds numbers ( )
7
10 Re 000 , 3 < < we can use
Konakovs relation:
( )
2
5 , 1 Re lg 8 , 1

. (7.62)
In turbulent flow through conduits, coefficient
no longer depends on Reynolds number, and it
can be determined with the help of Prandtl
Nikuradses relation:
2
0
74 , 1 lg 2

,
_

r
. (7.63)
Some of the most important formulae for the
calculus of coefficient are given in table 7.1, the
validity field of each relation being also shown [7].
Table 7.1
No.
a
I
Relation Regime Field
I III IV V
1 Poisse
uille Re
64

Laminar
2320 Re <
2 Prandtl
( )
2
8 , 0 Re lg 2


Smooth
turbulen
t
7
10 Re
000 , 3 Re
<
>
3 Blasius
25 , 0
Re 3164 , 0


5
10 Re
000 , 4 Re
<
>
4 Konako
v
( )
2
5 , 1 Re lg 8 , 1


7
10 Re
000 , 3 Re
<
>
5 Nikura
dze
237 , 0
Re 221 , 0 0032 , 0

+
6
5
10 2 Re
10 Re
<
>
6
Lees
35 , 0 3
Re 61 , 0 10 714 , 0

+
6
3
10 3 Re
10 Re
<
>
102
II
Auth
or
7 Colebr
ook-
White
Re
51 , 2
72 , 3
lg 2
1
+


d
Demi-
rugous
Universal
8 Prandtl
-
Nikurd
ze
2
0
74 , 1 lg 2

,
_

Turbulen
t rugous
8 5
10 Re 10 < <
9 Sifrins
on
25 , 0
11 , 0
,
_

500 Re >

d
103
7.5.2 Nikuradzes diagram
On the basis of experiments made with
conduits of homogeneous different rugosity, which
was achieved by sticking on the interior wall some
grains of sand of the same diameter, Nikuradze has
made up a diagram that represents the way
coefficient varies, both for laminar and turbulent
fields (fig.7.8).
Fig.7.8
We can notice that in the diagram appear five
areas in which variation of coefficient , distinctly
differs.
Area I is a straight line which represents in
logarithmic co-ordinates the variation:
104
Re
64
, (7.64)
105
corresponding to the laminar regime ( ) 2320 Re < . On
this line all the doted curves are superposed, which
represents variation ( ) Re f for different relative
rugosities
0
/ r
.
Area II is the shift from laminar regime to the
turbulent one which takes place for
( ) 2300 Re 4 , 3 Re lg .
Area III corresponds to the smooth hydraulic
conduits. In this area coefficient can be
determined with the help of Blasius relation (7.61),
to which the straight line III a corresponds, called
Blasius straight. Since the validity field of relation
(7.61) is limited by
5
10 Re , for higher values of
Reynolds number, we use Kanakovs formula, to
which curve III b corresponds. It is noticed that the
smaller the relative rugosity is, the greater the
variation field of Reynolds number, in which the
smooth turbulent regime is maintained.
In area IV each discontinuous curve, which
represents dependent ( ) Re f for different relative
rugosities becomes horizontal, which emphasises
the independence of on number Re . Therefore
this area corresponds to the rugous turbulent
regime, where is determined by (7.63).
It is noticed that in this case the losses of load
(7.59) are proportional to square velocity.
For this reason the rugous turbulent regime is
also called square regime.
Area V is characterised by the dependence of
the coefficient both on Reynolds number and on
the relative rugosity of the conduit.
106
It can be noticed that for areas IV and V,
coefficient decreases with the decrease of
relative rugosity.
107
8. FLOW THROUGH CIRCULAR
CONDUITS
In this chapter we shall present the hydraulic
calculus of conduits under pressure in a permanent
regime.
Conduits under pressure are in fact a
hydraulic system designed to transport fluids
between two points with different energetic loads.
Conduits can be simple (made up of one or
several sections of the same diameter or different
diameters), or with branches, in this case, setting
up networks of distribution.
By the manner in which the outcoming of the
fluid from the conduit is made, we distinguish
between conduits with a free outcome, which
discharge the fluid in the atmosphere (fig.8.1 a)
and conduits with chocked outcoming (fig. 8.1 b).
Fig.8.1a, b
108
If we write Bernoullis equation for a stream
of real liquid, between the free side of the liquid
from the tank A and the end of the conduit, taking
as a reference plane the horizontal plane N N, we
get:
f
h z
p
g
v
z
p
g
v
+ + + + +
2
2
2
2 2
1
1
2
1 1
2 2

