You are on page 1of 56

Department of Mechanical Engineering

Course Code: ME- Fall 2022 F.M 2 Assignment 2


01323
Student Name: Abdul Hannan Sandhu Registration Number: 70111683 Section: A
Last Date for Submission: 7-1-2023

a) Write down a brief note on following about pump Analysis (use


diagrams and equation where necessary)
 Pump Performance Curves and Matching a Pump to a
Piping System
 Pump Cavitation and Net Positive Suction Head
 Pumps in Series and Parallel
 Positive-Displacement Pumps, Dynamic Pumps,
Centrifugal Pumps, Axial Pumps Marks Mapping
 Pump scaling laws
b) Write down a brief note on following about Turbine Analysis
(use diagrams and equation where necessary)
 Positive-Displacement Turbines, Dynamic Turbines
 Impulse Turbine, Reaction Turbine
 Turbine scaling laws

Solution:
PLO 3
Pump Performance Curves and Matching a
Pump to a Piping System:
CLO 3
Pump performance curves are important
drawings produced by the pump manufacturer. Pump
performance curves are primarily used to predict the
variation of the differential head across the pump, as the
flow is changed. But in addition variation of efficiency, 10
power, NPSH required etc , as the flow is changed, can also
be represented on the pump performance curves by the
manufacturer. It is important to be able to read and Taxonomy
understand the pump curves for selection, testing, Level 4
operation and maintenance of pumps. (Cognitive)
Department of Mechanical Engineering

Typically a pump performance curve will carry information


about the following points.

Variation of differential head Vs flow:


This is the primary information reported in the pump
performance curves and very important information
regarding most of the pump calculations related to
differential pressure across the pump. As shown in the
sample performance curves, usually 3 curves of differential
head Vs. volumetric flow are reported.

Differential head Δh is related to differential pressure ΔP


by the equation, ΔP = ρgΔh.

1. Curve of differential head for Rated Impeller


Diameter represents the variation of differential
head with volumetric flow for the impeller with
rated diameter which will actually be provided with
the pump.
2. Variation of differential head with volumetric flow
for Maximum Impeller Diameter is plotted for the
Department of Mechanical Engineering
impeller with the maximum diameter that can be
accommodated within the pump. This impeller can
be used in case flow through the pump is increased
or if more differential head is required in the future,
with the same pump.
3. Variation of differential head with volumetric flow
for Minimum Impeller Diameter is plotted for the
impeller with minimum possible diameter. If the
flow or differential head requirement is reduced in
future, this impeller can be used with lower power
consumption.

Although the 3 curves are plotted for a wide range of


volumetric flow rates, the actual operation is to be limited
within the Maximum and Minimum allowable flow rates as
indicated in the sample pump performance curve. Values
of the maximum and minimum flow limits are given by the
pump manufacturer.

The point on differential head axis (Y-axis) where each of


these 3 curves terminates, represents the shut-off
differential head for that particular impeller diameter. For
normal intended operation, the shut-off differential head
for rated impeller diameter is important.

It should be note that the pump curves for differential head


Vs. volumetric flow rate are plotted for a particular liquid
density. If in the future the process liquid or even just liquid
density is changed, that effect has to be considered to
finally determine the differential pressure. In such as case,
revised volumetric flow should be calculated and located on
the pump curve and corresponding differential head should
be then determined from the curve for the appropriate
impeller diameter. This differential head should then be
used along with the changed liquid density to determine
the differential pressure across the pump.
Department of Mechanical Engineering

Pump Efficiency:
As indicated in the sample pump performance curve above,
the plot of pump efficiency against volumetric flow rate is
also commonly reported on the pump performance curves.
When the theoretical pumping power requirement is
divided by this efficiency for the corresponding flow, the
result is pump shaft power requirement. The efficiency
curve typically has a maximum within the allowable
operating range. This maximum is also known as the Best
Efficiency Point (BEP) as indicated in the sample curves. The
normal operation should be preferably done close this best
efficiency point for minimum power requirements.
Sometimes a plot of Pump Shaft Power requirement is also
done against the volumetric flow rate on the performance
curves. This curve readily gives the value of power
requirement for a particular flow rate.

How to read a pump curve:


Pump curves are usually plotted for different impeller
diameter values. For large impeller diameter, you get more
differential head. For a given impeller size, when more fluid
is pushed through the pump, it is going to develop less
differential head (given the mechanical and power
constraints).

We can see that on the pump performance curve. For any


given impeller diameter, pump generates the maximum
differential pressure or differential head near the shut off
point, when there is very little fluid getting pushed through
the pump.
Department of Mechanical Engineering

As the flow rate is increased, we move to the right of this


graph. And you can find that for the same impeller
diameter, differential head starts to drop as the flow is
increased.

But, even as the diff. head drops down the resulting output
given by the pump in terms of flow rate multiplied by
differential head goes up because of increasing flow rate.
As this happens, the pump also consumes more and more
power to push more fluid while trying to maintain similar
level of differential pressure (or differential head).

As a result, there is a point of optimal efficiency where the


pump can operate at highest ratio of output power / input
power. This point is known as the best efficiency point for
that pump and clearly visible on 'Pump Efficiency
Curve' plotted on the same chart, against operating flow
rate.
Department of Mechanical Engineering

Values of NPSH required (NPSHr) are also provided by the


pump maker, in the form of another graph plotted on the
same chart.

Examples of pump curves:


Different pump manufacturers have different
formats for their pump curves diagram. Therefore,
sometimes the pump curves can look a bit confusing
at first. Sometimes there will be multiple lines on a
single chart and you may not be sure about the
meaning of each line. In such cases, always look for
the label of each curve.
For example consider the following pump curve chart.
Department of Mechanical Engineering

There are multiple curves here, on a single chart. But the


good thing is that they are properly labeled. The green
curve clearly represents the variation of "head Vs flowrate".
The head is maximum at zero flow and drops down as the
flow is increased. Notably, the "head vs. flowrate" curve is
only plotted for a single impeller size.

Further, there is another U shaped curve of brown color.


This represents the variation of pump efficiency against the
flowrate. Peak point on the efficiency curve is clearly
marked on the chart as the - best efficiency point (BEP).

Finally, there is an orange color curve at the bottom


indicating the required NPSH (NPSHr) for different
flowrates. As the flowrate increases, so do the frictional
losses. Hence at NPSHr goes up with increasing flowrate.

