Вы находитесь на странице: 1из 42

Table of contents

3. Hydrodynamics
3.1  General considerations �����������������������������������������������������������������������������001

3.2  Basic conceptions �������������������������������������������������������������������������������������001

3.3  Continuity equation �����������������������������������������������������������������������������������002

3.4  Energy equation (Bernoulli equation) ������������������������������������������������������003

3.4.1.  Static pressure energy px  ���������������������������������������������������������������004

3.4.2.  Potential energy (energy of the situation or height) �������������������������004

3.4.3.  Kinetic energy or speed energy �������������������������������������������������������004

3.5  Flow out of an open vessel �����������������������������������������������������������������������006

3.6  Flow from a nozzle �������������������������������������������������������������������������������������007

3.7  Impulse equation ����������������������������������������������������������������������������������������008

3.7.1.  Vertical push against a wall �������������������������������������������������������������009

3.8  Viscosity of a fluid �������������������������������������������������������������������������������������010

3.9  Reynolds law of similarity �������������������������������������������������������������������������014

3.10  Flow patterns ��������������������������������������������������������������������������������������������015

3.11  Flow in piping systems ����������������������������������������������������������������������������016

3.12  Friction losses at laminar flow ����������������������������������������������������������������017

3.13  Friction losses at turbulent flow �������������������������������������������������������������018

3.14  Special resistors in pipes ������������������������������������������������������������������������019

3.15  Valves as pipe resistors ��������������������������������������������������������������������������022

3.16  Throttling of incompressible fluids ��������������������������������������������������������023

3.17  Energetic consideration of the throttling process ��������������������������������025

3.18  Operation of pumps in a pipe line system ���������������������������������������������026

3.19  Pump and System Curves �����������������������������������������������������������������������033

3.19.1.  Parallel pumping ����������������������������������������������������������������������������033

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Table of contents

3.19.2.  Series Pumping �����������������������������������������������������������������������������035

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
General considerations Page 001 of 040

3.  Hydrodynamics
3.1  General considerations
Mechanics is the teaching of strengths and effects on bodies which are ex-posed to these
strengths. Mechanics is part of natural science and builds up on theoretical laws and
experiences by scheduled observations of nature’s phenomena. The bodies which play a
roll in the analysis of mechanics (e. g. the effects on a motor vehicle at a crash test), have,
in this case, solid structures.
Hydromechanics considers the effects and laws of flowing liquid media. The name
hydromechanics is derived from the Greek language and means exactly "the teaching
of the mechanics of water". Since water an outstanding meaning and powerful place
in engineering, and because of the fact that many results of experiments are based on
measuring with water, the term hydromechanics is well earned.
Hydromechanics contains two head areas: Hydrostatics and hydrodynamics. Hydrostatics
deals mainly with static strengths of a liquid which acts on a vessel or, foe example,
the wall of a pool. The teaching of hydrodynamics describes the theoretical bases and
equations which are needed to be able to calculate the strengths in a fluent body or
medium.
This includes the liquid flow in pipes fittings and valves. The flow in gutters or open channels
and even rivers can be also described and calculated with the help of hydrodynamics.
This book concentrates on linear and/or one-dimensional flow conditions which appear,
for example, in pipes with a constant cross-sectional area.

3.2  Basic conceptions


The laws of liquid movement in technical systems, e. g. in pipes, heat ex-changers, pumps
or valves, are combined under the concept of hydraulics. It is about the interactions of
flow rate, pressure and fluid velocity which can be determined. The measuring and/or
recording of the fluid friction within a technical system is more difficult. The static pressure
energy produced by a pump or an overhead reservoir is reduced by the friction of the
liquid at the boundaries of a pipe or valve according to the flow rate. This behavior is
comparable with another law of nature defined by the German physicist Simon Ohm and
which was named after him.
The Ohm law reads:
U
I=
R
It means that the current I depends on the voltage U and the resistance R. A definite current
- comparable with the flow rate of a liquid - can either be reached by a higher voltage or
by a lower resistance. The electrical voltage corresponds to the pressure of a pump in a
hydraulics system. The hydraulic resistance is proportional to the reciprocal value of the
free cross-sectional area of a pipe or a valve body. It is therefore plausible that a large
free cross-sectional area produces less resistance against a fluent medium in a valve as
an arbitrary restriction to a fraction of the free cross-sectional area corresponding to the
nominal valve size.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 002 of 040 Chapter 3.  Hydrodynamics

Added to this, is the fact that the fluid medium in a valve body, is usually forced into
multiple direction changes which causes additional friction losses and increases the total
resistance.
It is normal to consider liquid flow at first, as an ideal model which doesn’t cause any friction.
A customization of the model is then carried out to examine a more realistic relationships
by the introduction of hydraulic friction coefficients. There are two basic relationships on
which all calculations of resistances build up in hydraulic systems: The flow or continuity
equation and Bernoulli’s equation named after the Swiss physicist Daniel Bernoulli.

3.3  Continuity equation


The continuity equation means that the product of cross-sectional area and velocity
remains constant in a pipe with an arbitrary cross-section. This fact is explained below in
more detail (Figure 3.3.-1).

Cross section A Velocity v

Figure 3.3.-1:  Flow in a pipe with a constant cross-sectional area

The resulting flow rate Q at a velocity w can be calculated according to equation 3-1:
Q = A⋅v (3-1)

If the physical unit is m2 for the cross-section and m/s for the fluid speed, the flow volume
unit becomes m3/s. When the pipe cross-section is changed at an arbitrary value, the
flowing speed of the liquid must change too, since the liquid is incompressible. The mean
velocity of the fluid at constant flow can be calculated at any location, if the accompanying
cross-sections are known (Figure 3.3.-2).

Cross section A1
Cross section A2
Velocity v2
Velocity v1

Figure 3.3.-2:  Liquid flow in a pipe with non-constant cross-sectional areas

At a reduced cross-sectional area Figure 3.3.-2) the fluid velocity must inevitably increase,
since the flow rate is constant at any point of the piping system. Hence the continuity
equation is:

A1 ⋅ v1 = A 2 ⋅ v 2 = constant (3-2)

Analogous to the law of Ohm the same flow appears in every cross-section of a closed
piping system.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Energy equation (Bernoulli equation) Page 003 of 040

3.4  Energy equation (Bernoulli equation)


The total mechanical energy of a body remains constant provided that energy isn’t
supplied or extracted from it.
p1
v1

v2
p2

h1
p3
h2 v3

h3

Figure 3.4.-1:  Schematic explanation of Bernoulli’s equation

This fundamental knowledge of the preservation of energy was formulated by the German
Robert Mayer for the first time. Daniel Bernoulli showed that these relationships apply
for a flowing liquid as well. Therefore one speaks about Bernoulli’s equation as far as
the energy content of liquid flow is concerned. (Figure 3.4.-1) shows the relationship
schematically.
„„ A liquid with the density ρ fills out the pipe completely and streams from above to
below. If looks one, for example, at the energy of a liquid particle in position 1, three
different energy forms can be distinguished:
„„ Static pressure energy (p1) which is inherent in the liquid and can be measured by
means of a pressure gauge in the respective position,
„„ Speed (velocity) energy (v1) which depends on the mass of the particle and the flowing
speed of the liquid,
„„ Potential energy (h1) which results from the geodesic height h1. The reference line by
which the respective height is measured is fixed in this example.

