You are on page 1of 375

Free Engineering Books http://boilersinfo.

com/
Centrifugal Pump
User's Guidebook
PROBLEMS AND SOLUTIONS

Free Engineering Books http://boilersinfo.com/


Centrifugal Pump
User's Guidebook
PROBLEMS AND SOLUTIONS

SAM YEDIDIAH

CHAPMAN & HALL

lOOP An International Thomson Publishing Company


New York' Albany· Bonn' Boston' Cincinnati· Detroit· Madrid' Melbourne
Mexico City· Pacific Grove • Paris • San Francisco • Singapore • Tokyo • Toronto • Washington

Free Engineering Books http://boilersinfo.com/


Cover design: Robert Freese

Copyright © 1996 by Chapman & Hall


Softcover reprint of the hardcover I st edition 1996

Chapman & Hall Chapman & Hall


115 Fifth Avenue 2-6 Boundary Row
New York. NY 10003 London SE 1 8HN
England

Thomas Nelson Australia Chapman & Hall GmbH


102 Dodds Street Postfach 100263
South Melbourne. 3205 0-69442 Weinheim
Victoria. Australia Germany

Nelson Canada International Thomson Publishing Asia


1120 Birchmount Road 221 Henderson Road #05-10
Scarborough. Ontario Henderson BuUding
Canada M1K 5G4 Singapore 0315

International Thomson Editores International Thomson Publishing-Japan


Campos EIiseos 385. Piso 7 Hirakawacho-cho Kyowa BuUding. 3F
Col. Polanco 1-2-1 Hirakawacho-cho
11560 Mexico O.F. Chiyoda-ku. 102 Tokyo
Mexico Japan

All rights reserved. No part of this book covered by the copyright hereon may be reproduced or used
in any form or by any means-graphic. electrOniC. or mechanical. including photocopying. recording.
taping. or information storage and retrieval systems-w1thout the written permission of the publisher.

1 2 3 4 5 6 7 8 9 10 XXX 01 00 99 98 97 96 95

Libnuy of Congreu CatalogiDg-in-PubUcatiOD Data

Yedidiah. S.
Centrifugal Pump User's Guidebook: problems and solutions / Shmariahu Yedidiah.
p. cm.
Includes bibliographical references and index:.

ISBN-13:978-1-4612-85 16-8 e-ISBN-13:978-1-4613-1217-8


001:10.1007/978-1-4613-1217-8

1. Centrifugal Pumps-Handbooks. manuals. etc. I. Title.


TJ919.Y39 1996
621.6'7--dc20 96-22405
CIP
Visit Chapman & Hall on the Internet: http://www.chaphall.comlchaphall.html

To order this or any other Chapman & Hall book. please contact Internatlooal ThOmsoD PubUshiDg,
7625 Empire Drive. Florence. KY 41042. Phone: (606) 525-6600 or 1-800-842-3636.
Fax: (606) 525-7778. e-mail: order@chaphall.com.

Fbr a complete listing of Chapman & Hall titles. send your request to Cbapman II: Hall, Dept. BC,
115 Fifth Avenue, New York, NY 10003.

Free Engineering Books http://boilersinfo.com/


Contents

Preface vii

PART I. INTRODUCTION 1

Chapter 1. Rotodynamic Pumps 3


Chapter 2. Performance Chracteristics of Centrifugal Pumps 13
Chapter 3. Classification of Centrifugal Pumps 19

PART II. PERFORMANCE FACTORS 29

Chapter 4. Effects of Free Air in the Pumped Liquid 31


Chapter 5. Cavitation 33
Chapter 6. Cavitation in Centrifugal Pumps 45
Chapter 7. Losses of Energy 61
Chapter 8. Effects of Temperature and Viscosity on Pump Performance 75
Chapter 9. Recirculation 79
Chapter 10. Axial and Radial Thrust and Balancing 97
Chapter 11. Miscellaneous Factors that Affect Pump Performance III

PART III. PROBLEMS ENCOUNTERED WITH CENTRIFUGAL PUMPS 115

Chapter 12. Testing 117


Chapter 13. Pump Performance at Reduced NPSH 145
Chapter 14. Pumping System Layout 163
Chapter 15. Installation. Handling, and Operation of Pumps and
Pumping Systems 193
Chapter 16. Problems with Bearings 209
Chapter 17. Sealing Rotating Parts 221
Chapter 18. Miscellaneous Studies 229
Chapter 19. Problems Related to Specific Circumstances 243
Chapter 20. Special Cases that Have Proven Very Difficult to Solve 263

Free Engineering Books http://boilersinfo.com/


VI Contents

PART IV. SOLVING PUMP PROBLEMS 271

Chapter 21. Solving Problems Prior to Visiting the Site 273


Chapter 22. Conclusions Drawn from Visual Inspection of Failed Parts 281
Chapter 23. On-Site Inspection and Testing 289

PART V. ELIMINATING PUMP PROBLEMS AND MODIFYING


PERFORMANCE 297

Chapter 24. Remedial Methods 299


Chapter 25. The Effects of Speed and Impeller OD on Pump Performance 305
Chapter 26. The Effects of Reducing the Impeller Width 313
Chapter 27. Moditying the Casing Geometry 317
Chapter 28. Various Methods of Altering Pump Performance 323

PART VI. A GLIMPSE OF THE FUTURE 331

Chapter 29. New Findings Concerning the Mode of Operation


of Rotodynamic Impellers 333
Chapter 30. Some Future Applications of the Presented Theory 343

Appendix A. Checklist of Problems with Centrifugal Pumps and


their Causes 347

1. Pump Does Not Develop Any Head, Nor Does It Deliver Liquid 348
2. Pump Develops Some Pressure, but Delivers No Liquid 348
3. Pump Delivers Less Liquid Than Expected 349
4. Pump Does Not Develop Enough Pressure 349
5. Shape of Head-Capacity Curve Differs from Rated Curve 349
6. Pump Consumes Too Much Power 350
7. Pump Does Not Perform Satisfactorily, Although Nothing
Appears to be Wrong with Pumping Unit or System 350
8. Pump Operates Satisfactorily During Start, but Performance
Deteriorates in a Relatively Short Time 351
9. Pump is Operating with Noise, Vibrations, or Both 351
10. Stuffing Box Leaks Excessively 352
11. Packing Has Short Life 353
12. Mechanical Seal Has Short Life 354
13. Mechanical Seal Leaks Excessively 354
14. Bearings Have Short Life 354

Free Engineering Books http://boilersinfo.com/


Contents Vll

15. Bearings Overheat 355


16. Bearings Operate with Noise 355
17. Pump Overheats, Seizes, or Both 356
18. Impeller or Casing, or Both, Has Short Life 356
19. Loud Blow is Heard Each Time Pump is Started or Stopped 357
20. Casing Bursts Each Time Pump is Started or Stopped 357
21. Gaskets Leak During Pump Operation 357
22. Flow-Rate Periodically Decreases, or Stops Completely, then
Returns to Normal 357
23. Pump Develops Cavitation when the Available NPSH is
Increased 357

Appendix B. Tables 359

Vapor Pressures of Water at Different Temperatures 359


Gall Resistance of Material Combinations 360
Packing Selections 361
Tables of Chemical Compatibility of Materials 362

Bibliography 375

Index 379

Free Engineering Books http://boilersinfo.com/


Preface

This book discusses two classes of activities: preventive procedures and corrective
steps. Sometimes these two activities are so intimately connected that it is impossible
to separate one from the other. Also, to execute these activities properly, an in-depth
knowledge and understanding of centrifugal pump operation and diagnostic
procedures is required.
The pump designer must adopt preventive steps to keep the possibility of failure at
a minimum. A pump must be selected carefully for each application and the system
of which it is destined to be a part must also be designed carefully. Preventive damage
control also requires great care in the manufacture, installation, and maintenance of
a pump.
The corrective procedures are related to diagnosing the cause of the malfunction
and taking the proper steps to correct the problem. Implementing corrective and
preventive procedures requires an in-depth knowledge of what makes a pump
operate satisfactorily and what affects its operation. It also requires a wide knowledge
of how different factors can affect the performance of a given pump.
The objective of this book is to provide the reader with the information required to
prevent, diagnose, and solve the widest possible spectrum of problems encountered
with centrifugal pumps. Because of the wide range of problems dealt with in this
book, it may prove to be of great assistance to a very diverse group. It is intended to:

1. Explain to pumping system designers how the layout of the suction fump, the
suction line, and the discharge line can affect pump performance.
2. Identify problems pump deSigners must take into account in order to incorporate
proper preventive and safety features that will make their designs more com-
petitive.
3. Describe the characteristics of materials and workmanship that are crucial for
turning out a good product. This will help quality-control and production
personnel establish and maintain the delicate balance between cost and quality.
4. Help those concerned with the selection and purchase of pumps make the best
choice.
5. Make it easier for maintenance personnel to determine and locate the source of
any problem and provide gUidelines for handling these problems. The book also
outlines steps that may be taken when a pump has to operate under
unanticipated conditions.
ix

Free Engineering Books http://boilersinfo.com/


x Preface

6. Caution personnel responsible for installing pumps and the pumping system
about problems they may create by not following correct installation procedures.
7. Provide pump attendants with a better understanding of how the equipment
functions and how to identifY signs of developing malfunctions.

This book may be of interest and assistance to people with different levels of
education and training. We therefore start with a review of some of the most basic
principles on which centrifugal pump operation is founded. This has been done for
the purpose of making this book accessible to those who may not be familiar with
these principles. People who are familiar with the problems discussed in the
introductory part of this book may safely skip it and start directly with the more
advanced topics.
Troubleshooting is, to a great extent, the ability to deal with unforeseen events.
This book tries to deal with as many situations as possible. Most problems can be
solved simply by consulting the checklists and following the instructions without
worrying about the analytical background.
On the other hand, unique problems are always possible. These cases are often
very difficult to analyze and solve. In such cases, an in-depth understanding of the
most basic principles can be of great help in finding the proper solution.

Sam Yedidiah

Free Engineering Books http://boilersinfo.com/


Centrifugal Pump
User's Guidebook
PROBLEMS AND SOLUTIONS

Free Engineering Books http://boilersinfo.com/


Introduction

Part I explains the principles and modes of operation of centrifugal pumps.


It begins with the basic laws of mechanics and continues with a discussion
of how these laws govern the operation of centrifugal pumps. It also
explains the methods by which centrifugal pumps are classified and defines
their operational characteristics. Finally, this part of the book explains how
a knowledge and understanding of the performance characteristics of
centrifugal pumps can help the pump user choose the best pump for a
given range of duties.

Free Engineering Books http://boilersinfo.com/


CHAPTER
1
Rotodynamic Pumps
The term rotodynamic pump refers to a class of devices that use rotating blades to
increase the head of the pumped liquid, by altering its state of motion. This class of
pumps includes the following: Axial flow, Centrifugal and mixed-flow, Regenerative pumps.
In this book, we are concerned primarily with the second group, but for ease in
explaining the basic principles. we shall relate certain details to the effects that these
prinCiples may have on pumps axia.l-flow pumps.

POWER TRANSMIITED BY A MOVING BLADE

Assume that a blade is totally submerged in a liquid and that both the blade and the
liquid are at rest. No transfer of power from the blade to the liquid (or vice versa) can
occur. Now assume that a person, who is moving with a uniform veloCity (- U). is
observing the behavior of this blade and the surrounding liquid. In relation to that
observer, both the blade and the liquid are moving with the same velocity (+U). This
velocity will not and cannot cause any exchange of power between the blade and the
liquid. The only time such an exchange can occur is when the blade is moving with a
velocity U and the liquid is moving in the same direction with a velocity C u , which is
different from U.
Assume that a blade of length L is moving with a velocity U in a liquid moving in
the direction marked by the arrow C in Fig. 1-1. Let the blade be inclined to the
direction of U by an angle f3. and let the velocity C consist of the sum of two velocity
components: CuI, which is parallel to U, and Cm• which is perpendicular to U.
If CUI is smaller than U (which is accepted practice in pumps), the blade L will
displace. in a unit of time, a volume of liqUid equal to

Figure 1-1 Symbols used in


explaining the interaction between a
moving blade and the surrounding
liquid. The distance A is the axial
distance from the inlet tips of the
blades.
3

Free Engineering Books http://boilersinfo.com/


4 Introduction

(1-1)

where B is the width of the blade in the direction perpendicular to the plane of the
illustration.
When the liquid comes into direct contact with the blade, the liquid must move
along the blade's surface. This change in direction produces a change in the
magnitude of Cu from CUI to Cu2. From the magnitudes and directions of the velocity
components and the blade it follows that when the liquid is being forced to flow along
the blade, the magnitude of Cu2 assumes the value of:

Cu2 = U- C m cot f3 (1-2)

To bring about such a change in the magnitude of the velocity C u , the blade must
exert a force in the direction of U. This force acts while the blade is moving in the same
direction. In other words, it is acting along a certain distance and, consequently,
performing work. The amount of work Wb the blade performs on the liquid per unit of
time is given by

Wb = yQb U (Cu2 - Cud. (1-3)


9

The term yQb/ 9 is the mass m of the liquid acted on in a unit of time. Consequently,
we get

Introducing this expression into Eq. 1-3 and writing U as distance divided by time,
we get:

liT
vVb =
[
m x (C U2 - Cud] x distance . (1-4)
time time

The expression [(Cu2-Cu d/time] signifies acceleration. Consequently, the expression


[m(Cu2 -Cu d/time] means force. Because, by definition, U and C u are acting in the
same direction, Eq. 1-4 defines the dot product of force multiplied by distance per
unit of time. In other words, Eq. 1-4 expresses the amount of power transmitted from
the blade to the liquid. (Power is defined by the amount of work performed in a unit
of time.)
Now let us replace the term Qb in Eq. 1-3 with the value given in Eq. 1-1. In that
case, we get

Wb = y(U - Cud x B x L x sinf3 x u( C u2 ~ CUI). (1-5)

This equation is particularly useful when the blade is curved, as shown in Fig. 1-2. In
this case, we can subdivide the blade into smaller elements LlL. Having done this, we
calculate the amounts of liquid LlQb displaced by each of the elements LlL using

Free Engineering Books http://boilersinfo.com/


Rotodynamic Pumps 5

Figure 1-2 Symbols used in the


study of power transfer from a curved A
blade to the surrounding liquid. The I
distance A is the axial distance from
the inlet tips of the blades.

Eq. 1-1. We can then calculate the power d Wb transmitted by each blade element.
Finally. by adding these values. we find the total amount of power transmitted by the
blade to the liquid.
Thus far. we have discussed only a simple case in which all points of the blade are
moving with the same velocity U. What happens when each point of the blade is
moving with a different velocity?
Referring to Fig 1-3. assume that the blade is rotating around the center 0 with an
angular speed w. In that case. the peripheral speed of any point of the blade depends
on its distance R from the center 0 of rotation and is equal to wR.
Let us consider the effect of a blade element dL on the liquid. Applying the method
used in Fig. 1-2. we conclude that the amount of liquid displaced by any given blade
element dL is equal to (Fig. 1-3)

dQb=B(wR-CulXdL sin B (1-6)


=B(wR-CuldR

. llRI2

I R \

o~l/
'--1---'--'--'--'--'--
llR-1

Figure 1-3 Symbols used in


the definition of power transfer
from a rotating blade to the I
surrounding liquid

Free Engineering Books http://boilersinfo.com/


6 Introduction

where B is the width of the blade element LlL in the direction perpendicular to the
plane of the illustration.
The amount of power transmitted by the blade element LlL is given by

(1-7)

where U2= w(R+O.5 LlR)


U I = w(R-O.5 LlR)
C u2 = magnitude of Cu at (R+O.5 LlR)
CuI = magnitude of C u at (R-O.5 LlR)

It is beyond the scope of this book to go into the details of how the head
calculations are carried out in practice. This is of primary interest to pump designers.
(For more details, see Part VI.) For pump users, the significance of this discussion
lies in the fact that, according to Eq. 1-7 the amount of power transmitted from a
blade to the liquid depends not only on the blade's peripheral velocity U but on the
angle of its inclination to the direction of U and its area. We have also demonstrated
that the total amount of power transmitted by a blade to the liquid is equal to the
sum of the amounts of power transmitted by each of the blade elements. The amount
of power transmitted from a blade to the liquid depends on the shape of the total
blade, not just on the shape of its inlet and outlet tips.
This conclusion differs dramatically from the earlier, generally accepted notion that
the power transmitted by a blade depends only on the angles of its inlet and outlet
tips and is unaffected by the shape of the rest of the blade. In the past, this theory
often led to incorrect conclusions, as demonstrated by the test results shown in
Fig. 1-4.
An in-depth understanding of the manner in which power is transmitted from a
blade may, in some instances, help us understand a problem encountered with a
particular centrifugal pump.

Centrifugal Pump Operation


A centrifugal pump consists, essentially, of a rotating impeller I installed inside a
stationary casing C, as shown schematically in Fig. 1-5.
The impeller consists of blades B that are usually cast integrally with the shroud
Sh and the hub hb. The other side of each blade usually ends with a second shroud
Ss, which has an opening (eye) E at its center, surrounded by a sealing ring Sr. The
casing C consists of a stationary collector prOvided with a discharge nozzle Dn and a
suction nozzle Sn, which is connected to the suction line SL. If the pump is located
above the level of the liqUid in the suction tank, the line SL is usually prOvided with a
nonreturn valve Nr.
Now, let us fill the suction line and the pump with liqUid. When the pump is
started, the centrifugal forces generated by the rotating blades force the contained
liquid outward and, at the same time, reduce the pressure at the pump inlet. The total
head developed by the impeller is equal to the power transmitted by a single blade
(which can be calculated using Eq. 1-7). multiplied by the number of blades, and
divided by the mass flow of the pumped liquid.