It is common practice in the pump industry to offer several


choices of impeller diameter for a single pump casing.
There are several reasons for this:
(1) to save manufacturing costs
(2) to enable capacity increase by simple impeller
replacement
Department of Mechanical Engineering
(3) to standardize installation mountings
(4) to enable reuse of equipment for a different
application.
When plotting the performance of such a “family” of
pumps, pump manufacturers do not plot separate curves of
H, hpump, and bhp for each impeller diameter in the form
sketched in Fig. 14–8. Instead, they prefer to combine the
performance curves of an entire family of pumps of
different impeller diameters onto a single plot (Fig. 14–14).
Specifically, they plot a curve of H as a function of V . for
each impeller diameter in the same way as in Fig. 14–8, but
create contour lines of constant efficiency, by drawing
smooth curves through points that have the same value of
hpump for the various choices of impeller diameter.
Contour lines of constant bhp are often drawn on the same
plot in similar fashion. An example is provided in Fig. 14–15
for a family of centrifugal pumps manufactured by Taco,
Inc. In this case, five impeller diameters are available, but
the identical pump casing is used for all five options. As seen
in Fig. 14–15, pump manufacturers do not always plot their
pumps’ performance curves all the way to free delivery.
This is because the pumps are usually not operated there
due to the low values of net head and efficiency. If higher
values of flow rate and net head are required, the customer
should step up to the next larger casing size, or consider
using additional pumps in series or parallel. It is clear from
the performance plot of Fig. 14–15 that for a given pump
casing, the larger the impeller, the higher the maximum
achievable efficiency. Why then would anyone buy the
smaller impeller pump? To answer this question, we must
recognize that the customer’s application requires a certain
combination of flow rate and net head. If the requirements
match a particular impeller diameter, it may be more cost
effective to sacrifice pump efficiency in order to satisfy
these requirements.
Department of Mechanical Engineering

Pump Cavitation and Net Positive


Suction Head:

Cavitation is not a new phenomenon that can


impact a pump system, but it is an issue that is
growing. While no official figures exist, it is not
misleading to say that in the last five years, cases of
pump cavitation have increased markedly. Often
the pump itself is unfairly blamed. Pumping system
problems, including cavitation, often manifest
themselves at the pump but are rarely caused by it.
In fact, nine out of 10 pump problems are not
caused by the pump itself but by issues such as
cavitation, poor system design and lack of
maintenance. Additional issues caused by
cavitation, such as vibration, can be severe and may
lead to mechanical damage to the pump. Cavitation
related problems also have the potential to reduce
Department of Mechanical Engineering

pump life from 10-15 years, down to just two years


in extreme cases.
Cavitation can have a serious negative impact on
pump operation and lifespan. It can affect many
aspects of a pump, but it is often the pump impeller
that is most severely impacted. A relatively new
impeller that has suffered from cavitation typically
looks like it has been in use for many years; the
impeller material may be eroded and it can be
damaged beyond repair.
Cavitation occurs when the liquid in a pump turns to
a vapor at low pressure. It occurs because there is
not enough pressure at the suction end of the
pump, or insufficient Net Positive Suction Head
available (NPSHa). When cavitation takes place, air
bubbles are created at low pressure. As the liquid
passes from the suction side of the impeller to the
delivery side, the bubbles implode. This creates a
shockwave that hits the impeller and creates pump
vibration and mechanical damage, possibly leading
to complete failure of the pump at some stage.
vapor pressure:
At a specific combination of pressure and
temperature, which is different for different liquids,
the liquid molecules turn to vapor. An everyday
example is a pot of water on the kitchen stove.
When boiled to 100o Celsius, atmospheric pressure
bubbles form on the bottom of the pan and steam
rises. This indicates vapor pressure and
temperature have been reached and the water will
begin boiling.
Vapor pressure is defined as the pressure at which
liquid molecules will turn into vapor (see Figure 1.0).
Department of Mechanical Engineering

It should be noted that the vapor pressure for all


liquids varies with temperature. It is also important
to understand that vapor pressure and temperature
are linked. A half full bottle of water subjected to a
partial vacuum will begin to boil without the
addition of any heat whatsoever.

Properties of water at different


temperatures:
By regulating the pressure at which water is
subjected its vapor pressure can be changed and it
will eventually boil at room temperature. Figure 2.0
illustrates a tabulation of temperatures from 5 to
100oC with variations in vapor pressure, density and
specific gravity for water.

At 5ºC, the density is 999.9, which an engineer will


round to 1000. The specific gravity is 0.9999, or 1.
At the bottom of the table one can see that at 100ºC
the density has changed to approximately 958, a
significant but not a big change. If one looks at the
specific gravity, it has changed from 1 to
approximately 0.96.
Department of Mechanical Engineering

The variability of vapor pressure:


At 5ºC, water vapor pressure is 872
in round figures, but at 100ºC it is 101,000 which is
a major change in pressure. This means that for any
combination of pressure and temperature in the
table, the water will begin boiling and turn into
vapor (see Figure 3.0). For example, a glass of water
subjected to a pressure of 2337 N/m2 at 20ºC will
begin boiling.
Department of Mechanical Engineering

At 100ºC the vapor pressure is 101325 N/m2, which


is atmospheric pressure. With cavitation one must
deal with absolute pressures, not gauge pressures.
By regulating the temperature and pressure, it is
possible to boil water at different points. Other
liquids have similar charts, but the values will be
different.
The influence of atmospheric
pressure:
Outlined in Figure 3.1, water will start to boil at
100ºC atmospheric pressure (1 bar) at sea level.
Other liquids such as hexane, carbon tetrachloride,
pentane and butane are very different and will boil
off at much lower temperatures than water.
Butane, for example, will start to boil at negative
temperatures.
Pressure is dependent on altitude. In Figure 3.1 a
line is plotted for the top of Ben Nevis, the highest
mountain in the United Kingdom at 1,344 meters,
and a line for Mount Everest in Nepal at 8,848
Department of Mechanical Engineering

meters. The table demonstrates that at the top of


Ben Nevis the vapor pressure of water has dropped
sufficiently to allow it to start to boil at 80ºC. While
it is unlikely that pumps will be installed on the top
of these mountains, it is the case that pumps will be
used in various locations around the world with
similar altitudes, including Johannesburg, South
Africa, which is approximately 1,700 meters above
sea level. In such places it is important to determine
the change in vapor pressure for that location.

The impact of cavitation on a pump:


Cavitation causes pump performance deterioration,
mechanical damage, noise and vibration which can
ultimately lead to pump failure. Vibration is a
common symptom of cavitation, and many times
the first sign of an issue. Vibration causes problems
for many pump components, including the shaft,
bearings and seals.
Department of Mechanical Engineering

These photographs show cavitation damage to an impeller with segments


chipped away and severe mechanical damage.

How to avoid cavitation:


Assuming no changes to the suction conditions or
liquid properties during operation, cavitation can be
avoided most easily during the design stage. The key
is to understand Net Positive Suction Head (NPSH)
and take it into account throughout the design
process.
In order to understand this term more easily it is
helpful to break it down: • Net refers to that which
is remaining after all deductions have been made
• Positive is obvious
• Suction Head refers to the pressure at the pump
inlet flange.