Assuming a friction free flow, the energy content remains unchanged in positions 2 and
3. The potential energy has, for example, reduced by the amount (h1-h3). On the other
hand the velocity energy at the same point has considerably increased, so that the total
energy remains unchanged.
If one expresses the different energy contents by equations, the following relationships
apply:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 004 of 040 Chapter 3.  Hydrodynamics

3.4.1.  Static pressure energy px


The pressure in a piping system produced by a pump can be replaced by the weight
of the liquid column. If the liquid is water with a standard density of 1000 kg/m3 the
corresponding height is 10 meter equivalent to about 1.0 bar or 100,000 N per square
meter which corresponds to the unit of Pascal (Pa). Hence px becomes in accordance
with equation 3-3:
p x = hx ⋅ g (3-3)

If the unit of pressure height hx is in meter (m) and the density of the liquid is in kilogram
per cubic meter (kg/m3):
p x = 10 m ⋅ 1000 kg m3 ⋅ 9.81 m s2 ≅ 100,000 Pa

The pressure unit Pascal (Pa) will be used in all theoretical considerations. For actual
process values, however, and for firm examples, the unit (bar) is chosen. The conversion
is as follows: 1.0 bar = 1∙105 Pa or 100 kPa (kilopascal).
For the following considerations it is even simpler to assume water as medium and use
for the pressure heights, i.e. energy, the unit (m). By rearrangement of the equation 3-3
we get the equivalent pressure height hx which corresponds to the static pressure px:

px
hx = (3-4)
ρ ⋅g

3.4.2.  Potential energy (energy of the situation or height)


Equation 3-3 is naturally also applicable for the definition of the potential energy. Since
the different energies will be expressed only as heights (in meter), the energy of the
situation h1 can be directly used in the final equation.

3.4.3.  Kinetic energy or speed energy


A general physical equation of mechanics can be applied here. The kinetic energy E of a
liquid is:
1
E= m ⋅ v2 (3-5)
2

If one assumes a mass unit and looks at the kinetic energy of 1.0 kg of the liquid, the
kinetic energy becomes:
v2 (3-6)
2⋅g

After using the common unit for the fluid velocity (m/s) the kinetic energy becomes also a
pressure height. Hence the total energy of a liquid particle at any place within the piping
system is:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Energy equation (Bernoulli equation) Page 005 of 040

px v2
E= + hx + = const.
ρ ⋅g 2⋅g (3-7)

Unfortunately the conditions aren’t ideal in practice, i. e. in reality the amount of total
energy is not constant but reduces itself by the amount of the friction losses. This means,
that the total energy content of the liquid at position 2 of Figure 3.4.-1 is less than in
position 1 and the medium at position 3 has a lower energy content than in position 2.
Equation 3-7 must therefore be corrected under consideration of the unavoidable friction
losses hv and develops to be:

px v2
E= + hx + + hv = const. (3-8)
ρ 2⋅g

When summarizing this knowledge, one can make the following statements:
„„ The different energy forms are equal and can be expressed in pressure heights.
„„ Static pressure energy can simply be changed into velocity energy. In accordance with
the continuity equation the fluid velocity rises and the static pressure falls through a
narrowing pipe or valve.
„„ Velocity energy can, be changed into static pressure energy again, if the speed is
slowly reduced by a progressive extension of the pipe or valve cross-section.
„„ All energy conversions are at all times accompanied by losses. If the fluid velocity is
increased at the expense of the static pressure by a pipe reduction and afterwards
reduced by a pipe extension again, a lasting loss of static pressure appears.

All control valves make use of this effect, in order to reach a reduction of the static pressure
or a regulation of flow through the valve. The energy difference in front of and behind
the control valve arises from internal friction (turbulence) of the medium and is always
associated with noise generation. The acoustic power is proportional to the dissipation of
mechanical energy in the throttling valve.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 006 of 040 Chapter 3.  Hydrodynamics

3.5  Flow out of an open vessel


Based on the continuity equation (3-1), the flow through a pipe or valve can be calculated,
if the cross-sectional area (A) of the narrowest place and the fluid velocity (w) are known.
The determination of the cross-section is a relatively simple matter provided that the
geometric dimensions of the opening are available. The calculation of the speed, however,
is not so simple. As a practical application of the continuity equation the flow out of an
open vessel shall be considered:

Figure 3.5.-1:  Flow out of an open vessel (schematic)

An open container or vessel with a filling height h, measured from the surface of the liquid
to the end of the short outlet pipe is being emptied. Which short-term flow occurs, if the
inside diameter of the outlet pipe is 100 mm and the filling height h is 2.0 m?
The determination of the cross-section area turns out as follows:

D2 ⋅ π 0.1⋅ 0.1⋅ 3.14


A= = = 0.079 m2
4 4

The fluid velocity at the vessel outlet arises from the equation for the “free fall”, which is
applicable for the flow out of an open vessel:

v = 2⋅g⋅h
(3-9)

This yields an outlet velocity as follows:

v = 2 ⋅ 9.81⋅ 2 = 6.26 m s

Without certain corrections, which will be explained later, the actual flow Q becomes:

Q = A ⋅ v = 0.079 m2 ⋅ 6.26 m s ~ 0.5 m3 s


To express the flow rate as normal, namely in cubic meters per hour (m3/h), the flow
works out to be - under consideration of the corresponding conversion factor - a value of
approximately 1,800 m3/h.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Flow from a nozzle Page 007 of 040

3.6  Flow from a nozzle


The basic knowledge from the previous example can also be applied to closed systems.
As an example a gardener, who wants to irrigate his plants with a garden hose. Even the
layman knows, in such a case, how to increase the penetration of water: he reduces the
cross-section of the hose and increases hereby the speed of water - which results in a
larger spraying distance.
In reality the facts are more complicated. A considerable pressure drop appears already
from a fully open water tap up to the front end of the hose, so that the speed of the
flowing water is reduced. Due to the narrowing of the tube not only the cross-section
decreases but the pressure drop also increases. These relationships will be explained in
detail later. A numerical example will clarify the facts. Given that the hose has a 30 mm
inside diameter and a nozzle of 5 mm diameter and the internal pressure is 3.5 bar. The
initial hose position and final hose position lie on the same geodesic height so that the
potential energy becomes zero.

Velocity v1
Velocity v2

Pressure p2
Pressure p1
Figure 3.6.-1:  Flow from a nozzle (schematic)

The pressure outside the nozzle corresponds to atmospheric pressure = 1.0 bar. How
much is the outlet velocity and the flow rate in liters per minute (l/min)? Also in this
example static pressure energy and velocity energy are first converted into heights. The
pressure height hx becomes, in accordance with equation 3-3 for water with a density of
1,000 kg/m3,
px 350000
hx = = ≅ 35 m
ρ ⋅ g 1000 ⋅ 9.81

The velocity energy in the tube still cannot be calculated since the flow rate is unknown,
but starting out from equation 3-7 one can, ignoring the friction losses, assume a pressure
height:
v 22 v12
= hx +
2⋅g 2⋅g

The pressure height at the nozzle outlet must be the sum of the energies which are in
the hose. Since the ratio of the cross-sections is very big (A1/A2 = 302/52 = 36), the term
w12/2g is ignored since it represents only a very low pressure height. The velocity w2 is
thus:
v 2 = 2 ⋅ 9.81⋅ 35 = 26.2 m s

When considering the continuity equation (3-2) the velocity in the hose becomes:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 008 of 040 Chapter 3.  Hydrodynamics

2
A2 D 
=v1 v=
2 v2  2 
A1  D1 

2
 5 
v1 =⋅
v2   = 26.2 ⋅ 0.028 =
0.73 m s
 30 

The flow rate results finally from equation 3-1:

0.0052 ⋅ π ⋅ 26.2
Q = A⋅v = = 0.00051m3 s = 30.9 l min
4

3.7  Impulse equation


The impulse equation describes the dynamic effects and reaction forces of a streaming
liquid. Mass and velocity are authoritative.

