Download as pdf or txt
Download as pdf or txt
You are on page 1of 94

KSB

Start-up process ............................................................................................................................................ 3


Drive .............................................................................................................................................................. 4
Torque ........................................................................................................................................................... 4
Drive starting torque..................................................................................................................................... 4
Nominal speed .............................................................................................................................................. 5
Operating point............................................................................................................................................. 5
Hsys/Q change: ....................................................................................................................................... 5
Starting method ............................................................................................................................................ 7
DOL starting .......................................................................................................................................... 7
Star-delta starting ................................................................................................................................. 8
Auto transformer ................................................................................................................................ 11
Shaft coupling ............................................................................................................................................. 11
Rigid shaft coupling ............................................................................................................................. 11
Examples of rigid shaft coupling types ............................................................................................... 11
Shaft efficiency............................................................................................................................................ 15
Pump efficiency........................................................................................................................................... 15
Mechanical power ...................................................................................................................................... 16
Power input................................................................................................................................................. 17
Special terms pertaining to pump input power .................................................................................. 17
Electrical power .......................................................................................................................................... 17
Switchgear and controlgear ........................................................................................................................ 18
Application-specific switching and control functions ......................................................................... 20
Additional monitoring functions to safeguard reliable operation of the pumps ............................... 20
Apparent power .......................................................................................................................................... 20
Effective power ........................................................................................................................................... 20
Starting torque ............................................................................................................................................ 21
Factors influencing torque: ................................................................................................................. 21
Breakaway torque ....................................................................................................................................... 22
System characteristic curve ........................................................................................................................ 22
Location-specific subscripts ................................................................................................................ 23
Run-up time ................................................................................................................................................ 24
Rotational speed ......................................................................................................................................... 26
Specific speed ............................................................................................................................................. 27
Similarity conditions ................................................................................................................................... 29
Geometric similarity............................................................................................................................ 29
Kinematic similarity............................................................................................................................. 30
Dynamic similarity ............................................................................................................................... 30
Efficiency scale-up....................................................................................................................................... 32
Performance coefficient ............................................................................................................................. 33
Head coefficient .......................................................................................................................................... 33
Flow coefficient........................................................................................................................................... 34
Strouhal number ......................................................................................................................................... 35
Affinity laws ................................................................................................................................................ 35
Establishing characteristic coefficients by means of the affinity laws:............................................... 36
Affinity laws......................................................................................................................................... 36
Flow velocity in a cross-section .................................................................................................................. 38
Piping .......................................................................................................................................................... 38
Open system ....................................................................................................................................... 41
Pump power output.................................................................................................................................... 44
Economic efficiency .................................................................................................................................... 45
Life cycle costs ............................................................................................................................................ 45
LCC analysis − Key financial factors ..................................................................................................... 47
Pump system............................................................................................................................................... 48
Accumulator................................................................................................................................................ 48
Principal standards, directives and regulations applicable to accumulators...................................... 53
NPSH ........................................................................................................................................................... 53
NPSH required by the pump (NPSHr).................................................................................................. 57
When specifying the NPSHr, it is also necessary to provide information on the relevant cavitation
criterion. Criteria include: ................................................................................................................... 57
Head ............................................................................................................................................................ 59
Fitting .......................................................................................................................................................... 63
Bellmouth.................................................................................................................................................... 66
Intake chamber ........................................................................................................................................... 67
Head loss ..................................................................................................................................................... 68
Recommended flow velocities ............................................................................................................ 68
Head loss in a pipe .............................................................................................................................. 69
The equation for the head loss of a flow in a straight length of piping with circular cross-section is:
............................................................................................................................................................ 69
Head losses for plastic and smooth drawn metal pipes ..................................................................... 72
Head losses in valves and fittings ....................................................................................................... 72
For all fittings a differentiation must be made between two forms of pressure loss: ....................... 76
Influence of highly viscous fluids on the system characteristic curve ................................................ 77
Influence of non-Newtonian fluids on the system characteristic curve ............................................. 77
Axial thrust .................................................................................................................................................. 77
Mechanical axial thrust balancing....................................................................................................... 81
Design-based axial thrust balancing ................................................................................................... 81
Axial thrust balancing at the impeller ................................................................................................. 83
Axial thrust balancing via balancing devices ....................................................................................... 86
Inlet conditions ........................................................................................................................................... 87
Absence of swirl .................................................................................................................................. 88
Uniform velocity distribution .............................................................................................................. 90
Absence of vortices ............................................................................................................................. 91
Measures to be taken to prevent air-entraining vortices ................................................................... 93
Measures taken to prevent submerged vortices ................................................................................ 93
Disturbance-free approach flow examples ......................................................................................... 94

Start-up process
The start-up process refers the process by which the drive is activated and accelerated until the
operating point of the pump is reached. There are several starting methods to make this stage of
operation as low in impact as possible.

The torque (Tp) of the pump that is transferred by the shaft coupling is directly related to power
(P) and rotational speed (n). When the centrifugal pump is started up, the torque versus speed
curve is almost a parabola. The starting torque provided by the asynchronous motor, however,
must be greater for the rotor to reach operating speed.

This motor torque, together with the voltage, has a direct impact on the current drawn by the
motor and thus also characterises the build-up of heat in the motor winding. Impermissible heat
build-up in the motor should be avoided by limiting the run-up time and/or current.

Drive
Centrifugal pumps are usually driven by electric motors. Piston engines (e. g. diesel engines) and
gas and steam turbines also provide drive power, however. Hydraulic engines are rarely used.
Electric motors and turbines generate uniform torque, whereas piston engines produce non-
uniform torque. This irregularity is largely compensated by implementing appropriate design
measures (e. g. flywheels, changing the number of cylinders and their arrangement).

In the low power range (up to 1 kW), single-phase AC motors (see Alternating current) with
squirrel-cage rotors (see Asynchronous motor) are the preferred choice of drive.

The medium to high power range (up to 8000 kW) is dominated by three-phase motors (e. g.
asynchronous motors) with squirrel-cage rotors.

Another option in the top power range are synchronous motors due to their relatively high level
of efficiency and ability to compensate reactive power. In storage power stations, they operate as
generators in turbine mode.

Torque
Torque is a physical quantity and describes a force couple that acts on a body (e.g. impeller of a
centrifugal pump) and accelerates (start-up process) or decelerates it, or works against an
opposing torque of equal magnitude (starting torque). It is the product of force and lever arm. Its
symbol is T and the unit is N m.

Drive starting torque


In the context of drive systems, starting torque refers to the torque that is applied to accelerate
the machine to the nominal speed when it is switched on (see also starting method).
Nominal speed
Nominal speed (nN) is a suitable, rounded speed value for classifying the speed range (see
Nominal value). Nominal speed is no longer the speed previously agreed on in the supply
contract, but typically a deviating, rounded, and generally applicable speed such as 2,900, 1,450,
or 980 min–1 (rpm) for a mains frequency of 50 Hz or 3,500, 1,750, 1,180 min–1 (rpm) at 60 Hz.

Operating point
The operating point of a centrifugal pump is the intersection of the pump characteristic curve
(H/Q curve) and the system characteristic curve Hsys /Q. H/Q is the pump-based variable, Hsys/Q
the the system-based variable. See Fig. 1 Operating point

Fig. 1 Operating point:


Definition of a centrifugal pump's operating point

The operating point's position shifts if the position or the gradient of the pump characteristic
curve H/Q and/or the system characteristic curve.

Hsys/Q change:
H/Q changes but Hsys/Q remains unchanged:

• This takes place in the case of variable speed centrifugal pumps (see Closed-loop control) (See
Fig. 2 Operating point) or
Fig. 2 Operating point:
Operating point’s position changes from B1 to B3 on the system characteristic curve Hsys/Q,
resulting from an increase in pump speed from n1 to n3

• When centrifugal pumps of the same size are started up and operated in parallel. See Fig. 3
Operating point

Fig. 3 Operating point:


Operating point’s position changes from B1 to B3 on the system characteristic curve
Hsys/Q when starting a second and third identical pump in parallel

Hsys/Q changes but H/Q remains unchanged:

• The system characteristic curve may change during operation as a result of increased head
losses (e.g. throttling via check valves, pipe incrustations) or changes in static head (e.g. fluid
level fluctuations in tanks).
• Exact correspondence between the design and duty points (the latter referring to those
specified by the customer) and the operating points only exists in rare cases. The operating
point is often matched to the required data by throttling. See Fig. 4 Operating point

Fig. 4 Operating point:


Operating point’s position changes from B1 to B3 on the pump characteristic curve H/Q as a
result of increased throttling

Starting method
The squirrel-cage motors used for centrifugal pumps (see Asynchronous motor) have high
starting currents.
For motor ratings below 4 kW, the DOL starting and soft starting methods are used, while the
star-delta, auto transformer, soft starter and frequency inverter methods are preferred for motor
ratings above 4 kW.

DOL starting

For DOL starting, the three motor winding connections are wired in delta configuration from the
outset. This means that the full mains voltage is immediately applied to the stopped motor, i.e.
the entire starting torque is available right from the beginning. The operating speed will be
reached within a very short period of time.

This starting method is the most favourable one for the motor, even if the starting current
increases to 8 times that of the nominal current. Since this can place a demanding load on the
power supply mains when larger motors are involved and cause voltage dips for adjacent
devices, it is important to observe the provisions issued by the energy supply companies for DOL
starting of motors with ratings above 5.5 kW in public low-voltage mains (400 V).

In actual practice, motors with ratings up to 7.5 kW are also started directly. See Fig. 1 Starting
method
Fig. 1 Starting method: Motor
terminal board of a three-phase asynchronous motor: matching the supply voltage by selecting
the configuration via jumpers

Star-delta starting

Star-delta starting is used to drive machines with a high moment of inertia and limit the starting
current of an asynchronous motor connected in delta configuration.
For star-delta starting, the armature winding is initially connected to the power supply mains in a
star configuration, and the motor is brought up to speed in this configuration. At switchover,
delta current is theoretically all that is required and corresponds to the current rotational speed.

The result is a reduction in starting current of 1/3 as compared with delta DOL starting. The
same relation applies to torque. See Figs. 2, 4 Starting method
Fig. 2 Starting method: Starting
curve for current I and torque T of squirrel-cage motors in star-delta configuration (Y = star
configuration; ∆ = delta configuration; P = pump)
Fig. 4 Starting method: Starting
curve for current I and torque T for canned motors (designations as in Fig. 1)

Star-delta starting can only be used for three-phase motors whose winding connections are not
connected internally, but are routed separately outwards. The delta connection may only be
established after the machine has run up to speed in order for the targeted reduction in starting or
inrush current to be realised. The torque produced with the star configuration must be sufficient
to accelerate the driven machine to about its nominal speed. Switching from star to delta can be
effected manually or automatically.

In practice, the star-delta configuration comprises a contactor circuit that allows the motor
winding connections to be switched between the external conductors and the star point. Both
switching states are interlocked in operation. Automatic switchover is possible if additional
control relays are used. See Fig. 3 Starting method
Fig. 3 Starting method: Star-
delta configuration with contactors; simplified schematic without control and safety equipment

Switching the configuration from star to delta will cause current and torque peaks, which
increase the mechanical load placed on coupled components. Very smooth starting and stopping
can only be achieved via electronic solutions such as a soft starter or frequency inverter.

Auto transformer

An auto transformer typically finds application in high-powered motors to ease starting. To this
end, it reduces the voltage (and thus the starting current) supplied to electric asynchronous
motors. The transformation ratio of the transformer further reduces this current by the square of
the reduction. Auto transformers are the most frequently used type of starting transformer for
cost reasons.