Bubbles form during cavitation. As the pressure in


the pump increases, those bubbles collapse in the
form of an implosion – equally as violent as an
explosion. The implosion causes shockwaves to
travel through the liquid and hit the impeller causing
mechanical damage.
Department of Mechanical Engineering

Net Positive Suction Head Available


(NPSHa):
NPSH is defined as the difference
between the pressure available at the pump inlet
and the vapor pressure of the liquid. Vapor pressure
is different for different liquids and varies with
pressure and temperature.
The pressure available at the pump inlet is what
remains after friction loss, velocity head loss and
inlet and outlet losses have been taken into account
within the suction pipework of the pumping system.
Because of this, during the design phase, it is
necessary to calculate these losses and process unit
Department of Mechanical Engineering

losses in the suction pipework and then deduct


those losses from the suction head available to the
pump. By doing this, at the point where the pump is
installed, one is left with net pressure remaining and
available for the pump.
Net Positive Suction Head available (NPSHa) has
nothing to do with the pump; it is a system value
specific to the system design being considered.
NPSHa is the head available at the pump suction
flange pipework connection for the pumping system
in question, and is completely independent of the
pump to be installed there. It is the actual difference
between the pressure at the pump inlet flange and
the vapor pressure of the liquid for the installation
and is determined by the design, configuration and
relative levels for the suction side of a particular
system.
NPSHa = P pump inlet ¬– vapor pressure(m)
Pressure available at the pump inlet is that which
remains after allowances have been made for all the
losses as described above
Net Positive Suction Head Required
(NPSHr):
Net positive suction head required (NPSHr) is a
pump characteristic that is not related to the
system. All pump NPSHr’s are different and the
values can be obtained from the pump
manufacturer.
NPSHr is defined as the difference between the
pressure at the pump inlet and the vapor pressure
required for the pump in question. This value should
not be taken as sufficient to ensure that cavitation
will not occur, because it is measured at during
Department of Mechanical Engineering

testing when the pump is just starting to cavitate.


The onset of cavitation is measured as being the
point at which the delivery head on the pump drops
by 3 percent.
It is therefore necessary to ensure that NPSHa is
greater than NPSHr, and for that, a margin must be
built in to the equation.
NPSHa ≥ NPSHr + margin
The golden rule is to ensure that there is always
sufficient margin to avoid cavitation. The value of
the margin is often specified in-house by design
consultants, but pump manufacturers will always
offer advice. Typically, a margin of approximately
1.5 meters will be sufficient.
NPSHa is a characteristic of the system design that
can be controlled. Every effort should be made at
the design stage to ensure that there is sufficient
NPSHa within a system. The consequences of not
doing so may prove to be costly and disruptive.
Department of Mechanical Engineering

Pumps in Series and Parallel:


Pump in Series:
Many pumping applications require fluids to be
transported over long distances and against high
static heads or total heads which are well in excess
of the head that can be developed by a single pump.
Examples of such scenarios would be, pumping
tailings, power station ash, underground fill and
pumping concentrates. Centrifugal pumps are
occasionally installed in series to increase the
operating range and standby capacity of the plant.
Multi-stage pumps can be considered as a series
type installation of single-stage pumps. However,
single stages in multi-stage pumps cannot be
decoupled.

If one of the pumps in a series type installation is not


operational, it causes considerable resistance to the
system. In an effort to overcome this, a bypass with
non-return valve could be installed as shown in
Figure 3.
Department of Mechanical Engineering

The head associated with a given flow rate for


pumps installed in series is can be determined by
adding the single heads vertically as shown in
Figure 4.

The composite pump chart for pumps installed in


series can be plotted in FluidFlow.
Department of Mechanical Engineering

Composite Pump Chart - Pumps in Series.

Advantages of Series Pumping:


1. Series connected pumps generally require less
horsepower than one large pump meeting the same
condition point.
2. In most cases, a higher efficiency point is achieved.
3. Centrifugal pumps in series allows for operation at
lower speeds. This, in and of itself, has additional
ancillary benefits:
o Reduced wear
o Lower energy consumption
o Increased reliability and longer life
4. Pumps are smaller, easier to maintain, and more
manageable, versus a larger centrifugal pump.

Disadvantages of Series Pumping:


Of course there are some disadvantages to series
pumping. More equipment is required for sure.
Department of Mechanical Engineering

You’ll need more pumps, motors, motor starters,


circuit breakers, and drives. You also need to be
careful not to exceed the maximum working
pressure of the pumps, especially the ones further
down the line in series. This over-pressurization
could damage the pump assembly or adjacent
equipment. All that additional equipment is going
to take up more floor space as well, but given the
advantages above, it’s likely the pluses will offset
the minuses.

5. Pumps in Parallel:
Pumps are often installed in Parallel in systems
which exhibit considerable variations in flow or
when the system has variable flow requirements
and when these requirements can be achieved by
switching the parallel pumps on and off. Typically,
pumps installed in parallel are of a similar type and
size. The pumps however can be of a different size
or the pumps can be speed-controlled which of
course means the pump performance curves will be
different.

Figure 1 provides an illustration of a system with


two identical pumps installed in parallel. The image
shows how non-return valves are installed in series
and downstream of each pump. This is to prevent
bypass circulation when a pump is not running.

The total system performance curve for parallel


pumps can be determined by adding Q1 and Q2 for
every value of head which is the same for both
Department of Mechanical Engineering

pumps, i.e. H1 = H2. As the pumps, in this case, are


identical, the resultant composite pump curve has
the same maximum head value, Hmax but the
maximum flow Qmax is twice as large. So, for each
value of the head, the flow is double that of a single
pump in operation.

Pumps in Parallel.
Pumps are sometimes installed in parallel in
pressure booster systems for water supply
and for water supply in large buildings.
Operational advantages can be achieved by
connecting two or more pumps in parallel
instead of installing just one single large
pump. The total pump output is usually only
required for a limited period. If a single pump
was used in this instance, it would operate at
a comparatively lower efficiency.

When a number of smaller pumps are


installed in parallel, the system can be
controlled to optimise or minimize the
number of pumps operating and the pumps
which are on will operate at or close-to the
Department of Mechanical Engineering

BEP (best efficiency point). To operate at the


most optimal point, one of the pumps
installed in parallel must have variable speed
control.

The composite pump chart for pumps


installed in parallel can be plotted in FluidFlow
as shown in Figure 2.