I = m⋅v (3-10)

The dynamic effects will be explained at two examples.


Reaction forces at the outflow from an open container or vessel

d
R v

Figure 3.7.-1:  Flow from a container

Under the assumption of a stationary flow, the impulse acts as a backward push R against
the direction v of the flowing medium (Figure 3.7.-1). Typical is the reaction force which
can be seen in the handling of the heavy hose of a fire brigade when they direct the hose
at a fire.
Mass is the product of volume and density. Mass per time unit pouring from the vessel
outlet is the product of the cross-sectional area and the velocity of the fluid:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Impulse equation Page 009 of 040

2
 = d ⋅π ⋅ v ⋅ ρ
m
4
The outlet velocity can be calculated according to equation 3-9:

v = 2⋅g⋅h

When combining both equations the reaction force (backward push) can be calculated:

d2 ⋅ π d2 ⋅ π
FI
= ⋅ ρ ⋅ v ⋅ 2 ⋅ g=
⋅h ⋅ ρ ⋅ v2 (3-11)
4 4

3.7.1.  Vertical push against a wall


Under the assumption that fluid velocity changes only its direction (90 degrees deflection)
but not its value, the same pressure height and the same reaction forces result, as in the
previous example Figure 3.7.1.-1.

FI
d v

Figure 3.7.1.-1:  Vertical push against a wall

The pushing force FI can be calculated as follows :

d2 ⋅ π
FI
= ⋅ ρ ⋅ v2 (3-12)
4

Since there is always a balance of forces the following rule applies:


Pushing force and reaction force are equal.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 010 of 040 Chapter 3.  Hydrodynamics

3.8  Viscosity of a fluid


Viscosity describes a fluid's internal resistance to flow and may be thought of as a measure
of fluid friction.
Various liquids show considerable differences with respect to their viscosity. Oil, honey
or liquid tar are, for example, considerably more viscous than water and therefore show
also a different flow behavior. The viscosity of a fluid arises from inner friction of the liquid
particles which must be overcome in order to flow. The viscosity strength F becomes
measurable, when a flat disk with an area A is pulled at a constant speed v along the
surface of the liquid with a layer thickness y (Figure 3.8.-1). In this experiment the liquid
is exposed to a shear stress or shear strain τ.
The speed reaches a maximum in the immediate proximity of the disk, i. e. the medium
has the same speed as the disk. On the other hand the speed of the medium becomes
zero in the immediate proximity of the firm counterpart or plate. A proportional distribution
of the speed is found between the extremes as shown in the illustration. The required force
for moving the disk along the firm plate can be calculated, if the area A, the viscosity η
and the quotient of dv/dy are known. In this case dy is the layer thickness of the liquid.

Plate
F
y τ vmax
dy
Fluid
dv
v0

Solid boundary

Figure 3.8.-1:  Definition of shear stress τ and dynamic viscosity η

F= A ⋅ τ (3-13)

The shear stress τ is defined as the friction force related to the area A.
dv
τ= η ⋅ (3-14)
dy

Dynamic viscosity η
The expression dv/dy is the quotient of velocity v and distance y and is described as the
dynamic viscosity η. The recommended SI-unit is Ns/m2.
The SI physical unit of dynamic viscosity is the pascal-second (Pa·s), which is identical
to kg·m−1·s−1.
If a fluid with a viscosity of one Pa·s is placed between two plates, and one plate is
pushed sideways with a shear stress of one pascal, it moves a distance equal to the
thickness of the layer between the plates in one second.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Viscosity of a fluid Page 011 of 040

The cgs physical unit for dynamic viscosity is the poise (P), named after Jean Louis
Marie Poiseuille. It is more commonly expressed, particularly in ASTM standards, as
centipoise (cP). Water at 20 °C has a viscosity of 1.0020 cP.
More common, however, are the units Poise (P) or Centipoise (cP).

1P =1 g ⋅ cm−1 ⋅ s−1 1cP = 0.001Ns m2

N⋅ s mN ⋅ s
1Poise =
(P) 0.1Pa=
⋅ s 0.1 1cP
= 1mPa=
⋅s 1
m2 m2

The kinematic viscosity ν


The kinematic viscosity ν of a fluid is frequently obtained through a relation of its density.
The expression

η
ν=
ρ (3-15)

is described as kinematic viscosity ν. The SI unit of kinematic viscosity is m2/s. But


the old units Stokes (St) and Centistokes (cSt) are still frequently used:

cm2 m2
1St 1= 10−4
=
s s
mm2 m2
1cSt 1= 10-6
=
s s

The viscosity of a liquid is determined by means of special measuring equipment where


the outflow time from a standardized receptacle is measured.

Non-standard units
Englergrad E (named after the inventor Carl Oswald Viktor Engler) is a unit of kinematic
viscosity and defines the ratio of outflow times of the test fluid compared to that of distilled
water. The volume of the receptacle is in this case 200 cm3, the temperature shall be 20
°C. A conversion into the recommended SI-unit is carried out as follows:

6.31
ν= ( 7.37 ⋅ E ) − ⋅ 10−6 m2 s (3-16)
E

Other common viscosity units are e. g. Saybolt-Universal and Redwood-Viscosity.


Saybolt-Universal is a unit which is common in the USA. Here also the outflow time t of
a certain quantity of the test liquid is determined by means of a special viscosity meter.
A conversion into the recommended SI-unit is carried out as follows:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 012 of 040 Chapter 3.  Hydrodynamics

 185 
=ν  0.266 ⋅ t −  ⋅ 10−6 m2 s
 t  (3-17)

The equation above is limited, however, to the following slowing times: Minimum 32
seconds, maximum 100 seconds. For slowing times above 100 seconds the following
conversion equation applies:

 135  −6 2
ν  0.22 ⋅ t −
=  ⋅ 10 m s (3-18)
 t 

In Anglo-American countries the Redwood-Viscosity is also common. The outflow time


is measured by a special viscosity meter at standard conditions. Similar to Saybolt-
Universal Viscosity (SUV) two different equations apply to convert this unit into the
recommended SI-unit.
 179  (3-19)
= ν  0.26 ⋅ t −  ⋅ 10−6 m2 s
 t 

On condition that: 34 s < t < 100 s


For outflow times more than 100 s applies:
 50  (3-20)
=ν  0.247 ⋅ t −  ⋅ 10−6 m2 s
 t 

The following explanations will help to illustrate the properties of a liquid or a gas
with regard to their viscosity:
„„ The viscosity is usually not a constant value. It depends rather on the temperature and
sometimes also on the pressure of the fluid.
„„ The temperature effect has a larger influence on the viscosity than the static pressure
and is different for liquids and gases. The viscosity of liquids decreases at rising
temperatures (e.g. oil). For gases and vapors however, the viscosity increases at
higher temperatures.
„„ The normally foreseeable effects on viscosity loses validity for certain media. Liquids
which behave according to equation (3-14) are described as Newtonian liquids. For
these media the shear stress remains constant, provided temperature and pressure do
not change. Most fluids used in the process automation industries (water, air, natural
gas etc.) belong to this group. Non-Newtonian fluids are liquids with solid particles
(slurries), slime, suspensions and highly viscous substances such as tar, special oils
or varnishes.
„„ Distinctive characteristics of a non-Newtonian fluid are either a variable viscosity or a
behavior which does not conform with equation 3-13. A typical example for a variable
viscosity is a so-called "thixotropic" varnish whose viscosity depends on the painting

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Viscosity of a fluid Page 013 of 040

speed with a brush. This effect is intended to avoid "tears" or "noses" when painting a
vertical body like a frame of a window. For such fluids a sufficiently low viscosity arises
in the context of a normal painting speed.
The high basic viscosity of such a varnish avoids, the slow vertical flow and the ugly
„tears“. Also special motor oils as used in motor vehicles today, fall into this category.
By adding certain components (additives) the oil has almost a constant viscosity
independent of its temperature. A mixture of liquids and solid substances is another
example for a non-Newtonian liquid which behaves differently compared with a pure
liquid.