, (8.1)
which, for the case presented in fig.8.1 a, when
0
1
v ,
0 2 1
p p p
, 1
2 1
, h z z +
2 1
, becomes:
f
h
g
v
h +
2
2
, (8.2)
where
2
v v is the mean velocity in the section of
the conduit , and h is the load of the conduit.
In the analysed case shown in fig. 8.1 b, by
introducing in equation (8.1) the relations
1 0 2 2 1 1 2 0 1 1
, , , , 0 h p p z h h z v v p p v + + +
and
1
2 1
, we shall get the expression (8.2).
From an energetic point of view, this relation
shows that from the available specific potential
energy (h), a part is transformed into specific
kinetic energy ( g v 2 /
2
) of the stream of fluid, which
for the given conduit is lost at the outcoming in the
atmosphere or in another volume. The other part
( )
f
h
is used to overcome the hydraulic resistances
(that arise due to the tangent efforts developed by
the fluid in motion) and is lost because it is
irreversibly transformed into heat.
109
Analysing the losses of load from the conduit
we shall divide them into two categories, writing
the relation:
' ' '
f f f
h h h +
. (8.3)
The losses of load, denoted by f h
'
are brought
about by the tangent efforts that are developed
during the motion of the fluid along the length of
the conduit ( l) and, for this reason, they are called
losses of load distributed. These losses of load
have been determined in paragraph 7.4.2, getting
the relation (7.54) which we may write in the form:
d
l
g
v
h f
2
2
'

, (8.4)
where the coefficient of losses of load, , called
Darcy coefficient is determined by the relations
shown in table 7.1 ; the manner of calculus being
also shown in that paragraph. Generally, in
practical cases, the values of coefficient vary in a
domain that ranges between
04 , 0 02 , 0
.
Being proportional to the length of the
conduit, the distributed losses of load are also
called linear losses.
The second category of losses of load is
represented by the local losses of load that are
brought about by: local perturbation of the normal
flow, the detachment of the stream from the wall,
whirl setting up, intensifying of the turbulent
mixture, etc; and arise in the area where the
conduit configuration is modified or at the meeting
an obstacle detouring (inlet of the fluid in the
110
conduit, flaring, contraction, bending and
derivation of the stream, etc.).
The local losses of load are calculated with the
help of a general formula, given by Weissbach:
g
v
h
f
2
2
' '

, (8.5)
where

is the local loss of load coefficient that is


determined for each local resistance (bends,
valves, narrowing or enlargements of the flow
section etc.).
Generally, coefficient

depends mainly on
the geometric parameters of the considered
element, as well as on some factors that
characterise the motion, such as: the velocities
distribution at the inlet of the fluid in the examined
element, the flow regime, Reynolds number etc.
In practice, coefficient

is determined with
respect to the type of the respective local
resistance, using tables, monograms or empirical
relations that are found in hydraulic books.
Therefore, for curved bends of angle
0
90 ,
coefficient

can be determined by using the


relation:
0
0
5 , 3
5 , 3
90
16 , 0 13 , 0

,
_

+
d
, (8.6)
where
and d
are the diameter and curvature
radius of the bend, respectively.
Coefficient

, corresponding to the loss of


load at the inlet in their conduit, depends mainly on
111
the wall thickness of the conduit with respect to its
diameter and on the way the conduit is attached to
the tank. If the conduit is embedded at the level of
the inferior wall of the tank, the losses of load that
arise at the inlet in the conduit are equivalent with
the losses of load in an exterior cylindrical nipple.
For this case,
5 . 0
.
If on the route of the conduit there are several
local resistances, the total loss of fluid will be given
by the arithmetic sum of the losses of load
corresponding to each local resistance in turn,
namely:

g
v
h
f
2
2
' '

, (8.7)
Using relations (8.4) and (8.7), we get the
total loss of load of the conduit:
g
v
d
l
h
f
2
2

,
_

+


, (8.8)
that allows us to write relation (8.2) in the form:
g
v
d
l
h
2
1
2

,
_

+ +


, (8.9)
where from the mean velocity in the flow section
will result:

+ +


d
l
h g
v
1
2
. (8.10)
The flow of the conduit is determined by:
112

+ +



d
l
h g d
v
d
Q
1
2
4 4
2 2
, (8.11)
which allows us to express the load of the conduit,
h, and diameter, d, with respect to flow Q; we get:

,
_

+ +


d
l
d
Q
g
h 1
8
4
2
2
, (8.12)
and respectively:
( )

+ +

d l d
h
Q
g
d
2
2
5
8
. (8.13)
Sometimes in the calculus of enough long
conduits, the kinetic term ( ) g v 2 /
2
and the local
losses of load are negligible with respect to the
linear losses of load.
In the case of such conduits, called long
conduits, relation (8.2) takes the form:
d
l
g
v
h h
f
2
2
'

, (8.14)
and relations (8.10), (8.11), (8.12) and (8.13)
become:
l
gdh
v

, (8.15)
l
gdh d
Q

2
4
2

, (8.16)
113
l
d
Q
g
h

5
2
2
8

, (8.17)
and, respectively:
l
h
Q
g
d

2
2
5
8

. (8.18)
With the help of the above written relations
all problems concerning the computation of
conduits under pressure can be solved. Generally,
these problems are divided into three categories:
a)to determine the load of the conduit, when
length, rugosity, flow and rugosity of interior
walls of the conduit are known;
b)to determine the optimal diameters when
flow, length, rugosity of the walls of conduit
as well as the admitted load are known;
c)to determine the flow of liquid conveyed
through the conduit when diameter, length,
nature of the wall of conduit and its load are
known.
114
9. HYDRODYNAMIC PROFILES
9.1 Geometric characteristics of
hydrodynamic profiles
A hydrodynamic profile is a contour with an
elongated shape with respect to the direction of
stream, rounded at the front edge-called leading
edge-and having a peak at the back edge, called
trailing edge.
In what follows we shall
stress on some of the
elements, which
characterise the profile.
a) The chord of the
profile is defined as the
straight line which joins
the trailing edge A, with
the point B, in which the
circle
Fig.9.1
with the centre in A is tangent to the leading edge;
the length of the chord will be noted by c (fig.9.1).
b) The thickness of the profile is measured on the
normal to the chord and is noted by e. This
thickness varies along the chord and reaches a
maximum in a section which is called section of
maximum thickness, situated at the distance
m
l

to the leading edge.
115
c) Relative thickness,

, and maximum relative


thickness,
m

, are defined by the relations:


c
e
and
c
e
m
m
. (9.1)
d) The framework of a profile, or the line of mean
curvature, is the curve that joins the mean
thickness points. The shape of the framework is
an important geometric parameter and is linked
to the curvature motion of the profile.
From this point of view, profiles can be with
simple curvature (fig.9.1) or with double
curvature (9.2).
e) The arrow of the profile, f, is the maximum
distance, measured on the normal to the chord,
between the framework and the chord of the
profile.
f) The extrados and intrados
of the profile represent the
upper and lower part of the
profile, respectively.
By the geometric shape of
the trailing edge, which
plays an important part in
the theory of profiles, we
may distinguish among
three categories of profiles:
Fig.9.2
- Jukovski profiles, profiles with a sharp edge,
for which the tangents to the trailing edge
at extrados and intrados superpose (fig.9.3
a)
116
- Karman-Trefftz profiles, or profiles with a
dihedral tip, for which the
tangents to the extrados and the intrados
make an angle in the
trailing edge (fig.9.3 b),
- Carafoli profiles, or profiles with the
rounded tip, for which the trailing
edge ends in a rounded contour, with a
small curvature radius.
(fig.9.3c).
It is generally studied the
plane potential motion
around the hydrodynamic
profile, considered as the
intersection of the complex
plane of motion with a
cylindrical object (called
wing), normal on this plane
and having an infinite length
(called span).
In reality, wings have a
finite span and, from a
geometrical point of view,
they are characterised by
the section of the wing,
which, generally, alters
Fig.9.3 a, b, c
the length of the wing and the shape of the wing in
plane.
By the shape of the wing in plane, there are:
rectangular wings (fig.9.4), trapezoidal wings (9.4
b), elliptical (9.4 c), and triangular wings (9.4 d).
117
Fig.9.4 a, b, c, d
An important parameter of the wing is the
relative elongation defined by the relation:
S
l
2
, (9.2)
where l and S represent the span and the surface of
the wing, respectively.
In the particular case of rectangular wing, the
length of the chord is constant
0
c c
and relation
(9.2) becomes:
0
/ c l
,
since:
0
c l S
.
We can classify wings by their elongation ;
into:
- wings of infinite span, when 6 > ;
- wings of finite span, when 6 < .
118
9.2 The flow of fluids around wings
Kutta-Jukovskis relation (5.62) can be applied
to any solid body in relative displacement with
respect to a fluid.
It indicates that whenever there is a
circulation around a body, there arises a lift force
y
R , whose value is determined, under the same
circumstances of environment (