FluidFlow Composite Pump Chart - Pumps in Parallel

Advantages of parallel pumps:


1. When one of the pumps running in parallel is
damaged, the other pumps can still continue to
supply water;
2. The flow rate and head of the pumping station
can be adjusted by the number of pumps on and
off to adapt to the changes in the user's water use;
Department of Mechanical Engineering

3. Increase the water supply. The total flow in the


water mains is equal to the sum of the water
output of each parallel pump.
Disadvantages of Parallel pumps:
Although parallel operation does increase the flow
rate.
It also causes greater fluid friction losses, results in
a higher discharge pressure.
Reduces the flow rate provided by each pump, and
alters the efficiency of each pump.
In addition, more energy is required to transfer a
given fluid volume
Department of Mechanical Engineering

Positive-Displacement Pumps,
Dynamic Pumps, Centrifugal Pumps,
Axial Pumps
Positive-Displacement Pumps:
People have designed numerous positive-
displacement pumps throughout the centuries. In
each design, fluid is sucked into an expanding
volume and then pushed along as that volume
contracts, but the mechanism that causes this
change in volume differs greatly among the various
designs. Some designs are very simple, like the
flexible-tube peristaltic pump (Fig. 14–26a) that
compresses a tube by small wheels, pushing the
fluid along. (This mechanism is somewhat similar to
peristalsis in your esophagus or intestines, where
muscles rather than wheels compress the tube.)
Others are more complex, using rotating cams with
synchronized lobes (Fig. 14–26b), interlocking gears
(Fig. 14–26c), or screws (Fig. 14–26d). Positive-
displacement pumps are ideal for high-pressure
applications like pumping viscous liquids or thick
slurries, and for applications where precise amounts
of liquid are to be dispensed or metered, as in
medical applications.
Department of Mechanical Engineering

FIGURE 14–26 :Examples of positive-displacement pumps: (a) flexible-


tube peristaltic pump, (b) three-lobe rotary pump, (c) gear pump, and (d)
double screw pump

FIGURE 14–27 : Four phases (one-eighth of a turn apart) in the operation


of a two-lobe rotary pump, a type of positive displacement pump. The
light blue region represents a chunk of fluid pushed through the top
rotor, while the dark blue region represents a chunk of fluid pushed
through the bottom rotor, which rotates in the opposite direction. Flow is
from left to right.
Department of Mechanical Engineering
To illustrate the operation of a positive-displacement pump, we sketch
four phases of half of a cycle of a simple rotary pump with two lobes on
each rotor (Fig. 14–27). The two rotors are synchronized by an external
gear box so as to rotate at the same angular speed, but in opposite
directions. In the diagram, the top rotor turns clockwise and the bottom
rotor turns counterclockwise, sucking in fluid from the left and
discharging it to the right. A white dot is drawn on one lobe of each rotor
to help you visualize the rotation.
Gaps exist between the rotors and the housing and between the lobes of
the rotors themselves, as illustrated (and exaggerated) in Fig. 14–27. Fluid
can leak through these gaps, reducing the pump’s efficiency. High-
viscosity fluids cannot penetrate the gaps as easily; hence the net head
(and efficiency) of a rotary pump generally increases with fluid viscosity,
as shown in Fig. 14–28. This is one reason why rotary pumps (and other
types of positive displacement pumps) are a good choice for pumping
highly viscous fluids and slurries. They are used, for example, as
automobile engine oil pumps and in the foods industry to pump heavy
liquids like syrup, tomato paste, and chocolate, and slurries like soups.

FIGURE 14–28 Comparison of the pump performance curves of a rotary


pump operating at the same speed, but with fluids of various viscosities.
To avoid motor overload the pump should not be operated in the shaded
region.
The pump performance curve (net head versus capacity) of a rotary pump
is nearly vertical throughout its recommended operating range, since the
capacity is fairly constant regardless of load at a given rotational speed
(Fig. 14–28). However, as indicated by the dashed blue line in Fig. 14–28,
at very high values of net head, corresponding to very high pump outlet
Department of Mechanical Engineering
pressure, leaks become more severe, even for high-viscosity fluids. In
addition, the motor driving the pump cannot overcome the large torque
caused by this high outlet pressure, and the motor begins to suffer stall or
overload, which may burn out the motor. Therefore, rotary pump
manufacturers do not recommend operation of the pump above a certain
maximum net head, which is typically well below the shutoff head. The
pump performance curves sup plied by the manufacturer often do not
even show the pump’s performance outside of its recommended
operating range. Positive-displacement pumps have many advantages
over dynamic pumps. For example, a positive-displacement pump is
better able to handle shear sensitive liquids since the induced shear is
much less than that of a dynamic pump operating at similar pressure and
flow rate. Blood is a shear sensitive liquid, and this is one reason why
positive-displacement pumps are used for artificial hearts. A well-sealed
positive-displacement pump can create a significant vacuum pressure at
its inlet, even when dry, and is thus able to lift a liquid from several
meters below the pump. We refer to this kind of pump as a self-priming
pump (Fig. 14–29). Finally, the rotor(s) of a positive displacement pump
run at lower speeds than the rotor (impeller) of a dynamic pump at
similar loads, extending the useful lifetime of seals, etc.

There are some disadvantages of positive-displacement pumps as well.


Their volume flow rate cannot be changed unless the rotation rate is
changed. (This is not as simple as it sounds, since most AC electric motors
are designed to operate at one or more fixed rotational speeds.) They
create very high pressure at the outlet side, and if the outlet becomes
blocked, ruptures may occur or electric motors may overheat, as
previously discussed. Overpressure protection (e.g., a pressure-relief
valve) is often required for this reason. Because of their design, positive-
Department of Mechanical Engineering
displacement pumps may deliver a pulsating flow, which may be
unacceptable for some applications.
Analysis of positive-displacement pumps is fairly straightforward. From
the geometry of the pump, we calculate the closed volume (Vclosed) that
is filled (and expelled) for every n rotations of the shaft. Volume flow rate
is then equal to rotation rate n . times Vclosed divided by n,

Dynamic Pumps There are three main types of dynamic pumps that
involve rotating blades called impeller blades or rotor blades, which
impart momentum to the fluid. For this reason they are sometimes called
rotodynamic pumps or simply rotary pumps (not to be confused with
rotary positive-displacement pumps, which use the same name). There
are also some nonrotary dynamic pumps, such as jet pumps and
electromagnetic pumps; these are not dis cussed in this text. Rotary
pumps are classified by the manner in which flow exits the pump:
centrifugal flow, axial flow, and mixed flow (Fig. 14–31). In a centrifugal-
flow pump, fluid enters axially (in the same direction as the axis of the
rotating shaft) in the center of the pump, but is discharged radially (or
tangentially) along the outer radius of the pump casing. For this rea son
centrifugal pumps are also called radial-flow pumps. In an axial-flow
pump, fluid enters and leaves axially, typically along the outer portion of
the pump because of blockage by the shaft, motor, hub, etc. A mixed-
flow pump is intermediate between centrifugal and axial, with the flow
entering axially, not necessarily in the center, but leaving at some angle
between radially and axially.
Department of Mechanical Engineering

FIGURE 14–31 The impeller (rotating portion) of the three main


categories of dynamic pumps: (a) centrifugal flow, (b) mixed flow, and (c)
axial flow.