Shaft coupling
The shaft coupling is the connecting element between the electric motor and the pump hydraulic
system. Slip-free shaft couplings employed in centrifugal pumps are divided into rigid and
flexible shaft couplings.

Rigid shaft coupling

Rigid shaft couplings are mainly used to connect perfectly aligned shafts. The slightest
misalignment results in considerable extra stress on shaft couplings and adjoining shaft ends.

Examples of rigid shaft coupling types

• Sleeve coupling
• Muff coupling
• Serrated (splined) coupling
• Split muff coupling (DIN 115)
• Disc coupling
• Flange coupling
• Gear coupling

Flexible shaft coupling

Shaft couplings to DIN 7400 are resilient (flexible), slip-free connecting elements fitted between
driving and driven components, capable of partially compensating axial, radial and angular
misalignment and shock loads.
See Fig. 1 Shaft coupling

Fig. 1 Shaft coupling:


Misalignment types

Flexibility is usually achieved by the deformation of dampening, rubber or metal-elastic spring


elements; their service life is heavily dependent on the extent to which misalignment has to be
compensated. Different flexible shaft coupling designs are available.
See Fig. 2 Shaft coupling
Fig. 2 Shaft coupling:
Coupling types

If shaft misalignment occurs between the driver and the pump as a result of, for example,
temperature fluctuations in the fluid handled (on heat transfer and hot water pumps), the double-
cardanic coupling type design is often employed.
See Fig. 3 Shaft coupling

Fig.3 Shaft coupling: Double-


cardanic coupling for compensating shaft offset

Gear couplings are flexible shaft connections for positive torque transmission and are
particularly suited to compensating axial, radial and angular shaft offsets.

The design principle employed by curved-tooth gear couplings (see Curved-tooth gear coupling)
prevents edge pressure when gears engage in the case of angular and radial offset, making these
couplings almost wear-free.
The double-cardanic operating principle of curved-tooth gear couplings ensures that the reaction
forces from angular and radial offsets are negligible and periodic fluctuations in the angular
velocity do not occur.
See Fig. 4 Shaft coupling

Fig. 4 Shaft coupling: Curved-


tooth gear coupling

A coupling with spacer sleeve (see Back pull-out design) allows the shaft seal and pump bearing
assembly to be removed without removing the pump casing and the drive.
See Fig. 5 Shaft coupling

Fig. 5 Shaft coupling: Spacer-


type coupling

If precise data on the influence of the frequency of starts and the ambient temperature is not
available, the load can be calculated on the basis of factors specified as reference values. No
fixed value is laid down for the ratio of maximum torque to operating torque. This means that all
coupling types can be taken into consideration according to their specific suitability. The
calculation of the loading by torque shocks relates therefore to the maximum torque (see Starting
torque).

The calculation method for sizing flexible couplings given in DIN 740 only applies on the
assumption that the coupling is the sole torsionally flexible element of a rotor, reducing the
installation to a linear two-mass system. In all other cases it is necessary to carry out a vibration
calculation.

Shaft efficiency
Shaft efficiency is also known as pump efficiency.

Pump efficiency
Pump efficiency (η) is also referred to as coupling or overall efficiency and characterises the
ratio of pump power output (PQ) to power input (P) for the operating point in question:

Pump efficiency (η) is the product of mechanical (ηm) and internal efficiency (ηi).

The best pump efficiency (ηopt) is the highest efficiency for the rotational speed and fluid
handled as specified in the delivery contract.

For centrifugal pumps whose mechanical design does not clearly separate the pump shaft from
the drive shaft, such as is the case with close-coupled pumps and submersible motor pumps, the
efficiency of the pump set (ηGr) is specified in place of pump efficiency (see DIN 24 260) (Gr
stands for group). This figure describes the ratio of pump power output (PQ) to the power
consumed by the driver (see Drive), which is measured at an agreed position (e.g. at the
terminals of the motor or where an underwater cable starts).
Achievable pump efficiency is very much a function of specific speed as well as the size and
type of the pump and increases as these two variables increase. Reference values for achievable
efficiency of modern centrifugal pump types are based on statistical analyses of the values for
existing
pumps. See Figs. 1 and 2 Pump efficiency

Fig. 1 Pump efficiency:


Attainable efficiency η of single-stage volute casing pumps without diffuser, and efficiency gain
Δη by using a diffuser, as a function of specific speed ns

Fig. 2 Pump efficiency:


Attainable efficiencies η of multistage high-pressure pumps (acc. to KARASSIK) as a function
of specific speed ns

Mechanical power
Mechanical power (P) is the quotient of mechanical work (A) by time (t). The SI unit of
measurement is watts (W).
Mechanical work (A) is performed when a force (F) causes movement in time (s) in the direction
of the force applied.

With respect to rotational movement, mechanical power is the product of torque (T) and angular
velocity (ω) (see Rotational speed).

Power input
The power input of a centrifugal pump (pump input power) is the mechanical power taken by the
pump shaft or coupling from the drive. The SI unit of measurement for power input is watts (W).
Input power

Calculation must be based on the flow rate (Q) at the inlet cross-section of the centrifugal pump
if the fluid handled exhibits substantial compressibility. Pump input power can also be defined
more precisely in conjunction with centrifugal pumps.

Special terms pertaining to pump input power

• Optimum pump input power (Popt): power input at the operating point of best efficiency
• Maximum pump input power (PG): highest pump input power of the operating range as defined
in the delivery contract
• Zero-flow pump input power (P0):
Pump input power when Q = 0 m3/s (see Characteristic curve)

Electrical power
Electrical power (P) is the product of voltage and amperage (active current). The unit is watts
(W).
Electrical power can be defined with respect to alternating and three-phase current, whereby a
distinction is made between the output types apparent power, effective power, and reactive
power (also see Power).

Electrical power is supplied to the consumer by the power company. Many electrical consumers,
such as alternating-current and three-phase motors, require effective (PW) and reactive power
(Pq). Whereas effective power is converted into mechanical power, reactive power is used to
build and dissipate the magnetic fields. It fluctuates periodically between generator and load. The
intensity of this energy over time is quantified by the reactive power.

Power companies must also provide apparent power (PS), since only active current Iw = I · cos φ
may be used to calculate power if a phase shift between current and voltage occurs about angle φ
in AC circuits. This current is in phase with the voltage, and the current component shifted by 90
degrees with respect to the voltage is the reactive current (Iq = I · sin φ). The product of voltage
(U) and (apparent) current (I) is apparent power (PS). Definitions:

Effective power Pw=U • I • cos φ


Reactive power Pq=U • I • sin φ
Apparent power Ps=U • I
Ps2=Pw2 + Pq2

Apparent power is typically specified not in watts (W), but in volt-amperes (VA), and reactive
power in volt-ampere reactives (var).

Fig. 1 Electrical power:


Correlation between reactive power, effective power and apparent power

Switchgear and controlgear


Switchgear and controlgear as defined by IEC 60947 for electronics applications encompass all
electrical components and devices that are used to activate, deactivate, and protect electrical
consumers and include switches, contactors, residual current devices, and motor protection
switches.

Pump control units are pump-specific electrical switchgears and controlgears that contain all
electric and electronic components required to safeguard proper operation of the connected
pumps. Models that start and stop one or more pumps as a function of the fluid level are typical
and are primarily used for single and dual-pump applications. See Figs. 1 and 2 Control unit

Fig. 1 Switchgear and


controlgear: Level control with level measured by float switches; dry-installed pumps (can also
be used for wet-installed pumps)

Fig. 2 Switchgear and


controlgear: Level control by continuous level measurement (pneumatic measurement
with/without compressor; wet-installed pumps)
Application-specific switching and control functions

• Equal distribution of operating hours


• Automatic pump changeover after a specified number of operating hours has been reached or
after every start
• Pump starting and stopping in response to service demand
• Pump changeover in the case of a pump fault
• Functional check run via battery-backed real-time clock (as a function of fluid level)
• Sequenced starting/stopping if both pumps have to be started or stopped, to prevent pressure
surges and minimise starting currents
• Freely selectable automatic re-start after a fault
• Adjustable after-run time
• Variable stop delays to prevent deposits in the tank

Additional monitoring functions to safeguard reliable operation of the pumps

• High water alert


• Operational availability
• Mains-independent alarm
• Programmable general fault/"in operation" message
• Phase monitoring
• Voltage monitoring and display
• Overload detection per pump
• Thermal monitoring of pump motors
• Sensor fault / Live zero
• Fault/warning per pump
• Low-load detection (e. g. for dry running or lack of water)
• Archiving of data of the last 30 faults
• Monitoring of adjustable service intervals

Apparent power
Apparent power (PS) is calculated based on effective power (PW) and reactive power (PQ):

Effective power
Effective power is the electrical power that can be converted to other mechanical, chemical or
thermal forms of power. The SIunit of measurement for effective power is watts (W).
Starting torque
Starting torque is the torque transferred by the shaft coupling during run-up (see Start-up
process). It is calculated based on the ratio of power (P) to angular velocity (ω) and is
represented as a rotational speed function.

Factors influencing torque:

• Progression of the characteristic curves for the head and pump power input in relation to flow
rate and rotational speed
• Position of the operating point on the characteristic curves
• Run-up behaviour of the drive as characterised by the run-up time (ta) of the unit (pump and
motor)
• System characteristic curve in relation to the valves fitted
• Run-up time (taQ) for accelerating the liquid mass in the filled piping

taQ Run-up time of the liquid mass in the piping in s


Q Flow rate in m3/s
H0 Shut-off head (head) in m
HA,0 Static part of the system characteristic curve in m
g Gravitational constant in m/s2
L Length of the piping in m
A Cross-sectional area of the piping in m2

To illustrate the possible progression of starting torque at low specific speeds, the head (H),
power input (P), and starting torque (TP) of a radial pump are examined under different operating
conditions. See Fig. 1 Starting torque

All starting torque curves (Tp) begin with the breakaway torque (TPL) to overcome bearing and
seal static friction. They reflect the increase in torque along with rising rotational speed (n) and
the increase in power input (P) as a function of increasing flow rate (Q). These processes occur
either at the same time or in succession.

In contrast to centrifugal pumps with low specific speeds, a different starting torque curve
develops at high specific speeds (e. g. propeller pumps) due to the increasing flow rate and
decreasing power input (characteristic curve). More starting torque is therefore required for
starting against a closed gate valve (points I and II would be located above line A-B) than for
starting against empty, unpressurised piping (point III would be located below the operating
point (B)). This characteristic must be observed.

Breakaway torque
Breakaway torque refers to the maximum torque required to set interconnected stator and rotor
components into motion. Instead of static friction forces, sliding friction forces will then apply.

System characteristic curve


The system characteristic curve (see Characteristic curve) represents the relationship between the
system head (Hsys) and the flow rate (Q). It is often parabola-shaped and does not generally pass
through the origin of the H/Q coordinate system. The curve becomes progressively steeper as
throttling increases. See Fig. 1 System characteristic curve and Fig. 4 Operating point

Fig. 1 System characteristic


curve: Diagram of a cooling tower system and system curves at variable water levels in the river

The intersection of the pump-specific H/Q curve with the system-specific curve Hsys/Q
determines the operating point. See Fig. 1 Operating point

The shape and position of the system characteristic curve result from the equation used to
determine the system head (Hsys):
p Static pressure
v Flow velocity
z Geodetic height
HL Head loss (see Pressure loss, Pressure head)
ρ Density of fluid handled
g Gravitational constant

Location-specific subscripts

e Defined inlet cross-section (suction tank)


See Head Fig. 2

a Defined outlet cross-section

s Pump suction nozzle

d Pump discharge nozzle

e,s Relate to the system's suction side, i.e. to the portion between the cross-sections e and s
See Head Fig. 2

d,a Relate to the system's discharge side, between the cross-sections d and a

The expression (va2 – ve2)/(2 ∙ g) is a negligible quantity if the system's cross-sections in e and a
are of adequate size or of approximately the same size.