v and ), by the
intensity of circulation.
To get a higher circulation around bodies, we
can act in two ways:
- for geometrical symmetric bodies: they are
asymmetrically placed with respect to

v
direction or a rotational motion is induced
(an infinitely long cylinder, sphere-Magnus
effect).
- for asymmetrical bodies: study of shapes
more proper to circulation.
On the basis of many theoretical and
experimental studies, we have come to designing
wings with a high lift, called hydrodynamic profiles.
Fig.9.5
119
In fig.9.5, the arising of circulation around the
hydrodynamic profile, alters the spectre of lines of
rectilinear stream, of velocity

v as follows: on the
extrados the sense of circulations coincides with
that of motion and is seen as a supplement of
velocity v , and on the intrados velocity is
decreased with v .
According to Bernoullis law, the velocities
asymmetry brings about the static pressures
asymmetry (high pressure on the intrados, low
pressure on the extrados) as well as the arising of
lift force.
Applying Bernoullis relation between a point
at

and a point on the profile, we get:


2 2
2 2
S
S
v
p
v
p

+ +

. (9.3)
The pressure coefficient is defined by the
relation:
2
2
2
1
2

v
v
v
p p
C
S S
p

. (9.4)
In fig. 9.6 it is shown the distribution of
pressure and of the pressure coefficient on a
hydrodynamic profile at a certain angle of
incidence,
*
.
120
Fig.9.6
The alteration of the incidence angle leads to
the shift in the pressures distribution.
9.3 Forces on the hydrodynamic
profiles
The forces which act upon hydrodynamic or
aerodynamic profiles: lift, shape resistance, friction
force or the force due to the detachment of the
limit layer give a resultant
R
which decomposes by
the direction of velocity in infinite and by a
direction which is perpendicular on it (fig.9.7).
121
The angle between and the chord of the p
rofile.
Component
x
R
is called resistance at advancement,
and component
y
R
, lift force.
They are usually written in the form:
.
2
;
2
2
2
S
v
C R
S
v
C R
y y
x x

(9.5)
where
x
C
is called the coefficient of resistance at
advancement, and
y
C
the lift coefficient (
l c S
for
profiles of constant chord).
Fig.9.7
122
Force
R
can also decompose by the direction
of chord (component
t
R ) and by a direction
perpendicular on the chord (component
n
R ).
These components may also be expressed
with the help of coefficients:
t
C
- the coefficient of tangent force and
n
C
- the
coefficient of normal force.
For a certain angle
s ,
is the distance
between the leading edge and the pressure centre
(the application point of hydrodynamic force).
The relation expresses the moment of the
force R with respect to the leading edge:
sin cos s R s R s R M
x y n
+
. (9.7)
Also, moment M can be expressed by an
analytic form similar to that used for the
components of hydrodynamic force:
S
v
c C M
m
2
2

. (9.8)
Using (9.5), (9.7), and (9.8), we get:
sin cos
x y
m
C C
C
c
s
+

. (9.9)
In the case of small incidence angles:
y
m
C
C
c
s

. (9.10)
123
The usage of coefficients
x
C
,
y
C
and
n
C
is
often met in actual practice. Their variation is
studied in different conditions and given in the
form of tables and graphics of great importance for
the calculus and design of systems, which deal with
profiles.
Coefficients
x
C
,
y
C
and
n
C
depend on the
following main elements:
- the shape of the profile;
- the span of the profile (finite or infinite,
finite of small span or great span);
- the type of the flow (Reynolds number);
- rugosity of surfaces;
- the angle of incidence.
For each shape of profile, at certain different
relative elongation, , (see paragraph 9.1), in the
case of certain flow velocities (numbers Re
variable), there are diagrams experimentally
established
( ) ( ) ( )
m y x
C and C C ,
.
Fig.9.8
124
In fig. 9.8 there are plotted the diagrams of
coefficients for resistance at advancement and for
lift force for a NACA 6412 profile, of relative
elongation 3, at a number Re = 85,000.
Another type of diagram often used is the
polar profile, namely the function
( )
x y
C C
at different
slanting angles (fig.9.9). The polar allows us to
define two characteristics of the profile:
- the floating or gliding coefficient:
y
x
C
C
tg
, (9.11)
- aerodynamic accuracy:
x
y
C
C
f