Centrifugal Pump:

A centrifugal pump is a mechanical device designed


to move a fluid by means of the transfer of
rotational energy from one or more driven rotors,
called impellers. Fluid enters the rapidly rotating
impeller along its axis and is cast out by centrifugal
Department of Mechanical Engineering

force along its circumference through the impeller’s


vane tips. The action of the impeller increases the
fluid’s velocity and pressure and also directs it
towards the pump outlet. The pump casing is
specially designed to constrict the fluid from the
pump inlet, direct it into the impeller and then slow
and control the fluid before discharge.
centrifugal pump working:
The impeller is the key component of a centrifugal
pump. It consists of a series of curved
vanes. These are normally sandwiched between
two discs (an enclosed impeller). For fluids with
entrained solids, an open or semi-open impeller
(backed by a single disc) is preferred

Fluid enters the impeller at its axis (the ‘eye’) and


exits along the circumference between the
vanes. The impeller, on the opposite side to the
eye, is connected through a drive shaft to a motor
and rotated at high speed (typically 500-
5000rpm). The rotational motion of the impeller
accelerates the fluid out through the impeller
vanes into the pump casing.

There are two basic designs of pump casing:


volute and diffuser. The purpose in both designs
is to translate the fluid flow into a controlled
discharge at pressure.

In a volute casing, the impeller is offset,


effectively creating a curved funnel with an
Department of Mechanical Engineering
increasing cross-sectional area towards the pump
outlet. This design causes the fluid pressure to
increase towards the outlet (Figure 2).

The same basic principle applies to diffuser


designs. In this case, the fluid pressure increases
as fluid is expelled between a set of stationary
vanes surrounding the impeller (Figure
3). Diffuser designs can be tailored for specific
applications and can therefore be more
efficient. Volute cases are better suited to
applications involving entrained solids or high
viscosity fluids when it is advantageous to avoid
the added constrictions of diffuser vanes. The
asymmetry of the volute design can result in
greater wear on the impeller and drive shaft.
Department of Mechanical Engineering

main features of a centrifugal pump:


There are two main families of pumps: centrifugal
and positive displacement pumps. In comparison
to the latter, centrifugal pumps are usually
specified for higher flows and for pumping lower
viscosity liquids, down to 0.1 cP. In some
chemical plants, 90% of the pumps in use will be
centrifugal pumps. However, there are a number
of applications for which positive displacement
pumps are preferred.
main applications for centrifugal pumps:
Centrifugal pumps are commonly used for
pumping water, solvents, organics, oils, acids,
bases and any ‘thin’ liquids in both industrial,
agricultural and domestic applications. In fact,
there is a design of centrifugal pump suitable for
virtually any application involving low viscosity
fluids.
Department of Mechanical Engineering

Axial Pumps:
Axial flow pumps, also called propeller pumps, are centrifugal
pumps which move fluid axially through an impeller. They
provide high flow rate and low head, but some models can be
adjusted to run efficiently at different conditions by changing the
impeller pitch.
Operation
Axial flow pumps are dynamic pumps, meaning they utilize fluid
momentum and velocity to generate pump pressure. Specifically,
they are centrifugal pumps, which generate this velocity by using
an impeller to apply centrifugal force to the moving liquid.

Axial flow pumps are one of three subtypes of centrifugal pumps,


the others being mixed flow and radial flow. Of these three types,
axial flow pumps are characterized by the highest flow rates and
lowest discharge pressures. They direct flow in a straight line
parallel to the impeller shaft (see image below) rather than
radially (perpendicular to the shaft). The impeller is shaped like a
propeller and contains only a few (typically three or four) vanes.
The impeller is driven by a motor that is either sealed directly in
the pump body or by a drive shaft that enters the pump tube
from the side. The impeller looks and operates similar to a boat
propeller, which is the reason why axial flow pumps are also
called propeller pumps.
Department of Mechanical Engineering

Specifications
When selecting an axial flow pump, there are a few key
performance specifications to consider:

 Flow rate describes the rate at which the pump can move fluid
through the system, typically expressed in gallons per minute
(gpm). The rated capacity of a pump must be matched to the flow
rate required by the application or system.
 Pressure is a measure of the force per unit area of resistance the
pump can handle or overcome, expressed in bar or psi (pounds per
square inch). As in all centrifugal pumps, the pressure in axial flow
pumps varies based on the pumped fluid's specific gravity. For this
reason, head is more commonly used to define pump energy in this
way.
 Head is the height above the suction inlet that a pump can lift a
fluid. It is a shortcut measurement of system resistance (pressure)
which is independent of the fluid's specific gravity, expressed
as a column height of water given in feet (ft) or meters (m).
 Net positive suction head (NPSH) is the difference between the
pump's inlet stagnation pressure head and the vapor pressure head.
The required NPSH is an important parameter in preventing
pump cavitation.
 Output power, also called water horsepower, is the power actually
delivered to the fluid by the pump, measured in horsepower (hp).
 Input power, also called brake horsepower, is the power that must
be supplied to the pump, measured in horsepower (hp).
 Efficiency is the ratio between the input power and output power. It
accounts for energy losses in the pump (friction and slip) to
describes how much of the input power does useful work.
Performance
The performance characteristics of axial flow pumps are different
from other pump types. Pump performance curves, which are
provided by a manufacturer to describe the correlation between
head and capacity of an individual pump, can be used
to describe these characteristics.
Department of Mechanical Engineering

The image above shows a typical performance curve for an axial


flow pump, depicting the relationship between head, flow rate,
power, and efficiency. As shown in the diagram, the shut-off
(zero flow) head of an axial flow pump can be as much as three
times the head at the pump's best efficiency point. Additionally,
the power requirement increases as flow decreases, with the
highest power draw at shut off. These trends are opposite radial
flow centrifugal pumps, which require more power as flow rate
increases.
Department of Mechanical Engineering

This collection of curves shows the change in performance at


different impeller pitch (angle). Power requirements and pump
head increase with increases in pitch, allowing pumps to be
tweaked to the conditions of the system to provide the most
efficient operation.
Media Type
Selecting the right pump requires an understanding of the
properties of the liquid in the addressed system. These
properties include viscosity and consistency.

 Viscosity is a measure of the thickness of a liquid. Viscous fluids


like sludges generate higher systems pressures and require more
pumping powerto move through the system. Low viscosity liquids
like water and oil which generate low head. Axial flow pumps are
designed to handle low viscosity fluids because they generate low
head and high capacities.
 Consistency is the material makeup of the liquid solution in terms
of chemicals and undissolved solids. Axial flow pumps are not
well-suited for handling media with solids, but can be used when
designed with the proper impeller type. Solutions with corrosive
chemicals should be handled by pumps with materials and parts
designed to withstand corrosion.
Department of Mechanical Engineering

Applications:
Axial flow pumps are used in applications requiring very high
flow rates and low pressures. They are used to circulate fluids in
power plants, sewage digesters, and evaporators. They are also
used in flood dewatering and irrigation systems. The applications
for axial flow pumps are not nearly as abundant however as for
radial flow pumps, so the equipment is not as common.

