You are on page 1of 6

Piping Easy23764

Net positive suction head


From Wikipedia, the free encyclopedia
Jump to navigationJump to search
This article is about net positive suction head. For the fire hose coupling thread,
see Glossary of firefighting equipment §  N.
In a hydraulic circuit, net positive suction head (NPSH) may refer to one of two
quantities in the analysis of cavitation:
1. The Available NPSH (NPSHA): a measure of how close
the fluid at a given point is to flashing, and so to
cavitation. Technically it is the absolute pressure head
minus the vapour pressure of the liquid.
2. The Required NPSH (NPSHR): the head value at the
suction side (e.g. the inlet of a pump) required to keep
the fluid away from cavitating (provided by the
manufacturer).
NPSH is particularly relevant inside centrifugal pumps and turbines, which are parts of a
hydraulic system that are most vulnerable to cavitation. If cavitation occurs, the drag
coefficient of the impeller vanes will increase drastically—possibly stopping flow
altogether—and prolonged exposure will damage the impeller.

Contents
 1NPSH in a pump
 2NPSH in a turbine
 3NPSH design considerations
 4Relationship to other cavitation parameters
 5Some general NPSH examples
 6References
NPSH in a pump[edit]

A simple hydraulic pumping circuit. Point O is the free suction surface, and point i is the inlet of the impeller.

In a pump, cavitation will first occur at the inlet of the impeller. [1] Denoting the inlet by i,
the NPSHA at this point is defined as:
where  is the absolute pressure at the inlet,  is the average velocity at the inlet,  is the
fluid density,  is the acceleration of gravity and  is the vapor pressure of the fluid. Note
that it is equivalent to the sum of both the static and dynamic heads – that is, the
stagnation head – from which one deducts the head corresponding to the equilibrium
vapor pressure, hence "net positive suction head".
Applying the first law of thermodynamics for control volumes enclosing the suction free
surface 0 and the pump inlet i, under the assumption that the kinetic energy at 0 is
negligible, that the fluid is inviscid, and that the fluid density is constant:
Using the above application of Bernoulli to eliminate the velocity term and local pressure
terms in the definition of NPSHA:
This is the standard expression for the available NPSH at a point. Cavitation will occur
at the point i when the available NPSH is less than the NPSH required to prevent
cavitation (NPSHR). For simple impeller systems, NPSHR can be derived theoretically,
[2]
 but very often it is determined empirically.[1] Note NPSHAand NPSHR are in absolute
units and usually expressed in "m" or "ft," not "psia".
Experimentally, NPSHR is often defined as the NPSH3, the point at which the head
output of the pump decreases by 3 % at a given flow due to reduced hydraulic
performance. On multi-stage pumps this is limited to a 3 % drop in the first stage head. [3]
NPSH in a turbine[edit]
The calculation of NPSH in a reaction turbine is different to the calculation of NPSH in a
pump, because the point at which cavitation will first occur is in a different place. In a
reaction turbine, cavitation will first occur at the outlet of the impeller, at the entrance of
the draft tube.[4] Denoting the entrance of the draft tube by e, the NPSHA is defined in the
same way as for pumps:
[1]

Applying Bernoulli's principle from the draft tube entrance e to the lower free surface 0,
under the assumption that the kinetic energy at 0 is negligible, that the fluid is inviscid,
and that the fluid density is constant:
Using the above application of Bernoulli to eliminate the velocity term and local pressure
terms in the definition of NPSHA:
Note that, in turbines minor friction losses () alleviate the effect of cavitation - opposite
to what happens in pumps.

NPSH design considerations[edit]


Vapour pressure is strongly dependent on temperature, and thus so will both
NPSHR and NPSHA. Centrifugal pumps are particularly vulnerable especially when
pumping heated solution near the vapor pressure, whereas positive displacement
pumps are less affected by cavitation, as they are better able to pump two-phase flow
(the mixture of gas and liquid), however, the resultant flow rate of the pump will be
diminished because of the gas volumetrically displacing a disproportion of liquid. Careful
design is required to pump high temperature liquids with a centrifugal pump when the
liquid is near its boiling point.
The violent collapse of the cavitation bubble creates a shock wave that can carve
material from internal pump components (usually the leading edge of the impeller) and
creates noise often described as "pumping gravel". Additionally, the inevitable increase
in vibration can cause other mechanical faults in the pump and associated equipment.

Relationship to other cavitation parameters[edit]


The NPSH appears in a number of other cavitation-relevant parameters. The suction
head coefficient is a dimensionless measure of NPSH:
Where  is the angular velocity (in rad/s) of the turbo-machine shaft, and  is the turbo-
machine impeller diameter. Thoma's cavitation number is defined as:
Where  is the head across the turbo-machine.