Free Engineering Books http://boilersinfo.com/


Rotodynamic Pumps 7

1 - - - - - - 235 Dia - - - - - t - -
13 1 = 18° 30'
13 2 = 30°
Number of blades:
z=6

10

~
<D
Qi
E-
:r:
-cell
<D
I

VI
1300 RPM
6 blades
13 1 = 18° 30'
Figure 1-4 Performances of six
13 2 = 30°
different impellers of the same inlet
and outlet angles, same number of
blades, and identical axial section,
but differing in blade shape.
RPM=Revolutions per minute.

To begin pumping, the head developed by the impeller must be able to develop
enough of a difference between the pressures at the inlet and outlet nozzles to do the
following

1. Overcome the pressure, elevation, velocity head, and the frictional and viscous losses in
the discharge line
2. Reduce the pressure at the suction nozzle Sn of the pump to below the pressure at the
water level Sl-Sl in the suction tank. The magnitude of this reduction should be
adequate to overcome the resistance to flow created by the sum of the elevation HL above
the liquid in the suction tank and the frictional and viscous losses in suction line. It
should also provide an excess H2 of head, which is necessary to accelerate the liquid
entering the suction line.

Free Engineering Books http://boilersinfo.com/


8 Introduction

Nr

Figure 1-5 Symbols


used in the description
of the pumping action
of a centrifugal pump.

SUCTION LIFT AND CAVITATION

In an open tank, the pressure at the water level SI-SI (Fig. 1-5) is equal to the
atmospheric pressure Pa. To make the liquid reach the centerline 0-0 of the pump,
the pressure at 0-0 must be less than -yHL, where -y is the specific weight of the
pumped liquid. Consequently, the head Hs must be equal to:
Pa
Hs = --HL
Y (1-8)
= Ha -HL.

In practice, all liquids can only be subjected to compressive forces. Consequently,


the minimum pressure that can exist at any point in the liquid cannot be smaller
than zero. This, in turn, means that if the suction tank is open to the air, the
maximum elevation of a liquid in the suction line cannot be greater than Pah=Ha.
In reality, the suction lift HL must be Significantly lower than the head H a , which
expresses the atmospheric pressure, divided by the specific weight of the liquid. First,
a part of the available head Ha must be used to accelerate the liquid that enters the
suction line SL. Another part must be used to overcome the resistance to flow in the

Free Engineering Books http://boilersinfo.com/


Rotodynamic Pumps 9

suction line. Finally, if the absolute pressure at 0-0 falls below the vapor pressure of
the liquid, the pumped liquid will flash into vapor, causing vapor-filled voids.
The development of vapor-filled voids within the pumped liquid is generally known
as cavitation. We will deal with this problem in more detail later.

PRIMING

As a general rule, a centrifugal pump cannot start pumping unless it has been
primed. Before starting, the pump and the suction line must be completely filled with
the liquid to be pumped. Priming is intimately associated with centrifugal pump
operation.
The head developed by a centrifugal pump can be defined as the amount of power
transmitted from the blades to the liquid (Eqs. 1-3 and 1-7), divided by the mass flow
"IQb. Therefore, for a given blade geometry and a given blade velocity U, an impeller
will develop the same head at the given flow rate, regardless of the specific weight of
the liquid.
For example, assume that a pump handling water develops a head of 231 ft. The
pressure gage reading will show a differential pressure of 100 psi. Now assume that
the same pump is used to handle a mineral oil with a specific weight of 0.75. The
head developed by the pump at the same flow rate will again be equal to 231 ft.
However, because the specific weight of the mineral oil is only 0.75 times the specific
weight of water, the pressure differential is only 75 psi.
Now let us return to Fig. 1-5 and to the description of the process of pumping.
First, let us assume that the system is primed. The pump will develop a total head of
Ht=Hd+Hs . Hd is the head necessary to overcome the resistance in the discharge line.
The head H s , on the other hand, is used to reduce the pressure at the centerline 0-0
of the pump by an amount equal to "I1Hs. This reduction of pressure should be great
enough to overcome the pressure "I1HL reqUired to lift the water through the height
HL and provide enough pressure to overcome the frictional resistance in the suction
pipe and accelerate the liqUid flOwing through the suction line SL.
When the pump is primed, the pressure drop at the centerline 0-0 is equal to "IHs,
where "11 is the specific gravity of the pumped liqUid.
Now assume that the pump has not been primed before starting. It has been filled
with air. When started, it still develops the same total head of Ht=Hd+Hs . This time,
however, the drop of pressure at the centerline 0-0 is the magnitude of the same
head H s , multiplied by the specific gravity of air.
Depending on the temperature, pressure, and moisture content, the specific gravity
of air is about 600 to 650 times less than that of water. Consequently, the decrease in
pressure drop will now be about 600 to 650 times smaller than that of the pump
when it has been primed. Such a minute drop in pressure will rarely be adequate to
lift the water through the elevation HL. That is why priming is necessary.

PRESSURE DIFFERENTIAL ON BOTH SIDES OF A BLADE

Equations 1-3-1-7 define the amount of power transmitted from the blades to the
liquid. This power must be supplied to the blades from the shaft. This in turn, means

Free Engineering Books http://boilersinfo.com/


10 Introduction

that the blade must overcome a resistance to its motion. In other words, the pressure
on the leading face of the blades must be higher than on the trailing face.
In Eq. 1-7, we can write

Compare this to Eq. I-I.


Again, the force ilF, that the blade element ilL must overcome to transfer power
from the shaft to the liquid can be expressed as

f1.F= f1.Wb
U
= ilWb
UJR

where wR is the mean radius of the blade element ilL.


This force ilF acts on an area M which is equal to B(ilLsin 13). Consequently, the
pressure differential between the leading and the trailing faces of the blade element
ilL can be expressed as

Pd = Il.F
B x Il.Lsin~
(1-9)
y(wR - C u ) x (U 2 C u2 - UjCud
gwR

where Cu is the magnitude of the Cu component of the liquid at the mean radius of
the blade element ilL.

BERNOULLI'S EQUATION FOR A ROTATING SYSTEM

When a liquid is acted on by an impeller, its total head increases from Hj to H2.
Mathematically, this is expressed by

This equation can be transformed into the following form:

H2 _ H j = (U~ - ur + W 2 - wi) + (C~ - Cr)


j
0-10)
29

where C and Ware the absolute and relative velocities of the liquid, respectively, and
1 relates to blade inlet and 2 relates to blade outlet.
This increase of the total head consists of the increase in the kinetic head

Free Engineering Books http://boilersinfo.com/


Rotodynamic Pumps 11

and the increase in the pressure head (Hp 2-HpI L which is equal to the difference
between the total head and the kinetic head, or

(1-11)

The increase in pressure head is therefore

H p2 H pI W( - W{ u~ - ul
- = +
2g 2g
or (1-12)

H p2 W{ -u~
+
2g
Equation 1-12 is often referred to as Bernoulli's equation for a rotating system.
Now assume that an observer is rotating with the impeller. In relation to that
observer, both UI and U2 are equal to zero. In relation to the rotating observer

This is Bernoulli's equation for a stationary system. This analysis indicates that the
flow with respect to a rotating observer is subject to the same laws as is our system.

Free Engineering Books http://boilersinfo.com/


CHAPTER 2
Performance
Characteristics of
Centrifugal Pumps

The energy transfer in a centrifugal pump is accomplished by changing the state of


motion of the liquid. The magnitude of this change depends upon the dimensions and
shape of the waterways and on the operating speed of the impeller. For a given pump
geometry and operating speed, the velocities of the liquid vary with the flow rate.
Consequently, centrifugal pumps develop different heads at different flow rates.
Similarly, efficiency, power requirements, and suction capability also vary with flow
rate.
In common parlance, we usually describe the performance-features of a centrifugal
pump by specifYing the head developed by the pump at a given flow rate. Conversely,
we often define the performance of a centrifugal pump by stating the flow rate
delivered against a given head.
In practice, however, a centrifugal pump rarely operates against a certain given
head, or delivers a constant amount of liqUid. Most pumps must operate over a
certain range of heads and flows. Consequently, it is often very important to know
how the performance features of a pump change with the flow rate.
When a new pump is developed by a manufacturer, it is tested under controlled
conditions. The results of these tests are then plotted as curves on what are
commonly known as peiformance charts. Figure 2-1 shows a typical performance
chart. The curve marked QH shows how the developed head changes with the flow
rate. The curve marked HP represents the power consumed at different flows, and the
curve marked EFF shows the ratio between the actual amount of power added to the
liqUid and the power consumed by the pump at the given flow rate.
Performance charts also show the minimum head required at the suction nozzle of
the pump to avoid cavitation. This is usually expressed by a curve marked NPSH (net
positive suction head).
Under normal operating conditions, a pump is expected to demonstrate the same
performance characteristics in the field as shown in the performance chart. If it
does not, something is probably wrong with either the pump or the pumping
system.
13
14 Introduction

11-----

t
tt

Figure 2-1 A typical performance


Flow - - - - chart of a centrifugal pump.

OPERATING CONDITIONS AND PERFORMANCE CURVES

Steep QH Curve versus Flat QH Curve


Figure 2-2 shows the performance of two pumps that are developing the same head
HI at the same specified flow QI. At the given operating point, both pumps operate
with exactly the same efficiency. Which pump will perform better in practice?
Assume that the pump must supply water to the spraying nozzles of a battery of
machines used to clean raw vegetables. The number of machines used at any given
time will, of course, depend on the abundance of the produce, and therefore the flow
rate will vary with the size of the harvest. However, to get the best results, the head
will have to be kept nearly constant over the total working range of the pump. In
such a case, the best choice is a pump that produces a flat QH curve, such as curve
A in Fig. 2-2.
Now assume that a pump that supplies water for irrigation draws from a well
in which the water level varies significantly within very short time periods. Such a
pump must operate against a variable head. To satisfy the demand under varying
head conditions, the best choice is a pump with a steep curve (curve B in
Fig. 2-2).
Another factor that must be considered is power consumption. In a pump with a
flat QH curve (curve A), the power consumption usually increases continuously with
the flow rate. This may lead to overloading of the power supply in the event of a
rupture in a discharge pipe or some similar accident. A pump with a very steep QH
curve, however, may overload the driver at reduced flow rates.

Drooping QH Curve
Another problem related to the shape of the QH curve is what is commonly called a
"drooping" curve. In general, the head developed by a centrifugal pump increases
Performance Characteristics of Centrifugal Pumps 15

HP for CUrve B

HP lor curve A

Figure 2-2 Comparison between


a flat and a steep QH curve. Flow rate - - -

with a reduced flow rate. In some pumps, however, the head reaches its maximum at
a partial flow rate. Any additional reduction in flow then causes the head to drop, as
shown in Figs. 2-3 and 2-4.
In general, a drooping curve is most frequently encountered in pumps with high
head coefficients (Le., a pump that is developing a relatively high head for its size and
speed). The advantages of pumps with higher head coefficients are as follows:
1. They enable the use of a smaller, less expensive pump.
2. In many cases, they can be designed for higher efficienCies (due to a reduction in the
percentage of losses caused by disc friction), which can reduce operating costs.

Despite these advantages, many users are reluctant to use such pumps for fear that
they may become a source of unstable performance. In theory, this is a legitimate
argument, but in practice such instabilities are very rare. This can be best explained
with the aid of Figs. 2-3 and 2-4.
Figure 2-3 shows two drooping curves: A and B. It also shows curve C that
demonstrates how the resistance of the pumping system varies with the flow rate.
This resistance consists of two parts: the elevation He to which the liqUid must be
lifted, and the frictional losses in the pipelines H.r. The operating point of the pump is
at the intersection I of the QH curve with the system head curve C.
Curve C intersects the QH curve A at two points, I and I-I. Therefore the pump can
deliver two different flow rates, Ql and fJ2. Even the slightest disturbance may cause
the pump to jump from flow Ql to Q2 and back. This, of course, can lead to very
unstable performance. However, if the shape of the QH curve is similar to curve B in
Fig. 2-3, no such instability will occur because curve C intersects curve B at only one
point.
As a general rule, whenever the shut-off head Ho of the pump is higher than the
elevation He, the droop in the QH curve cannot cause any fluctuation in the
performance of the pump (see Fig. 2-4).
16 Introduction

t
"0
co
CI>
J:
--- ---

Figure 2-3 The effect of a droop


on the stability of the performance
Flow-- of a centrifugal pump.

Figure 2-4 The drooping curve


cannot cause unstable performance
if the shut-off head Ho is higher
Flow _ _
than He.

The other areas in which the shape of a rating curve affects pump performance are
the shape of the cavitation curve and the shape of the efficiency curve.
The effects of the cavitation characteristics will be dealt with in the chapters
devoted to cavitation problems. Here. we briefly discuss the importance of the shape
of the efficiency curve.
Performance Characteristics of Centrifugal Pumps 17

Shape of the Efficiency Curve


One of main advantages of higher efficiency is the reduction in operating costs. It
has been estimated that an average increase of 3% in efficiency can repay the total
cost of the pump within a year. But cost savings are not the only advantage of higher
efficiencies: a more efficient pump may need a smaller motor. thus requiring less
floor space. Higher efficiency also means less power is spent on friction and other
activities that may cause excessive wear of the parts.
The importance of higher efficiency is indisputable. however. it is also important to
realize that a pump with a higher peak efficiency is not always the best choice for a
given application. This is best explained with the aid of Fig. 2-5.
Imagine that a pump must be selected to deliver a flow rate QI against a head HI.
The choice is between two pumps that show the same efficiency at the speCified flow
and head. but the peak efficiency of pump A is higher than the peak efficiency of
pump B. This does not necessarily mean that pump A is more suitable for the given
duty than is pump B.
As noted earlier. in the field a pump rarely operates at one fixed point of the rating
curve. It usually operates over a significant range of flow rates and heads. Now. if the
pump must operate within a range of flow rates that vary from Qa to Qb. pump A is
unquestionably the better choice. However. if the pump has to operate mainly
between Qc and Qd. pump B is the better choice.
The examples discussed here are by no means exhaustive. but they do illustrate
how shape of the rating curves determine the quality of the field performance of a
centrifugal pump.

al
Q)
I
H11-----------------------=~

t
EFF

Figure 2-5 The relationship between


the shape of the efficiency curve and
the operating range of a centrifugal Flow _ _
pump.
CHAPTER 3
Classification of
Centrifugal Pumps
Centrifugal pumps are used in a variety of applications, including the following: water
supply and irrigation, power-generating utilities, flood control, sewage handling and
treatment, food industries, chemical and petrochemical industries, process industries
(e.g., textiles and leather), domestic appliances, mining and ore processing, trans-
porting liqUid-solid mixtures, environmental control, spaceships, airplanes, and motor
vehicles. The list can go on and on.
Because of their wide spectrum of application, centrifugal pumps are manu-
factured in a variety of shapes and with a variety of performance characteristics. In
general, they can be classified in three distinct ways: by structural features, function,
and specific speed.

STRUCTURAL FEATURES
• A pump may consist of a single impeller operating in a casing that has been cast
integrally with suction and discharge nozzles (Figs. 3-1-3-6).
• It may consist of several impellers mounted on a single shaft and operate in a two-part
casing that is split along a plane passing through its centerline (Fig. 3-7).
• It may consist several impellers mounted on a common shaft, operating in an equal
number of casings assembled into one unit (Figs. 3-8 and 3-9).
• The centerline of the shaft may run horizontally (Figs. 3-1-3-4 and 3-6-3-8) or it may
have a vertical orientation (Figs. 3-5 and 3-9-3-11).
• The casing of each impeller may consist of a volute or of an array of diffusers (Figs. 3-8
and 3-9).
• The suction nozzle may be concentric with the centerline of the pump (end-suction
pumps; Figs. 3-1-3-3). or it may be perpendicular to the axis (Fig. 3-4).
• The impeller may be cast integrally with both shrouds (closed impeller; Figure 3-12). or it
may be cast with only one shroud (semi-open impeller; see Fig. 3-13).
• The impeller may have a single suction nozzle (Figs. 3-1-3-5) or a double suction nozzle
(Figure 3-6).

The list can go on indefinitely. Most centrifugal pumps belong to at least one of the
groups listed above.
19
20 Introduction

Figure 3-1 Horizontal, single-stage end-suction, frame-mounted pump.

Figure 3-2 Horizontal, single-


stage, center-mounted pump.
(Worthington).
Classification of Centrifugal Pumps 21

Figure 3-3 Close-coupled pump.

Figure 3-4 Horizontal, single-stage, side-suction pump.


22 Introduction

Figure 3-5 Single-stage, vertical,


in-line pump.
Classification of Centrifugal Pumps 23

Figure 3-6 Horizontal, split-case pump with double-suction impeller.

Figure 3-7 Multistage, split-case


pump.
24 Introduction

4610 2450.2 1413 6571 1450 6512 6700 3849 6545 2110
3011 4120 i 3200 I 1140 4610 2220 6576 3261 6700
I

ik.1 -L___ L_ -~--~--+ -~ -~ -~ ---- <~ nl~: \


. I g

3266 6580
2461.2 4130
l'cl'r,~Ttr.I~r:\~ , ~, 25~
4132 6700 6571 6700 4421 6553 2461.1 3012
""
2531
2483
I

Figure 3-8 Multistage, donut pump.

FUNCTION

In many cases, the structural features of a pump are determined by the pump's
application. For example, when water must be drawn from a deep underground source,
deep-well pumps are commonly used (Fig. 3-9). These pumps can be installed in
relatively small-diameter wells but are capable of supplying relatively high flow rates.
When a pump must handle hot liquids, its casing is usually mounted at its center
(Fig. 3-2) to compensate for the differences in the thermal expansion of the pump and
the driver.
Vertical pumps are often used to handle effluents (Fig. 3-10). These are installed
over an underground sump. The pumping end is submerged into the liquid, and the
driver is located above the ground.
To handle liquids containing a large amount of solid matter, pumps with extra-wide
impellers are often used (Fig. 3-12 and 3-13) to minimize the danger of clogging.
Sometimes the application for which a pump must be chosen mandates certain
special requirements related to the type of the sealing arrangement. In other cases,
the waterways must be made of special materials to prevent the corrosive or erosive
action of the pumped liqUid.
Here again, the list can go on and on.
Classification of Centrifugal Pumps 25

Figure 3-9 Multistage,


vertical, deep-well pump. Figure 3-10 Vertical,
(Worthington). wet-pit pump.