Pump scaling laws:


Dimensional Analysis:
Turbomachinery provides a very practical example of the power and
usefulness of dimensional analysis (Chap. 7). We apply the method of
repeating variables to the relationship between gravity times net head
(gH) and pump properties such as volume flow rate (V . ); some
characteristic length, typically the diameter of the impeller blades (D);
blade surface roughness height (e); and impeller rotational speed (v),
along with fluid properties density (r) and viscosity (m). Note that we
treat the group gH as one variable. The dimensionless Pi groups are
shown in Fig. 14–67; the result is the following relationship involving
dimensionless parameters:
Department of Mechanical Engineering

A similar analysis with input brake horsepower as a function of the same


variables results in

The second dimensionless parameter (or ) group) on the right side of both
Eqs. 14–28 and 14–29 is obviously a Reynolds number since vD is a
characteristic velocity,

The third on the right is the non -dimensional roughness parameter. The
three new dimensionless groups in these two equations are given
symbols and named as follows:

Note the subscript Q in the symbol for capacity coefficient. This comes
from the nomenclature found in many fluid mechanics and
turbomachinery textbooks that Q rather than V . is the volume flow rate
through the pump.
We use the notation CQ for consistency with turbomachinery convention,
even though we use V . for volume flow rate to avoid confusion with heat
transfer. When pumping liquids, cavitation may be of concern, and we
need another dimensionless parameter related to the required net
positive suction head. Fortunately, we can simply substitute
NPSHrequired in place of H in the dimensional analysis, since they have
identical dimensions (length). The result is
Department of Mechanical Engineering
Other variables, such as gap thickness between blade tips and pump
housing and blade thickness, can be added to the dimensional analysis if
necessary. Fortunately, these variables typically are of only minor
importance and are not considered here. In fact, you may argue that two
pumps are not even strictly geometrically similar unless gap thickness,
blade thickness, and sur face roughness scale geometrically.
Relationships derived by dimensional analysis, such as Eqs. 14–28 and
14–29, can be interpreted as follows: If two pumps, A and B, are
geometrically similar (pump A is geometrically proportional to pump B,
although they are of different sizes), and if the independent )’s are equal
to each other (in this case if CQ, A ! CQ, B, ReA ! ReB, and eA/DA ! eB/DB),
then the dependent )’s are guaranteed to also be equal to each other as
well. In particular, CH, A ! CH, B from Eq. 14–28 and CP, A ! CP, B from Eq.
14–29. If such conditions are established, the two pumps are said to be
dynamically similar (Fig. 14–68). When dynamic similarity is achieved, the
operating point on the pump performance curve of pump A and the
corresponding operating point on the pump performance curve of pump
B are said to be homologous.

The requirement of equality of all three of the independent


dimensionless parameters can be relaxed somewhat. If the Reynolds
numbers of both pump A and pump B exceed several thousand, turbulent
flow conditions exist inside the pump. It turns out that for turbulent flow,
if the values of ReA and ReB are not equal, but not too far apart, dynamic
similarity between the two pumps is still a reasonable approximation.
This fortunate condition is called Reynolds number independence. (Note
that if the pumps operate in the laminar regime, the Reynolds number
must usually remain as a scaling parameter.) In most cases of practical
turbomachinery engineering analysis, the effect of differences in the
roughness parameter is also small, unless the roughness differences are
Department of Mechanical Engineering
large, as when one is scaling from a very small pump to a very large pump
(or vice versa). Thus, for many practical problems, we may neglect the
effect of both Re and e/D. Equations 14–28 and 14–29 then reduce to

As always, dimensional analysis cannot predict the shape of the


functional relationships of Eq. 14–32, but once these relationships are
obtained for a particular pump, they can be generalized for geometrically
similar pumps that are of different diameters, operate at different
rotational speeds and flow rates, and operate even with fluids of different
density and viscosity.
We transform Eq. 14–5 for pump efficiency into a function of the
dimensionless parameters of Eq. 14–30,

Since hpump is already dimensionless, it is another dimensionless pump


para meter all by itself. Note that since Eq. 14–33 reveals that hpump
can be formed by the combination of three other )’s, hpump is not
necessary for pump scaling. It is, however, certainly a useful parameter.
Since CH, CP, and hpump are functions only of CQ, we often plot these
three parameters as functions of CQ on the same plot, generating a set of
nondimensional pump performance curves. An example is provided in
Fig. 14–69 for the case of a typical centrifugal pump. The curve shapes for
other types of pumps would, of course, be different.

The simplified similarity laws of Eqs. 14–32 and 14–33 break down when
the full-scale prototype is significantly larger than its model (Fig. 14–70);
the prototype’s performance is generally better. There are several
reasons for this: The prototype pump often operates at high Reynolds
numbers that are not achievable in the laboratory. We know from the
Moody chart that the friction factor decreases with Re, as does boundary
layer thickness. Hence, the influence of viscous boundary layers is less
Department of Mechanical Engineering
significant as pump size increases, since the boundary layers occupy a less
significant percentage of the flow path through the impeller. In addition,
the relative roughness (e/D) on the surfaces of the prototype impeller
blades may be significantly smaller than that on the model pump blades
unless the model surfaces are micropolished. Finally, large full-scale
pumps have smaller tip clearances relative to the blade diameter;
therefore, tip losses and leakage are less significant. Some empirical
equations have been developed to account for the increase in efficiency
between a small model and a full-scale prototype. One such equation was
suggested by Moody (1926) for turbines, but it can be used as a first-
order correction for pumps as well,

Pump Specific Speed:


Another useful dimensionless parameter called pump specific speed
(NSp) is formed by a combination of parameters CQ and CH:

If all engineers watched their units carefully, NSp would always be listed
as a dimensionless parameter. Unfortunately, practicing engineers have
grown accustomed to using inconsistent units in Eq. 14–35, which renders
the perfectly fine dimensionless parameter NSp into a cumbersome
dimensional quantity (Fig. 14–71). Further confusion results because
Department of Mechanical Engineering
some engineers prefer units of rotations per minute (rpm) for rotational
speed.
use rotations per second (Hz), the latter being more common in Europe.
In addition, practicing engineers in the United States typically ignore the
gravitational constant in the definition of NSp. In this book, we add
subscripts “Eur” or “US” to NSp in order to distinguish the dimensional
forms of pump specific speed from the nondimensional form. In the
United States, it is customary to write H in units of feet (net head
expressed as an equivalent column height of the fluid being pumped), V .
in units of gallons per minute (gpm), and rotation rate in terms of n .
(rpm) instead of v (rad/s). Using Eq. 14–35 we define

In Europe it is customary to write H in units of meters (and to include g !