In practice, this expression is seldom of any significance. The expressions (pa – pe)/(ρ ∙ g) and (za
– ze) are independent of the pump's flow rate (Q).

Therefore, the relationship between the system head (Hsys) and the flow rate (Q) is evidenced
mainly in the head losses (HL) which can be calculated by means of the following equation:

ζ Loss coefficient (Head loss)


v Flow velocity in a characteristic cross-section
(of cross-sectional area A)

As the flow velocity (v) is the quotient of the flow rate (Q) and the cross-sectional area (A), and
assuming a constant loss coefficient (ζ) and sufficiently high Reynolds numbers (see Model
laws), we have: HL ~ Q2.
The reason for the system curve's parabola shape becomes clear. For the vertex of the system
characteristic curve at Q = 0, we have:

From the above equation it follows that the system characteristic curve shifts vertically in the
H/Qsys coordinate system if the system's tank pressures (pa, pe) and the geodetic head Hgeo = za –
ze vary. Hsys,0 is often referred to as Hstat in the scholarly literature.

Thus for instance, we have the following equations for a cooling water system (see Cooling
water pump) comprising a pipe drawing water out of a river, a cooling water pump and a
discharge line leading into a cooling tower basin: See Fig. 1 System characteristic curve

HA = za - ze + Hv.e,s + Hv.d,a

HA,0 = za - ze

pa = pe = pb (see Atmospheric pressure)


va = 0 (negligible flow velocities at a)
ve = 0 (negligible inlet velocities in the intake structure from
the river)
HL Head loss (pressure losses at inlet and outlet, pressure
losses through valves or elbows,
pressure losses caused by pipe friction, passage
through the condenser
and by abrupt changes of cross-section etc.)
za – ze Difference in geodetic head of the water level in the
cooling water basin and in the river bed.

As the water level in the river (ze) fluctuates, the system characteristic curves will shift
accordingly.
See Fig. 1 System characteristic curve

Run-up time
The run-up time is the period of time during which the drive (see Drive) is started and
accelerated up to the operating point of the centrifugal pump. It is calculated from the
acceleration torque (Tb.mittel) averaged over the rotational speed (n):
The calculation made with this formula will only produce useful values, however, if the
acceleration torque is approximately constant across the entire speed range. Should the TM and
TP curves approach each other considerably at certain points, the run-up time must be calculated
in sequence using a computational or graphical method. To this end, the speed range is split into
sections (Δni) in which constant acceleration torque values (Tbi) are used for calculation. See
Fig. 1 Run-up time

The run-up time (ta) is the product of the sum of the individual steps: See Fig. 1 Run-up time

The run-down time (tdown) is calculated in the same way as the run-up time. The only difference
is that the acceleration torque (Tbi) is replaced with the starting torque TPi = f (n), which produces
a load, or deceleration, torque in this scenario:
Fig. 1 Run-up time:
Determining the run-up and run-down time of a centrifugal pump

Rotational speed
Rotational speed (also called speed, or speed of rotation) can be quantified as the number of
revolutions a rotating system makes within a defined period of time. The unit used for rotational
speed is s–1 (rev/s); pump speed is generally given in min–1 (rpm).

The rotating frequency of the pump shaft therefore characterises a pump's rotational speed (n). It
should not be confused with specific speed (ns) and is always defined as a positive figure.

The pump direction of rotation is specified as clockwise or anti-clockwise and is separate to the
defined direction of rotation of the impeller, which, when turning to the right with respect to the
direction of inflow, is clockwise.

The selection of pump rotational speed is closely related to the characteristics of the pump
hydraulic system (circumferential speed, impeller, specific speed), as the overall strength and
economic efficiency of the pump and drive system need to be taken into account.

Most pumps operate at rotational speeds between 1000 and 3000 rpm but frequently reach in
excess of 6,000 rpm with special gearing and turbine drives.

Larger centrifugal pumps (e.g. cooling water pumps for power stations), however, are typically
mated to slow-running electric drives that are very costly. Reduction gears between the drive and
pump maintain today's low pump speeds of just 200 rpm.

Rotational speed (n) is proportionate to angular velocity (ω), the latter of which is more
conducive to physical calculations and is the quotient of the plane angle and time interval. The
unit is rad/s. The rad (radiant) is equal to the plane angle (57.296 degrees), which intersects an
arc of 1 m in length as the centre angle of a circle with a 1 m radius.
This is represented with the number 1 in practice. The following relationship exists between
rotational speed (n) and angular velocity (ω):

Specific speed
The specific speed (ns) is a characteristic coefficient derived from the similarity conditions
which allows a comparison of impellers of various pump sizes even when their operating data
differ (flow rate and head at best efficiency point, rotational speed). The specific speed can be
used to classify the optimum impeller design and the corresponding pump characteristic curves.
See Fig. 1 Specific speed

Fig. 1 Specific speed:


Influence of specific speed ns on centrifugal pump impeller design; the diffuser elements
(casings) of single-stage pumps are outlined (blue).

Defined as the theoretical rotational speed at which a geometrically similar impeller would run if
it were of such a size as to produce 1 m of head (Hopt) at a flow rate (Qopt m3/s at the best
efficiency point, the specific speed is expressed in the same units as the rotational speed:
Qopt in m3/s Flow rate at ηmax
Hopt in m Head at ηmax
n in rpm Pump speed
ns in rpm Specific speed

A dimensionless characteristic coefficient in accordance with DIN 24260 can be established


using the following equation:

n in s rpm-1 Rotational speed


Qopt in m3/s Flow rate at η max
Hopt in m Head at η max
g = 9.81 m/s2 Gravitational constant

The following relationship exists between the numerical values of the dimensional and
dimensionless coefficients:

ns = 333 · ns*

The values to be inserted in the above equation are the optimum head Hopt for one stage in the
case of multistage pumps, and the optimum flow rate Qopt for one impeller half in the case of
double-entry impellers.

The fluid flow through the impeller changes with increasing specific speed, i.e. radial impellers
have low specific speeds, mixed flow ("diagonal") impellers have a higher specific speed range
and axial impellers have the highest specific speeds. Establishing the specific speed ns via a
graph:
See Annex, Specific speed, Fig 2
Fig. 2 Specific speed:
Establishing the specific speed nS via a graph

Diffuser elements on radial casings such as volute casings are also required to become larger and
larger with increasing specific speed as long as the flow can be guided through the impeller in a
radial direction. Eventually (i. e. at high specific speeds) the flow can only exit axially, e.g. via
tubular casings.

The specific speed's numerical value is also needed to select the influencing factors required for
the conversion of pump characteristic curves, for example, if fluids of higher viscosity or solids-
laden fluids are pumped).

In Anglo-Saxon countries the specific speed is called "type number K" in accordance with EN
12723 and ISO 9906. In the USA it is referred to as Ns (pump specific speed), with the flow rate
being specified in gallons/min, the head in foot and the rotational speed in rpm. The conversion
factors are:

K = ns / 52.9 and Ns = ns ∙ 51.6

Similarity conditions
The similarity theory requires that three essential conditions be met for hydraulic model tests:
geometric (length), kinematic (velocity) and dynamic (forces) similarity between the model (M)
and the full-scale version (G). Kinematic and dynamic similarity are grouped together under the
heading of physical similarity (see Affinity laws).

Geometric similarity

In order to fulfil the condition of geometric similarity, all the linear dimensions of the model
pumps (IM) and the corresponding dimensions of the full-scale version ("prototype") (IG) must
have the same ratio (mI; model scale):
The geometrically similar reproduction of a pump and its system environment in model form is
only required for sections relevant for the actual flow analysis. Establishing geometric similarity
on the discharge side of the system is irrelevant if the fluid flow analysis is confined to the
pump's suction side.

The full-scale version's wall surface roughness can only be reproduced in the model with limited
accuracy, which is insufficient to achieve a microscopic level of geometric similarity, meaning
that the boundary layer flow and the resultant pressure losses arising due to wall friction can only
be examined to a limited extent.

Kinematic similarity

Kinematic similarity requires the proportionality of the corresponding velocity vectors in the
model (vM) and the full-scale version (vG) (see Velocity triangle). The requirement of a constant
velocity scale can strictly speaking only be fulfilled in conjunction with geometric and dynamic
similarity:

Any deviation from geometric similarity will result in a roughly equal deviation from kinematic
similarity. In model tests, deviations from kinematic similarity often manifest themselves via
discrepancies between the degree of turbulence in the model flow and the flow in the full-scale
version. This degree of turbulence has an influence on the change from laminar to turbulent flow
(see Boundary layer, Fluid mechanics), on the possible occurrence of flow separation, and
therefore on flow losses. Often, these cannot be assessed with sufficient accuracy on a model.

Experience has shown that the different types and structures of the boundary layers in the model
and the full-scale version result in only minor deviations from kinematic similarity provided that
there are no significant differences in flow separation zones and investigations are not concerned
with areas close to surfaces such as those of vanes.

Dynamic similarity

In order to fulfil the requirement of dynamic similarity, a defined scale ratio (mf) must apply to
all forces (F) which determine flow phenomena in both the model (M) and the full-scale version
(G).

Apart from the two-phase effects in two-phase flow, the forces of significance in hydraulic pump
modelling are inertia, gravity, pressure and friction.
Dynamic similarity with regard to the inertia and gravity forces in a model and full-scale version
is expressed by the fact that the Froude number (Fr) is constant:

v Characteristic flow velocity


l Characteristic length
g Local gravitational constant

The same applies to the Reynolds number (Re):

ReM = ReG

v Kinematic viscosity of the fluid handled

Dynamic similarity with regard to the pressure and inertia forces present in the model and the
full-scale version is expressed by the same values of the Euler number (Eu):

EuM = EuG

P Characteristic pressure (pressure difference)


ρ Fluid density
g Local gravitational constant

In the case of centrifugal pumps, the Euler number expresses the relationship between the
pressure rise in the pump (i.e. the characteristic pressure difference) and the circumferential
velocity at the impeller outer diameter (u2) (i.e. the characteristic velocity) and is as such termed
the head coefficient. Achieving the same Euler number or head coefficients in both the model
and the full-scale version requires that both geometric and kinematic similarity are ensured and
that the same Froude and Reynolds numbers are present in both the model and the full-scale
version.
When transient flows with a frequency f are involved, the Strouhal number comes into play. In
hydraulic modelling, a frequent deviation from dynamic similarity arises from the fact that the
Froude or Reynolds numbers in the model and in the full-scale version are not the same due to
technical reasons relating to the tests. Many years of experience have enabled certain ranges of
these numbers to be obtained in the model and the full-scale version without substantially
impairing physical similarity (see Efficiency scale-up).

Efficiency scale-up
Efficiency scale-up takes account of the influence of the Reynolds number (Re number) on the
pump's efficiency. When two geometrically similar centrifugal pumps are compared, either the
larger one, the one with the higher speed or the one handling a fluid of lower viscosity usually
exhibits a higher pump efficiency. This means that when transferring data established on a test
model to a life-size prototype, the pump efficiency measured on the model must be scaled up.
The pre-condition is however that geometrical similarity has been maintained in all components
including surface roughness and clearance gap width (see Similarity conditions).