Figure 3-12 Closed impeller for a sewage-disposal pump.

Figure 3-11
Vertical, dry-
pit, sewage
pump.

Figure 3-13 Semi-open impeller


for a sewage-disposal pump.
26 Introduction

Specific Speed
This category is best explained by an example. Assume that there is a need for a
pump to deliver 1000 GPM against a head of 200 ft. Because of cavitation the inlet
velocity cannot exceed a certain limit. regardless of the operating speed of the pump.
For a flow rate of 1000 GPM, an impeller with an eye diameter of less than 4 in. is
rarely found.
The outer diameter of the impeller is chosen on the basis of the desired head. To
achieve a head of 200 ft, an impeller operating at 1770 RPM needs an outer diameter
of about 16 in. At 3560 RPM, the impeller can achieve the same head with an outer
diameter of only 8 in. In a pump designed to develop a head of 200 ft at 1770 RPM,
the radial distance between the inner and the outer tips of the blades is equal to
(16-4)/2=6 in. In the high-speed pump, the radial distance between the inlet and
the outlet tips of the blades is equal to (8-4)/2=2 in. In the slower pump, there will
be three times more radial room for the blades than in the high-speed pump.
As demonstrated earlier, the power transmitted from a blade to the liquid depends
not only on the blade's shape and peripheral speed, but also on its area. To achieve a
comparable blade area with the high-speed impeller, it is necessary to make it wider
than the low-speed impeller. Recall that a wider impeller will require a wider casing
and will cause many other changes in the design. These changes, in turn, will cause
other significant changes in the performance characteristics of the pump. In
conclusion, we see that the design and performance of a pump is greatly dependent
on the relation between its operating speed, the flow rate for which it has been
designed, and the developed head.
In practice, this relationship is defined as the specific speed. In the United States,
its magnitude is expressed by the equation

(3-1)

where N = operating speed, in RPM


Q = flow rate, in U.S. GPM
H = head, in feet

When expressing the specific speed in metric units, we usually express the flow
rate Q in m 3 / sec and the head H in meters. In that case, the relationship between the
specific speeds expressed in U.S. units and in metric units is
N s (U.S.)=51.64Ns (metric). (3-2)

Experience teaches us that there is a certain relationship between the specific speed
of a pump, its geometry, and its performance. Figure 3-14 shows the general trend of
how the geometry, efficiency, and shape of the head-capacity curve varies with the
specific speed.
It should be pointed out, however, that the data shown in Fig. 3-14 were compiled
about half a century ago. More recent studies and experimental findings have
demonstrated that significant departures from the recommendations presented in
Fig. 3-14 are not only permissible, but very often they are highly desirable. This
subject, however, relates to the design of the centrifugal pumps and is beyond the
scope of this book.
Classification of Centrifugal Pumps 27

o o
o o o o 0 o o
o o o o o o o
o
LO
o o
C\J
oC') o
'<t
0
0
LO
o to

90
---
~-
..
~oooft-
--- ----- - ---
-r-- --1--
:-- --
3000 to'1o~r-o~ r- ......
- - I - - ,000 G lIer 70~
... .. ......
90