Dynamic viscosity of various liquids at 25 °C:

Dynamic Viscosity η Dynamic Viscosity η


Pa·s cP
Acetone 3.06∙10-4 0.306
Benzene 6.04∙10-4 0.604
Castor oil 0.985 985
Corn syrup 1.3806 1380.6
Ethanol 1.074∙10-3 1.074
Ethylene glycol 1.61∙10-2 16.1
Glycerol 1.5 1500
HFO-380 2.022 2022
Mercury 1.526∙10-3 1.526
Methanol 5.44∙10-4 0.544
Nitrobenzene 1.863∙10-3 1.863
Liquid nitrogen 1.58∙10-4 0.158
Propanol 1.945∙10-3 1.945
Olive oil 0.081 81
* Pitch 2.3∙108 2.3∙1011
Sulfuric acid 2.42∙10-2 24.2
Water 8.94∙10-4 0.894
Table 3.8.-2:  Dynamic Viscosity η of various liquids

* Pitch is the name for any of a number of highly viscous liquids which appear solid. Pitch
can be made from petroleum products or plants. Petroleum-derived pitch is also called
bitumen. Pitch produced from plants is also known as resin. Products made from plant
resin are also known as rosin.
Dynamic Viscosity Unit Convertor:
See the Homepage: http://www.uniteasy.com/en/unitsCon/Dynamic_viscosity.htm

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 014 of 040 Chapter 3.  Hydrodynamics

3.9  Reynolds law of similarity


We owe the basic knowledge, gained from experiments on models, to the English
physicist Reynolds. If, for example, the friction of water has to be minimized for a big ship
or a large ocean steamer, no one would build a model of the original size but a model in
a reduced scale. On this model the flow pattern, friction of water and power consumption
needed, can be watched and measured. The main question after such a test is how far
the measurements at the model are transferable to the original. Reynolds found out, that
a geometric similarity between model and original must be fulfilled:
„„ Model and original must behave proportionally in all aspects of measurements exposed
to the medium (water).
„„ Also the surfaces exposed to the medium must be similar, i. e. should have the same
geometrical scaled „roughness“ as the original.
„„ Finally the flow pattern must be similar to and correspond with the actual conditions at
the experiment. This applies particularly to the mechanical parameters such as force,
speed and acceleration.

The greatest problem of the experiment is to reach a geometric similarity with respect to
the surfaces, since the original (e. g. a ship) is already relatively smooth on the outside
skin. Therefore a scaled reduction of the roughness is not always possible. The absolute
roughness k of a surface is usually defined as "middle peak-to-valley height". To make
a comparison between original and model possible, one makes use of the relative
roughness which results from the ratio of the absolute roughness k and a characteristic
length L. From this follows:
Two surfaces are geometrically similar if they show the same relative roughness
k/L.
If one examines the flow pattern of a liquid in detail, a differentiation between pressure
forces, inertial forces and friction forces is required. Utilizing these different forces and
corresponding equations we get a dimensionless expression defined as Reynolds number
Re. The following relation applies to a perfect circular pipe with a constant cross-section:
v ⋅ lq
Re = (3-21)
ν
In equation 3-21 v is the speed of the flow in (m/s), lq the length of the lateral dimension
(pipe diameter) in (m) and the kinematic viscosity ν in (m2/s). Since this book deals in the
first place with control valves, the flow behavior in valve bodies, fittings and pipelines are
of particular interest. The general rule is here:
Two hydrodynamic flow patterns are similar, if the contours and boundaries
exposed to the flow stream are geometrically similar and their Reynolds numbers
are the same.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Flow patterns Page 015 of 040

3.10  Flow patterns


Various flow patterns may appear in pipes and fittings with their specific characteristics.
The extremes are: Laminar flow and turbulent flow. The rules for the respective flow
pattern are in accordance with equation (3-21): Dimensions, velocity and viscosity of the
medium. The liquid particles move at laminar flow in well-ordered layers. The stream lines
in this classic test by Reynolds can be made visible, by using a transparent pipe to which
a thin colored beam (e. g. red ink) has been fed. The flow direction is clear in this case,
transverse motions and turbulences do not appear (Fig. 3.10.-1).

Injection of
red ink
Transparent glas pipe

Figure 3.10.-1:  Laminar flow pattern in a glass pipe (schematic)

If the fluid velocity is steadily increased, sooner or later sudden detachments and
turbulence of the medium occurs due to an increased skin friction at the pipe wall.

Injection of
red ink
Transparent glas pipe

Figure 3.10.-2:  Turbulent flow in a glass pipe (schematic)

Transverse motions now overcome the main flow direction, as represented schematically
in Figure 3.10.-2. A very small disturbance often causes a sudden change of the flow
pattern.
To make optimal use of pipes, fittings and control valves, turbulent flow predominates
in industrial applications. Laminar flow conditions occur only with very tough media and
control valves with very low throttling areas and small pressure differentials. The flow
pattern of pipes with a perfectly circular cross-section becomes predictable by introduction
of the critical Reynolds number Recrit = 2320. A flow below Recrit is (in pipes) always
laminar.
Also above this number a flow can still be laminar if no eddy causing tremors or other
disturbances appear. With control valves a transitional range can be frequently found
which is indicated by an overlapping of laminar flow and turbulence.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 016 of 040 Chapter 3.  Hydrodynamics

3.11  Flow in piping systems


To cause a flow of a liquid medium in a pipe, a driving force is required just as the electrical
analogy, to overcome the frictional resistance of the liquid particles. The necessary energy
for this is usually delivered by a driven pump or by a high static head.
The enlarged Bernoulli equation (3-8) is therefore completed by the so-called head loss hv.
The permanent expenditure of mechanical power is mainly converted into thermal energy
and cannot be re-gained. The pressure drop, however, comparable with the electrical
voltage drop at a resistance, can be calculated if the influencing parameters are known
(flow pattern, fluid velocity and pipe dimensions).
The head loss hv of straight pipes is obtained per unit of length e. g. per meter. This
dimensionless quotient Jg is obtained from the quotient of the head loss hv and the
considered pipe length l and is described as the slope of the hydraulic gradient:

hv (3-22)
Jg =
l

vmax

vmean
Figure 3.11.-1:  Velocity profile at laminar flow

vmax

vmean
Figure 3.11.-2:  Velocity profile at turbulent flow

The velocity profile in a pipe is not unique. In the immediate proximity of the pipe wall
the velocity is almost zero, while the liquid reaches a maximum in the center of the pipe.
In addition, the flow profile depends on the Reynolds number. As mentioned before the
laminar flow prevails at very low velocities. The flow profile for laminar flow conditions is
a parabola (Figure 3.11.-1). The highest fluid velocity is in the middle of the pipe while it
steadily decreases towards the outside. The mean speed of a parabolic velocity profile is
approx. 50% of the maximum value.
At turbulent flow conditions the velocity distribution is more proportionate than for laminar
flow as shown in Figure 3.11.-2. The maximum speed occurs - as for laminar flow - in the
center of the pipe. The mean speed, however, reaches fundamentally higher values than
at laminar flow with the exception of a very thin boundary layer near the wall, so that the
resultant mean speed reaches about 80 % to 87 % of the maximum speed in the center
of pipe.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Friction losses at laminar flow Page 017 of 040

The flow is, in the boundary layer, laminar. The friction losses in the boundary layer can
be calculated by means of equation (3-25). The friction losses outside the boundary layer
are of another nature and arise mainly from shock losses in the turbulent fluid.