Thus this change in efficiency is only a consequence of the change in the Reynolds number
resulting from the change in pump size, rotational speed and viscosity.

Larger centrifugal pumps generally have higher Re numbers. According to the laws of fluid
mechanics, these pumps exhibit lower flow losses within certain limits and therefore feature a
higher internal efficiency.

As it is almost impossible to achieve exact geometrical similarity, it is important to take into


account that the additional influence of the machine size has the same effect as that of the Re
number. The influence of pump size on pump efficiency is of practical significance in all cases
where efficiency measurements are performed on reduced scale models of larger pumps (see
Pump test facility) This allows an evaluation of the anticipated efficiency of the life-size pump to
obtain the pump input power, for instance, which would have exceeded the installed test facility
power.

There are as yet no universally applicable rules for efficiency scale-up. In all cases, the
efficiency scale-up method used must be clearly defined between user and manufacturer before
the model test takes place. Examples of approximation equations for efficiency scale-up are
those given by Pfleidererand Ackeret:
The Re number should be calculated on the basis of the circumferential speed of the impeller
outlet diameter.

Practical examples for the use of efficiency scale-up equations are found in international
standards: "Hydraulic Institute: Standards for centrifugal, rotary and reciprocating pumps. 14th
ed, Cleveland 1983" and "IEC 497: International code for model acceptance tests of storage
pumps. 1976" (12/2011 edition).

Performance coefficient
The performance coefficient (λ) is a characteristic coefficient identifying the pump input power.
The following equations are yielded in conjunction with the pump power output (PQ) and pump
input power (P):

ρ Density of fluid handled


η Pump efficiency
U2 Circumferential velocity at the impeller outlet
D2 Impeller diameter at the outlet
b2 Impeller outlet width
φ Flow coefficient
ψ Head coefficient

Head coefficient
The head coefficient (ψ) is a characteristic coefficient derived from the corresponding physical
quantity according to the affinity laws and used to characterise the operating behaviour. It
characterises the head (H) of the pump:

When the head varies at constant rotational pump speed, ψ ~ H. The head coefficient (ψ) is
therefore indicative of the ordinate (analogous to H) on H/Q curves plotted in a non-dimensional
representation. In conjunction with the specific energy (Y), this results in:

Flow coefficient
The flow coefficient (φ) is a dimensionless quantity used to describe the volume flow rate. It is
also referred to as the volume or volume flow coefficient and characterises the flow rate (Q).

When the flow rate varies at constant rotational pump speed, vm ~ Q and therefore φ ~ Q.

The flow coefficient (φ) is therefore indicative of the abscissa (analogous to Q) on H/Q curves
plotted in non-dimensional representation. With reference to the vane inlet and outlet diameters,
the following equations are obtained:
With regard to the head coefficient, it is preferable to adopt the flow coefficient related to the
impeller outlet.

Pump Lexicon

Strouhal number
The Strouhal number is defined as follows:

w velocity of flow around a body (for propellers, the relative


velocity at the impeller outlet is used) in m/s
f vortex-shedding frequency (sought excitation frequency) in s-1
d characteristic quantity of the separating vortices (thickness of
profile surrounded by the flow) in m

The dimensionless number plays an important role in hydro-acoustics and characterises transient
flow phenomena in all cyclically operating prime

Sr = 0.2 to 0.24 (flow along the plate)

Affinity laws
When investigating flow phenomena, cost factors often favour the use of models which are
geometrically similar to the original, full-sized equipment (see Similarity conditions). For this
type of testing it is necessary that models are not only geometrically similar, but are also
subjected to similar physical conditions.

The physical laws (differential equations including boundary conditions) applied must remain
invariant under similarity transformations. This is achieved by dividing all relevant physical
quantities by exponential products characteristic of the configuration to be tested so as to obtain
ratios of the unit 1.
Physical similarity is achieved if the ratios (see Characteristic coefficient) of the original and the
model are the same. The relationships established between the physical quantity of the original
and that of the model by means of the characteristic coefficients are called affinity laws.

Using characteristic pump parameters such as the impeller diameter (D), rotational speed (n ),
gravitational constant (g) and the density of the fluid handled (ρ), various characteristic
coefficients can be established for a centrifugal pump assuming frictionless, incompressible,
non-cavitating flow.

Establishing characteristic coefficients by means of the affinity laws:

• Flow velocity (v / (n · D))


• Pressure (p / (ρ · n2 · D2))
• Specific energy (Y / (n2 · D2))
• Head (g ∙ H / (n2 · D2))
• Flow rate (Q / (n · D3))
• Power input (P / (ρ · n3 · D5))

The following model laws thus apply to two geometrically similar centrifugal pumps operating
under physically similar conditions:

Affinity laws

• Flow velocity in a cross-section

• Pressure

• Specific energy

• Head

• Flow rate

• Power input (assuming identical pump efficiencies)


• Flow coefficient

• Head coefficient

If Δhs = Y (see Specific energy) is inserted, the pressure coefficient in its known form can be
established:

As pump efficiencies are more or less dependent on friction conditions, they are subject to other
conversion laws (see Efficiency scale-up). The selection of characteristic quantities to determine
the characteristic coefficients is largely arbitrary. For instance, when studying the theory of flow
in radial impellers (see Impeller), the impeller's circumferential speed (u), its outlet diameter (D)
and the outlet width (b) of the vane passage are selected as characteristic quantities. These are
used to establish two characteristic coefficients, where Δhs is the isentropic increase (see
Entropy) of the generalised specific enthalpy of the fluid handled.

As a general rule, a length (l) and a velocity (v) are selected as characteristic quantities for flow
investigations. Flows subject to friction are characterised by the kinematic viscosity (ν). The
Reynolds number (Re) can be derived from these and also gives the ratio of inertia force to
friction force.

If gravity has to be taken into account as an external force, the characteristic coefficient of the
acceleration due to gravity (gravitational constant) is g · l / v2. Its reciprocal value is the Froude
number (Fr).

It expresses the ratio of inertia force to the force of gravity. If further physical phenomena such
as compressibility, heat transfer and surface tension etc. have to be taken into account, further
characteristic coefficients must be introduced.

As characteristic coefficients are not independent of one another, it becomes impossible to


achieve physical similarity when multiple characteristic coefficients require consideration.

Model tests are widely used to investigate fluid mechanics, design strength and heat transfer
problems.
Flow velocity in a cross-section
The flow velocity (v) in a cross-section is the volumetrically averaged flow velocity in a specific
flow cross-section (e. g. a pipe cross-section).

The unit of measurement for flow velocity is m/s.

Q Flow rate in m3/h

A Cross-selectional area in m2

This relationship produces standard reference values for determining the flow velocity in piping
(see Piping).

Piping
Piping is used to transport fluids. The inside diameters of piping are classed according to the
nominal diameters (DN). The permissible load capacities as determined by the maximum
internal pressures are classed according to nominal pressures (PN). The recommended upper
limit for the flow velocity (v) is approx. 2.3 m/s for discharge lines and approx. 1.8 m/s for
suction lines.

Economic efficiency should be taken into account when selecting the discharge-side velocities in
the case of long piping and extended periods of operation. Due to the fact that suction-side
piping is shorter in length, the NPSH conditions are particularly important for the selection of the
suction-side velocities. The selected suction line's inside diameter is often larger than the pump
suction nozzle.

Expansion joints are built into the piping system to absorb movements in the piping, whatever
their cause may be. As well as compensating movements, they also separate the pump from the
piping in order to prevent vibration transmission. Sometimes expansion joints are also used with
pumps to ensure that their connection to the piping does not result in the transmission of any
stresses or strains. See Fig. 1 Piping
Fig. 1: Piping: Expansion
joints' compensating movements (lateral compensation)

If expansion joints are employed, those used to connect the pump and the piping should be
friction-type expansion joints for the transmission of the axial forces (also braced expansion
joints). This is necessary to ensure that the forces resulting from differential pressures do not act
upon the pump, as these are often considerably higher than the permissible flange forces. These
forces also shift the pump towards the suction side. This would severely affect the pump's
alignment, as its mounting is not designed for this movement.

A distinction is made between closed and open piping systems. Unbraced expansion and
dismantling joints turn a technically closed pipeline into an open one.

Closed system

The closed system features friction-type connections for the transmission of axial forces, e.g.
flanges and rigid dismantling joints. The axial force arising from the internal pressure is
absorbed by the pipe wall and the pipe connections. The supports and fasteners of a closed
piping system are only required to handle its weight and dynamic forces. Thermal expansion is
absorbed by flexible pipe supports or by expansion joints. See Fig. 2
Fig. 2 Piping: Articulated expansion joint (lateral
compensation)

The approximate spans (ls) for water-filled steel pipes should be established using the following
equation:

The wall thickness (b) of steel pipes subjected to internal pressure are calculated in accordance
with EN 13480-3. For pipes mainly subjected to static load conditions, the following roughly
applies:

In the case of changes in temperature, the change in length of a straight pipe is calculated
according to the following equation:

Open system

The open system has socket joints, flexible dismantling joints and axial expansion joints without
tie bolts that provide compensation for thermal expansion. See Fig. 3 Piping
Fig. 3 Piping: Axial expansion
joint

As an external force, the axial force arising from the internal pressure must be absorbed by
anchorage points at the beginning and the end of the piping and at any change of direction or
cross-section.

Adequate guidance in the form of clips or roller bearings must be provided to keep the piping
from buckling.
Fig. 4 Piping: Pipe end
anchoring
Fig. 5: Piping: Branch
anchoring

Fig. 6 Piping: Elbow


anchoring

Pump power output


The pump power output (PQ) is the useful power transmitted to the fluid handled by the
centrifugal pump. The unit is watt (W).
It is defined as:

If the fluid handled is compressible, the density (ρ) is conventionally agreed to refer to the
condition in the pump suction nozzle or the arithmetic mean (ρs + ρd)/2.

Economic efficiency
Economic efficiency is the ratio between the measurable increase in value and the expenditure
required to achieve this increase. This concept is extremely important within the context of life
cycle costs.

Life cycle costs


Life cycle costs (also see LCC) are the total costs incurred throughout the service life of a pump
system. They are used to compare the economic efficiency of various technical designs. The
LCC equation is determined in accordance with the guidelines of EUROPUMP and the
Hydraulic Institute.

The life cycle costs associated with the operation of a pump or pump system are determined by
calculating the annual costs of operation plus the interest and depreciation for the non-current
assets, like the machinery and buildings, for a variety of alternatives.
In the case of centrifugal pumps, energy (Ce), operating (Co) and maintenance costs (Cm)
account for the largest proportion of the life cycle costs.
See Fig. 1 Life cycle costs

Fig. 1 Life cycle costs: Cost


breakdown throughout service life (example)

The energy costs are calculated as follows:

The calculation applies to one specific operating point only. Given that pump operation usually
involves a broad flow rate range, a pro rata calculation must be performed for the various flow
rates involved. The individual results are then added together, taking the load profile into
account.