C 80 PA.4 - ,000...... 80
Q) Q\O~- Gp/l1--.
~ \OCJ G?~
Q)
0.
C \0
\000 1 ...,.....--
--
-;, 70 - <:,00 ~~ 70

~~~
u
c
Q) - - I--
.(3
Q?
7,,00
irl 60 7,,00 0«~ 60

L"r::s *" __
fJ'o
~o
00 I

/ ~
50 50
/ I I
I I
40 i I I 40
o o o o o 0 o o
o o o o o 0 o o
LO o o oC') o 0 o o
C\J '<t LO o to
RPMJGPM
Values of specific speed Ns = --'--=--
H3/4
..... c
2:8

__ D_J7_%~;wJA I
C ell
Q) .....

I Centrifugal ~ Mixflo ~ Propeller

Figure 3-14 Effect of specific speed, the impeller's-geometry, and flow-rate, on efficiency. RPM,
Revolutions per minute. GPM, gallons per minute. (Worthington.)

In general, the data presented in Fig. 3-14 are not always the best configuration;
however, they are a good illustration of the general influence of the specific speed on
the impellers geometry.
The data presented in Fig. 3-14 may be helpful to a prospective buyer of a
centrifugal pump. Imagine that a prospective buyer is considering a pump of a
specific speed of 2000 for a flow rate of 4000 GPM and a peak efficiency of 81 %. A
glance at Fig. 3-14 shows that the efficiency falls well below that expected from a
pump with these specific speed and flow rate values. Consequently, it is wise to
continue looking.
28 Introduction

Now assume that the prospective buyer is looking at a pump with the same specific
speed of 2000, but a flow rate of 200 GPM. If the manufacturer claims that the
efficiency of this pump is 81%, the buyer should be suspicious. As can be seen in Fig.
3-14, it is very doubtful as to whether this efficiency is correct. For safety, the buyer
should demand that the claimed efficiency be verified by a witnessed test.
PART/II

Performance
Factors

Various factors affect the performance of centrifugal pumps. The chapters


contained in this part explain the impact that these factors have on pump
operation and, wherever possible, identitY their causes.
CHAPTER 4
Effects of Free Air in the
Pumped Liquid
In general, the presence of free air (or any other gas) in the pumped liquid always
adversely affects centrifugal pump performance. One negative aspect of the presence
of free air has been discussed earlier in the section on priming (Chapter 1). However,
even when a pump has been primed properly, free air can appear during operation
and disrupt the performance of the pump.
When a mixture of liqUid and air enters the impeller, the centrifugal forces
generated by the rotating blades throw the heavier liqUid outward, thus trapping the
air in the center of the pump. Under certain conditions, so much air can accumulate
in the eye of the impeller that it separates the liqUid in the suction line completely
from the liqUid at the outer radii of the impeller. This, of course, disrupts pumping.
Whether such a disruption in the pumping action will occur depends on the source
of the air entering the pump, the flow rate of the entering air, and the temperature
and pressure at the impeller eye.

SOURCES OF FREE AIR

The most common sources of free air are as following:


1. Air enters at a low-pressure zone of the pumping system through a hole. crack. or some
other break that connects the low-pressure zone with the air of the atmosphere.
2. Air (or gas) may be dissolved in the pumped liqUid. The solubility of gases in liquids
decreases with an increase in the temperature of the liqUid and also with any reduction
in pressure. Consequently, air is apt to separate from the solution when entering the eye
of the impeller where the pressure is usually the lowest.
3. Air bubbles may occur in water that contains decaying or fermenting organic matter.
They may also be generated by chemical processes that continue to take place while the
liqUid is being handled by the pump.

Before forming an air pocket. the air usually appears in the liqUid in the form of small
bubbles. As long as the bubbles are relatively small and remain in the impeller for a
short time, the pump treats the mixture as a liquid of lower specific gravity. The pressure
gages record lower pressures. even if the pump is developing the same head. as in the
case of the pure liquid (see the discussion of priming in Chapter 1).
31
32 Performance Factors

IMPELLER SHAPE AND AIR-HANDLING CAPABILITIES

Figure 4-1 shows the performance of a 6-in. pump having a specific speed of 2270
(U.S. units) handling a mixture of air and water. The percentages that appear near
each curve signifY the volumetric percentage of the contents of free air within the
liquid. When the volumetric percentage rose above 10%, the pump lost its prime and
stopped delivering liquid.
Similar tests were performed with a pump having a specific speed of about 600. In
that case, the pump lost its prime at a volumetric percentage of air of 6%.
These tests highlight the importance of the impellers geometry on the air-handling
capabilities of a centrifugal pump. Referring to Fig. 3-14, we see that a pump having
a specific speed of 600 has a relatively small eye diameter and is very narrow. A small
volume of air can completely stop the flow of liquid in such a pump.
In addition, the distance between the radius of the inlet tips and the outlet tips is
relatively large. This gives the centrifugal forces generated by the blades enough time
to separate the air from the liquid.
By contrast, a pump with a specific speed of 2270 has a much larger eye diameter,
and its impeller is significantly wider. In such a pump, a significantly larger volume of
air is needed to stop the flow of the liqUid. The distance between the radius of the inlet
tips and the impeller outlet is also significantly smaller. Therefore, the impeller blades
engage the liqUid for a shorter time period. Consequently, the pump represented in
Fig. 4-1 was able to handle up to 10% of entrained air before losing its prime.
In general, the air-handling capability of an impeller increases with its specific
speed. For example, an axial-flow impeller is capable of handling a liquid-air mixture
as long as some liquid is present within the impeller blades.
The data presented in Fig. 4-1 and other publications can serve only as an
illustration of the effects of the presence of free air within the pumped liqUid.
In many cases, especially when air enters the pump in solution, the results are
highly dependent on the temperature and pressure of the liqUid. Despite these
difficulties, it is sometimes advantageous to admit a small amount of free air into the
pumped liqUid. For example, when a pump is subjected to heavy abuse owing to
cavitation, a small amount of air produces a cushioning effect, thus reducing the
severity of the blows of the collapsing vapor bubbles.

70
1450 RPM
60

~ 50
Q)
g40
:r:
-c:i
al 30
I

20 Figure 4-1 The effect of free air


in the liquid on the performance
10
of a 6-in. pump having a specific
o '----_ _----L-_ _---'-_ _ _--'---_ _- - ' -_ _ _L -_ _---.J speed of 2270. RPM, Revolutions
o 200 400 600 800 1000 1200 per minute; GPM, gallons per
GPM minute.
CHAPTER 5
Cavitation
It is difficult to draw the distinction between cavitation and boiling. The closest
definition might be "local boiling within a liquid that is otherwise in a nonboiling
state." This can be best illustrated by the example of the teapot, when heating water.
simmering noises. and cracking are heard long before the water boils. These noises
occur when the water in immediate contact with the bottom of the teapot starts to
boil while the rest of the water is still well below the boiling point. The vapor bubbles
generated at the bottom of the teapot. being lighter than the liquid. move upward
through the nonboiling water. The moment they reach a cooler zone of liquid they
collapse vigorously and create cracking noises.
When the temperature of the heated water reaches 212F (lOOC), vapor bubbles
start to form throughout the liquid. Any condensation of the generated vapor begins
to occur only outside the bulk of the liquid. after the vapor has reached the cooler air.
At this moment the phenomenon is simply defined as boiling.
The only reason the pumped liquid enters the impeller is due to the pushing force
created by the reduction of pressure at the pump inlet. Whenever the reduced
pressure falls below the vapor pressure of the pumped liquid. vapor-filled cavities
begin to appear within the liquid.
Generally. the reduction of pressure within the impeller eye is not uniform. but
varies with the location of a given space within the impeller inlet. At certain places in
the entrances to the blades, vapor bubbles start to appear earlier. Consequently. at
the earlier stages of cavitation. enough unobstructed space remains to allow the liquid
to continue to flow (although at a reduced rate). This flowing liquid carries away some
of the vapor bubbles into zones of higher pressure where they collapse vigorously. This
collapse manifests itself in noise. vibration. and a reduction of output. It may also
cause very serious damage to the surfaces against which the bubbles collapse.
Cavitation is usually caused by a local increase in temperature or a local drop in
pressure. Every liquid has a fixed relation between the temperature at which it starts
to boil and its pressure. Water. for example, boils at 212F (IOOC) under a pressure of
one atmosphere. However. under a pressure of 10 atm absolute, it boils at a
temperature of 355F (l79C). Similarly, under a pressure of 0.024 atm. it boils at 70F
(2IC).
In a stationary liquid. the pressure in any horizontal plane is constant in all
directions. However. when liquid is flowing through a centrifugal pump. there may be
great differences in the local velocities of the liquid. This. in turn. causes significant
pressure differences throughout the liquid.
33
34 Performance Factors

Figure 5-1 The location where cavitation is


most likely to begin to develop.

Pressure differences also occur as the impeller blades perform work on the liquid.
To transmit power, a difference of pressure must exist between the leading and the
trailing faces of the blades. Differences in pressure mean that the pressure is lower in
some zones of the pump than in others. Under certain conditions, the pressure in
some regions of the pump may drop below the vapor pressure of the liquid, giving rise
to cavitation.
Except for very high flow rates, cavitation usually occurs on the trailing face of the
blades in the vicinity of the inlet tips (Fig. 5-1). At very high flow rates, cavitation also
often occurs at the tongue of the casing (or at the inlet tips of the diffusor blades) or
at the leading face of the impeller vanes.
The effect of cavitation in centrifugal pumps manifests itself in two ways. First, the
vapor bubbles created within the impeller passages obstruct the flow of the pumped
liquid, reducing output. Often, this causes complete breakdown of both the head and
efficiency. Second, when the bubbles collapse upon reaching a zone of higher
pressure, they may exert enormous local stresses on the surfaces against which they
collapse, causing damage. Sometimes cavitation can ruin a pump within a few hours.
Externally, cavitation manifests itself by noise and vibration as well as by reduced
pump output.

NET POSITIVE SUCTION HEAD (NPSH)


The net positive suction head (NPSH) is the difference between the total absolute
head available at the pump inlet and the head that corresponds to the vapor pressure
of the pumped liqUid. It is calculated from the absolute pressure measured at the
pump inlet and from the velocity head of the pumped liqUid, in accordance with the
equation

NPSH = Ps - Pv + C2 (5-1)
Y 29
where Ps = absolute suction pressure measured at the pump inlet
Pv = vapor pressure of liqUid at the pumped temperature
C = velocity of the liquid at the section where measurements are taken
9 = acceleration owing to gravity
"y = specific gravity of the liqUid
Cavitation 35

When dealing with centrifugal pumps, there are two values of NPSH: available
NPSH and minimum required NPSH. The available NPSH is defined by Eq. 5-1 and
represents the actual suction pressure at the pump inlet.
The minimum required NPSH is the value of the available suction pressure that a
given pump needs to operate satisfactorily at a given speed. Whereas the magnitude
of the available NPSH is defined by Eq. 5-1, the minimum required NPSH cannot be
determined unequivocally.
The minimum required NPSH is the value of the available suction head a pump
requires to operate satisfactorily. However, determining whether a pump operates
satisfactorily depends largely on the duties for which the pump was chosen.
Rarely does a pump operate at a constant flow rate or head, or under constant
suction conditions. Usually, a pump must operate over a wide range of heads,
capacities, and available NPSH values. Depending on the conditions under which the
pump must operate, a pump that is the best choice for one application may be
completely impractical for another.
Even the definition of the minimum suction head required by a pump often cannot
be determined unequivocally. Figures 5-2 and 5-3 illustrate such a case.
In Fig. 5-2, there is little doubt as to what is the minimum allowable NPSH. In this
case, both the head (H) and efficiency (EFF) drop suddenly when the available NPSH
is lowered to 7 ft. Therefore, 7 ft is the obvious minimum required NPSH value for
this pump when it is operating at this flow rate.
In Fig. 5-3, the conditions are completely different. Here, the head begins to drop at
an NPSH value of 30 ft, whereas the efficiency begins to drop at 18 ft. However, as
long as the available NPSH is above 9 ft, the head and efficiency continue to drop
gradually. Only when the NPSH is reduced to 9 ft does a complete breakdown of head
and efficiency occur.
There is no absolute criterion for determining what the minimum allowable NPSH
values for such a case should be. Some pump experts define the minimum allowable
NPSH as the value at which the head drops by 3%, as compared with noncavitating

H
70 I-

EFF
- 50
~ 60 r-
~ - 40 'E
:r: HP Q)

-c ~
til 15- - 30 ~
-l!. 50 I- HP ()
10- /" - 20 ~
"<:5
5- - 10 ~
I I I
Figure 5-2 An NPSH characteristic with a clearly o 10 20 30
defined onset of cavitation. HP=Horse Power. NPSH (feet) ~
- - - - , l..
36 Performance Factors

100

60
80
~

Q)
Q) 50 :g-
~ 60 Q)

J:: ~
40 Q)

----
8
"C -9:
ctl
~ 40 6
HP 30 ~
c
Q)

4 20:2
::::
20 w
2 10
0 9 Figure 5-3 An NPSH characteristic without a
0 10 20 30 clearly defined onset of cavitation. HP=Horse
NPSH (feet) - - - - Power.

operating conditions. Others regard a drop of 5%, 8%, or even 10% as the significant
value. Still others regard a corresponding drop in efficiency or horsepower as the
significant magnitude. Alternatively, many pump experts take the significant NPSH
value as the suction pressure at which a complete breakdown of performance occurs.
The trouble is that all these approaches are correct. The circumstances depend on
the duties of the pump.

SUCTION CHARACTERISTICS

Methods of Defining the Intensity of


Cavitation
The NPSH requirements of a given centrifugal pump are usually determined by
testing it at a nearly constant temperature and speed. This is often accomplished by
varying the available NPSH at the tested flow rates and observing how the
performance at a given flow rate changes the available NPSH. For convenience, these
tests are usually done when pumping water (compare Figs. 5-2 and 5-3).
The data obtained from these tests are the only basis available to determine whether
the suction capability of a pump matches the actual operating conditions. In practice,
however, these data are not always sufficient. A pump must often operate under
conditions that are completely different from the conditions under which it was tested.
Certain applications may require that variations in head and NPSH follow a
prescribed pattern. Some may require that the pump be able to operate in the
presence of cavitation. Others may exclude the presence of cavitation even when
hydraulic performance is not affected by it. This list is endless.
Another spectrum of problems may arise when the pump operates under
conditions that vary with time. These variations may be linked directly to pressure
fluctuations in the suction tank and to fluctuations in the liquid level. There may also
be fluctuations in the operating speed, pumped-liquid temperature, and flow rate.
Cavitation 37

Cavitation in centrifugal pumps manifests itself in at least one of the following


phenomena: excessive noise and vibration, reduced output, or destroyed wetted
parts. However, each of these may also be caused by a host of other conditions that
have nothing to do with cavitation.
The only tool for determining whether the problem is due to cavitation is the NPSH
data that were determined during tests carried out under completely different
operating conditions. These differences in operating conditions may include
differences in liquid temperature, operating speed, and the physical and chemical
properties of the pumped liquid. There may also be differences in the amount of gas
or air dissolved in the liquid, as well as differences in the nature and amount of
different impurities. Also, the pumped liquid may have completely different thermal
properties and a different viscosity than the liquid with which the NPSH tests have
been performed.
Under such conditions, only in-depth knowledge and an understanding of the
process of cavitation in centrifugal pumps may enable one to arrive at the correct
conclusions.
Many attempts have been made to establish general definitions that quantitatively
describe the suction characteristics. This need arose from necessity to predict or
estimate performance in four instances: when operating a pump at other than the
tested speed, when only the test data of a reduced model are available, when
pumping liquids at other than the tested temperature, and when comparing the
suction capability of different pumps.
Thoma's Cavitation Number. Humans began to use water turbines much
earlier than centrifugal pumps. In the early stages of their development, centrifugal
pumps were looked upon as turbines operating in reverse. Consequently, it appeared
quite natural to apply concepts used to describe water turbines to centrifugal pumps.
In terms of the suction capability of centrifugal pumps, this turned out to be very
liquid depends oma's cavitation number was originally developed in connection with
water turbines. Mathematically, it is expressed as

NPSH (5-2)
0=--
H

where H is the total head developed by the pump.


This parameter was originally developed in connection with water turbines. Here,
the inlet pressure is the total head H available, and the minimum required NPSH
is the least amount of the total head H that must remain unused to suppress
cavitation. Its use relative to centrifugal pumps is rather artificial and may often lead
to incorrect conclusions.
For example, if the outside diameter of an impeller is cut down, its discharge head
is reduced without reducing its suction capability (except at very high flow rates, at
which cavitation may develop at the volute tongue or at the inlet to the guide vanes).
A different value of the NPSHj H ratio does not necessarily mean a difference in the
suction capability of a pump (Fig. 5-4).
When this deficiency of the application of Thoma's cavitation number to centrifugal
pumps became evident, an attempt was made to relate the value of (]" to the specific
speed of a pump. Figure 5-5 presents data based on that approach. These data were
38 Performance Factors

4.25 in. Dia

70

60

3.75 in. Dia


3l 50
LL
- - -/::r - - __ ~
......
'b. .....
.....
40 0----0 4.25 in. Dia ~ ..... ,
b:- - -I::.. 3.75 in. Dia 'I::.. ,
30 ~
EFF 't::..

70
60

50 25

~ 40 20
~ ~
~ 30 15 ~
::c Figure 5-4 Test results with full
20 10 U)
a.. and cut-down impeller diameters
z demonstrate that Thoma's cavit-
10 5
ation number has no effect on the
O-----~----~----L-----L-----L-----L-~O
50 100 150 200 250 suction capability of a centrifugal
GPM pump. GPM, Gallons per minute.

taken from the standards published by the Hydraulic Institute in 1947. In practice,
these data have sometimes produced ridiculous situations, as demonstrated by the
following case history.
A customer ordered a pump that had to operate against a total head of 135 ft and
be capable of handling water located 15 ft below the pump's centerline. The customer
also requested that the suction capability of the pump be verified by actually lowering
the water level in the suction sump. During the test, the pump not only satisfied the
specified suction condition, but was even able to continue to operate satisfactorily
until the water level in the suction sump dropped 20 ft below the centerline of the
pump.
To the manufacturer's surprise, the customer refused to accept the pump. The
specific speed of that pump happened to be equal to 1910 (in U.S. units). The
customer argued that, according to the Hydraulic Institute (Fig. 5-5). a pump with
such a specific speed should be capable of lifting water from a depth of only 15 ft. If
it is capable of lifting water from a depth of 20 ft while delivering against a head of
135 ft, its specific speed should have been only 1580.
Cavitation 39

Pumps with double-suction impellers


1500 2500 3500 4500 Ns
(U.S.
units)
500
400

300

200

150
Qi
~
:r::
'0
100
'"
<Il
I 80
70
60
50
40

30

20
11001200 1400 1600 1800 2000 2500 3000 3500 4000 Ns
Pumps with single-suction impellers (U.S. units)

Figure 5-5 A suction lift chart based on the assumption that Thoma's cavitation
number defines the suction capability of a centrifugal pump (Standards of the
Hydraulic Institute, 1947.)

The manufacturer had a very hard time convincing the customer that the pump
had much better suction capabilities than those specified by the Hydraulic Institute.
Only in those infrequent cases in which two different pumps develop exactly the
same heads at exactly the same flow rates and speeds, does Thoma's cavitation
number present a true comparison between the suction capability of two different
pumps. In this case, however, such a comparison can be made directly by comparing
the NPSH requirements.
A second instance in which Thoma's cavitation number has limited application is
that it proves that the (theoretical) suction requirements of a pump increase as the
square of its operating speed.
Suction Specific Speed. The mathematical expression for the suction specific
speed Sis:

s= N x QO.5
(5-3)
NPSHo. 75

where N = operating speed of the pump


Q = flow rate, in GPM
NPSH = minimum required positive suction head, in feet
40 Performance Factors

Generally, the higher the S value for a given pump, the better is its suction perfor-
mance. However, when using the S values for comparing the suction capabilities of
different pumps, certain pitfalls are observed.
One pitfall from Eq. 5-3 is that the magnitude of S varies with the flow rate Q. In
fact, the S value of each pump is zero at shutoff. It then rises quickly to a certain
maximum at a certain partial flow rate. Afterwards it starts to decrease gradually,
again asymptotically approaching the value of zero.
Now imagine that a pump user must choose between two pumps. The maximum S
value of one is 22,000 (in U.S. units). and the value of the other is 18,000. This does
not necessarily mean that the first pump is better than the second when operating
under reduced suction heads. On the one hand, the first pump may be able to attain