3.12  Friction losses at laminar flow


The mean velocity, head loss hv and the drop of the hydraulic gradient Jg can be calculated,
if the velocity profile is known. At laminar flow conditions with a parabolic velocity profile,
the mean velocity results in the following equation (3-23):

1 (3-23)
v mean= ⋅ v max
2
Without examining the derivation of the following equation more clearly, the hydraulic
gradient Jg becomes, in such a case:
32 ⋅ν ⋅ v
Jg = (3-24)
g ⋅ d2

The necessary hydraulic gradient Jg at laminar flow is proportional to the kinematic


viscosity ν and the mean speed v, as well as reciprocally proportional to the square of the
pipe diameter. The head loss hv = Jg ∙ l in a pipe is therefore:
32 ⋅ν ⋅ v
=hv ⋅l (3-25)
g ⋅ d2

One can express the head loss by inclusion of the Reynolds number in a different fashion:

64 l v 2
hv = ⋅ ⋅ (3-26)
Re d 2 ⋅ g
The expression 64/Re is called the pipe friction coefficient λ.
As already mentioned laminar flow hasn't a great meaning in process industries, since it
doesn’t often appear. Examples for laminar flows are: Highly viscous fluids, suspensions
and flow through very narrow cross-sections. Due to the thick boundary layer with laminar
flow, the small grooves or scores on the pipe walls are covered so that the pipe roughness
doesn't have any influence on the head loss. This effect is contradictory to turbulent
flow where only a very thin boundary layer exists and near the wall there is already a
considerable velocity and head loss.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 018 of 040 Chapter 3.  Hydrodynamics

3.13  Friction losses at turbulent flow


At turbulent flow the main flow direction is superimposed by irregular cross-and push
movements. This causes collisions of liquid particles with different velocities. This again
causes considerable shock losses and consequently a measurable head loss hv. In
turbulent flow conditions the head loss increases with the square of the fluid velocity in
pipes, fittings, valves or other components. Under inclusion of the dimensionless friction
coefficient λ the head loss becomes:

l v2 (3-27)
hv = λ ⋅ ⋅
d 2⋅g
The friction coefficient λ is, however, not a constant figure but depends, on the one hand
on the Reynolds number Re, and on the other hand on the relative roughness k/d of the
pipe. With very smooth pipes where roughness can be practically neglected, there are
distinguishable flow patterns. For laminar flow holds:
64 (3-28)
λ=
Re
In the turbulent range two different equations apply depending on the actual Reynolds
number. If Recrit ≤ Re ≤ 105 applies:
0.316
λ= 4
(3-29)
Re

At even higher Reynolds numbers 105 ≤ Re ≤ 108 the following applies:


0.221 (3-30)
=λ 0.0032 +
Re0.237
Normally pipes are usually rough which requires, at turbulent flow, consideration of this
relative roughness. In such a pipe the peaks of the roughness rise above the thin boundary
layer at turbulence so that the head loss exclusively results from eddy causing effects. In
this area the friction coefficient λ becomes a constant and the head loss increases (in a
square law) with the fluid velocity. The following equation is used to determine λ:
1
λ= 2
 d  (3-31)
 2 ⋅ lg⋅ k + 1.138 
 

For rough pipes there are many possible conditions. At very low Reynolds numbers the
boundary layer is bigger than the roughness peaks. The flow pattern is therefore as in a
smooth pipe and the friction coefficient λ can be relatively simply determined. As already
mentioned the pipe roughness plays no role in laminar flow so here the equation (3-28)
is applied. The other extreme is a completely rough pipe in which the friction coefficient λ
becomes a constant of approx. 0.02. Values in between usually range from 0.02 to 0.04
but are very difficult to determine. They are therefore normally estimated.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Special resistors in pipes Page 019 of 040

3.14  Special resistors in pipes


An industrial piping system consists usually of straight pipes with various cross-sections,
elbows, T-pieces, diffusers and so forth and all of them represent a certain resistance to
the fluent medium. The process requires also adjustable resistances like control valves
or other throttling devices. All these components cause a pressure drop which has to be
taken into consideration during the layout of a plant. Authoritative for such a layout are:
flow per time unit, cross-section of the various pipe components, fluid velocity and the
characteristic individual friction coefficients (Zeta) of the various pipe components. The
total head loss of a piping system arises from the sum of the head losses of the single
components:

hvtotal = hv1 + hv2 + hv3 + ... hvn (3-32)

An exact determination of the total head loss requires a summation of all single resistances.
One can calculate the head loss in straight pipes according to the equation of chapter
3.13. Special resistances are mostly determined from corresponding tables which show
the empirically determined friction coefficients (Zeta) for the most common standard parts
like T-pieces, elbows or valves. Generally, the head loss of a component can be calculated
according to equation (3-33) when the individual friction coefficient ζ is known:

v2
hv= ζ ⋅ (3-33)
2⋅g

The frequently required pipe components for matching the pipe diameter to the size of the
corresponding valve inlet and/or outlet (e. g. diffusers) represent a special problem for the
determination of the head loss.
One must differentiate here between inlet and outlet head losses. The corresponding
values ζ1 and ζ2 are empirically determined by tests. The total head loss arises from the
algebraic sum of all resistance coefficients of the fittings or customization pieces:

Σζ = ζ 1 + ζ 2 + ζ B1 − ζ B2 (3-34)

If the pipe diameter at the valve inlet and outlet are the same, then the Bernoulli
coefficients B1 and B2 can be ignored, since they fall out from equation (3-34). When
utilizing customary reducers the pipe friction coefficients ζ1 and ζ2 can be determined
with the help of the following equations.
(a) Pipe reduction at the valve inlet only:
2
  DN 2 
ς 1 = 0.5 ⋅ 1 −    (3-35)
  D  

(b) Pipe extension at the valve outlet only:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 020 of 040 Chapter 3.  Hydrodynamics

2
  DN 2 
ζ 2 =1⋅ 1 −    (3-36)
  D  

(c) Pipe reduction at the valve inlet and pipe extensions at the valve outlet for the same
nominal pipe size before and behind the control valve:
2
  DN 2 
ζ 1 + ζ 2 = 1.5 ⋅ 1 −    (3-37)
  D  

If the nominal pipe sizes are different before and behind the control valve, then the
Bernoulli coefficients must be calculated separately and included in equation (3-34):
4
 DN 
ζ=
B 1.0 −   (3-38)
 D 
Sudden pipe extensions or reductions, as schematically shown in Figure 3.14.-1 and
Figure 3.14.-2, should be avoided to keep turbulences and head losses to a minimum.
Diffusers and extenders should have flat openings and reduced angles instead.