The costs of operation and maintenance must be determined on a case-by-case basis and are
dependent on the level of automation, operating period and maintenance requirement of the
system.
LCC analysis − Key financial factors

• Energy price increase (inflation)


• Interest and discount rate
• Expected system life (calculation period)

Calculating the current costs associated with a specific cost element:

Based on empirical evidence, the following factors can be derived for dimensioning an
economically efficient pump system or piping:
See Fig. 2 Life cycle costs

Fig. 2 Life cycle costs:


Diagram (example) illustrating the payback period of flow adjustment by variable speed pump
drives (y) and by throttling of discharge-side valves (x), taking into account all life cycle costs
involved

• Having high flow velocities in narrow piping reduces the cost of the system, but increases energy
requirements and wear.
• If operating periods are long, the energy costs are the dominant factor. This means that any
extra costs for energy-saving measures such as speed adjustment (see Closed-loop control) pay
off very quickly.
• If operating periods are short, investments should be low and flow velocities can be high.
• A smaller number of fairly large pumps often produces higher levels of efficiency, resulting in
lower energy costs. However, as the number of redundant pumps increases so does the capital
expenditure.

Pump system
In centrifugal pump technology, the pump system, or simply the system, encompasses the space
through which the fluid handled flows, excluding the pump itself. A differentiation is made in
the system between the suction and the discharge side, where each side is equipped with an
appropriate tank and piping including all requisite valves.

The suction side of the system is situated between the system's inlet cross-section (Ae) and the
pump's inlet cross-section (As); the discharge side is located between the pump’s outlet cross-
section (Ad) and the system's outlet cross-section (Aa), which has to be specifically defined. See
Fig. 2 Head

The system's energy consumers include accumulators, coolers, condensers, high-level


distributing tanks, all of which account for a significant proportion of energy consumed, but also
piping as well as fittings and valves.

Accumulator
An accumulator is a vessel which is partly filled with liquid and partly with gas (often air); its
internal pressure is generally higher than atmospheric pressure. Accumulators store fluids to be
handled under increased pressure (e. g. in pressure booster systems) in order to attenuate surge
pressures and serve as energy storage devices to prolong the run-down time of centrifugal
pumps. A transient flow analysis determines the accumulators' size and the valves, compressed
air supply connections and instrumentation used.

Accumulators for automatic pressure control in water supply systems (see Pressure booster
system) are usually installed vertically; horizontal installations are rare. See Fig. 1 Accumulator
Fig.1 Accumulator: Automatic pressure control in water supply systems

Accumulator size is determined by the pump set's number of starts per hour (Z). The number of
start-ups depends on a variety of factors; information on the frequency of starts should be
obtained from the electric motor suppliers (see Frequency of starts).

At start-up pressure (pe), the lowest water level selected must ensure that air can under no
circumstances enter the discharge line. The accumulator volume (V) should therefore be selected
so that it is 25 to 40 % larger than the effective accumulator volume (J) required. A compressed
air shut-off valve may be provided as an additional component. Its purpose is to prevent
compressed air entering the discharge line. In the case of unfavourable piping layouts (e. g. in
domestic water supply systems) and horizontal vessels, the water level must be checked; if
necessary, the connection must be placed at a lower level (e. g. dome).

A safety allowance of 25 % is included in the equation given below for accumulator sizing.
Fig. 2 Accumulator:
correction value K

The proportion of usable water volume (S) in relation to total volume (V) depends solely on the
start-up and stop pressures and can be calculated as follows:
In set-ups with more than one of the same pump, increasing the number of starts and stops by
periodically switching between the pumps allows a reduction of accumulator size. Membrane-
type accumulators are often provided for smaller units; these eliminate the need for a compressed
air shut-off valve or a compressor. In this case, an extra 25 to 40 % of volume in addition to the
effective volume (J) is not required.

The number of pumps in a pressure booster system has no bearing on the calculation of the
accumulator volume. If several pumps with different flow rates are employed, the mean flow rate
of the largest pump should be used in the equation. For systems in which several pumps are
flow-controlled, and only the base load pump is started and stopped as a function of pressure, the
accumulator size should be calculated in relation to this base load pump.

A sub-division of the calculated accumulator volume between several accumulators is desirable


if such smaller vessels can be accommodated more easily in the available space, and the system
costs are thereby reduced. When dividing the volume between two accumulators, the pressure
settings for pump start-up and stopping can be set in such a way that the second accumulator is
filled with air only.

If the volume is divided between more than two accumulators, these must be connected via the
gas (air) side to ensure that each accumulator is evenly used. See Fig. 3 Accumulator
Fig. 3 Accumulator: Schematic for a water supply system as pressure booster system

As a proportion of the accumulator's air content is gradually absorbed by the water under
pressure, the compressed air in the vessel must be topped up from time to time, usually by means
of a compressor. The compressor size is determined by its suction capacity (Qk). Compressor
selection depends on the time (T) required to fill the whole accumulator volume. It is assumed
that only two thirds of the accumulator volume (which corresponds to the water level at stop
pressure) must be filled with compressed air. The filling time should not exceed eight hours.
The suction capacity in m3/h is:

The compressor's operating pressure should as a minimum correspond to the pump's maximum
stop pressure. The safety valve on the compressor must be pre-set so that the maximum
permissible operating pressure of the accumulator is not exceeded.

In accordance with the accident prevention regulations for pressure vessels (German Gas and
Waterworks Professional Association, Düsseldorf), fitting a safety valve on accumulators for
centrifugal pumps is not mandatory as long as the H/Q curves (see Characteristic curve) of the
pumps do not exceed 1.1 times the maximum permissible operating pressure for the vessel, and
steps are taken to prevent critical overspeeding of the pumps.

The accumulators are welded, cast, riveted and, occasionally, finished in strip-wound
construction (for very high pressures and temperatures in the chemical industry). The materials
used are steel plate (boiler plate), non-ferrous metal plate, cast steel and plastic. The design and
operating data of commonly used accumulators are standardised.

Principal standards, directives and regulations applicable to accumulators

• American Petroleum Institute: API 610


• American Society of Mechanical Engineers:
ASME-Boiler and Pressure Vessel Code Section I-X
• German Pressure Vessel Society:
AD regulations
• Federal Ministry of Economics: Protection of Labour Act (Federal Bulletin 4/1980) and Steam
Boiler and Pressure Vessel Act
• DIN 3171, DIN 4661, DIN 4810 and EN 962
• German Organisation for Technical Standards in the Gas and Water Industries (DVGW): DVGW
Worksheet W 314
• TRD Technical Rules for Steam Boilers
• German Federation of Technical Supervision Associations
• Regulations of shipbuilding classification societies, e. g. German Lloyd (GL)

NPSH
The term NPSH is the abbreviation of "net positive suction head" and is an important factor in
evaluating the suction characteristics of a centrifugal pump. It allows a prediction to be made
regarding the safety margin required to avoid the effects of cavitation during operation.

In the EN 12723 standard the German term "Haltedruckhöhe (retaining pressure head)" is used
as a synonym for NPSH. As different reference levels are defined for the two terms, their
numerical value can differ by zs (difference in geodetic head between reference levels s and s').
In practice, only the NPSH value is used.

As the fluid flows through the centrifugal pump's impeller, the static pressure – relative to the
pressure upstream of the impeller – will drop, especially at the inlet to the vane passage. The
extent of the pressure drop depends on the rotational speed, the fluid density and viscosity, the
impeller’s inlet geometry, the operating point and the velocity profile of the approach flow.

In order to avoid cavitation or to limit it to an acceptable level, the pressure upstream of the
impeller must exceed the vapour pressure level of the fluid handled by a specified minimum
margin. Assessing the likelihood of occurrence, extent and impact of cavitation in a centrifugal
pump requires comparison of two NPSH values: the NPSH required by the pump NPSHr and the
NPSH available in the system, NPSHa.

The system's NPSH, i.e. NPSH available (NPSHa) is defined as

Point s refers to the suction nozzle's centre. If the pump's design does not feature a suction
nozzle, as is the case with in-line pumps with welded-in pipes (i.e. welded-in pumps) or
submersible pumps with bellmouths, a location s which corresponds to the point s in the suction
nozzle's centre must be defined and clearly specified when specifying the NPSHa value. See Fig.
1 NPSH
Fig. 1 NPSH: Position of
reference points s' for the NPSH value and s for the "retaining pressure head" (in this example
the flow approaches the impeller from below)

The total pressure at point s can be expressed as:

The reference point s' for the NPSH value is the impeller’s centre, i.e. the intersection of the
pump shaft axis and a plane situated at right angles to the pump shaft passing through the outer
points of the vane leading edge. See Fig. 2 NPSH
Fig. 2 NPSH: Position of
reference point s' for various types of pump impellers

The system's NPSH is thus established as follows:

Inserting the values at the system's inlet cross-section gives:

The head loss also includes any entry losses and pressure drops across valves and fittings etc.
NPSH required by the pump (NPSHr)

The definition of the NPSH required by the pump (NPSHr) is similar to that of the system's
NPSH, i.e. the symbols in brackets have the same meaning:

However, a significant difference is that the sum of the parameters defined by the terms in the
brackets must not fall below a minimum value (min) specified for a given pump and application.

If this condition is not met, the occurrence of cavitation cannot be ruled out.

When specifying the NPSHr, it is also necessary to provide information on the relevant cavitation
criterion. Criteria include:

• Incipient cavitation, NPSHi


• A certainextent of the cavitation zone on the vanes
• Start of head drop as a result of cavitation (NPSH0)
• Cavitation-induced head drop by 3 % (NPSH3)

The first three criteria are less common, and providing evidence for NPSHi requires demanding
and expensive testing. For this reason, it is commonly agreed that NPSHr = NPSH3.

The cavitation criteria listed above and their related NPSH values are dependent on the operating
point. See Fig. 3 NPSH

Fig. 3 NPSH: NPSHr for


various criteria as a function of the relative flow rate Q/Qshock-free
The illustration shows the curves for NPSHR of a specific impeller as a function of the relative
flow rate. The parameters shown are the cavitation phenomena, e. g. the length (Lcav) of the
resulting bubble trail (cavitation zone) in relation to the vane spacing or pitch (t) (see Vane
cascade).

If the NPSHa curve is also displayed in the diagram, it is possible to determine the type of
cavitation to be expected as a function of flow rate.

The upper curve (NPSHi) indicates incipient cavitation. If NPSHa is higher than NPSHi,
cavitation will not develop and the impeller will rotate without the formation of bubbles. The
lower the NPSHa value drops, the longer the bubble trail (cavitation zone) will become.

From a minimum level represented in the graph as the intersection of the lines denoting suction-
side and discharge-side cavitation, the bubble trail length will increase under low-flow/overload
conditions at a constant NPSHa. The flow rate at this minimum level corresponds to the flow
direction of shock-free entry which causes the lowest increases in fluid velocity on the pressure
side and suction side of the vane. It is therefore referred to as the shock-free flow rate (Qshock-
free).

If the flow rate (Q) is lower than the shock-free flow rate (Qshock-free), then cavitation will
develop on the vane's suction side; if the flow rate is higher than the shock-free flow rate, then
cavitation will develop on the vane's pressure side.

Establishing NPSHr is largely a matter of testing, in particular when:

• Converting the NPSH of the pump from one rotational speed to another
• On similar pumps, converting NPSH from one pump size to another
• Converting NPSH from one fluid to another (in particular if the fluid contains dissolved or
undissolved gas) (see Gas content of fluid handled)

A centrifugal pump's operating point can only be operated at continuously if:

The following relationship exists between the NPSH value and the German concept of
"Haltedruckhöhe” (retaining pressure head) which is no longer used:

Retaining pressure head of the system


HHA = NPSHa – zs

Retaining pressure head of the pump


HP = NPSHr – zs

In the case of horizontal pumps, there is no difference in height (zs = 0) between the reference
points for NSPH and "retaining pressure head", making the two terms identical. The following
coefficients are sometimes used in connection with the NPSH value:
When hydrocarbons or high-temperature water are handled, the NPSH3 value measured is lower
than that measured for cold water. This means that the required NPSH value for hydrocarbons or
hot water can actually be reduced when performing acceptance tests with cold water:

• Hydrocarbons in accordance with HI


(standards laid down by the Hydraulic Institute, New York)
• Hot water See Fig. 4 NPSH

Fig. 4 NPSH: Correction


factor f for NPSH3 when handling hot water (based on KSB measurements)

Head
This term is an important energy concept (EN 12723) in centrifugal pump engineering. A
distinction must be made between the pump head and the system head.