1
. (9.12)
Fig.9.9
125
9.4 Induced resistance in the case of
finite span profiles
For wings of great span, considered infinite
l , the motion around the profile is plane.
Circulation may be replaced by a whirl.
In reality, at
the tips of the
wing, because of
the difference in
pressure, there
arises a motion
of fluid from
intrados to
extrados (9.10).
The greater the
weight of this
motion, the
smaller the wing
span is.
Fig. 9.10
As a
consequence,
circulation is
no longer
constant; at the
tips there is a
minimum.
(fig.9.11).
This leads to an
alteration of
hydrodynamic
parameters,
through the
126
arising of the so-
called induced
resistance.
Fig.9.11
In fig.9.12 the scheme of hydrodynamic forces
for the wing of finite span is plotted.
Due to the arising of an induced velocity
i
v ,
created by the free whirl, perpendicular on the
velocity in infinite

v , the resultant velocity


becomes:
i v v v +
. (9.13)
Fig.9.12
As a consequence there will appear an
induced incidence angle
i

, which thus decreases


the incidence angle

.
The alteration of direction and value of
velocity bring about the corresponding alteration of
lift, which, as we have already shown, is
perpendicular on the direction of stream velocity.
127
If
y R is the lift of the infinite profile and
F
is
the lift under the circumstances of an induced
velocity (perpendicular on the direction of velocity
v
), then:
. cos
; sin
i y
i i
F R
F R

(9.14)
In the conditions of very small values of
i

, we
may assume that
F R
y

, namely lift does not alter.
Component
i
R
acting on the direction Ox is
called induced resistance and may be written in the
form:
S
v
C R
xi i
2
2


. (9.15)
The total resistance of the wing of infinite
span is the sum between the resistance of wing of
infinite span
x
R
and the induced resistance
i
R
.
9.5 Network profiles
Several profiles that are in the stream of fluid
are in reciprocal influence, behaving in a different
manner within the assembly, rather than solitary.
Networks of profiles are often met in practice in the
hydraulic or pneumatic units, propellers, etc.
To study the behaviour of profiles in network,
let us consider a system made up of several
identical profiles, of span l and control contour
ABCD (fig.9.13). The pitch of the network is t.
128
Fig.9.13
129
Velocities v in points 1 and 2 have the
components
x
v
and
y
v
, according to the system of
axes shown in the figure. Assuming that the
density of fluid doesnt alter in a significant way
when passing through the network,
2 1
, then
2 1
x x
v v
.
Indeed, applying the equation of continuity:
l t v l t v m
x x
2 1
.

, (9.16)
it results
x x x
v v v
2 1
.
We have denoted by m the massic flow.
Applying the theorem of impulse, we get
component
y
R
of the lift force in the network:
( ) ( )
2 1 2 1
.
y y x y y y
v v l t v v v m R
. (9.17)
The circulation of velocity on the control
contour will be:

+ + +
ABCD
C
B
D
C
A
D
y
B
A
y ds v ds v ds v ds v ds v
2 1
. (9.18)
The integrals on the segments of contour BC
and AD cancel reciprocally. There only remains:
( ) t v v ds v ds v
D
C
y y y
B
A
y


2 1 2 1
. (9.19)
Therefore:
130
t
v v
y y


2 1
. (9.20)
Replacing (9.20) into (9.17), we get:
l v R
x y

. (9.21)
The axial component
x
R
is due to the
difference of pressure:
( ) t l p p R
x 2 1

. (9.22)
Applying Bernoullis equation between the points 1
and 2, we get:
2 2
2
2
2
2
1
1
v
p
v
p

+ + , (9.23)
or else:
( ) ( )
2 2 2
1 2 2
1
2
2
2
1
2
2 2 1
y y
y y
v v
t
v v v v p p
+



. (9.24)
Replacing (9.24) into (9.22). we get:

+
l
v v
R
y y
x
2
1 2
. (9.25)
The resultant force will be:
( )
t l
v
C
v v
v l R R R
r
y y
x y x r
2 4
2
2
2 2 2 2 1

+
+ +
. (9.26)
In relation (9.26) we have denoted by
r
C the
coefficient of the network and by

v the mean
velocity in the network (fig.9.14).
131
( )
4
2
2 2 1
y y
x
v v
v v
+
+

. (9.24)
Fig.9.14
The lift force is perpendicular on

v .
Coefficient
r
C is different from the hydro-
aerodynamic coefficient corresponding to a
separate profile.
132

You might also like