v1 v2

Figure 3.14.-1:  Sudden pipe reduction (schematic)

v1 v2

Figure 3.14.-2:  Sudden pipe extension (schematic)

The head loss in pipe diffusers or extenders depends further on the ratio of the cross-
sections of valve and pipe respectively and the ratio valve size/pipe size is also authoritative
for the resistance coefficient ζ of the fittings. An exact calculation of the head loss requires,
therefore, the knowledge of the typical resistance coefficients which can be taken from
corresponding tables. As an example the typical resistance coefficients at sudden pipe
reductions for different ratios of cross-sections A1/A2 are shown in Table 3.14.-3 below.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Special resistors in pipes Page 021 of 040

Ratio A1 / A2 0.1 0.2 0.3 0.4 0.6 0.8 1.0


ζ 0.46 0.42 0.37 0.33 0.23 0.13 0.0
Table 3.14.-3:  Resistance coefficients for sudden pipe reductions

Continuity equation
In a pipeline of variable cross-section the volume flow rate Q is constant throughout all
cross-sections over time t.

A1 ⋅ v1 = A 2 ⋅ v 2 or A ⋅ v = cons tan t

By smaller cross-sections the liquid streams faster and the other way round.

Example:

A1
A2

v1 Q1 Q2 v2

Figure 3.14.-4:  Gradual pipe contraction (schematic)

Pipeline with A1 = 19.6 cm2, A2 = 8.04 cm2 and Q = 120 l/min.


Calculate the speed v1 and v2.
Solution:

Q 120000 cm3 / min cm m


v1 =
= = 6122 = 1.02
A1 19.6 cm2 min s
v1 ⋅ A1 1.02 m / s ⋅ 19.6 cm2 m
=v2 = = 2.49
A2 8.04 cm2 s

v1 = 1.02 m/s
v2 = 2.49 m/s

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 022 of 040 Chapter 3.  Hydrodynamics

3.15  Valves as pipe resistors


Pipe systems and process plant include, usually, a number of manual On/Off valves.
Pipe systems with these shut-off devices can perhaps be for turning off certain branches
in emergencies without the need to have a shutdown of the whole plant. Due to limited
cross-sections and multiple changes of the flow direction these components cause
considerable pressure losses. Globe or gate valves are treated exactly the same way
as other components installed into piping systems. The basic equation (3-33) with
corresponding resistance coefficients ζ is also applied here. Table 3.15.-1 gives an idea
of frequently used valve types and their ζ values. As can be seen from Table 3.15.-1 the
resistance coefficients become higher the more deflections in the valve body occur. Ball
valves with a straight flow path behave more favorably than globe valves.

Name Description ζ-value


Symmetrical valve body with clearance volume
Standard globe valve 4.1
at the valve inlet.
Streamlined valve body without clearance
Streamlined globe valve 2.7
volume and optimized flow path.
Globe valve body with an inclined seat and
Y-style globe valve body 2.5
minimum deflection of flow path.
Optimized Y-style valve body with full bore seat
Full bore valve 0.6
ring and straight flow path.
Full bore ball valve with smooth and short
Ball valve 0.13
straight flow path.
Table 3.15.-1:  Resistance coefficients of various On/Off valves

Nowadays mostly the Kv or Cv value is used instead of the resistance coefficient ζ. The
relationship between these parameters will be explained later.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Throttling of incompressible fluids Page 023 of 040

3.16  Throttling of incompressible fluids

Inlet pressure p1 Differential pressure

Vapor pressure pv3


Inlet pressure drop
Oulet pressure p2
Oulet pressure drop
Vapor pressure pv2

Vapor pressure pv1


Lowest pressure vena contracta

Figure 3.16.-1:  Internal pressure distribution of a throttling valve (schematic)

Process automation requires accurate control of temperature, flow, pressure or level.


These tasks are mostly handled by control valves. A loss of pressure is caused deliberately,
in order to bring the process variable in line with the set point. When a pressure is reduced
one speaks about „throttling“. In this process the available static pressure energy is
changed into heat as the biggest portion. In order to understand the throttling process
better, the extended Bernoulli equations is re-stated again:

px v2
E= + hx + + hv = const.
ρ 2⋅g

This equation says, that the total energy consisting of pressure height + energy of situation
(geodesic height) + velocity energy + the total head loss caused by inner friction always
remains constant. Since the potential energy (geodesic height) can be mostly neglected
in this consideration, the following happens in a pressure reduction process by throttling:

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 024 of 040 Chapter 3.  Hydrodynamics

„„ The static pressure energy is changed into velocity energy by a deliberate cross-section
pipe reduction at which it reaches a maximum in the narrowest point (vena contracta).

„„ The accelerated flow meets, behind the throttling area, fundamentally slower liquid
particles or, in an extreme case, the solid walls of the valve body.
„„ This effect causes pressure losses through intense turbulence and an exchange of
energy. The liquid heats. This fact is easy to prove in a pump circuit with a control
valve as throttling element - as e.g. in the determination of the flow coefficient.
„„ Depending on the construction of the control valve, a certain part of the velocity energy
is again converted into static pressure energy. This effect is usually not desired and
can cause cavitation on. The events at the throttling process of a liquid are represented
schematically in Figure 3.16.-1.
„„ The liquid enters the valve with a pressure of p1. A small pressure drop appears already
at the inlet part of the valve body through deflection of the medium. The medium is
then strongly accelerated in the restricted seat area and this is accompanied by a
sharp decline of the static pressure. Behind the throttling area a certain pressure
recovery occurs, depending on the amount of deceleration. Finally, a small pressure
drop at the valve outlet occurs caused by repeated deflections of the fluid. The actual
measured differential pressure is the total pressure loss between inlet and outlet of
the valve.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Energetic consideration of the throttling process Page 025 of 040

3.17  Energetic consideration of the throttling process


Processes normally require accurate control, in order to keep the pressure on the valve
outlet absolutely constant independent of occurring disturbances. This requires a certain
differential pressure at the control valve and causes a non-retrievable loss of pressure
energy. On the other hand this energy loss should be limited to a minimum for ecological
and economic reasons. To cope with both demands skilled lay-out and design of the
control equipment are necessary. The mechanical power is mainly converted into heat
during the throttling process and can be calculated when the mass flow rate W, inlet
pressure p1, outlet pressure p2 and the density ρ of the medium are known. Under the
assumption of a complete transformation of pressure energy into velocity energy and vice
versa, the converted power Pth with W in (kg/h), p1 and p2 in (bar) and the fluid density ρ
in (kg/m3) for an incompressible media (liquid) will be:

W ⋅ ( p1 − p2 )
Pth ≅ (3-39)
36 ⋅ ρ

A concrete example will explain this principle idea of the converted power in a control
valve. A throttling valve is regulating the feed water supply in a closed loop on a boiler.
Given are the following process data:
Flow rate W : 120,000 kg/h
Inlet pressure p1 : 80 bar
Outlet pressure p2 : 2 bar
Valve size : DN 100
Density ρ : 1000 kg/m3
Converted power P : Approx. 260 kW
As can be seen from the example above, almost the complete power consumption of the
pump is used up by the valve itself during the start-up of the boiler. The extremely high
energy density within a relatively small control valve is remarkable and requires special
measures to keep wear and tear caused by cavitation and/or erosion to a minimum.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 026 of 040 Chapter 3.  Hydrodynamics

3.18  Operation of pumps in a pipe line system


A centrifugal pump is a kinetic machine converting mechanical energy into hydraulic
energy through centrifugal activity.