Some general NPSH examples[edit]


(based on sea level).
Example Number 1: A tank with a liquid level 2 metres above the pump intake, plus
the atmospheric pressure of 10 metres, minus a 2 metre friction loss into the pump (say
for pipe & valve loss), minus the NPSHR curve (say 2.5 metres) of the pre-designed
pump (see the manufacturers curve) = an NPSHA (available) of 7.5 metres. (not
forgetting the flow duty). This equates to 3 times the NPSH required. This pump will
operate well so long as all other parameters are correct.
Remember that positive or negative flow duty will change the reading on the pump
manufacture NPSHR curve. The lower the flow, the lower the NPSH R, and vice versa.
Lifting out of a well will also create negative NPSH; however remember that
atmospheric pressure at sea level is 10 metres! This helps us, as it gives us a bonus
boost or “push” into the pump intake. (Remember that you only have 10 metres of
atmospheric pressure as a bonus and nothing more!).
Example Number 2: A well or bore with an operating level of 5 metres below the
intake, minus a 2 metre friction loss into pump (pipe loss), minus the NPSH R curve (say
2.4 metres) of the pre-designed pump = an NPSH A (available) of (negative) -9.4 metres.
Adding the atmospheric pressure of 10 metres gives a positive NPSH A of 0.6 metres.
The minimum requirement is 0.6 metres above NPSH R), so the pump should lift from the
well.
Using the situation from example 2 above, but pumping 70 degrees Celsius (158F)
water from a hot spring, creating negative NPSH, yields the following:
Example Number 3: A well or bore running at 70 degrees Celsius (158F) with an
operating level of 5 metres below the intake, minus a 2 metre friction loss into pump
(pipe loss), minus the NPSHR curve (say 2.4 metres) of the pre-designed pump, minus a
temperature loss of 3 metres/10 feet = an NPSH A (available) of (negative) -12.4 metres.
Adding the atmospheric pressure of 10 metres and gives a negative NPSH A of -2.4
metres remaining.
Remembering that the minimum requirement is 600 mm above the NPSHR therefore this
pump will not be able to pump the 70 degree Celsius liquid and will cavitate and lose
performance and cause damage. To work efficiently, the pump must be buried in the
ground at a depth of 2.4 metres plus the required 600 mm minimum, totalling a total
depth of 3 metres into the pit. (3.5 metres to be completely safe).
A minimum of 600 mm (0.06 bar) and a recommended 1.5 metre (0.15 bar) head
pressure “higher” than the NPSHR pressure value required by the manufacturer is
required to allow the pump to operate properly.
Serious damage may occur if a large pump has been sited incorrectly with an incorrect
NPSHR value and this may result in a very expensive pump or installation repair.
NPSH problems may be able to be solved by changing the NPSH R or by re-siting the
pump.
If an NPSHA is say 10 bar then the pump you are using will deliver exactly 10 bar more
over the entire operational curve of a pump than its listed operational curve.
Example: A pump with a max. pressure head of 8 bar (80 metres) will actually run at 18
bar if the NPSHA is 10 bar.
i.e.: 8 bar (pump curve) plus 10 bar NPSHA = 18 bar.
This phenomenon is what manufacturers use when they design multistage pumps,
(Pumps with more than one impeller). Each multi stacked impeller boosts the
succeeding impeller to raise the pressure head. Some pumps can have up to 150
stages or more, in order to boost heads up to hundreds of metres.
References[edit]
1. ^ Jump up to:      Frank M. White Fluid Mechanics, 7th Ed., p. 771
a b c

2. ^ Paresh Girdhar, Octo Moniz, Practical Centrifugal Pumps, p. 68


3. ^ "Welcome to the Hydraulic Institute". Archived from the original on
2010-03-23.
4. ^ "Cavitation in reaction turbines". Archived from the original on 2016-
03-10.

Categories: 
 Hydraulics
 Fluid mechanics
Navigation menu
 Not logged in
 Talk
 Contributions
 Create account
 Log in
 Article
 Talk
 Read
 Edit
 View history
Search
Search Go

 Main page
 Contents
 Current events
 Random article
 About Wikipedia
 Contact us
 Donate
Contribute
 Help
 Learn to edit
 Community portal
 Recent changes
 Upload file
Tools
 What links here
 Related changes
 Special pages
 Permanent link
 Page information
 Cite this page
 Wikidata item
Print/export
 Download as PDF
 Printable version
Languages
 ‫العربية‬
 Deutsch
 Español
 Français
 한국어
 日本語
 Português
5 more
Edit links
 This page was last edited on 17 May 2021, at 04:31 (UTC).
 Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation,
Inc., a non-profit organization.
 Privacy policy

 About Wikipedia

 Disclaimers

 Contact Wikipedia

 Mobile view

 Developers

 Statistics
 Cookie statement

You might also like