its maximum S value only at a flow rate that the pump is never expected to achieve
under the given operating conditions. On the other hand, while operating within its
assigned duties, the second pump may possess better suction capabilities than the
first.
Another pitfall may result because the NPSH term in Eq. 5-3 refers to the minimum
reqUired NPSH. There are cases in which there is no absolute way to determine what
is the minimum required suction head. This, again, may lead to different S values
without proving that the pump with the higher S value has better characteristics
than pumps with nominally lower S values.
To avoid such pitfalls, carefully check how the value of the minimum reqUired
NPSH was determined, as well as the flow rate to which the specified S value refers.
Dimensionless Cavitation Number K. There is a growing tendency to relate the
NPSH requirements of a pump to a certain velocity at the impeller inlet. For this
purpose, many experts use the cavitation number. The exact definition of this
concept is, as yet, not finalized. In the literature it appears in several forms. Some
authors use it to express the ratio between the NPSH and the velocity head of the
relative velocity of the liqUid at the inlet tips. In this case, K is expressed as

K = NPSH
(5-4)
W t 2 /2g

where W t is the relative velocity of the liqUid with respect to the inlet tips of the blades.
Others use Ut in Eq. 5-3 instead of W t, where Ut is the peripheral velocity of
the inlet tips of the blade.

Generally, the cavitation number K is less popular than other concepts. Its
application is limited mostly to papers and articles of more theoretical interest than
practical value.
The terms suction specific speed, Thoma's cavitation number, and so on were all
established to compare the suction performance of centrifugal pumps and to predict
the suction performance of a pump under other than tested conditions.
It seems, however, that these measures are equally suitable for each of the cases
for which they are used. Thoma's cavitation number is best suited for turbines, but
suction specific speed seems to be more appropriate for centrifugal pumps. However,
even this concept may lead to incorrect conclusions if the influences of additional
factors are ignored, such as size, speed, temperature, specific heat, latent heat of
evaporation, and thermal conductivity of the liqUid.
Cavitation 41

Ways to Represent Suction Characteristics


Graphically
To evaluate the different suction characteristics of a pump, it is convenient to
express them in graph form. This can be accomplished in several different ways:

1. Head, horsepower, and efficiency at a constant flow rate and varying NPSH. Figures 5-2
and 5-3 present the behavior of a pump at a constant flow rate and speed, while the
available NPSH varies. A complete set of such graphs provides the pump designer with
valuable data about the suction characteristics of centrifugal pumps. However, the
complexity of these graphs makes them inconvenient for most pump users. It is more
convenient to have all the data combined into a single chart that provides an overview of
the suction performance of the pump at different flow rates.
2. NPSH and suction specific speed at differentJlow rates. The NPSH data presented in Fig.
5-6 are the minimum NPSH values below which it is recommended not to allow a pump
to operate. These values were derived on the basis of tests performed under laboratory
conditions. During field operation, conditions may be less favorable. Because of this, a
certain factor of safety should be used and the pump should be provided with somewhat
higher available NPSH than is specified in the chart. From Fig. 5-6 it is possible to
calculate the S values of the given pump at different flow rates. This is done using Eq. 5-3.
These data are sometimes also entered on the graph.
3. Head, capacity, horsepower, and NPSH requirements plotted against different flow rates.
This is the most popular way of presenting the overall performance characteristics of a
centrifugal pump. It is used almost exclusively in all price books and catalogs published
by various pump manufacturers (Fig. 5-7).
4. Suction specific speed versus the ratio oftheJlow rate. To analyze the suction performance
of a pump at different speeds, the S values are plotted against some dimensionless ratio
of flow rates. As a basis for such a ratio, choose a flow rate Qopt at which the pump
exhibits its highest effiCiency. The values of S are then drawn against the ratio of Q/Qopt

s S
16,000
15,000
14,000
13,000 I
12,000 - I
11,000 - I
I
10,000 - I
I
I
I
I
I
I
I
I 30
I
I 201)5 ~
I 10~,gs
Figure 5·6 Suction specific speed
Sand NPSH plotted against the flow 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 On
rate On. Flow rate
42 Performance Factors

(Fig. 5-8). When using such a graph, the value Qopt is assumed to vary directly with the
operating speed of the pump. Thus, the theoretical value of S remains constant for any
given ratio of Q/Qopt for a significant range of operating speeds. This allows calculating
the NPSH requirements of a pump when operating at other than the tested speed.
Although this is presently the only way to predict the NPSH requirements of a pump at
different speeds, it is not foolproof. Experience has shown that it often leads to incorrect
data, particularly when the NPSH requirements at reduced speeds are calculated. Thus,
it is always advisable to provide a pump with higher available NPSH than obtained on the
basis of the assumption that S is constant for the given ratio Q/Qopt.
5. A set of QH curves at different NPSH values. A less popular, but useful, way to represent
the suction performance of a pump is shown in Fig. 5-9. Here, the QH curve changes at
different values of the available NPSH. This graph can be compiled either by directly
measuring the QH performance of the pump at discrete, constant NPSH values, or from
test data obtained when measuring the variations of head at constant flow rates and
varying NPSH values. In this last case, the H-NPSH curve is similar to Figs. 5-2 and 5-3.

EFF

Figure 5-7 Head, efficiency, power consumption,


and minimum required NPSH plotted against flow
Capacity (GPM) rate.

S
12,000

10,000

8000

6000

4000

2000

Figure 5-8 Suction specific speeds S


o 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Q/Qopt plotted against flow-rate ratios Q/Qopt.
Cavitation 43

To convert these curves into a curve similar to that in Fig. 5-9, draw a composite chart
of all available H-NPSH curves on one common chart (Fig. 5-10). Now draw a series of
vertical lines at different values of NPSH. (In Fig. 5-10, these values are for 23, 18, 12, 9,
and 5 ft.) If all points of intersection of the vertical line at 23 ft in Fig 5-10 are transferred
to Fig. 5-9 and joined by a smooth curve, the QH performance of the pump at 23 ft is
obtained. In this case the curve happens to coincide with the noncavitating operation of
the pump. If the procedure is repeated with the vertical line drawn at 18 [t NPSH, the
head starts to fall rapidly at 347 GPM. All other lines are drawn Similarly.
A performance curve such as that in Fig. 5-9 is very useful when analyzing a pump
operating within a fixed system that is subject to variations in suction conditions. When
a pump is operating within a fixed system under noncavitating conditions, it is easy to
determine the resulting flow rate with a graph such as the one in Fig. 5-11.
Curve A represents the QH relations of the pump. Line B, which is parallel to the Q
axis, represents the static head Hst against which the pump must deliver. Curve C shows
the frictional losses Hr in the system at the various flow rates. These losses are zero at
shut-off and increase as the square of the flow rate increases.
Curve D is the sum of the static head Hst and the frictional losses at different flow
rates. This sum represents the total resistance He of the system at various flow rates.
Point M where curves A and D intersect, shows the flow rate that the pump will deliver
under the present conditions. This point tells us that, when working within the given
system, the pump will deliver the flow rate Qa against a total head of Ha. Imagine now
that the suction pressure is reduced so that cavitation develops within the pump when it
is operating at flow rate Qa. From Figs. 5-2 and 5-3, it can be concluded that, when

500
'$
~400
,
I
:t I
. 300
1iOJ NPSH (feet) = 5 9 18
I 200
Figure 5-9 A graph showing
the deterioration of the QH 100
curve at different available Q
NPSH values. 200 200 200 200 200 200 GPM

5ft 9ft 12ft 18 ft 23 ft

'I 100 GPM 150 GP M

(/
550

7-'-
200 GPM
I I I I I 250 GPM
CD
~500 276 GPM
:t I 300 GPM
·450 I 347 GPM
1i I
~ 400 1 1 1
I I I I 1
350
1 1 1 1 1
Figure 5-10 A composite chart of
H-NPSH curves for various flow 5 10 15 20 25 30
rates. NPSH (feet)
44 Performance Factors

Figure 5-11 The effect of system


resistance on the flow rate delivered
1..------0.-------..·1 by a pump.

Head, H 1 _____ ~Q~H~c::!u~rv::e _ _ _~

.....
'\
I
Hf I
I I I
r:---=====;:::====:i~ili _f--_N_P_S_H2 __' _ NPSH 1
Figure 5-12 The flow in a given
pumping system at different
Flow rate Q available NPSH values.

delivering the same flow rate Qa, the head developed by the pump will be reduced or will
break down completely. The pump will now be unable to overcome the total resistance of
the system at Qu, and the flow will move to a lower capacity.
Now the problem is how to determine in advance what the reduced capacity will be. To
do this, rearrange the available performance and NPSH data as in Figs. 5-9 and 5-12,
and draw a curve representing the total resistance of the system (Figs. 5-11 and 5-12).
Points A, B, C, etc. (Fig. 5-12), where the QH curves for different NPSH values intersect
the Hr curve L, are the different operating pOints of a given pump when pumping liquid
within this system under the different suction conditions.
CHAPTER 6
Cavitation in
Centrifugal Pumps
The suction characteristics of a pump are determined by a series of tests performed
at nearly constant speed. Often. however, a pump is not used at its test speed. Other
pumps may be too big to test on any available rig, and all tests must be performed on
a scale model. In both cases, one must resort to using model laws to estimate
whether the pump will perform satisfactorily under actual working conditions.
An older approach to this problem was to estimate the suction performance of the
pump, assuming that its suction specific speed at a constant ratio of Q/Qopt (see Fig.
5-8) was constant. However, recent findings have revealed that this is not always the
case. The suction specific speed clearly varies with the size of the pump and its
operating speed. This is repeated here because of its importance in choosing the right
pump for low-NPSH applications. This problem was first presented at an ASME
(American Society of Mechanical Engineers) Symposium [11.

CAVITATION AT DESIGN FLOW RATE

When a pump operates at its design flow rate, cavitation is most likely to develop on
the trailing face of the impeller blades. This usually occurs at an NPSH value that is
Significantly higher than the vapor pressure P v of the pumped liqUid, divided by its
specific weight y.
Cavitation may begin at an NPSH value higher than the value of Pv/y for several
reasons. Part of the available NPSH may be consumed by surface friction between the
flowing liqUid and the surfaces with which it comes into contact or by energy losses
from eddies in dead spaces. There may be local zones of high relative velocities. Part
of the available NPSH will be converted into velocity head owing to the flow of the
liqUid into the impeller. A pressure drop at the trailing face of the blades may occur
when the blades of a pump transfer energy from the pump shaft to the liqUid.
In a well-designed conventional pump, the frictional and separation losses at the
impeller inlet are relatively small when the pump operates at its design flow rate. At
this flow rate, they have little effect on the NPSH requirements of a pump.
The drop in pressure caused by high local velocities requires a bit more consider-
ation. Assume that a flat plate P moves through a liquid with a uniform velocity Win
45
46 Performance Factors

B C

.. w

Figure 6-1 Section through the


B C inlet part of a blade.

a direction parallel to its faces (Fig. 6-1). Let the absolute pressure in the
undisturbed liquid be equal to NPSH. Relative to an observer situated on the moving
plate, the total head H of the liquid is equal to:

W2
H=NPSH+- (6-1)
29

Along the faces C the liqUid moves in relation to the observer at a uniform velocity.
However, at the front tip of the plate a stagnation point where the liquid is at rest
occurs. At this point, the pressure is equal to the total head expressed by Eq. 6-1,
multiplied by the specific weight of the liqUid.
Between points A and B, the relative velocity of the liqUid increases continuously,
assuming local velocities much higher than W. The magnitude of these local velocities
depends largely on the shape of the curves that connect points A and B. If the curve
is a quarter of a circle, the highest velocity would be at point B and would be equal to
twice the velocity W. For a quarter of an ellipse, the magnitude of the local velocity Wb
is approximately

(6-2)

Because Wb is larger than W, the absolute pressure at B falls considerably below


the available NPSH, because part of the pressure at this point is converted into a
velocity head equal to Wb2/2g, that is larger than W2 /2g. In the case of a blunt edge,
the theoretical value of Wb may approach infinity.
Here, the shape of the inlet edge may have a significant effect on the NPSH
requirements of a pump. From these considerations it would follow that the inlet tips
of the blades should have a sharp edge to reduce the NPSH requirements.
Most conventional pumps have semicircular inlet edges, primarily to facilitate
manufacturing and to enable the impeller to operate satisfactorily over a wider range
of flow rates. Consequently, in conventional pumps, the effects of the shape of the
inlet edge are almost identical in all cases.
The NPSH at the inlet to the impeller consists of a combination of an absolute
pressure head and a velocity head. The latter results because the liqUid entering the
pump needs a certain velocity to supply the pump with the required flow rate.
Consequently, the velocity of flow reduces the amount of pressure available to
suppress the vapor pressure of the liquid. The magnitude of the velocity with which
the liqUid enters the impeller may have a serious effect on its NPSH requirements.
Cavitation in Centrifugal Pumps 47

However. because of the many design considerations. the magnitude of this velocity
at the design flow rate is chosen within rather narrow limits. As a result. the
influence of this velocity on the NPSH requirements does not vary significantly from
one pump to another. The main factor that creates significant differences between the
suction requirements of pumps is the pressure differential between the leading face
and the trailing face of the blades.
At the design flow rate Qn. the pressure differential Pd is proportional to the
product of the angular speed <u of the impeller and the square root of the design flow
rate [I]:

(6-3)

Centrifugal pumps are designed for flow rates as low as 10 GPM (2.3 m 3 /hr) and as
high as 100.000 GPM (23.000 m 3 /hr) and even higher. The speeds for which they are
designed range from 200 to over 20.000 RPM. Therefore. the values of Pd may vary
within wide limits from one pump to another.
At the inlet parts of the impeller blades. the amount of energy transferred from the
shaft to the liqUid is still small. Consequently. the pressure on the leading faces of
the blades is only a little higher than in the impeller eye. The pressure differential Pd
expresses itself. therefore. in a drop of pressure on the trailing face of the blades to a
magnitude well below that in the impeller eye.
Figure 6-2 presents the results of actual tests performed in 1957 by Acosta and
Bowerman [2]. In this illustration. Cp is the pressure coefficient related to the
peripheral velocity of the impeller outlet (Pstat - Hot! pU22/2). The term convex side
refers to the trailing face of the blade. The value Cp=O signifies the pressure in the
suction inlet of the pump before the liquid has been acted on by the blades. <l>e is the
flow coefficient at the design point, <I> is the flow coefficient at which the test data
were taken.
The data in Fig. 6-2 demonstrate that for flow rates greater than the design point,
the pressure at the trailing face of the blade may fall well below the pressure at the
impeller inlet.

EFFECT OF DESIGN FLOW RATE AND OPERATING SPEED ON NPSH


REQUIREMENTS

Pressure differences bring about the greatest variations in suction capabilities among
pumps. Therefore. based on Eq. 6-3. the NPSH requirements of a pump is also propor-
tional to the product wJQ,:. or

NPSH rx wJQ,:. (6-4)

Experience and theoretical studies have found that certain optimal parameters
govern the design of conventional centrifugal pumps. These parameters converge
around the design feature on which Eq. 6-4 is based.
48 Performance Factors

Cp
1.0
<j>=0
<j>/<j>e = 0
0.8

It

.. •
0.6
It"

"
;. ".
0.4
",;

0.2
Cp
e- •• 1.0
<j> = 0.167
<j>/<j>e = 1.43
O 0.8
X Convex side
-0.2 • Concave side
0.6
o Theoretical
-0.4
R/R2 0.6 0.7 0.8 0.9 1.0 0.4

Cp
1.0 0.2
<j>=0.113

. / ....
<j>/<j>e = 0.966 Co
0.8
0

L .... - ..- -...


0.6 f--------------,~---___j
-0.2 "'-.--e-"
,-
,- • ,,
0.4 f------~~+_-------<_~ -0.4 '--------'...L------*--'--------'
R/R2 0.6 0.7 0.9 1.0

0.2 f----U/----+_-------7~-___j

-0.2 f---~r"'----------___j

-0.4 '---_ _ _ _ _ _ _ _ _ _ _ _-"


RIR2 0.6 0.7 0.8 0.9 1.0

Figure 6-2 Pressure distribution around an impeller blade. R/R2 =Radial distance from center of
impeller, expressed as a fraction of the outer radius.
Cavitation in Centrifugal Pumps 49

Therefore, Eq. 6-4 can illustrate how NPSH requirements of a pump vary with
design flow rate and operating speed. In this respect, the accuracy of Eq. 6-4 can be
compared with the accuracy of the different popular tables that show the relation
between flow rate, specific speed, and maximum efficiency of a pump (Fig. 3-14).
Many pumps on the market show efficiencies that are higher or lower than the
efficiencies in Fig. 3-14. Nevertheless, statistical data prove that Fig. 3-14 presents a
fair picture of the maximum efficiencies of conventional pumps.
When a pump operates at a constant nominal speed, Eq. 6-4 can be rewritten as

NPSH=KQg·5 (6-5)

where K is a constant. This equation shows that the NPSH requirements of a pump
that operates at a given speed depend on the flow rate for which the pump was
designed.
In practice, pumps designed for the same speed and different flow rates also vary in
size. Consequently, Eq. 6-5 establishes a relationship between size and NPSH
reqUirements. According to this equation, the NPSH requirements of a pump can be
expected to vary with the square root of the design flow rate Qn. In practice, however,
several factors cause deviations from this rule.
In small pumps, the hydraulic radius of the waterways is relatively very small, and
the relative roughness is great. This consumes a significant part of the available
NPSH. Consequently, NPSH requirements for small pumps are higher than that given
by Eq. 6-5. It is easier to keep frictional losses low in large pumps than in smaller
ones. Also, more care is devoted to the manufacture of large pumps than smaller ones.
Differences in NPSH requirements are also caused by shape and design features.
All these factors result in two things. First, the actual NPSH requirements show a
wide scattering of values. Second, the general trend of the NPSH requirements does
not increase with the square root of the design flow rate, but according to some lower
value.
Figures 6-3 and 6-4 show the NPSH requirements of over 600 different pumps.
They represent the suction capability of pumps manufactured by 12 of the world's
largest and most prestigious pump companies.
As expected, the scatter is great. However, the NPSH requirements of the pumps
tends to increase as the 0.424 power of the design flow rate of the pumps. This value is
a little smaller than the theoretical value of 0.5 (Le., the square root) given by Eq. 6-5.
Figures 6-3 and 6-4 reveal another important detail. The coefficient 0.67 that
appears before Qg.424 in Fig. 6-4 is almost twice as large as the value of 0.34 in Fig.
6-3. Also, the operating speed of the pumps in Fig. 6-4 is twice as great as that in
Fig. 6-3. This confirms that the NPSH requirements of a pump designed for a given
flow rate are directly proportional to the designed operating speed.
As was mentioned earlier, in-depth knowledge and an understanding of the factors
that affect the suction performance of centrifugal pumps may reduce costly errors
relating to the suitability of a pump for a given duty. Still worse is the case in which
one has incorrect notions about the influence of such factors.
The previous discussion seems to dispute an older notion that the design flow rate
has no effect on the suction capability of a pump. According to this notion, the NPSH
requirements of a pump are affected only by its specific speed. The test data
presented in Fig. 6-5 prove unequivocally that these notions are far from true.
50 Performance Factors

1760 RPM
10
8 •

. .:.-
6
5 • •• •
4 • e.. ..:.. • • ••••
~
Q) 3 •
...... . ..",-
.. .

..
..1.. . ._:. •••
Ql
g 2 • •


•••••
• ...:. . . .. ~...

.. ... ,. ..: ..·i . •••


I
en
a.. ••
•• ••• • •
•• ...... ......
: :••: '"• .• •
z
•• -.:: ....
.-.-... .. :....
. ...
• .::. .
• ••••••
••
0.5

2.5 5 10 25 50 100 250 500 1000


Flow rate (m 3/hr)

Figure 6-3 NPSH requirements of various pumps at 1760 RPM.

3500 RPM

10
8 • •
••
•••
•• •
6 •
..........
: .......... .
5 • • •• •
• ••
: • .:. • e •••
~

~
Q)
Ql 3
4


.:

:.~
• ••• ••• 1
.
g • •••-••••••• 1.•••••: .. • •
I
en 2 A'l-A · ..... :_: . e. .
a.. e,1 a
().
• • • •• •
z ..., Q. • •
~,?S'0 ....

• •

0.5

2.5 5 10 25 50 100 250 500 1000


Flow rate (m 3/hr)

Figure 6-4 NPSH requirements of different pumps at 3500 RPM.


Cavitation in Centrifugal Pumps 51

Group III
5 • Ns = 32.5 1jJ = 1.05
Group I Ns = 23.2 1jJ = 1.09
o Ns = 11.3 1jJ = 1.15 o Ns =16.8 1jJ=1.16
Ns = 8.5 1jJ = 1.14 • Ns = 11.5 1jJ = 1.26
4-

~ Group IV
!!?
23 • Ns = 34.5 1jJ = 1.00
Q)
Ns = 25.0 1jJ = 1.00
E- o Ns = 16.4 1jJ = 1.08
I
(f)
Cl.. o
z
2 Group II
• Ns = 34.5 1jJ = 1.00
Ns = 25.0 1jJ = 1.00
o Ns = 16.4 1jJ = 1.08
1- • Ns = 11.2 1jJ = 1.23

OL-________ ~ __________ ~ __________ _ L _ _ _ _ _ _ _ _ _ _~ _ _ _ _ _ _ _ _ _ _~ _ _ _ _ _ _ _ _

o 25 50 75 100 125
Flow rate (m3/hr)
Figure 6-5 Experimental evidence that the NPSH requirements of a pump are independent of the
specific speed. (\]I is the head coefficient at design flow).

By modifying Eq. 6-5. we can establish a general chart of the expected NPSH
requirements of a pump designed to deliver a given flow rate Qn at a given speed S
(Fig. 6-6). In practice, the actual NPSH requirements of a particular pump may differ
significantly from data presented in Fig. 6-6. Nevertheless, these data still present a
good overall picture of the effects of speed and scale on the suction capability of a
pump, especially in the case of conventional end-suction pumps of good design.