9.81 m/s2 in the equation), V . in units of m3/s, and rotation rate n . in
units of rotations per second (Hz) instead of v (rad/s) or n . (rpm). Using
Eq. 14–35 we define

Affinity Laws:
We have developed dimensionless groups that are useful for relating any
two pumps that are both geometrically similar and dynamically similar. It
is convenient to summarize the similarity relationships as ratios. Some
authors call these relationships similarity rules, while others call them
affinity laws. For any two homologous states A and B,

Equations 14–38 apply to both pumps and turbines. States A and B can be
any two homologous states between any two geometrically similar
turbomachines, or even between two homologous states of the same
machine. Examples include changing rotational speed or pumping a
different fluid with the same pump. For the simple case of a given pump
in which v is varied, but the same fluid is pumped, DA ! DB, and rA ! rB. In
such a case, Eqs. 14–38 reduce to the forms shown in Fig. 14–74. A
“jingle” has been developed to help us remember the exponent on v, as
indicated in the figure. Note also that anywhere there is a ratio of two
rotational speeds (v), we may substitute the appropriate values of rpm (n
. ) instead, since the conversion is the same in both the numerator and
the denominator. The pump affinity laws are quite useful as a design tool.
Department of Mechanical Engineering
In particular, suppose the performance curves of an existing pump are
known, and the pump operates with reasonable efficiency and reliability.
The pump manufacturer decides to design a new, larger pump for other
applications, e.g., to pump a much heavier fluid or to deliver a
substantially greater net head. Rather than starting from scratch,
engineers often simply scale up an existing design. The pump affinity laws
enable such scaling to be accomplished with a minimal amount of effort.
Department of Mechanical Engineering
Write down a brief note on following about
Turbine Analysis (use diagrams and equation
where necessary)
Positive-Displacement Turbines, Dynamic
Turbines
Positive-Displacement Turbines:
ment pump running backward—as fluid pushes into a closed volume, it
turns a shaft or displaces a reciprocating rod. The closed volume of fluid is
then pushed out as more fluid enters the device. There is a net head loss
through the positive-displacement turbine; in other words, energy is
extracted from the flowing fluid and is turned into mechanical energy.
How ever, positive-displacement turbines are generally not used for
power pro duction, but rather for flow rate or flow volume
measurement.
The most common example is the water meter in your house (Fig. 14–80).
Many commercial water meters use a nutating disc that wobbles and
spins as water flows through the meter. The disc has a sphere in its center
with appro priate linkages that transfer the eccentric spinning motion of
the nutating disc into rotation of a shaft. The volume of fluid that passes
through the device per 360o rotation of the shaft is known precisely, and
thus the total volume of water used is recorded by the device. When
water is flowing at moderate speed from a spigot in your house, you can
sometimes hear a bubbly sound coming from the water meter—this is
the sound of the nutating disc wobbling inside the meter. There are, of
course, other positive-displacement turbine designs, just as there are
various designs of positive-displacement pumps.
Department of Mechanical Engineering

Dynamic Turbines:
Dynamic turbines are used both as flow measuring devices and as power
generators. For example, meteorologists use a three-cup anemometer to
mea sure wind speed (Fig. 14–81a). Experimental fluid mechanics
researchers use small turbines of various shapes (most of which look like
small pro pellers) to measure air speed or water speed (Chap. 8). In these
applications, the shaft power output and the efficiency of the turbine are
of little concern. Rather, these instruments are designed such that their
rotational speed can be accurately calibrated to the speed of the fluid.
Then, by electronically counting the number of blade rotations per
second, the speed of the fluid is calculated and displayed by the device.

A novel application of a dynamic turbine is shown in Fig. 14–81b. NASA


researchers mounted turbines at the wing tips of a Piper PA28 research
aircraft to extract energy from wing tip vortices (Chap. 11); the extracted
energy was converted to electricity to be used for on-board power
requirements.
In this chapter, we emphasize large dynamic turbines that are designed to
produce electricity. Most of our discussion concerns hydroturbines that
Department of Mechanical Engineering
uti lize the large elevation change across a dam to generate electricity.
There are two basic types of dynamic turbine—impulse and reaction,
each of which are discussed in some detail. Comparing the two power-
producing dynamic turbines, impulse turbines require a higher head, but
can operate with a smaller volume flow rate. Reaction turbines can
operate with much less head, but require a higher volume flow rate.
Impulse Turbine, Reaction Turbine

Impulse Turbines:
In an impulse turbine, the fluid is sent through a nozzle so that most of its
available mechanical energy is converted into kinetic energy. The
high speed jet then impinges on bucket-shaped vanes that transfer
energy to the turbine shaft, as sketched in Fig. 14–82. The modern and
most efficient type of impulse turbine was invented by Lester A. Pelton
(1829–1908) in 1878, and the rotating wheel is now called a Pelton wheel
in his honor. The buck ets of a Pelton wheel are designed so as to split
the flow in half, and turn the flow nearly 180° around (with respect to a
frame of reference moving with the bucket), as illustrated in Fig. 14–82b.
According to legend, Pelton modeled the splitter ridge shape after the
nostrils of a cow’s nose. A portion of the outermost part of each bucket is
cut out so that the majority of the jet can pass through the bucket that is
not aligned with the jet (bucket n $ 1 in Fig. 14–82) to reach the most
aligned bucket (bucket n in Fig. 14–82). In this way, the maximum amount
of momentum from the jet is utilized. These details are seen in a
photograph of a Pelton wheel (Fig. 14–83). Figure 14–84 shows a Pelton
wheel in operation; the splitting and turning of the water jet is clearly
seen.
Department of Mechanical Engineering

We analyze the power output of a Pelton wheel turbine by using the


Euler turbomachine equation. The power output of the shaft is equal to
vTshaft, where Tshaft is given by Eq. 14–14,

We must be careful of negative signs since this is an energy-producing


rather than an energy-absorbing device. For turbines, it is conventional to
define point 2 as the inlet and point 1 as the outlet. The center of the
bucket moves at tangential velocity rv, as illustrated in Fig. 14–82. We
simplify the analysis by assuming that since there is an opening in the
outermost part of each bucket, the entire jet strikes the bucket that
happens to be at the direct bottom of the wheel at the instant of time
under consideration (bucket n in Fig. 14–82). Furthermore, since both the
size of the bucket and the diameter of the water jet are small compared
to the wheel radius, we approximate r1 and r2 as equal to r. Finally, we
assume that the water is turned through angle b without losing any
speed; in the relative frame of reference moving with the bucket, the
relative exit speed is thus Vj # rv (the same as the relative inlet speed) as
sketched in Fig. 14–82. Returning to the absolute reference frame, which
is necessary for the application of Eq. 14–39, the tangential component of
velocity at the inlet, V2, t , is simply the jet speed itself, Vj . We construct
a velocity diagram in Fig. 14–85 as an aid in calculating the tangential
Department of Mechanical Engineering
compo nent of absolute velocity at the outlet, V1, t . After some
trigonometry, which you can verify after noting that sin (b # 90°) ! #cos b,

Obviously, the maximum power is achieved theoretically if b ! 180°.