Figure 3.18.-1:  Centrifugal Pump (Foto KSB Type Etanorm)

Figure 3.18.-2:  Mechanical construction of Centrifugal Pump (Foto KSB Type Etanorm RSY)

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Operation of pumps in a pipe line system Page 027 of 040

Pump fundamentals:

System Curves
„„ Static Head
„„ Dynamic Head
„„ Pipe Friction
„„ Fitting Losses
„„ System Head = Static Head + Dynamic Head

Pump performance parameters


„„ Head
„„ Flow Rate
„„ Power
„„ Efficiency
„„ Net Positive Suction Head (NPSH)
„„ Characteristic Curves

Total differential head (TDH)


„„ TDH = Total Discharge Head - Total Suction Head
„„ Total Head = Discharge Pressure + Velocity head + Static head

NPSH
„„ Net Positive Suction Head (NPSH)
ƒƒ NPSH required (NPSHr)
ƒƒ NPSH available (NPSHa)
„„ NPSH is a measure of the energy (pressure) in a liquid above the vapor pressure
„„ If the pressure drops below the vapor pressure the liquid boils
ƒƒ That condition is called cavitation
„„ All pumps require the NPSHa to be > 0
„„ How much, is called the NPSHr

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 028 of 040 Chapter 3.  Hydrodynamics

Pump Curves
When the pump head discharge curve and the system head curve are plotted on the same
axes (as shown in Figure 3.18.-3), only one point lies on both the pump characteristic
curve and the system head curve. This intersection defines the pump operating point,
which represents the discharge that will pass through the pump and the head that the
pump will add. This head is equal to the head needed to overcome the static head and
other losses in the system.

Pu m p c h a r a c t
eristic
C u rv
e

Pump Operation Point


Head H

Head Losses

Cu rve
System

Static Lift

Flow rate Q
Figure 3.18.-3:  System operating point

In addition to the pump head-discharge curve, other curves representing pump behavior
describe power, water horsepower, and efficiency (see Figure 3.18.-4). Since utilities want
to minimize the amount of energy necessary for system operation, the engineer should
select pumps that run as efficiently as possible.
Another issue when designing a pump is the net positive suction head (NPSH1) required.
NPSH is the head that is present at the suction side of the pump. Each pump requires that
the available NPSH exceed the required NPSH to ensure that local pressures within the
pump do not drop below the vapor pressure of the fluid, causing cavitation. Cavitation is
essentially a boiling of the liquid within the pump, and it can cause tremendous damage.
The NPSH required is unique for each pump model, and is a function of flow rate.

1 NPSH = Net Positive Suction Head

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Operation of pumps in a pipe line system Page 029 of 040

Net positive suction head is the term that is usually used to describe the absolute pressure
of a fluid at the inlet to a pump minus the vapour pressure of the liquid. The resultant value
is known as the Net Positive Suction Head available.
The term is normally shortened to the acronym NPSHa, the ‘a’ denotes ‘available’. A
similar term is used by pump manufactures to describe the energy losses that occur
within many pumps as the fluid volume is allowed to expand within the pump body.
This energy loss is expressed as a head of fluid and is described as NPSHr (Net Positive
Suction Head requirement) the ‘r’ suffix is used to denote the value is a requirement.
Different pumps will have different NPSH requirements dependant on the impellor design,
impellor diameter, inlet type, flow rate, pump speed and other factors.
A pump performance curve will usually include a NPSH requirement graph expressed
in metres or feet head so that the NPSHr for the operating condition can be established.

Eff

Pump Efficiency, %
ic i
en
cy
cu
r

ve
Pu m
p dis BEP
cha
rge (Best Efficiency Point)
h ea
Discharge head

)
Head
on
e Sucti
NPSH

tiv
(Net Posi
NPSH

Flow rate Q
Figure 3.18.-4:  Pump efficiency curve

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 030 of 040 Chapter 3.  Hydrodynamics

System curve or characteristic


The system curve (or characteristic) is the graph of the system head HA required for the
installation shown as a function of the flow rate Q. It represents the flow resistance of the
system for the pump. It is made up of static and dynamic components.
The static components (static head Hgeo and pressure difference between the inlet and
outlet vessel (pout-pin)) are independent of the flow rate. The dynamic component consists
of the flow resistance in the piping system and grows as the square of the flow velocity
(Hv = f(Q2)).

Discharge

Suction
Engine

Pump
Figure 3.18.-5:  Pump Operating

Duty or operating point


Intersection of the system curve with the pump characteristic curve (for the actual speed
of rotation). The only possible working condition of the pump in a non-controlled system.
Each pump’s duty point is selected according to the mode of operation, considering the
efficiency of each pump and the pump system, as well as the NPSH of the pumps.
The choice of pump size for pumps working in parallel is therefore largely dependent on
the expected operating time. If it is planned that most often three pumps will be working
in parallel, then the best efficiency point (BEP) of the pumps should be chosen for this
condition (Qtot = 3 • Qopt, individual pumps).
However, it must be ensured that if a single pump is operated on its own, the increased
flow of that pump still falls within the permissible range, in particular with respect to NPSH
performance (see Figure 3.18.-4).

Net Positive Suction Head (NPSH)


Net positive suction head is the term that is usually used to describe the absolute pressure
of a fluid at the inlet to a pump minus the vapor pressure of the liquid. The resultant value
is known as the Net Positive Suction Head available. The term is normally shortened to
the acronym NPSHa, the ‘a’ denotes ‘available’.
A similar term is used by pump manufactures to describe the energy losses that occur
within many pumps as the fluid volume is allowed to expand within the pump body.
This energy loss is expressed as a head of fluid and is described as NPSHr (Net Positive
Suction Head requirement) the ‘r’ suffix is used to denote the value is a requirement.
Different pumps will have different NPSH requirements dependant on the impellor design,
impellor diameter, inlet type, flow rate, pump speed and other factors. A pump performance

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Operation of pumps in a pipe line system Page 031 of 040

curve will usually include a NPSH requirement graph expressed in metres or feet head so
that the NPSHr for the operating condition can be established.

Pressure at the pump inlet


The fluid pressure at a pump inlet will be determined by the pressure on the fluid surface,
the frictional losses in the suction pipework and any rises or falls within the suction
pipework system.
Fluid surface
pressure

Discharge

Positive
head

Negative
head
NPSHr energy loss
occurs within the
pump inlet system
Energy loss occurs
due to friction in the
pipework or within the
fluid.
Figure 3.18.-6:  Schematic for NPSHa

Fundamentals
„„ Gage Pressure (psig) Pressure above surrounding atmospheric pressure.
„„ Atmospheric pressure at sea level is 14.7 psig
„„ Absolute Pressure (psia) Pressure above an absolute vacuum.