For end-suction pumps, the chart may be a helpful gUide for a preliminary study of
the suitability of a given pump to operating under certain suction conditions. It is
especially useful when no test data are available. Of course, the final decision should
always be based on the actual test.
Another direct use of Fig. 6-6 is illustrated by the following case. Assume that a
pump designed for and delivering 800 GPM at 1750 RPM develops serious cavitation
problems. If these problems appear at 15 ft of available NPSH, the chart shows that
there is a good chance of finding another pump that will perform better. However, if
these problems develop at 7 ft of NPSH, the chart shows that the chance of finding a
pump with better suction capabilities is slim. At the same time, there might be a
pump that delivers 800 GPM at 1150 RPM that will operate satisfactorily under these
suction conditions.
52 Performance Factors

2 /~... ", :J.Vf-"'"


. . . ;.t:-~
1.5 .............. "" .....
f'...
... ~
I........

1 /..... ? ......
10 15 20 30 405060 80100 150200 300400 600 1000 2000 4000 6000 10,00t
Design flow rate (GPM)

Figure 6-6 Expected NPSH requirements of pumps designed for a specific flow rate at a specified speed.

EFFECT OF SCALE AND SPEED ON SUCTION SPECIFIC SPEED

Suction specific speed is often used to estimate the NPSH requirements of a pump
operating under other-than-tested conditions. The suction specific speed is also
commonly used to estimate the suction capabilities of a pump on the basis of tests
performed on another pump that is similar geometrically. Unfortunately, the results
of such estimates are rarely accurate or reliable.
Among other reasons, the lack of accuracy stems from the fact that the suction
specific speed of a pump, like its suction requirements, is affected by the pump's
scale and speed. If we replace NPSH in Eq. 5-3 by Eq. 6-5, the values of S become
solely dependent on the values of the speed N and the design flow rate Qn.
Using the expressions in Figs. 6-3 and 6-4 for the curve fits of the variations of
NPSH requirements with flow rate, the following expressions for S are obtained:

At 1740 RPM, S = 288 QR182 (metric units),


At 3500 RPM, S = 343 QR182 (metric units).

(These equations are derived from Ref. 1.)


To check how well these relations agree with the actual suction specific speed of
end-suction conventional pumps, the S values of the pumps in Figs. 6-3 and 6-4
Cavitation in Centrifugal Pumps 53

were calculated and entered in Figs. 6-7 and 6-8. Here. the deviations of the data
from the calculated values are even smaller than for the variations of NPSH
requirements with flow rate.
The data in Figs. 6-7 and 6-8. combined with the earlier discussions, permit us to
compile a general chart that shows how the suction specific speed varies with speed
and flow rate (Fig. 6-9). Similarly, the S values of individual pumps may differ
considerably from data in the graph. Nevertheless, this graph presents a good picture
of the general trend.
This graph is important for the pump user because familiarity with it may guard
against certain common misconceptions regarding the suction specific speed of a
pump. This, in turn, may prevent serious errors in choosing a pump for low-NPSH
applications.

1740 RPM
400
300
CfJ
-g 200
'5}" 150
<0= "
.~ 100
c.
'" 80
g 70
U
:J
60
en 50

Figure 6-7 S values


of centrifugal pumps 0.001 0.0020.003 0.006 0.010 0.0200.030 0.060 0.100 0.2000.300
at 1750 RPM. Flow rate (m3/sec.)

3500 RPM
800
600
500
400

.
CfJ
-0 250

....\-.-. :.: ....:: .....


3l 200
5} 180
1S8 .: ....... ..- ....
.-.:. .. .
~
al 120

..
5} 100
c 80
o
U
:J
70
en 60
50

Figure 6-8 S values


of centrifugal pumps 0.001 0.002 0.003 0.006 0.010 0.020 0.030 0.060 0.100 0.200 0.300
at 3500 RPM. Flow rate (m3/sec.)
54 Performance Factors

60,000
40,000

- -:::
~-

30,000
~0,500R~ ~ ::::
__ ~~oOo~~ ~ ::::
-
r:: ::::
::::
Cf) 20,000
-0
Ql
Ql
g.10,000
:::: f-
~ -t::::-;-OO RPM -~-t:: ~
I-- ~ ""!"~M ~__ I-- -I-
() 8000
~ 6000 .- -:::: 1--1--1""" r--",,1.:-~7§2~.- ~
c. :::: ~ 1--
~'-::;:;~8~~ -
-- -;::::...
<Jj 4000 440 RPM
c
~ 3000
a5 2000

1000
Figure 6-9 S values
800 of centrifugal pumps
600 designed to deliver a
6810 203040601100 2003001 60011000 1300016000 1 20,000140,000
80 400 800 2000 4000 10,000 30,000 specific flow rate at a
Design flow rate (GPM) specifed speed.

When choosing a pump for low-NPSH applications, many users are misguided by
the notion that a higher S value means better suction performance. Figure 6-9 shows
that this assumption is not always correct.
Take, for example, a pump for 1000 GPM at 3500 RPM with a suction specific
speed of 10,000. According to Fig. 6-9, its suction performance is lower than the
expected average. Now select a pump for 100 GPM at 1750 RPM with an S value of
7000. Although the S value of the second pump is lower than that of the first, it is
higher than the expected average (Fig. 6-9). This means that the second pump has
better suction characteristics than the first one.
Figure 6-9 is also useful when estimating the suction performance of a pump on
the basis of test data obtained from a model that is similar geometrically. In the past,
this was usually done by assuming that the S values of the models are constant.
Figure 6-9 shows that this is not correct.
The discussions and test data in Figs. 6-3 through 6-9 lead to equations that are
expected to have a better potential to give more correct estimates of NPSH
requirements based on model tests than does the assumption that S is constant:

NPSH p = (Np )1.424 ( Dp )1.272


(6-6)
NPSH m Nm Dm

and

Sp = (Dp )0.546( Np )0.432 (6-7)


Sm Dm Nm
In both equations, the subscript p refers to the pump and m to the model. Again, D
refers to any specific linear dimension of the pump, such as the impeller diameter.
Although suction capability is affected by more factors than scale and speed, these
equations give more significant values than do those that assume the S values remain
unaffected by size.
Cavitation in Centrifugal Pumps 55

EFFECT OF SPEED ON A PUMP OF GIVEN SIZE

The effects of flow rate and speed are intimately related to the size and geometry of
the impeller inlet. Although during the design stage a pump is chosen to deliver a
certain flow rate at a certain speed, it is often used at a different speed. For example,
a pump designed to deliver 2000 GPM at 3500 RPM may be used to deliver 1000
GPM at 1750 RPM. In this case, the impeller inlet will be significantly larger than in
the case of a pump that was originally designed to deliver 1000 GPM at 1750 RPM.
Figure 6-9 shows that the S values of a pump designed for 2000 GPM at 3500 RPM
are higher than those for a pump designed for 1000 GPM at 1750 RPM. This means
that, when operating at 1750 RPM, the pump designed for 2000 GPM at 3500 RPM
will require less NPSH at 1000 GPM than the pump designed for the lower speed.
Therefore, under extremely low suction heads, it may be better to use a high-speed
pump at 1750 RPM than to use a pump designed for the same flow rate at the lower
speed.
Mter a pump is designed, its suction specific speed remains almost constant over a
wide range of operating speeds. This is confirmed by tests for NPSH requirements at
two or more different speeds that were performed on many pumps.
Although S is usually constant for any given pump, there are exceptions [3). In
many cases, the S value of a pump seems to vary erratically with the pump's speed.
However, even in this case, there is a certain pattern: as the speed increases, the S
value increases.
Presently, little is known about what causes these differences. However, there are
strong indications that these differences are due simply to inadequate testing
practices. Most pumps that are tested confirm the assumption that S is constant over
a wide range of operating speeds. However, there are certain exceptions to the rule. In
each case, the required NPSH became very low at reduced speeds. At low NPSH
values, it is difficult to keep the pumping system absolutely tight. Also, it is very
difficult to measure very low NPSH values accurately with conventional instruments.
Mter the difficulties were overcome, the suction capability of each pump once again
followed the rule that S is constant. Thus, the suction specific speed may be constant
in any given centrifugal pump. It is always safer to assume, however, that the value
of S decreases when the speed is reduced.

EXTRAPOLATING TEST DATA

The manner in which the NPSH requirements of a pump operating at a constant


speed vary with the flow rate can be expressed by [4)

NPSH = KQ'2+L (6-8)

where K and L are constants.


The values of K and L may be found when the NPSH requirements for at least two
flow rates are know. Consequently, Eq. 6-8 allows us to estimate the NPSH
requirements for a given pump at flow rates for which no test data are available.
Assume that the highest flow rate Q for which test data are available is a. Let the
required NPSH at this flow rate equal A, the lowest flow rate for which the test data
56 Performance Factors

are available equal b, and the corresponding NPSH value equal B. In this case, the
values of K and L are given by

and

from which

(6-9)

Equation 6-9 enables us to estimate the NPSH requirements of a pump over most
of its total operating range, but it is rather dangerous to use it at very high or low
flow rates.
At high flow rates, the NPSH requirements of a pump increase with the flow rate at
a much higher rate than that expressed by Eq. 6-8. At very low flow rates,
particularly near shut-off, a strong interchange develops between the liqUid entering
the impeller inlet and the liquid in the suction pipe (see Chapter 9). This sets up
unstable performance at low NPSH values.
Even if there are no instabilities at low flow rates, Eq. 6-9 may produce misleading
results. At very low flow rates, the efficiency of a pump is very low. The additional
consumed power is mainly used to raise the temperature of the pumped liqUid. This,
in turn, raises the NPSH requirements of a pump.

NPSH REQUIREMENTS AT VERY HIGH FLOW RATES

At flow rates significantly higher than the design point, the Q-NPSH curves cease to
follow Eq. 6-8. Instead, they rise rapidly with any small increase in Q. (Fig. 6-10).
There are two reasons for this behavior. The first is cavitation in the volute throat (or
in the passages of the diffuser vanes), caused by the low total head developed by an
impeller at excessive flow rates. The high velocity of liqUid through the throat
converts part of the pressure energy into velocity head, reducing the pressure. At
adequately large flow rates, this reduction may cause cavitation.
In pumps with full-diameter impellers, cavitation in the volute occurs only at very
high flow rates. However, when a pump needs to operate with a cut-down impeller,

-gl---__ QH I
<D
I
I I
I

\
NPSH \
\ Q Figure 6-10 An increase in NPSH require-
ments with flow rate, in pumps operating at a
constant speed.
Cavitation in Centrifugal Pumps 57

the increase in its cavitation requirements occurs at a significantly lower flow rate.
This is because the pressures developed by an impeller with a cut-down aD will be
significantly lower than in a pump with a full-size impeller.
The NPSH requirements of a pump, published by the manufacturer, are often
based on tests performed with a full-size impeller. As long as these data are used for
cut-down impellers, at low or moderate flow rates, the pumps usually operate
satisfactorily. However, when a pump with a cut-down impeller operates at high flow
rates, cavitation may develop in the throat of the volute (or at the inlet to the diffuser
vanes).
The magnitude of the throat area At of a volute (or the passage at the entrance to
the gUide vanes) is usually chosen for a given design flow rate Qn and a given design
head Hn. Its magnitude depends on a large number of design parameters, such as the
head coefficient, shape of the cross-sectional area of the volute, and so on. In general,
however, it is possible to relate the magnitude of the throat area to the magnitude of
Qn and Hn by means of the following equation:

(6-10)

where K is a coefficient that takes into account all the relevant design parameters.
The total head H t of the liquid passing through the throat area At consists of the
velocity head V?/2g and the pressure head Pt/'Y:

H t = Pt + Vt2 . (6-11)
Y 29

The head H t is equal to the head Hi developed by the impeller, minus the losses HL
in the volute. Consequently, the pressure head of the liquid flowing through the
throat of the volute is equal to

(6-12)

At any flow rate Qc, the velocity of the liquid flowing through the throat area At is
given by

V t = Qc = Qc'K~
At Qn

or

(6-13)

Equation 6-13 tells us that the pressure head in the throat area decreases as the
square of the ratio of the actual flow rate Qc to design flow rate Qn. On the other
hand, the head Hi decreases significantly with the cut-down of the impeller diameter,
and with the increase in flow rate. This leads to the conclusion that with a cut-down
58 Performance Factors

impeller, cavitation in the volute will set in earlier than with a full-size impeller.
The second reason for the rapid increase of the NPSH requirements at high flow
rates lies in the manner in which power is transferred from the impeller blades to the
liquid. The process of energy transfer from a blade to the liquid is much more complex
than the discussion presented below (compare the discussion of Eqs. 1-1-1-7).
However, as far as the effect of high flow rates on the NPSH requirements is con-
cerned, the approach presented below is simpler and easier to understand.
Figure 6-11 presents schematically a particle of liquid P moving along an impeller
blade with a relative velocity W. If the magnitude of the blade angle at P is denoted
by ~ and the liqUid flows parallel to the blades, the following relations can be
established:

1. Peripheral velocity of blades at P:

U=wR

where w = angular speed of the impeller


r = radius of point P

---- ----tL--
I R

,,
,
',C,
,,
~-----:'---!'...J..._ _
C'L ___
...J...-...J...ooI ___ " ~
Figure 6-11 Velocity component
of a particle P moving along a
rotating blade.
Cavitation in Centrifugal Pumps 59

2. Velocity component Cm of the liquid in the radial direction

Q
C =-- (6-14)
m 2nkRB

where B = width of impeller in the direction parallel to the axis


k = effect of blade thickness on the through-flow area.
3. Relative velocity of liquid W=c,n/sin B. The component of W in the peripheral direction
(Fig. 6-11) is

(U - C LL ) = Cm x cot 13 (6-15)

where Cu is the tangential component of absolute velocity C of the liquid.

From equations 6-14 and 6-15, when the flow rate increases from Qa to Qb (Figs.
6-10 and 6-12), the (U-C u) increases from (U-Cula to

(6-16)

To simplifY the following discussions, assume that the tangential velocity component
CUI which the liquid has before entering the blades, equals zero. In this case, the
increase in total dynamic head of the liquid as it reaches radius R (Fig. 6-11) will be

HR = (UCLL)R (6-17)
9

where R is any specific radius


Whenever the liquid enters the impeller vanes without prerotation, a certain flow
rate exists where the relation between U I and C ml is expressed by

Cm1 =Ul tan 131 (7-18)

where the subSCript 1 denotes the inlet tips of the vanes.


At the flow rate Qa at which Eq. 6-18 is fulfilled, the variations of Cu with the radius
can be represented by curve Cua, which begins with zero at the inlet tips of the vanes
and increases continuously with the radius (Fig. 6-12). In practice, the shape of the
Cu curve varies from design to design. However, for ease of explanation, in Fig. 6-12
we have chosen to show it as a straight line.
Assume that the flow rate increases from Qa to Qb. The value of (U- Cul will increase
Qb/ Qa times its original value, at all radii. This produces negative values of ell at the
inlet portions of the impeller blades. Consequently, the inlet portions act as a turbine
instead of a pump (see Fig. 6-2 for actual test data).
The inlet portions remove energy from the liquid instead of adding it. This, in turn,
reduces the available NPSH at the blade inlets. This is one of the major causes of the
rapid rise in the NPSH requirements at flow rates significantly higher than the design
point.
60 Performance Factors

R--

I....c--~~ R ~--=---I Figure 6-12 Changes in the magnitude


I ....c--~~~R/~~~~
of the Cu component with increase in flow
1------~--R2~~~~~~~
rate.

In Fig. 6-12, a radius RI exists at which Cu=O, even for the curve expressing the C u
values at Q= Qb. Under ideal conditions, the NPSH available at this radius is identical
to that at the pump inlet. At this radius, the turbine action of the blades has
terminated. Still, the conditions here are less favorable than those at the inlet tips, as
far as the NPSH requirements are concerned.
The relationship between the difference of pressure Pd on both sides of the blades,
the radius R, and the velocity of whirl Cu can be expressed by the equation

Q( dCu
y ~-+-
Cu)
(6-19)
Pd = dR R
gZB

where Z is number of blades


Compare the value of Pd at the two flow rates Qa and Qb when Cu=O, the value of Pd
is Significantly higher at Q=Qb than at Q=Qa. This follows because Qb> Qa (Eq. 6-19)
and because (dCu/dR)b > (dCu/dR)a (because the angle </>b > </>a: Fig. 6-12). This
increase in Pd causes an additional increase in the NPSH requirements in addition to
the other effects.
All these factors result in very fast growth in the NPSH reqUirements of a pump
when the flow rate exceeds a certain critical value.
CHAPTER 7
Losses of Energy
LOSSES DUE TO RESISTANCE TO FLOW

When a pump must deliver a certain amount of liquid to an elevation He, it must
develop a head H that is higher than He because part of the head developed by the
pump is used up by the resistance of the pipelines to flow. This resistance is caused
by friction between the liquid and the wetted surfaces of the pipes (or ducts), and by
changes in the direction of flow. It is also caused by the changes in the diameter of
the pipes and the resistance rendered by the various fittings.
Extensive studies have demonstrated that the frictional losses in a pipe increase
directly with the length of the pipe, and inversely with its diameter. In a pipe of a
given length and diameter, frictional resistance increases (in most cases encountered
in practice) as the square of the average velocity of the liquid through the pipe.
Frictional resistance also depends on the viscosity of the flowing liquid and upon
the roughness of the wetted surfaces: the higher the viscosity and/or the roughness
of the pipes, the greater the resistance to flow.
In noncircular pipes or open channels, the pipe diameter used in the equations for
calculating the losses caused by friction is usually replaced with the ratio of the
cross-sectional area of the liquid, divided by the wetted perimeter of the duct.
In the past, attempts to calculate the losses in centrifugal pumps have used the
same methods that are used to calculate the losses in ducts. This usually led to very
disappointing results, because in centrifugal pumps the losses caused by friction are
much more complex than those in a stationary pipeline.

Frictional Losses in an Impeller

Disk Friction. When a disc that is submerged in a liqUid is put into rotation, its
motion will continue only as long as power is supplied to the rotating disc. It was
previously generally assumed that this power constituted a pure loss of energy with
no effect on the output of the pump.
Later findings have demonstrated, however, that this conclusion is incorrect. When
a rotating disc constitutes a shroud of an impeller, part of the power consumed by
the disc can be-and often is-returned to the pumped liqUid in the form of an
increase in head [5].
61
62 Performance Factors

200 QHcurves 7.7 m

180

+ Smooth shrouds
~ 160 • Sand-coated sh rouds
2Q)
.s 3000 RPM
~
Q)
.<:
(ij 140
(5
I- 70 0.7
60:g- 0.6
Q)

120 50 ~ 0.5
Q)

40 S 0.4 3:
~ (1J
30 a5 0.3
20g 0.2
100 W Figure 7-1 The effect of roughen-
10 0.1
ing impeller shrouds on pump per-
100 200 300 400 500 600 700 formance. (Worster and Thorne).
Flow rate (m 3/hr) BHP=brake horse power.

60
c
Q)
u
li5 50 ,.. ". ...... ---
EFF
......
.3: '"
~
55
·u
40 /
/ '" '"
I
ffi I

1400 RPM
<p

c
1.2
-- ...... .......
Q)
·u 1.0
....
"-
"-
~o - - - - Smooth shrouds ....
()
- - - Sand-coated shrouds ........
-g
Q)
0.8 .... 4.0
0...
I I
(1J

0.6 2.0

0.4 '--_ _----"_ _ _-'-_ _ _"""'----_ _ _'---_ _--" Figure 7-2 The effect of coating the
o impeller shrouds with 0.9 mm sand.
Flow coefficient BHP=brake horse power. (Varley).
Losses of Energy 63

Figures 7-1 and 7-2 show [6,7] the effects of increasing the disc friction by coating
the shrouds with sand. The increased roughness of the shrouds requires greater
power consumption, thus reducing the efficiency. However, not all of the additional
power consumed by the shroud is used up in reducing the efficiency. Part of it is
returned to the pumped liquid in the form of an increase in head.
Figure 7-3 illustrates the action of a rotating disc on the surrounding liquid. Owing
to the adhesive forces between the liquid and the disc surfaces, the liquid acquires a
velocity component CuI in the direction of rotation. However, the magnitude of C4f' is
not constant at any given radius R, but varies in the axial direction Z from the
rotating surface of the disc to the stationary surface of the casing [5,6]. Within the
boundary layer the value of CuI decreases very rapidly as the axial distance Z from the
rotating disc increases. Within the remaining liquid that is contained between the
rotating and stationary boundary layers, the value of C 4f continues to decrease as the
distance Z, but at a significantly reduced rate [5,6].
The differences between the CUI values near the rotating and stationary surfaces
produce a recirculatory motion of the liquid [7-10], as shown by the arrows in Fig. 7-3.
When we replace the rotating disc with an impeller (Fig. 7-4), some of the liquid
that had been acted on by the shrouds may start to interact with the liquid issuing
from the impeller, thus setting up a secondary loop or recirculation. This is shown
schematically by the dotted arrows in Fig. 7-4.
The causes and effects of such a secondary loop can be illustrated by means of the
following, Simplified model [7]. At the impeller outlet, the average value of the velocity

Figure 7-3 Distribution of Cur at the outer radius R2 , within the space
between the rotating disc and the stationary walls (schematic).
64 Performance Factors

Figure 7-4 The effect of change in Cui on the amount of


power recovered from disc friction (schematic).

component C u of the liquid that has been acted on by the impeller blades is given by:
C. = gHi (7 -1)
Ul U2

where Hi = head developed by the impeller blades


U2 = peripheral speed of the outer radius R2 of the blades
It may happen that within a certain portion of the gap between the shroud and the
casing, the values of Cuf (at the radius R 2 ) will be greater than CUi (Fig. 7-4). This will
cause a flow from this region toward the impeller outlet.
The velocity V at which the liquid flows from the cylindrical surface of width dZ
toward the impeller is given by
Vr=F [2g(Hrmlo. 5 =F [2U(C uj -Cu JIO. 5 (7-2)

where F is the coefficient that takes into account real liquid effects. Consequently,