The pump head is the hydraulic power or pump output power (PQ) transmitted to the fluid
handled relative to ρ · g · Q.
The sum of all power (positive input, negative output) represented by the pump power output
(PQ) must be zero within the boundaries of the system. See Fig. 1 Head

Fig. 1 Head: Explanation of


the pump power output PQ = PQ.d – PQ.s = P – Pv.i – Pm
If the expression PQ.d – PQ.s represents the pump power output (PQ), the useful power output is as
follows:

According to BERNOULLI (see Fluid mechanics), the equation for useful power output is:

For the pump head, this means:

If the fluid handled is compressible, the value for density (ρ) should be defined as the arithmetic
mean of the density at the pump discharge nozzle and the density at the pump suction nozzle:
The system head can be established in a similar manner, taking into account the head losses (HL):

The term geodetic head (Hgeo) is sometimes used to designate the system head. It refers to the
difference in elevation, or height, between the system's outlet cross-section (Aa) and the system's
inlet cross-section (Ae):

Under steady-state conditions (rotational speed (n) = constant), the pump head is equal to the
system head.

The unit of head is metres (m). The following expressions are also used in conjunction with the
term head.

Heads and their significance

• Head at BEP (Hopt): pump head at the best efficiency point


• Nominal head (HN): pump head for which the pump has been designed
• Upper head limit (Hmax): max. permissible head at which the pump can be continuously operated
without suffering damage
• Lower head limit (Hmin): min. permissible head at which the pump can be operated without
suffering damage
• Shut-off head (H0): head for a flow rate Q = 0 m3/s
• Peak head (Hpeak): head at apex (relative maximum) of an unstable H/Q curve, see Fig. 4
Characteristic curve
• Static head (HA,0 or Hstat): the portion of the system head (see System characteristic curve and
Characteristic curve) which is independent of the flow rate (Q)
Fig. 2 Head: Illustration of the
magnitudes relating to the system head Hsys

Fitting
Fittings in a centrifugal pump system comprise all piping components which function to change
the piping's direction, to install piping branches, and/or to provide a transition between different
pipe cross-sections.

Fittings should be shaped to offer the least possible resistance to flow in order to minimise
pressure losses (see system head); where this involves increased manufacturing costs, these
should be weighed against the corresponding gains in economic efficiency.

Common fittings:
Pipe bends

• Pipe bends should have a radius of curvature of R > 2 2∙ D + 100 +100 mm (D = pipe diameter)
particularly if they are fitted immediately upstream of pump suction nozzles. Pipe bends
fabricated from cylindrical segments welded together should consist of at least six segments for
a 90° bend.
See Fig. 1 Fitting
Fig. 1 Fitting: 90° pipe bend

Y-branch

• Y-branches' fluid dynamic characteristics make them preferable to tees.


See Fig. 2 Fitting

Fig. 2 Fitting: Y-branch

Diffusor

• Its face-to-face length (L) should be approximately L = 5 ∙ (D2 – D1)


(D = pipe diameter) when used as a diffuser in flow direction. Diffuser outlets (e.g. in the case of
low-lift pumping stations and pumps for use in low-lift pumping stations should be sized such
that the discharge velocity (v) (see Flow velocity) is 1.0 to 1.5 SPdL m__s SPdL.
See Fig. 3 Fitting
Nozzle-shaped reducer

• In contrast to a diffuser, the face-to-face length of a fitting used as a reducer can be much
shorter. A nozzle-shaped reducer features favourable flow characteristics.
See Fig. 4 Fitting

Fig. 4 Fitting: Nozzle-shaped


reducer

Reducer for avoiding air pockets

• Eccentric reducers should be installed in horizontal suction lines to avoid the formation of air
pockets (see Formation of air pockets).
See Fig. 5 Fitting

Fig. 5 Fitting: Reducer for


avoiding air pockets
Branch fitting for avoiding air pockets

• Eccentric branch fittings should be installed in horizontal suction lines to avoid the formation of
air pockets.
See Fig. 6 Fitting
• For further fittings see bellmouth, intake chamber and intake elbow (see Inlet conditions).

Pump Lexicon

Bellmouth
A bellmouth, also called a suction bellmouth in connection with centrifugal pumps, is a nozzle-
shaped inlet casing component (see Fitting), often employed with vertical tubular casing pumps.
The flow acceleration resultant of the bellmouth's shape minimises irregularities in the velocity
distribution. Even velocity distribution ensures a uniform approach flow (see Inlet conditions);
this is especially important for high specific speed pumps (see Specific speed).

In the case of vortex flow at the inlet, a flow straightener should be fitted to provide a degree of
flow straightening. See Fig. 1 Bellmouth

Fig. 1 Bellmouth: High


specific speed pump with bellmouth and flow straightener

If pre-swirl control has been provided downstream, there is no need to fit a flow straightener in
the bellmouth. See Fig. 2 Bellmouth
Fig. 2 Bellmouth: High
specific speed pump with bellmouth and pre-swirl control

Intake chamber
The intake chamber is often referred to as a pump sump. It is a collecting chamber situated
directly upstream of a centrifugal pump, through which the fluid handled, usually water, flows
towards the pump. This ensures that the approach flow towards the centrifugal pump is evenly
balanced on all sides and free of turbulence (see Inlet conditions). Such a smooth, disturbance-
free approach flow is indispensable for high specific speed tubular casing pumps with propellers
or mixed flow impellers because these pumps respond immediately to irregularities and
disturbances in the approach flow. A simple intake chamber design is all that is required to avoid
damage from cavitation and vibrations, and a possible drop in pump power output or pump
efficiency caused by irregular approach flows. The risk of air-entraining vortices being sucked in
from the water surface is avoided by ensuring that water levels in the intake chamber are
sufficient.
The required excavation depth depends on the intake chamber's design and shape. See Fig. 1
Intake chamber Fig. 1 Intake
chamber: Four different intake chamber designs

Intake chambers have a simple structural shape with a rectangular floor plan. A comparison of
the four different intake chamber designs reveals that, given an identical flow rate, design variant
I requires the highest minimum water level, variant IV the lowest. The designs I, II and III are
open intake chambers suitable for axially parallel approach flow. Design variant IV with a
splitter is also suitable for perpendicular approach flow. In the case of complex inlet conditions,
model tests are advisable.

A disturbance-free approach flow can also be achieved using intake elbows. Economic efficiency
should be calculated when deciding whether an intake chamber should be provided. They are
often built for vertical cooling water pumps.

In power stations, operational reliability is of crucial importance for pump availability. The
intake chamber therefore represents a structural unit which must be designed and built with great
care.

Intake chambers are also employed in irrigation and drainage stations where simple designs can
significantly reduce construction costs

Head loss
Head losses are a result of wall friction in all types of pipelines and of local resistance to flow,
for example in valves and fittings (see also Pressure loss).

Recommended flow velocities

• For cold water:


Suction line 0.7-1.5 m/s
Discharge line 1.0-2.0 m/s
• For hot water:
Suction line 0.5-1.0 m/s
Discharge line 1.5-3.5 m/s

Head loss in a pipe


The equation for the head loss of a flow in a straight length of piping with circular cross-section
is:

λ Pipe friction factor


L Pipe length in m
d Pipe inside diameter in m
v Flow velocity in a cross-section in m/s
(= 4 Q / π d2 with Q in m3/s)
g Gravitational constant in m/s2

see Fig. 1 and 4 Head loss

The pipe friction factor was established experimentally. It is only dependent on the state of flow
of the fluid handled and of the relative roughness (d/k) of the pipes through which the fluid is
flowing. For non-circular pipe cross-sections the equivalent diameter in fluid-mechanical terms
(d) applies:

A Cross-section in m2
U Wetted cross-section circumference in m
(the free surface of an open channel is not considered)

The state of flow is determined by the Reynolds number (Re) according to the affinity laws. The
following applies to circular pipes:

v Flow velocity in a cross-section in m/s


(= 4 Q / π d2 with Q in m3/s)
ν Kinematic viscosity in m2/s
(for water at 20 °C: 1.00 · 10 - 6 m2/s)
d Pipe inside diameter in m
See Fig. 4 Head loss

For hydraulically smooth pipes such as smooth drawn metal or plastic piping (e. g. PE or PVC),
or in the case of laminar flow, the pipe friction factor (λ) can be calculated. For laminar flow in a
pipe with a Reynolds number smaller than 2320 the pipe friction factor is independent of
roughness:

If flow is turbulent, or the Reynolds number higher than 2320, the pipe friction factor in
hydraulically smooth pipes can be represented by an empirical equation according to Eck (due to
the fact that deviations are below 1 % if the Reynolds number is lower than 108).

The pipe friction factor (λ) also depends on a further dimensionless parameter, i.e. on the relative
roughness of the pipe's inner surface (d/k). Both must be specified in the same unit (e. g. mm).

See Fig. 1 Head loss

(k) is the mean absolute roughness of the pipe inner surface for which approximate values are
available depending on the material and manufacturing processes. See Fig. 2 Head loss
Fig. 2
Head loss: Estimates of mean peak-to-valley heights k (absolute roughness) of pipes

Above the limit curve, the pipe friction factor (λ) is solely dependent on the pipe's relative
roughness (d/k). See Fig. 1 Head loss

The following empirical equation by Moody can be used for this region:
For practical use, the head loss (HL) per 100 m of straight steel pipe is shown in the diagram as a
function of the flow rate (Q) and pipe inside diameter (d).
See Fig. 3 Head loss

The values are valid only for cold, clean water or for fluids with the same kinematic viscosity,
for completely filled pipes and for absolute roughness of the pipe inner surface of k = 0.05 mm.
Dimensions, weights, water fill for new seamless or longitudinally welded steel pipes
See Annex, Head loss, Fig. 4

The effect of an increased surface roughness k will be demonstrated in the following for a
frequently used set of parameter ranges (nominal diameter DN = 50 to 300, flow velocity v = 0.8
to 3.0 m/s). See Fig. 3 Head loss
The light blue region corresponds to the similarly marked region for an absolute mean roughness
of k = 0.05 mm.
See Fig. 1 Head loss

For a roughness increased by a factor of 6 (slightly incrusted old steel pipe with k = 0.30 = 300
μm (0.30 mm), the pipe friction factors (and the associated proportional head losses) in the dark
blue region are only 25 - 60 % higher than before.
See Fig. 1 Head loss

For sewage pipes the increased roughness caused by soiling must be taken into consideration.
For pipes subject to extreme incrustation, the actual head loss can only be determined
experimentally. Deviations from the nominal diameter change the head loss considerably, as the
pipe inside diameter features in the equation to the 5th power.

A 5 % reduction of the inside diameter, for example, leads to an increase in head loss by as much
as 30 %. It is therefore important that the internal diameter is not simply replaced with the
nominal diameter in the calculations.