How ever, if that were the case, the water exiting one bucket would
strike the back side of its neighbor coming along behind it, reducing the
generated torque and power. It turns out that in practice, the maximum
power is achieved by reducing b to around 160° to 165°. The efficiency
factor due to b being less than 180° is

When b ! 160°, for example, hb ! 0.97—a loss of only about 3 percent.


Finally, we see from Eq. 14–40 that the shaft power output W . shaft is
zero if rv ! 0 (wheel not turning at all). W . shaft is also zero if rv ! Vj
(bucket mov ing at the jet speed). Somewhere in between these two
extremes lies the optimum wheel speed. By setting the derivative of Eq.
14–40 with respect to rv to zero, we find that this occurs when rv ! Vj /2
(bucket moving at half the jet speed, as shown in Fig. 14–86).
Department of Mechanical Engineering

For an actual Pelton wheel turbine, there are other losses besides that of
Eq. 14–41: mechanical friction, aerodynamic drag on the buckets, friction
along the inside walls of the buckets, nonalignment of the jet and bucket
as the bucket turns, backsplashing, and nozzle losses. Even so, the
efficiency of a well-designed Pelton wheel turbine can approach 90
percent. In other words, up to 90 percent of the available mechanical
energy of the water is converted to rotating shaft energy.
Reaction Turbines:
From wind mills to hydro-power plants we have reaction turbines all
around the world to generate electricity efficiently. Almost 60% of
turbines used in hydro-power plants are reaction turbine. Unlike impulse
turbines they remain dipped in water and use the pressure energy of
water to generate power.
The idea of dipping the whole turbine into water came from Germany in
mid twelfth century. The basic thinking was that we can use the weight of
water to turn the turbine blades more efficiently compared to the water
just striking the water wheel’s base. But it took a look time to built an
efficiently working reaction turbine, which was built in mid 18th century
by Benoit Fourneyron, who is said to be the inventor of modern hydraulic
turbines, since then we are putting in our efforts to further improve our
designs, to make them more and more efficient. So let’s just figure it out
what we have in our modern reaction turbines.

Working Principle:
Department of Mechanical Engineering

The working of the reaction turbine can be well


understand by taking a rotor having moving nozzles and
water of high pressure is coming out of the nozzle. As the
water leaves the nozzle, a reaction force is experienced
by the nozzle. This reaction force rotates the rotor at very
high speed.

In the same way in reaction turbine, a reaction force is


generated by the fluid moving on the runner blades. The
reaction force produced on the runner blades makes the
runner to rotate. Fluid after moving over the runner
blades enters into draft tube and finally to the trail race.

Main Components:
Department of Mechanical Engineering

1. Spiral casing

It is a spiral casing, with uniformly decreasing cross-


section area, along the circumference. Its decreasing
cross-section area makes sure that we have a uniform
velocity of the water striking the runner blades, as we
have openings for water flow in-to the runner blades
from the very starting of the casing, so pressure would
decrease as it travels along the casing. So we reduce its
cross-section area along its circumference to make
pressure uniform, thus uniform momentum or velocity
striking the runner blades.

2. Guide vanes

Guide vanes are installed in the spiral casing, their most


important function is to make sure that water striking the
runner blades must have a direction along length of the
axis of turbine otherwise the flow would be highly
Department of Mechanical Engineering

swirling as it moves through spiral casing, making it in-


efficient to rotate runner blades. The angle of these guide
vanes is adjustable in modern turbines, and we can adjust
the water flow rate by varying the angle of these guide
vanes according to the load on the turbine.

3. Runner blades

Runner blades are said to be heart of a reaction turbine.


It is the shape of the runner blades which uses the
pressure energy of water to run turbine. Their design
plays a major role in deciding the efficiency of a turbine.
In modern turbines these blades can pitch about their
axis, thus can vary the pressure force acting on them
according to the load and available pressure.

4. Draft tube

Draft tube connects the runner exit to the tail race. Its
cross-section area increases along its length, as the water
coming out of runner blades is at considerably low
pressure, so its expanding cross-section area help it to
recover the pressure as it flows towards tail race.
Department of Mechanical Engineering

Turbine scaling laws:


Dimensionless Turbine Parameters:
We define dimensionless groups (Pi groups) for turbines in much the
same way as we did in Section 14–3 for pumps. Neglecting Reynolds
number and roughness effects, we deal with the same dimensional
variables: gravity times net head (gH), volume flow rate (V . ), diameter of
the runner blades (D), runner rotational speed (v), output brake
horsepower (bhp), and fluid density (r), as illustrated in Fig. 14–102. In
fact, the dimensional analysis is identical whether analyzing a pump or a
turbine, except for the fact that for turbines, we take bhp instead of V . as
the independent variable. In addition, hturbine (Eq. 14–44) is used in
place of hpump as the nondimensional effi ciency. A summary of the
dimensionless parameters is provided here:

The affinity laws (Eqs. 14–38) can be applied to turbines as well as to


pumps, allowing us to scale turbines up or down in size (Fig. 14–103). We
also use the affinity laws to predict the performance of a given turbine
Department of Mechanical Engineering
oper ating at different speeds and flow rates in the same way as we did
previ ously for pumps.
The simple similarity laws are strictly valid only if the model and the
pro totype operate at identical Reynolds numbers and are exactly
geometrically similar (including relative surface roughness and tip
clearance). Unfortu nately, it is not always possible to satisfy all these
criteria when performing model tests, because the Reynolds number
achievable in the model tests is generally much smaller than that of the
prototype, and the model surfaces have larger relative roughness and tip
clearances. When the full-scale proto type is significantly larger than its
model, the prototype’s performance is generally better, for the same
reasons discussed previously for pumps. Some empirical equations have
been developed to account for the increase in efficiency between a small
model and a full-scale prototype. One such equation was suggested by
Moody (1926), and can be used as a first-order correction,

Note that Eq. 14–49 is also used as a first-order correction when scaling
model pumps to full scale (Eq. 14–34). In practice, hydroturbine engineers
generally find that the actual increase in efficiency from model to
prototype is only about two-thirds of the increase given by Eq. 14–49. For
example, suppose the efficiency of a one tenth scale model is 93.2
percent. Equation 14–49 predicts a full-scale effi ciency of 95.7 percent,
or an increase of 2.5 percent. In practice, we expect only about two-thirds
of this increase, or 93.2 $ 2.5(2/3) ! 94.9 percent. Some more advanced
correction equations are available from the Interna tional
Electrotechnical Commission (IEC), a worldwide organization for
standardization.

You might also like