NPSHa calculation
The elements used to calculate NPSHa are all expressed in absolute head units.
The NPSHa is calculated from:
Fluid surface pressure + positive head – pipework friction loss – fluid vapor pressure
or
Fluid surface pressure - negative head – pipework friction loss – fluid vapor pressure.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 032 of 040 Chapter 3.  Hydrodynamics

Understanding NPSHa and NPSHr


Positiv Positiv
Head Head

Pipework Pipework
friction friction
Fluid Fluid
surface surface
pressure Fluid pressure
vapour
pressure
Fluid
Zero NPSHa Zero vapour
pressure NPSHr pressure pressure NPSHr
absolute absolute NPSHa

NPSHa is sufficient to avoid cavitation NPSHa is too low cavitation will occur

Figure 3.18.-7:  Schematic for NPSHa with positiv head

In a system where the fluid needs to be lifted to the pump inlet , the negative head reduces
the motive force to move the fluid to the pump.
In these instances it is essential to size the supply pipe work and isolating valves
generously so that high frictional losses do not reduce the NPSHa below the NPSHr .

Negativ Negativ
Head Head

Fluid Pipework Fluid Pipework


surface friction surface friction
pressure pressure
Fluid
vapour
pressure Fluid
vapour
Zero NPSHa Zero
pressure
pressure NPSHr pressure NPSHr
NPSHa
absolute absolute

NPSHa is sufficient to avoid cavitation NPSHa is too low cavitation will occur

Figure 3.18.-8:  Schematic for NPSHa with negativ head

Comparison of NPSHa and NPSHr


All calculated values must be in the same units either m head or ft head.
If the NPSHa is greater than the NPSHr cavitation should not occur.
If the NPSHa is lower than the NPSHr then gas bubbles will form in the fluid and cavitation
will occur.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Pump and System Curves Page 033 of 040

3.19  Pump and System Curves


3.19.1.  Parallel pumping

Abstract:
At first sight, parallel operation of pumps does not seem to pose any
problems. Under real conditions, however, parallel pumping proves
to be rather more complex than typically portrayed in the relevant
literature.
So as to ensure that pumps and systems produce the required
flow rates and pressures detailed information is needed about the
real curve of both the pumps and the system. The duty point of the
individual pumps operated in parallel will shift and this will affect both
efficiency and NPSH performance. Only if these changes are taken
into account will a parallel pumping system function properly.

Figure 3.19.1.-1:  Pump system schematic (2 pumps operating in parallel)

Characteristic cu
H0 rve Pump
I + Pump
Chara II
cteri
stic c Operation Point
urve
Pum parallel
p I or
H Pu
mp
II Operation Point
single

System curve HA
Head H

Qsingle

Qparallel

Qp a ra lle l Qp a ra lle
=l QI + QII
Q
=I Q=
II
2
Flow rate Q
Figure 3.19.1.-2:  Parallel operation of two identical centrifugal pumps with stable characteristic curve

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 034 of 040 Chapter 3.  Hydrodynamics

Too many pumps in parallel


The form of the system characteristic must be carefully considered when an existing
facility is to be expanded, for example by installing further pumps in parallel. As shown
in Figure 3.19.1.-4 (presenting the system curve of a cooling circuit), the total flow rate
cannot be significantly increased by installing more pumps. This could result in the final
pump adding only a fraction of its capability to the system output as is indicated.

Figure 3.19.1.-3:  Three parallel pumps

3 pumps
in para
2 pum llel
ps in
paral
lel
Head H

Pu
mp
I

QI QI+II QI+II+III Flow rate Q


Figure 3.19.1.-4:  Pumps operating in parallel: low increase in flow rate with steep system curve

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Pump and System Curves Page 035 of 040

3.19.2.  Series Pumping


One arrangement is a Series Operation, where one pump discharges directly into the
suction of the second pump, with both pumps delivering the same flow rate, but sharing
in the development of the combined pressure.(see Figure 3.19.2.-1)
Fluid surface
pressure

Discharge

Suction
Main

Discharge
Suction
Booste

Figure 3.19.2.-1:  Pumps series operation

Series Operation
This arrangement is frequently used where a larger pump cannot operate with the NPSH
being made available from the system. A smaller pump is therefore installed upstream of
the larger one to boost the Suction Pressure to the larger pump.(see Figure 3.19.2.-2
and Figure 3.19.2.-3)
It should be noted that in the series operation, the only essential similarity between the
two pumps is that they must both be able to operate at the same flow rate. The pumps
can deliver totally different levels of head as long as they operate at the same capacity.
The ultimate example of series operation is the multistage pump where the first impeller
pumps into the second and then the third, etc. This results in a high pressure pump with
all the impellers operating at the same capacity.
In some pumping arrangements we can have multiple pumps that operate on two systems
that are closely combined, but are not in a true series operation and cannot be considered
as such.
One such example would be when one pump is supplying the flow and pressure for one
system, while the second pump is bleeding off the first system to deliver a lesser flow to
another system. While one pump is indeed discharging into the suction of the second
pump, it is doing so at a different flow rate. Consequently, they cannot be treated as
though they were operating in series.

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 036 of 040 Chapter 3.  Hydrodynamics

Pumps series operation


Falling Flow rate with outage of one pump with and without geodetic heights

Head H
Pump I + I
I

η
ien cy
fic
Pump I Ef
H dynamic

rve
cu NPSH
tem
Sys
~70% 100%
Flow rate Q
Falling Flow rate with
outage of one pump
Figure 3.19.2.-2:  Pumps series operation with geodetic heights = 0
Head H

Pump I + II

rve
cu
H dynamic

m
ste
Pump I Sy

η
y
H geodesic

nc
ie
fic

NPSH
Ef

~40% 100%
Flow rate Q
Falling Flow rate with
outage of one pump

Figure 3.19.2.-3:  Pumps series operation with Hgeodetic ≥ Hdynamic

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Pump and System Curves Page 037 of 040

Variable Speed Difficulties


In some rare cases pumps can be run in parallel when they do not have the same shutoff
head. As shown in Figure 3.19.2.-4, satisfactory operation can only be achieved when
the shutoff head of the smaller pump is not exceeded. For smaller flow rates the same
problem occurs as for a performance curve instability, i.e. the smaller pump is running but
cannot open its check valve, and the motor energy heats up the fluid in the pump.
A similar mode of operation and risk are present when a second pump is started up with
a variable speed drive in parallel to a pump already running.
Through flow in the second pump will only start when its shutoff head exceeds the operating
pressure of the first pump (intersection with the system curve). Except for this problem,
a variable speed pump is an ideal partner to a fixed speed pump for parallel operation.
Sufficient instrumentation with flow meters and pressure gauges is helpful in this case.
Shutoff head:
Head developed by the pump for zero flow rate (operating against a closed valve).

Head H

H0 II Pum
p II

H0 I Pum Pum
pI p I+I
I

System characteristic

Qmin

QI QII QIII Flow rate Q


Figure 3.19.2.-4:  Variable speed pump curves

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Page 038 of 040 Chapter 3.  Hydrodynamics

Summary the operating modes of Pumps

Series Pumping
„„ Heads H add at the same flow rate Q
„„ Second stage pump must be rated for discharge pressure
„„ Start up and shutdown procedures are critical
Head H

2 Pumps in series

Single Pump

Flow rate Q

Parallel Pumping

„„ At the same head H flow rates Q add


„„ Pumps must be matched for effective operation
„„ Provision must be made to observe minimum flow criteria
„„ Can be a good way to handle wide flow rate variations
Head H

2 Pumps parallel

Single Pump

Flow rate Q

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
Pump and System Curves Page 039 of 040

SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany


Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de
SAMSON AG · MESS- UND REGELTECHNIK · Weismüllerstraße 3 · 60314 Frankfurt am Main · Germany
Phone: +49 69 4009-0 · Fax: +49 69 4009-1507 · E-mail: samson@samson.de · Internet: www.samson.de

Вам также может понравиться