The head losses in plastic pipes or smooth drawn metal piping are very low thanks to the smooth
pipe surfaces. The head losses established are valid for water at 10 °C. At other temperatures, the
loss for plastic pipes must be multiplied by a specified temperature correction factor to account
for their larger thermal expansion. For sewage or other untreated water, an additional 20-30 %
head loss should be taken into account for potential deposits.

Head losses for plastic and smooth drawn metal pipes

See Annex, Head loss, Fig. 5

Head losses in valves and fittings

The head loss (HL) in valves and fittings is given by:


ζ Loss coefficient
See Figs. 6 to 12 Head loss
v Flow velocity in a characteristic cross-sectional area A
(e. g. at the nozzle) in m/s
g Gravitational constant 9.81 m/s2

Fig. 6 Head loss: Schematic


diagram of valve designs
Fig. 11 Head loss: Influence
on the loss coefficient ζ of rounding off the inner and outer side of elbows in square ducts

Fig. 12 Head loss: Loss


coefficients ζ for butterfly, globe and gate valves depending on the degree of opening

The losses attributable to the straightening of the flow disturbances over a pipe length equivalent
to 12 x DN downstream of the valve are included in the loss coefficients in accordance with the
VDI/VDE 2173 guideline. The values apply to valves which have a steady approach flow, are
fully opened and operated with cold water. Depending on the inlet and outlet flow conditions, the
valve models and development objectives (i. e. inexpensive or energy-saving valves), the loss
values can vary dramatically. See Annex, Head loss, Fig. 7
Often the kv value is used instead of the loss coefficient (ζ) when calculating the pressure loss for
water in valves:

The kv value is the flow rate in m3/h which would result from a pressure drop pv = 1 bar through
the valve for cold water. It describes the correlation between the pressure loss (pL) in bar and the
flow rate (Q) in m3/h. Conversion to flow coefficient ζ for cold water:

d Reference (nominal) diameter of the valve in cm

For the calculation of head losses in fittings, branch fittings and adapters require a different
approach. See Figs. 9 and 10 Head loss
Fig. 9 Head loss: Loss

coefficients ζ for fittings Fig.


10 Head loss: Loss coefficients ζ for adapters

For all fittings a differentiation must be made between two forms of pressure loss:

• Irreversible pressure losses (reduction in pressure)

pv Pressure loss in Pa
ζ Loss coefficient
ρ Density in kg/m3
v Flow velocity in a cross-section in m/s

• Reversible pressure changes of the frictionless flow according to Bernoulli's equation

For accelerated flows such as reductions in the pipe diameter, (p2 − p1) is always negative; for
decelerated flows such as pipe expansions, it is always positive. When calculating the net
pressure change as the arithmetic sum of pL and (p2 − p1), the irreversible pressure losses must
always be subtracted.

Influence of highly viscous fluids on the system characteristic curve

As the laws of fluid dynamics retain their validity for all Newtonian fluids, the equations and
diagrams for calculating the pipe friction factors and loss coefficients for valves are also
applicable to viscous fluids with a higher viscosity than water.

When calculating the Reynolds number Re = v · d / ν , one must simply substitute the kinematic
viscosity of the viscous fluids νz for the water viscosity νz.

This yields a lower Re number and, according to Fig. 1 Head loss, a larger pipe friction
coefficient λz (Note: the influence of the wall roughness can now often be ignored because of the
larger boundary layer thickness in the flow).
All of the pressure losses in the pipes and valves calculated for water are to be extrapolated using
the ratio λz/λw.

Figure 13 Head loss is also suitable for general practical use: the pipe friction factor λz can be
determined quickly as a function of the flow rate Q, pipe inside diameter d and kinematic
viscosity νz. It must be kept in mind, however, that the coefficient λw in this diagram is only valid
for hydraulically smooth pipes (i.e. not for rough pipes)! The corresponding λw can be used to
calculate the ratio λz/λw.

As the static component of the system characteristic curve Hsys , see Fig. 1 System characteristic
curve and Fig. 2 Head, is not affected by viscosity, the dynamic component of the system
characteristic curve for water can be redrawn as a steeper parabola for a viscous fluid.

Influence of non-Newtonian fluids on the system characteristic curve

As the flow curves are not straight lines of constant linear viscosity, the calculation of the head
losses is very cumbersome. In this case, loss calculation is based on experience with particular
fluids.

Axial thrust
The axial thrust is the resultant force of all the axial forces (F) acting on the pump rotor. See Fig.
1 Axial thrust

Fig. 1 Axial thrust: Axial


forces in a single-stage centrifugal pump

Axial forces acting on the rotor in the case of a single-stage centrifugal pump

• The axial impeller force (F1) is the difference between the axial forces on the discharge-side (Fd)
and suction-side (Fs) impeller shroud
F1 = Fd – Fs
• Momentum (FJ) is a force which constantly acts on the fluid contained in a defined space (see
Principle of conservation of momentum, Fluid mechanics). It is calculated as follows:
• FJ = ρ · Q · ΔVax
Q Flow rate
ρ Density of the fluid handled
ΔVax Difference between the axial components of the absolute velocity at the impeller inlet and
outlet
• The resultant pressure forces arising from the static pressures up- and downstream of the shaft
seal (ss) on the relevant shaft cross-section Ass: FWd = AWd · ΔpWd
• Special axial forces, e.g. when changes to the vortex conditions in the clearances between
impeller and casing (side gaps) occur during the start-up process (see Disc friction)
• Other axial forces such as the force of the rotor weight (FW) on non-horizontal centrifugal pumps
or magnetic pull in the electric motor (Fmech), e.g. in close-coupled pumps

The axial thrust component (F1 + FJ) of closed impellers (i. e. with suction-side shrouds) which
are not hydraulically balanced is:
The axial thrust coefficient is essentially dependent on the specific speed (ns). For radial and
mixed flow impellers, the following equation applies in the range of 6 < ns < 130 rpm :

See Fig. 2 Axial thrust

Fig. 2 Axial thrust: Non-


balanced impeller design with conical impeller outlet area

This equation applies to flow rates (Q) of 0.8 · Qopt to 1.0 · Qopt and to the clearance gap width s
= 0,1 mm. If the clearance gap width is doubled, α increases by 8 %.
In the case of multistage pumps with diffusers (e. g. boiler feed pumps), the axial impeller force
(F1) is largely determined by the impeller's axial position in relation to the diffuser. In the case of
open radial impellers with no shrouds on the suction side, the axial force (Fs) is much lower than
on closed impellers, meaning that the axial impeller force (F1) is higher.

Open impellers with cut-outs in the impeller shroud between adjoining impeller vanes develop a
lower pressure force (Fd), and, consequently, a lower axial force (F1) than impellers with a full
discharge side shroud. See Fig. 13 Impeller

For axial propellers, the axial thrust coefficient (α) is almost equal to the degree of reaction (rth).
The axial thrust can then be roughly calculated using the propeller's outside diameter (OD):

The following proportionality applies to the F1 component of the axial thrust (See Fig.1 Axial
thrust) in the case of geometrically similar pumps at a defined rotational speed (n) and at the
largest impeller diameter (D2):

The rotation of the fluid handled in the discharge-side and suction-side clearances between
impeller and casing exerts a strong influence on the axial pressure forces (Fd) and (Fs). The mean
angular velocity (see Rotational speed) of the rotating fluid handled reaches approx. half the
impeller speed.

In addition, as a result of Coriolis accelerations, the inward directed clearance flow in the
suction-side (i.e. outer) clearance between impeller and casing (side gap) further increases the
side gap turbulences. In the discharge-side (i.e. inner) side gap of multistage pumps whose
impellers are not hydraulically balanced, the process is reversed as a result of the outward-
directed gap flow. The vortex motion is decelerated resulting in an increase of the axial force Fd,
and hence of F1.

The axial impeller force is higher during the start-up process than during steady-state operation,
as during start-up rotation of the fluid handled begins slowly due to disc friction caused by the
action of the impeller shrouds or the braking effect of the stationary casing surfaces.

Various forms of axial thrust balancing

• Mechanical: complete absorption of the axial thrust via a thrust bearing (e. g. plain bearing,
rolling element bearing)
• Design-based: back-to-back arrangement of the impellers or stages (see Back-to-back impeller
pump) and through the absorption of the residual axial thrust via a thrust bearing
• Balancing or reduction of the axial thrust on the individual impeller via balancing holesSee Figs.
7, 9 Axial thrust
• Balancing of the complete rotating assembly via a balancing device with automatic balancing (e.
g. balance disc and balance disc seat) or partial balancing via a balance drum and double drum
• Reduction at the individual impeller by back vanes (dynamic effect)
See Fig. 8 Axial thrust

Mechanical axial thrust balancing

The absorption of the axial thrust by a rolling element bearing is the most efficient, cost-effective
solution. However, if the absence of special balancing equipment requires the use of particularly
complex thrust bearings, these benefits in terms of efficiency and costs may be eliminated.

Design-based axial thrust balancing

In the case of an impeller arrangement in a pipeline pump with four stages, each featuring a 2 x 2
back-to-back arrangement, a maximum of twice the normal axial thrust per stage can occur in the
event that system conditions cause cavitation in two stages. See Fig. 5 Axial thrust

If, however, a more complex, parallel-coupled back-to-back impeller arrangement is chosen,


only the normal axial thrust per stage occurs. See Fig. 6 Axial thrust

Both pump types must be equipped with thrust bearings of appropriate strength.

Axial thrust elimination

• Double-entry impeller arrangement (impeller, double suction pump)


See Fig. 3 Axial thrust
• Two-stage, back-to-back impeller arrangement (back-to-back and multistage pump, impeller)
See Fig. 4 Axial thrust
• Multistage, back-to-back mpeller arrangement
See Fig. 5 Axial thrust
• Parallel-coupled back-to-back impeller arrangement (e. g. pipeline pumps) See Fig. 6 Axial thrust
Fig. 3 Axial thrust: Axial
thrust balancing by double-entry impeller arrangement

Fig. 4 Axial thrust: Axial


thrust balancing by two-stage, back-to-back impeller arrangement
Fig. 5 Axial thrust: Axial
thrust balancing in a four-stage pipeline pump with two opposed sets of two series-coupled
impellers each

Fig. 6 Axial thrust: Axial


thrust balancing in a four-stage pipeline pump with two sets of parallel-coupled opposed
impellers

Axial thrust balancing at the impeller

This is the oldest method for balancing axial thrust and involves reducing the pressure in a
chamber equipped with a throttling gap, usually down to the pressure level encountered at the
impeller inlet. The pressure is balanced via balancing holes in the impeller.

These balancing holes may lead to variations in axial thrust balancing as a result of varying inlet
conditions. See Fig. 7 Axial thrust
Fig. 7 Axial thrust: Axial
thrust balancing in a single-stage centrifugal pump with discharge-side sealing clearance and
balancing holes

The angular velocity has a dynamic influence on the magnitude of the axial thrust (see Rotational
speed). An increase in angular velocity is mostly achieved by back vanes which are radially
arranged on the rear side of the impeller.