the flow rate of the liquid issuing from the cylindrical surface of width dZ is:
(7-3)

and the flow of energy from the surface element dZ toward the impeller outlet will be
equal to
dEfi=9 P M CUj U2 dQj (7-4)

where p is the density of the pumped liquid.


Losses of Energy 65

In Eq. 7-4, M is a coefficient that accounts for different losses as well as the fact
that part of the flow rate dQr may flow directly toward the stationary casing, without
having interacted with the pumped liquid. The total amount of energy added to the
pumped liquid by one shroud is given by

Eft = 29pltFMR2JHrl2U(Cuj -Cu dlo. 5 dZ. (7-5)

°
Consequently, the head H that results from the interaction between the liquid that
has been acted on by both shrouds of the impeller and the liquid that came into
contact with the blades are equal (approximately) to

2Eft + gpQ;Hi
H= . (7-6)
gp(Qi + 2QI)

It is beyond the scope of this book to discuss all the parameters that affect the
magnitudes and the distribution of Cui along the axial distance Z. However, tests
reported in Refs. 11 and 12 imply that CUi increases with the relative surface
roughness of the shrouds. This seems to be responsible for the increases in the head
illustrated in Figs. 7-1 and 7-2.
Figures 7-1 and 7-2 demonstrate that the rate of power recovery increases with the
flow rate. For example, in Fig. 7-1, the amount of water power returned to the liqUid
at Q=Qd is twice the power returned to the liqUid at Q=O.5XQd.
When we increase the flow rate of the pumped liqUid to Qia, the head H developed
by that impeller decreases to Hill. This results in a decrease of CUi to the value of CUia
(Fig. 7-4). This decrease in the magnitude of CUi increases both the difference (of the
velocity components (CLf] Cud and the width of the cylindrical surfaces that extend
between the shrouds and the coordinates Zc, at which the value of CUi becomes equal
to the value of Cui. This, in turn, increases the magnitude of Eft (Eq. 7-5), that is, the
amount of power returned from the disc friction to the pumped liquid. This explains
why the amount of energy returned to the pumped liquid increases with the flow rate.
Roughness of the Watenvays. Roughness of the inner waterways of the impeller
always reduces effiCiency. As far as head is concerned, the results are not conclusive.
According to my own experience with pumps of medium and high specific speeds,
polishing the waterways of the impeller always improves both efficiency and head.
However, in one case of a pump with a very low specific speed, polishing the leading
face of the blades near the outlet drastically reduced both head and efficiency. More
corroborating test data ar needed to confirm the validity of these results.

Frictional Losses in Casing


In theory, the frictional losses in the casing should have exactly the same effect as
they do in any other stationary duct. Both the head and efficiency should decrease
with increased roughness. This has also been confirmed experimentally on many
occasions. In fact, it has almost become an industry standard to coat the casings of
deep-well pumps with ceramics or epoxy paint to achieve higher efficiencies.
Although this practice is usually successful in pumps of medium and high specific
66 Performance Factors

600 - o

550

500 3550 RPM

450

~400
0--0 Original casing 70
_ Casing coated with epoxy paint
~
0--0 Paint removed by sandblasting 60
J: 350
""0
ro
:!l! 300 50 C
~
OJ
250 403
~
c
200 30.~
:E
UJ
150 20
Figure 7-5 The effect of painting
10
the casing of a pump of low
specific speed on performance
o 100 200 300 400 GPM and efficiency.

speeds, the effect of casing roughness on the performance of pumps of low specific
speeds requires more study.
Figure 7-5 presents the results of tests performed at 3550 RPM on a single-suction
volute pump with a 1.5-in. discharge nozzle and an impeller with a 11.25-in. outside
diameter. The test with the original casing produced the QH and efficiency curves
shown by the solid lines. To improve the pump's performance, the waterways of the
casing were coated with epoxy paint. To everyone's surprise, this reduced both the
efficiencies and the heads developed by the pump. The pump and the test loop were
carefully examined, but nothing wrong could be found. The test was then repeated
several more times, always with the same results. When no other reasons could be
found for the poor performance of the pump with the painted casing, the paint was
removed by means of sandblasting and the test repeated. The pump returned to its
original performance (Fig. 7-5). Again, it is premature to arrive at any conclusions on
the basis of a single case history. This case indicates only that the matter requires
additional study.

LOSSES DUE TO LEAKAGE

Figure 7-6 presents a typical cross-section through a single-suction volute pump


with a closed impeller. The liqUid enters the impeller I through the suction nozzle.
The rotating impeller adds energy to the liqUid, which then enters the stationary
casing C. From there, most of the liqUid enters the discharge line. A certain amount
of it returns to the suction nozzle of the pump through the gap G, that exists between
the rotating impeller and the stationary casing.
Losses of Energy 67

DISCHARGE

t
Figure 7-6 Clearance
in a single-stage, end-
suction pump with a
SUCTION
closed impeller.

The amount of liquid that leaks through this gap depends on the difference in
pressure at the inlet and the outlet of the gap G, and on the width a (Fig. 7-7) of the
distance between the wearing ring of the rotating impeller and that of the stationary
casing. The amount of leakage through the wearing rings also depends on the ratio of
the gap's width a to its length b (Fig. 7-7) and on a number of additional factors, such
as the shape of the wearing ring (Fig. 7-7), the eccentricity of the impeller in relation
to the casing, the viscosity of the liquid, and so on.
In general, however, the amount of liquid QL that is lost to leakage can be
expressed as
0.5
QL = KJtaDw [ 29(P~ - PsI ] (7-7)
68 Performance Factors

Figure 7-7 Different shapes of wearing rings.

where K = a coefficient whose magnitude is determined by the factors discussed


above
a =width of the gap
Dw = mean diameter of the annulus that forms the gap
Pc = pressure in the casing at the inlet to the gap
P s = pressure in the suction nozzle

From Eq. 7-7, it follows that the amount of leakage can be expected to vary directly
with the width a of the gap G. Experience teaches us that this is correct, but only
within certain limits. Whenever the clearance a is reduced below about 0.010 in. to
about 0.020 in. (depending on the pump), any additional reduction in the clearance
seems to have no effect on the performance of the pump. This seems to be due to the
fact that the relative motion of the stationary and rotating faces of the sealing gap are
generating boundary layers on both faces of the passage. Under the conditions that
normally exist in a pump, the combined thicknesses of the two boundary layers
completely block off the leakage whenever the width of the gap becomes smaller than
about 0.010 in.
Whenever the width a of the gap is increased by so much that the overall efficiency
of the pump is reduced by 25% to 35%, any further increase in the width of the gap
seems to have no additional effect on the performance. In that case, the liquid
entering from the gap (Fig. 7-6) in the suction nozzle mixes with the main flow of the
pumped liqUid and returns part of the energy that it acquired before flOwing through
the sealing gap, to the incoming liqUid. When the ratio of leaking liqUid to pumped
liquid becomes large enough, an equilibrium is established between the amount of
energy lost owing to leakage and the amount of energy returned to the incoming
liqUid.
Figure 7-8 presents the overall losses that can be expected owing to leakage through
a cylindrical gap of a given width a, and of a length b of between 0.5 and 1.0 in. Of
course, in practice, these data vary because of many other factors. In general,
however, this illustration may serve as a useful gUide for carrying out rough
estimates.
Losses of Energy 69

The data presented in Fig. 7-8 are supported by the test results presented in Figs.
7-9 and 7-10. The relative amount of energy returned to the incoming liquid can be
expected to increase with the increase of the ratio of the amount of leakage to the
amount of pumped liquid. According to Eq. 7-3, the amount of leakage through a
wearing ring is proportional to the square root of the developed head. In most cases
encountered in practice, this means that the amount of power returned to the
incoming liquid owing to leakage through a given wearing ring will not be significantly
affected by the variations in the head developed by a given pump at different flow
rates.
Consequently, the ratio of the power returned owing to leakage of the water
horsepower of the pumped liqUid will increase primarily with the increase of the flow-
rate ratio of the leaking liqUid QL to the flow rate delivered by the pump Qp.
The relation (1- Qd Qp) will show us the effect of the ratio Qd Qp on the relative loss
of water horsepower. Figures 7-9 and 7 -lO demonstrate that the losses expressed by
the magnitude of (1- Qd Qp) are really decreasing at reduced flows of the pumped
liqUid.

~
~ 100
""C0> 90
.0;

~E
'" ~ 80
'"
>-~

'-'
co.
"'- 70
·13
~
Figure 7·8 The effect .~ 60
co
of gap width between a;
the wearing rings on 0:: o 0.010 0.020 0.030 0.040 0.050 0.060 0.070 0.080
overall efficiency. Width of gap between wearing rings (in.)

t-...:J 0.01 0 in. gap width

0--00.012 in. gap width

! ::~
(1 - QL/Qp )
_0.140 in. g~a~p:,w~id!!th~_ _ _ _ _':"'-""":""::"-----

______~QH
1750 RPM
~ 5
2°t:::~---<~----
o ""0"
'"
Qi
.s:r: 15

__ __
-0
''""

~::~I':~ ~ ~~----1-~~~~~~~~B~H~P~
Figure 7·9 The effect of wearing o 00 10 20 30
ring clearance on performance. Flow rate (Lisee)
70 Performance Factors

30

20
C
<IJ 160
~
(1 - QL/Qp )
cf 10
140
0
'§'
<IJ

1120
l:
-cro
~ 100
60

80
50

~ 40
tIl

30

20

5 10 15 20 Figure 7-10 The effect of a missing


Flow rate (Usee) wearing ring on performance.

Effects of Leakage on the Shape of the


QHCurve
In general, leakage can be expected to change the shape of a QH curve in the
manner shown in Fig. 7-11. At any given head. the flow rate should be reduced by an
amount q that corresponds to the additional loss owing to leakage. This has been
confirmed by testing the same pump with different clearances. In fact. I have often
used these data to determine the amount of the increase in leakage caused by
increasing the clearance between the wearing rings. This method. however. has
proven to be reliable only within a range of about plus or minus 30% of the design
flow rate. At lower partial flow rates. a reduction in head. which seems to be affected
by parameters other than the leakage flow. occurs. No test data are available. as yet,
for large flow rates.

EFFECT ON Q-NPSH CURVE

For relatively moderate increases in leakage. the effects of these losses are analogous
to their effect on the QH curve. The required NPSH moves to a flow rate reduced by
the amount of the increase in leakage. Figure 7-12 shows such a case. When the
Losses of Energy 71

Figure 7-11 The effect of leakage on


the shape of the QH curve. Flow rate - -

6 /
~ /
25
Q) 8---...-"
E /'
/'
~4 /

8--- ---
(J)
8 __ / /
~3

---
2
_;:....-- b=0.12mm

Figure 7-12 The effect of wearing


ring clearance on NPSH require- o 10 20 30 40 50 60 70
ments. Flow rate (m3/hr)

clearance between the stationary and the rotating rings increased from 0.12 to 0.25
mm, the NPSH required for each flow rate moved to a flow rate reduced by 8 m 3 /hr.

Leakage in a Semi-Open Impeller


Practice has shown that an increase in the gap a between the impeller blades and
the wear surface of the casing can reduce the output of a pump very significantly.
The conditions here are somewhat similar to leakage through the wearing rings of a
closed impeller, except for three important differences:

1. Instead of the difference (Pc-Ps) from Eq. 7-7, we must take into account the difference of
pressure on both sides of the blades as defined by Eq. 6-19.
2. The leakage from the leading face of the blades will not always move perpendicular to the
blade surface, but will move in the direction of least resistance. This direction will vary
72 Performance Factors

not only among impellers of different geometries, but along the total length of the same
blade.
3. In addition to the leakage caused by the difference of pressure on both sides of the
blades, additional leakage will occur in a tangential direction because the liquid in
contact with the stationary casing will tend to remain stationary while the liquid in
contact with the rotating blades will tend to rotate.

Many attempts have been made to handle this problem analytically, but with little
success. Figure 7-13 illustrates [13] how the performance of an impeller can change
with the increase in the ratio between the width of gap a and the width of the
impeller b at the blade outlet. As in the case of a closed impeller, when the magnitude
of the gap falls below a=O.O 10 in., leakage seems to cease. When the magnitude of
the gap exceeds a certain limit, the leakage no longer affects pump performance.
There is, in fact, a special line of pumps (so-called vortex pumps) in which the
impeller blades are located completely out of the way of the pumped liquid.

Head coefficient of total head

1.5 -
\)I2T

Efficiency
1.0 - 90

0.5 - 80~
>="

- 70 Static head coefficient

alb
• 0.007
0.020
.
0

0.5 - 0.033
0 0.062
• 0.095

'"
()
0.130
0.170

o~ ____________ ~ ____________ ~ ____________ ~ ________________


o 0.2 0.4 0.6
<I>

Figure 7-13 The effect of clearance on the performance of a semi-open impeller. <!>=Flow
coefficient. {Ishida and Senoe).
Losses of Energy 73

EFFECTS OF TIME

Our discussion of losses caused by surface roughness and leakage brings to light an
often overlooked detail: no pump can maintain its performance over its total life-
span. With time, the wetted surfaces can get rougher due to corrosion or erosion.
This will usually increase the power consumption of the pump and, in the case of
medium and high specific speeds, it will usually reduce the heads developed at a
given flow rate. The liquid may contain fine particles of abrasive material that may
polish the waterways. In certain instances, these abrasive particles may even
increase the size of the waterways. Both cases may increase the output of the pump.
Wear of the sealing surfaces of the wearing rings will always increase the amount of
leakage, thus reducing the output. Finally, both the roughness of the wetted surfaces
and the amount of leakage can-and usually will-affect the suction capability of the
pump.
Other time-related factors that may affect the suction capability include the
following:
1. When a pump takes water directly from a river or lake, the strainer often becomes filled
with weed or grass, which obstructs the flow.
2. Solid matter may accumulate within the foot valve. thus increasing the resistance to flow.
3. The water table may fall, thus reducing the available NPSH.
There is no way to prevent the "aging" of a pump. In this respect, a pump that
operates intermittently is more likely to be subjected to wear and to changes in the
texture of the surface than a pump that operates continuously. When a pump is at a
standstill, solids that were suspended in the liqUid may separate and settle inside the
waterways, thus reducing the size of the passages. If a pump were handling
concentrated solutions, chemicals dissolved in the liqUid may crystallize. This, in
turn, may cause serious erosion of the waterways when the pump is restarted.
However, it is possible to delay the process of increased wear by selecting special
materials for the wetted parts of the pump. In general, the resistance of a given
material to wear depends on a variety of different factors, such as temperature,
chemical affinity of the pumped liqUid, and so forth. Generally speaking, however, the
wear resistance of materials used in centrifugal pumps increases in the following
order:
1. Aluminum and some of its alloys
2. Plastics (with the exception of some special materials)
3. Cast iron
4. Cast steel
5. Certain varieties of bronze
6. Manganese steels
7. Low-alloy chrome steels
8. High-alloy chrome steels
9. Certain varieties of rubber (used in conjunction with stainless steel for the mating
parts)
10. Certain ceramics
11. Certain special alloys
12. Certain carbides
74 Performance Factors

The list above serves only as a very rough guide. The wear resistance of a material
may vary not only with its composition, but also with the manner in which it has
been produced. For safety, it is always good to seek the advice of the manufacturer of
the materials that have been used in the pump.
The wear resistance of a given material is not the only criterion that determines the
longevity of a pump. Other factors are corrosion resistance and resistance to galling.
The first property relates to the chemical action of the pumped liquid, as well as to
the galvanic action that often occurs when two different materials are submerged in
an electrolyte. Galling is the tendency of certain materials to wear off particles from
the surface of another material and to fuse them to their own, when two surfaces rub
against each other.
Other time-related factors that can affect the performance of a centrifugal pump
are setting of foundations or of other structural components of the pumping system
and changes in the operating conditions of the pump, due to changes in demand, and
so forth (see also Chapter 19).
CHAPTER 8
Effects of Temperature
and Viscosity on
Pump Performance
EFFECTS OF ELEVATED TEMPERATURES

Problems related to elevated temperatures can be subdivided into two classes: over-
heating caused by mechanical faults and problems related to pumping hot liqUids.

Overheating Pump Parts


The parts most frequently subjected to overheating are the bearings, bearing housing
(see also Chapter 16), stuffing box, and mechanical seals (see also Chapter 17).
Common causes of overheating are faults in handling, installation, assembly, and
machining (see also Chapter 15). In addition, when the bearings or the seal requires
cooling, overheating may be caused by inadequate cooling arrangements.
Overheating is dangerous because it can trigger a chain reaction or a series of
chain reactions. Overheating the bearings may destroy the lubricating properties of
the lubricant. This may heat the bearings even more, leading to oxidation and
scorching of the surfaces and complete breakdown. Overheating the bearings will
also overheat and expand certain parts of the bearing housing. This may put parts
out of alignment and ultimately lead to breakdown.
An overheated packing or mechanical seal can lead to uneven expansion of
different pump parts. This, in turn. may lead to scorching of the shaft surface that
comes into contact with the packing. This can also ruin the mating faces of
mechanical seals. In both cases, the sealing capacity of the stuffing box may be
reduced considerably. which can result in excessive leakage of liqUid or allow air to
enter the pump.

Pumping Hot Liquids


Five problems are common to all hot-liqUid pump applications:

1. Reduced material strength


2. Higher thermal expansion in areas closest to the hot liquid
75
76 Performance Factors

3. Packing material and mechanical seals may fail on contact


4. Special gasket material may be required
5. Wetted pump parts may be less corrosion-resistant

This list is not comprehensive, but it demonstrates one important point: never pump
hot liquids with a pump purchased for normal temperature applications, without
checking with the manufacturer.
Even when a pump is specially designed and built for high-temperature applic-
ations, it may experience "thermal shock." When a surface is heated qUickly, it
expands suddenly, imposing enormous stresses on adjacent cooler parts. Depending
on the rate of the temperature increase, its final value, the conductivity of the heated
material, and its mechanical properties, such a thermal shock can be detrimental.
For example, a cold glass will break if it is quickly filled with hot water. This is the
most drastic case. In more ductile materials, a thermal shock can deform part of it
beyond its elastic limit. In other cases, thermal shock may cause fatigue.
The severity of thermal shock increases with the speed of heating, temperature,
and size of the pump. In smaller pumps, the effects of thermal shock can often be
counteracted by a suitable design and a choice of suitable materials. In severe cases,
however, it is very difficult-and often impossible-to counteract the effects of
thermal shock in large pumps. The only remedy is to heat the pumping unit
gradually, according to the manufacturer's recommendations.
Another problem encountered when pumping hot liqUids is that changes of
temperature will also cause pipelines to expand and contract. This may impose high
stresses on the pump casing. The imposed forces may also push the pump out of
alignment with the driver. Thus the alignment must be checked both at room
temperature and after heating.
A special class of problems may occur when a pump handles overheated liquid
within a closed pressurized loop. Loop breakdown may flash the liqUid into vapor and
convert the pump into a steam turbine. This may raise the operating speed of the
rotating unit beyond safe limits, damaging bearings and structural elements of the
pump and overheating and binding the sealing faces. In extreme cases, the centrifugal
forces can literally pulverize the impeller. Also, the increased impeller speed may cause
cavitation in certain pump zones, resulting in additional heavy damage.

EFFECTS OF VISCOSITY

Effects on Performance
Viscosity is defined as the resistance of a real liqUid to flow. The higher the
viscosity, the greater the force required to make the liqUid flow. Viscosity is generally
expressed as the force required to move a plane of a unit area of liqUid over another
plane of equal area in one second.
The viscosity of liquids is measured by an instrument called a viscometer. Unfortun-
ately, various industries use different standards for expressing the magnitude of
viscosity, and viscometers are manufactured with scales that are calibrated to these
differing standards. Often, conversion tables must be used to adapt the results of the
tests with a specific viscometer to the standards of a particular industry.
Effects of Temperature and Viscosity on Pump Performance 77

-
Performance Correction Chart
25 to 10,000 GPM
1.00
r-- ~

'"
0.90

al 0.80
OJ
I
~ I 0.70

----
o
ti
co 0.60
LL
§ 1.00
:g
~ 0.90
r-..... ~f'..
i'"'"-- .......
o~
() a5 0.80
i', ~Q
.(3
~ '\

'"
iE 0.70
w
-g 0.60 i \
co
.~0.50
, I~E \
()
co I
\
~0.40
()
0.30 I
!
\
0.20
I
/ ~~ u<
I cP. ~~~
t

t.fu
Ji>
/
o
/
\5' "b <f-> v 6J- ~ <f-> 6' "b 'b '0
Ji> 6]
C5Q ~ ~ <SO 'b <t, Centistokes

\ \\ \ \\ \ ,\ 1\\\ 1\1\ \l\ \' ~\ \ 600

\ \ \\ \\ \\ \\ l\ 1\ \\ i\\ I\~ \
400
300

....-: ~ ~~
~ 200
~ 150

\\,\ \ \\ \\ \ 1\l\ 1\ \\ '\ \\) ~ ~ ~ :...-:~ ~:::


100
\ ~
~8
\ \ 40

1\ \ 1\\ \\ \ \ 1\1\\\\i\\ \ ~ ~ ~ ~ %/::/ ' ~V


\ \ \\ \ 1\ \ 1\ 1\ ~ ~ ~ ~ ~ ~

1\ \\ \ I\,V ~ ~ ~:;.; ~ ~ ~ ~ \
\ .\~ ~ ~ ~~ ~ ~ ~ ~ 1\'\.\ 1\
~ 1\\\\ l'-..\ ~
\
~~~~~
j;~ ~~