The higher mean angular velocity of the vortices in the clearance between impeller shroud and
casing results in a lower static pressure on the discharge-side impeller shroud. This results in a
lower axial force Fd and thus a lower F1. See Fig. 8 Axial thrust

Most radial back vanes are designed with diameters (Dbv o, Dbv i), side space depth (a), vane
height (h) and vane number (z) which vary according to requirements. The power absorbed by
this method of axial thrust balancing depends on the sizing of the back vanes. The pump
efficiency may drop by up to three points due to the back vanes. See Fig. 8 Axial thrust
Fig. 8 Axial thrust: Axial
thrust balancing in a single-stage centrifugal pump with back vanes

A comparable effect is achieved when the impeller is balanced via the provision of balancing
holes at defined areas on the discharge side without fitting a second discharge-side joint ring.
The gap flow directed toward the inside creates an angular momentum in the space between the
impeller shroud and the casing which increases the local angular velocity and, as a consequence,
reduces the static pressure. See Fig. 9 Axial thrust

Fig. 9 Axial thrust: Axial


thrust balancing in a single-stage centrifugal pump with balancing holes only

All hydraulic balancing devices are fully effective at optimal flow rate Qopt. Residual forces
occurring under low-flow and overload conditions must be absorbed by thrust bearings.
See Figs. 7 to 9 Axial thrust
Axial thrust balancing via balancing devices
Available options

• Balance disc with balance disc seat and balancing flow return line
See Fig. 10 Axial thrust
• Balance drum with balancing flow return line and thrust bearing
See Fig. 11 Axial thrust
• Double drum with balancing fluid return line and thrust bearing
See Fig. 12 Axial thrust

For all three types, the balancing flow (see Bypass) is returned to the pump suction nozzle (after
being cooled if necessary) or to the centrifugal pump's inlet tank.

In the case of the balance disc, the gap flow (see Clearance gap loss) is low because the self-
adjusting axial gap (s) remains very narrow which means that the pump's efficiency is only
slightly reduced. However, in the case of the balance drum, the radial clearance gaps are wider
and therefore the gap flows are higher, causing a greater drop in efficiency which is then further
compounded by the fact that an additional thrust bearing
is required. See Fig. 11 Axial thrust

Labyrinth-type gap seals are fitted to minimise the high gap flow (e. g. double drum balancing
devices).
Thanks to its large axial clearance (s), double drum balancing devices allow an additional thrust
bearing to be fitted which is mainly designed to prevent mechanical seizure in the balancing
system. See Fig. 12 Axial thrust

Fig. 10 Axial thrust:


Balancing device with balance disc
Fig. 11 Axial thrust:
Balancing device with balance drum and thrust bearing

Fig. 12 Axial thrust:


Balancing device with double drum and thrust bearing

Seizing may occur during start-up, during operation at extreme overloads (see Operating
behaviour) or when cavitation takes place. The pump rotor's position and, consequently, the wear
of the balancing equipment or thrust bearing can be ascertained by simple devices indicating
seizure during operation.

Pump Lexicon

Inlet conditions
To ensure trouble-free operation of a centrifugal pump, the approach flow to the impeller must
be disturbance-free and uniform.

While energy transfer from the vanes to the fluid handled is based on the centrifugal effect in the
case of low specific speed radial flow pumps, it is effected via flow deflection at the blades in the
case of high specific speed propeller pumps. For this reason, pumps with higher specific speeds
are more susceptible to disturbances in the approach flow than those with lower specific speeds.

Depending on the impeller type involved, the conditions for disturbance-free approach flow vary
and must be strictly complied with. The three major criteria are absence of swirl, a uniform
velocity distribution and absence of vortices.

Absence of swirl

The swirl (see Vortex flow) at the impeller inlet represents a disturbance of the ideal approach
flow to the impeller, unless it is deliberately used to control the head or improve the pump's
suction characteristics (see Inducer). Vortex flows at the inlet cross-section of a pump in most
cases are the result of asymmetries in the inlet (transverse approach flows, flows across an
elbow, asymmetrical flow separation) or suction recirculation (see Operating behaviour).

If the direction of the tangential components at the pump's inlet cross-section coincides with the
pump direction of rotation, the swirl rotates in the same direction. As a swirl in the same
direction of rotation as that of the pump increases (i.e. larger tangential components of the vortex
flow), the head, pump input power and efficiency will decrease at constant flow rate. The reason
is the reduced deflection performed by the impeller vanes in comparison with their design point
capability.

If a swirl in the opposite direction to the pump's direction of rotation increases, the head will
increase at constant flow rate to the point where the vanes are overloaded (flow separation at the
suction side of the vane, mechanical vibrations). The pump's efficiency will however drop faster
than in the case of a swirl in the same direction of rotation. In the case of a swirl in the opposite
direction, the resultant additional pump input
power requirement may lead to drive overload.

The disturbance factor "swirl" in the approach flow is established via flow velocity
measurements with regard to both magnitude and direction, using probes, hot wire or laser
measurement equipment or (as often employed in modelling) a rotameter, i.e. a speed-monitored
paddle wheel of the size of an impeller arranged in the suction nozzle.

The swirl can also be established using CFD analysis.

Honeycomb flow straighteners offer the best solution for reducing existing swirl components,
but simple diffuser plates in the form of baffles, cruciform flow straighteners and centrally
arranged longitudinal baffles (splitters) are also effective. See Fig. 1 to 3 Inlet conditions
Fig. 1 Inlet conditions: Intake
chamber upstream of a radial centrifugal pump designed for disturbance-free approach flow to
the impeller with central guide baffle (marine pump with inducer)

Fig. 2 Inlet condition: Intake


accelerating elbow for disturbance-free approach flow from a 90° change in direction
Fig. 3 Inlet conditions: Intake
chamber examples

On the basis of tests on pumps equipped with pre-swirl control (specific speeds ns of 70 to 200
rpm), tangential components of the flow velocity at the suction nozzle's outer edge which amount
to less than 7 % of the corresponding axial components do not represent a major disturbance.
This corresponds to a swirl angle of approx. 4 degrees.

The percentage rate above must be reduced in the case of pumps with specific speeds exceeding
200 rpm due to the need to comply more strictly with the inlet conditions.

Complete absence of swirl at the pump's inlet is practically impossible to achieve.

Uniform velocity distribution

When designing a pump impeller, a uniform velocity distribution in the cylindrical portion of the
suction line is generally assumed. This is understood as comprising all the profiles of the axial
flow velocities between the rectangular profile and the profile for fully developed turbulent flow
in the pipe.

Distortions of the uniform velocity distribution pattern mainly occur as a result of flow around
obstacles (wake depressions), and from any form of flow deviation and separation.
See Fig. 2 Intake elbow

The greater the deviation from uniform velocity distribution, the less likely it is that the pump
can reach the required performance data as the individual vanes are exposed to an approach flow
under low flow and overload conditions in the region of the velocity distortion.

If the non-uniformity of velocity distribution is not rotationally symmetric, mechanical vibrations


will occur as a result of the transient flow along the vanes.
Obstacles in the shape of screens or perforated plates arranged uniformly across the flow cross-
section, a straight length of piping or considerable acceleration via a nozzle of an appropriate
length will have a stabilising effect on the distorted velocity distribution.

If such devices cause the NPSH of the system (NPSHa) to drop to an unacceptable level making
cavitation more likely, the only effective (but often very expensive) remedy is to improve the
hydraulic characteristics of the intake chamber. Widening of the intake space results in lower
velocity peaks and lengthening it will reduce the wake depressions. Deflections and flows around
obstacles should be avoided or their effects minimised.

Rotationally symmetric deviations of the axial velocity from the volumetric average by more
than ±10 10 % and non-rotationally symmetric deviations of the local axial velocity on a
circular segment by more than ± 5 5 % are generally considered unacceptable. However,
deviations of the order of ± 5 % in the case of pumps with a very high specific speed (ns > 200
rpm) may already prove to be unacceptably high.

Absence of vortices

Vortices develop in shear flows and at locations with high gradients in the velocity profile. They
may occur in the approach flow as a result of flow separation, deceleration, acceleration,
branching off and a flow around installed structures. A differentiation is made between surface
and submerged vortices; a more detailed presentation of different vortices is provided in the
classification according to Hecker. Steam and gas may develop in the vortex core at a
sufficiently high speed of rotation. See Fig. 4 Inlet conditions
Fig. 4 Inlet conditions: Vortex
classification according to Hecker

Vortices which reach as far as the pump inlet impair the pump's performance data and operating
behaviour. They include but are not limited Unbenannt-15 1 28.09.15 14:54 to air-entraining
surface vortices and air/gas-filled submerged vortices.

Swirl may lead to changes in the power, performance, head and flow rate. The vortices are often
transient and, as a result of fluctuating pump performance, may in some cases be a cause of
increased vibration and noise values which are associated with mechanical stress on the impeller
or impeller vanes. Vortices' air and gas content additionally results in reduced performance data
and lower efficiencies.

The most important prerequisite for disturbance-free continuous pump operation is the
prevention of air-entraining vortices and air-/gas-laden submerged vortices. For this reason,
surface vortices from type 3 and submerged vortices of types 2 and 3 (according to Hecker) are
not acceptable for practical applications. Various measures can be taken to prevent air-entraining
intake vortices (see Intake chamber).

Measures to be taken to prevent air-entraining vortices

• Improving the approach flow to avoid rotation and gradients in the velocity profile
• Increasing submergence (h1)
See Fig. 3 Inlet conditions
• Covering the suction water level vulnerable to air vortices by means of a raft
• Installing anti-vortex baffles in the region of suction water levels vulnerable to air vortices

Measures taken to prevent submerged vortices

• Improving the approach flow to avoid rotation and gradients in the velocity profile
• Using inlet cones with baffles
See Fig. 1 Intake chamber
• Influencing the flow close to the respective walls by fitting anti-vortex vanes or similar structures

Standardised intake elbows and chambers optimised in model tests offer the best conditions for
an approach flow that is as uniform and vortex-free as possible. See Fig. 3 Inlet conditions

In the case of pumps installed in pipes, the inlet line upstream of the pump should not have any
turbulence-inducing structures. Elbows, valves and pipe branches may cause unacceptable inlet
flow conditions for the pump. Several elbows installed one after another and/or elbows installed
asymmetrically must be avoided. An unfavourable arrangement of these components may
severely affect the pump's operating behaviour.

In the case of double-entry pumps, it is important to pay particular attention to the inlet
conditions. When the flow leaving the elbow is asymmetrical (resulting from elbows installed
asymmetrically), the impeller halves are not uniformly subjected to load, i. e. the impeller halves
operate at different load points. In such cases, the negative impact on the NPSH value of the
pump (NPSHr), the vibration behaviour and the load on the bearings may be much more
pronounced than the effect on the pump characteristic curve.

If sources of interference upstream of the pump cannot be avoided in the system, it is necessary
to stabilise the inlet flow to an acceptable level.

This can be achieved by installing a sufficiently long, straight length of piping (approx. 5 to 8
times the nominal diameter DNs between the pump and the point of interference), elbows with a
large radius, elbows with deflection vanes
See Fig. 3 Inlet elbow and acceleration nozzles.

Disturbance-free approach flow examples

• The intake chamber design for a marine pump with inducer provides a relatively wide and long
intake (deflection at low flow velocity). It is additionally equipped with a longitudinal baffle
(splitter) to prevent larger swirl components. See Fig.1 Inlet conditions
• Disturbance-free inlet flow in an intake elbow with circular cross-sections can be achieved if in
the 90° deflection zone (elbow) the flow velocity is increased to approx. 2 to 4 times the pipe
flow See Fig. 2 Inlet conditionsSee Fig. 2 Inlet conditions. Accelerating elbows of this type are
also successfully manufactured from concrete with adapter cross-sections from rectangular
shapes to circles. See Fig. 6 Cooling water pump
• Recommended intake chamber designs for disturbance-free approach flows to vertical pumps
are shown in Fig. 3 Inlet conditions. The covered intake chamber with inlet cone and splitter also
allows cross flows to the pump, e. g. emergency operation in the event of a failure of associated
travelling screens.

Pump Lexicon

You might also like