~ ~ ~ ~ ~ ~ ~ PI: \\ \ \\ \\ \\ \' \\ \
./
)~ ~ ~ ~ ~ K\ 1\ \ \'\\ \'\ \ \\ \' .\ \\ ~\
~ ~ ~ ? V\i '\ 1\1\ '\\ \\ \ \\ \ l\ \\ l\ i\
\\
\ 1\ \\ \ ,\ \\\ \\ \ ,\ \ \\
600
8 \
~~~
-- 40
OJ 30

1\ \ ~ \\ \\ \ \\ t\\ l\ \ \ \
200
2 ,
~V I 1\
150
- 100
"0 80
co 60 \
\ 'tl\\ \\ \ \ \ 1\
OJ 40 ,

~
I
I ,
30 ,
i\
rI
20 I 1\ \ \ \
10
Viscosity (SSU) I
I
'v
II
6! Ol.-': /.;! v- 6! Ol /
v 't> I<f, 't> v 't> 't> 't>'t2
/, ~
'f1 't2 ~ ~
VVVVV
<.il v-

0.25 0.50 1.5 2 3 4 5 6 7 8 10 15 20 30 40 5060 80 100


Capacity in 100 GPM
Figure 8-1 The effect of viscosity on performance of a centrifugal pump.
78 Performance Factors

In the United States the most commonly used units are the centistoke and the
Second Sable Universal (SSU). For comparison, the viscosity of water at 60°F is
1.1 centistoke or 31.5 SSU.
Figure 8-1 shows how viscosity affects the performance of a pump. All rating
curves published by pump manufacturers are based on tests performed with water.
The data shown in Fig. 8-1 demonstrate how the performance of a viscous liqUid
compares with the performance of a pump handling water. To use this data, find the
point of intersection of the line that expresses the flow rate with the line that
expresses the head. From this point, draw a horizontal line until it intersects the
proper viscosity line. Now draw a vertical line upward until it intersects the curves
CQ, CE, and CH. The points of intersection of this vertical line with the capacity curve
Co, the efficiency curve CF, and the head curve CH will show you by how much the
pump characteristics published by the manufacturer will drop due to the viscosity of
the liqUid. However, the data presented in Fig. 8-1 are based on a very limited
number of tested case histories and should be used only to get a rough idea of what
can be expected from a given pump, and not as a basis for quantitative predictions.
The viscosity of a liqUid decreases with an increase in temperature. Very viscous
liqUids are frequently preheated before being handled by the pump. Certain
precautions are necessary if the pump is to be operated intermittently. First, to avoid
sudden cooling of the liquid at each restart (when it comes into contact with parts of
the pump that have had time to cool om,
it may be necessary to keep the pump hot,
even when it is not in operation. When pump operation is to be interrupted for a
longer period of time, the liqUid must be drained from the pump before its
temperature decreases and its viscosity increases beyond tolerable limits.

Effect on NPSH Requirements


It is a common belief that viscosity has no significant effect on the NPSH require-
ments of centrifugal pumps. It is hard to find any data that contradicts this opinion,
but while stUdying the performance of gear pumps I came across data that indicate, for
viscosities higher than about 100 SSU, NPSH reqUirements increase with the viscosity
of the liqUid. The manufacturer of the pumps I studied did not specifY whether the
data are the results of actual tests, based on some sort of theory, or copied from data
published by other manufacturers. Consequently, it is not known whether these data
are really valid or if their validity is general or limited only to these seven pumps.
According to these data, the NPSH reqUirements of a pump handling viscous
liquids increases K times over the NPSH requirements of a pump handling liquids of
a viscosity lower than 100 SSU. For viscosities higher than 100 SSU, the
manufacturer has provided data for viscosities up to 70,000 SSU. The data provided
by the manufacturer were in the form of sets of straight lines, drawn on logarithmic
paper, making it possible to express the values in the form of an equation:
4 3 2

K = -0.4776[IOg( ~)] + 3.8688[IOg( ~)] - 9.875[ log( ~)] + 8.1772 IOg( ~) + 1 (8-1)

where VI relates to a liqUid with a viscosity of 100 SSU and V relates to the viscosity
of the pumped liqUid.
At present, I know of no experimental data that would confirm or deny the validity
of this equation.
CHAPTER 9
Recirculation
For many years, recirculation was regarded as some sort of mysterious factor
responsible for every problem that could not be explained. As such, it had no exact
definition suitable for analytical discussion. This sometimes led to inaccurate or
misleading conclusions, as shown by the case history presented here.
The habit of attributing to recirculation any problem for which no explanation
could be found led to the belief in a direct link between recirculation and the head
coefficient for which the impeller has been designed.
It has been demonstrated in Ref. [l4) (see also Chapter 11) that impellers designed
for high head coefficients are more apt to develop a drooping curve than those
designed for lower ones. This effect is a direct result of hydraulic losses that are not
necessarily related to recirculation. Unfortunately, many people are unfamiliar with
the study presented in Ref. [14). For them, the cause of the droop constitutes some
sort of mystery that is most conveniently attributed to recirculation. This belief
resulted in a proliferation of (often very expensive and time-consuming) studies aimed
at establishing a relationship between the head coefficient and recirculation.
Today, we know that recirculation can be linked only indirectly to a drooping curve
[8, 9, 15). We also know that this link has nothing to do with the magnitude of the
head coefficient, for which an impeller has been designed (compare the discussion
related to Figs. 9-17 and 9-18). In fact, when used properly, recirculation can change
a drooping curve into a steadily rising one, as demonstrated by the test results
presented in Fig. 9-3 and the accompanying discussion. It has also been confirmed
by curve A in Fig. 9-13, as well as by hundreds of practical applications.
In fact, thanks to recirculation the industry was able to come up with a complete
class of rotodynamic pumps (the so-called regenerative pumps) that are capable of
developing unusually high head coefficients and very steep QH-curves that rise
steadily up to shut-off. These features were achieved by using the effect of
recirculation [16). Regenerative pumps have been on the market for nearly six
decades and always exhibit the features listed above.
In real life, I know of cases in which the practice of attributing unexplained
problems to recirculation has caused many problems to remain unsolved. In some
cases, it even resulted in needlessly scrapping costly equipment solely because it had
been diagnosed with a "terminal case of incurable recirculation."
Today much more is known about recirculation [8-10, 16-20). For example, it has
been found that, when properly understood and handled, recirculation can be
successfully used to improve certain operational features. In particular, it can convert
a drooping QH curve into a steadily rising one (see Figs. 9-3, 9-11, and 9-13).
79
80 Performance Factors

Let us begin by defining the exact meaning of the term recirculation. We define it as
the effect in which a liquid, after having exited the impeller, reverses its direction and
flows back toward the impeller. In general, we distinguish two zones of recirculation:
one at the impeller inlet and the other at the impeller outlet. We will also consider
certain combinations of these effects.

RECIRCULATION AT THE INLET OF A PUMP

The principal causes of recirculation at the pumps-inlet can be best explained by


Figs. 9-1 and 9-2.
Referring to Fig. 9-1, let us consider the following:

1. A control volume B of liquid that is being acted on directly by the inlet tips of the rotating
blades
2. 1\vo cylindrical shells of mean radii Ra and Rb, respectively. each of radial thickness dR.
3. A disc-shaped volume A located too far upstream from the blades to be affected by their
rotation

Figure 9-1 Recirculation at the inlet


of a centrifugal pump.

Pressure head
in absence of
recirculation

Pressure head
Figure 9-2 The effect of recirculation
distribution due
to recirculation on pressure distribution in a suction
pipe.
Recirculation 81

Let us begin our study with the assumption that the pump is operating against a
closed discharge valve. Under the action of the blade tips, the liquid contained within
zone B acquires different peripheral velocity components Cu at different radii. As a
result, a pressure gradient appears within zone B of the control volume. This pressure
gradient is equal to

dP = yC~ (9-1)
dR gR

Head measurements by Peck [21] and Schweiger [22] indicate that it is possible to
approximate the distribution of the pressure P along the radius R within the zone B
by the equation

(9-2)

In Eq. 9-2, K is a coefficient (not necessarily a constant) that accounts for the factors
that determine the magnitudes of C u at the given radii R (within zone B), and Po is the
pressure at the center of the suction pipe.
From the outset, we have assumed that there is no prerotation within zone A.
Consequently, in zone A, the pressure within the liquid will remains the same as
before the pump was started. This is shown schematically in Fig. 9-2. In zone A,
where prerotation is absent, a constant pressure equal to the datum pressure D-D,
shown in Fig. 9-2, exists. In zone B, however, the pressure varies in a parabolic
manner, as expressed by Eqs. 9-1 and 9-2. This pressure distribution is represented
by parabola C in Fig. 9-2, which tells us that in the presence of prerotation, the
pressures at the outer diameters of zone B are higher than the pressure in zone A.
Similarly, at the inner diameters of the suction pipe, the pressures in zone A are
higher than in zone B.
As a result of these pressure differences, the liquid begins to flow at the outer radii
from zone B toward zone A (i.e., away from the impeller), and within the inner radii of
the suction pipe, it flows from zone A toward zone B (i.e., toward the impeller). These
two flows form a recirculatory motion upstream of the impeller blades, as shown
schematically in Fig. 9-1.
The pressure differentials within zone B are created when the blades impart the
peripheral velOCity components Cu to the liquid (see Fig. 9-1). Because of inertia, the
liquid retains some of that peripheral velocity after it moves away from zone B toward
zone A. This produces the well-known effect of prerotation, an inherent part of
recirculation at the inlet of the impeller.
A glance at Fig. 9-2 tells us that at radii larger than Re, the pressures within zone B
of the suction pipe are higher than the pressure in zone A. Between the radii Re and
Rs of the suction pipe, the liquid flows away from the impeller with a velocity equal to
(9-3)

where Ha=Pal"{ is the difference in the pressure heads between zones B and A at any
given radius. Through any cylindrical section of the suction pipe of mean radii Ra and
width dR, a flow rate equal to

(9-4)
82 Performance Factors

is recirculated back into the suction pipe. From this it follows that the total flow rate
of the liquid recirculated back from the impeller into the suction pipe is equal to
R"
Qa = 2:JtfRCbadR. (9-5)
He

Similarly, between the radii 0 and Rc (Fig. 9-2), the liquid returns to the impeller
with a velocity equal to Cab= [2gHb]O.5, and the flow rate is equal to
He
Qb = 2:JtfRCbadR. (9-6)
o
The law of continuity mandates that at the closed discharge valve

(9-7)

Consequently, Eqs. 9-1-9-7 make it possible to calculate the intensity of recirculation


whenever the magnitudes of the Cu components are known. Later, we will see how the
inlet geometry of the impeller blades can affect the intensity of recirculation.
Until now, our study has been limited to the case of a pump operating against a
closed valve. Now let us see what happens when we start to open the discharge valve.
As long as the flow is less than the design flow rate, the value of the coefficient Kin
Eq. 9-2 decreases [8,9]. In addition, the liqUid begin to flow toward the impeller eye
with an average velocity of

(9-8)

where Qp is the flow rate of the pumped liqUid.


Now because the direction of Co is opposite that of Cba, the velocity at which the
liqUid returns to the suction pipe due to the pressure differential generated by
prerotation is reduced.
The new back-flow velocity CR is now equal to
(9-9)

Equation 9-9 tells us that the intensity of back-flow decreases with an increase in
the velocity Co (Le., with an increase in flow rate).
Whenever the flow of the pumped liquid is increased by so much that
(9-10)

at all radii of the pump inlet, the back-flow from the impeller toward the pump inlet
vanishes completely.
The effects of recirculation on the head developed by a centrifugal pump can be
compared with the action of a multistage centrifugal pump [16]. In a multistage
pump, it is possible to attain very high heads because the liqUid, after being
subjected to the action of one impeller, enters consecutive impellers, which add more
energy every time the liqUid comes into contact with the impeller blades.
Recirculation 83

During recirculation, some of the liquid that has already been acted on by the
blades once, returns to the same impeller, where it is acted on again. This increases
the energy contents of the liquid.
The recirculated liquid mixes with the main stream of the pumped liquid, thus
increasing the energy contents of the total flow of the pumped liquid. The amount of
energy added to the total flow of the pumped liquid depends on the ratio of the
amount of recirculated liquid to that of pumped liquid. This ratio increases with any
reduction in the amount of the pumped liquid (Le., with any reduction in flow rate
delivered by the pump).
Mathematically, this can be proven as follows. Assume that when a volume Qp of
liquid has been acted on by the impeller once, its head is increased by the magnitude
Hp. Now assume that a flow rate QR has been recirculated into the impeller for a
second time. As a result, its head has increased to Hp+ HR. When the recirculated
liquid mixes with the main flow delivered by the pump, the resultant head developed
by the pump at the flow rate Qp is equal to

or

(9-11)

Equation 9-11 tells us that the increase in total head caused by recirculation is the
largest at Qp=O (Le., at shut-off) and decreases with an increase in the flow rate Qp.
Indirectly, this assumption can be corroborated by installing an inducer at the inlet
of the pump (Fig. 9-3).
The relationship between inlet recirculation and the developed head is best
illustrated by the following case. Figure 9-3 presents the results of tests performed
on an end-suction centrifugal pump with a specific speed of 27 (metric units). The
impeller of that pump had cylindrical (single curvature) vanes, whose inlet edges were
inclined to the axis by about 10 degrees. The pump was tested with a horizontal
suction pipe having a 105-mm inner diameter, whose centerline was submerged 500
mm below the surface of the liquid in the suction tank.
At shut -off, when tested at 3560 RPM, the manometer showed a positive head of
1050 mm at the center of the suction line. This means that it has a head 550 mm
higher than the submergence of the center of the suction pipe below the level of the
liquid. The only possible source of this excess head is the action of the impeller. This
means that some of the liquid in the vicinity of the suction tap after being acted on by
the impeller blades has been recirculated back into the suction pipe.
Next, an inducer was installed in the suction nozzle of the pump. This time, at
shut-off, the manometer read a suction head of 4150 mm above the atmospheric
head. (Le., the recirculation has added 3560 mm of head to the liquid in the suction
pipe.
Now let us see how this enormous increase of recirculation at the impeller inlet
affects the total head (Fig. 9-3). At high flow rates where recirculation is absent, the
addition of the inducer has made no difference in performance. At partial flow rates,
84 Performance Factors

-----.. .....
1.2 - _ - - -__ ' , ....With inducer
Impeller only .... :..
:sJ: 1.0-
0)
C\J

0.8 3560 RPM

EFF
80

70

60
'E
~ 50
Qi
0.
;: 40 ~
() Q)
c Q)
.s
Q)
·030
:E
LlJ I
(f)
20 c..
z
10
Figure 9-3 The effect of installing an
o inducer (dashed lines) on the perfor-
o 0.2 0.4 0.6 0.8 1.0 1.2
Q/QoPt mance of a 3-in. end-suction pump.

however, the head developed with the inducer began to increase over the head
developed by the impeller alone. This difference increased with each reduction in flow
rate, as could be expected from our earlier discussion. The validity of our study of the
effects of recirculation on the developed head has also been verified in an additional
way [20]. For this study, the impeller used in the tests represented in Fig. 9-3, was
remachined, as shown in Fig. 9-4 and retested. The removal of part of the blades that
projected into the impeller eye immediately reduced the intensity of recirculation, as
shown in Fig. 9-5. At the same time, however, it lowered the heads developed at
reduced flow rates enough to produce a drooping curve (Fig. 9-6).
Note that to compare the intensities of the induced recirculation, we have applied
the following criteria. As mentioned earlier, the maximum absolute head that can
exist in the suction pipe in the absence of recirculation is equal to the atmospheriC
pressure plus (or minus) the (positive or negative) elevation of the liquid in the
suction tank, as related to the centerline of the impeller. In our particular case, this
elevation was 500 mm positive. Whenever the head measured in the suction pipe
exceeded this value, we accepted it as a measure of recirculation. This assumption
has also been double-checked visually by installing a transparent pipe section with
thin threads attached to its inner walls upstream of the pump inlet. The onset and
increases in intenSity of recirculation could be seen from the deviation of these
threads from the direction parallel to the axis of the suction pipe. The magnitudes of
these deviations conformed well to the magnitudes of the increases in the excess
pressure at the pump inlet.
Recirculation 85

Figure 9-4 An original and a re-


machined impeller used in the tests
shown in Figs. 9-5 and 9-6.

450

400

350
A.Hs = Excess of suction-head
0>300 readings over maximum possible
J: in absence of recirculation
~ 250
::to)

100

50
OL-~ __~__~~~~~~__~__L-~_ _~_ _~~
Figure 9-5 The effect of inlet o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
geometry on recirculation. aQd

This is not an isolated case. I have observed similar effects in hundreds of separate
and unrelated cases.

Hysteresis and Other Forms of Instability


The study that led to Eq. 9-lO indicates that when we increase the flow rate so that
the velocity Co of the incoming liquid exceeds a certain limit, recirculation is
eliminated. In reality, however, recirculation is likely to disappear at two flow rates.
Figure 9-7 shows the flow pattern of the liquid in the suction nozzle when
recirculation first starts to appear (23]. If we begin testing a pump at a closed valve,
the pump will certainly start to operate within its recirculation range. Now, when the
discharge valve is opened gradually, the flow rate increases gradually until at a
86 Performance Factors

I, Original impeller
II, With inducer
1.3
III, Remachined impeller

1.2

1.1
C,)::r.

1.0

0.9 I + II

0.8 II
EFF /

/
80
'E 70 /
~ 60 III
21. 50
-;: 40
g
Q)
30
'(3 20
:E
lJ.J 10 Figure 9-6 The effects of impeller
O~-L __~__~__L-~_ _-L__~__~__L-~_ _-L~
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 modifications on the QH and EFF
Q/Qd curves.

Figure 9-7 Flow pattern at the


onset of recirculation at the inlet of a
rotodynamic pump [23).

certain flow rate Qo, the velocity Cai of the incoming liquid becomes so great that
recirculation vanishes. The magnitude of the flow rate Qo at which this occurs is
given by

(9-12)

(see Fig. 9-7).


Now let us repeat the same test, starting with an open valve. Let us also assume
that recirculation appears when the velocity of the liquid falls below the same value
Recirculation 87

Cai. In this case, the flow rate at which recirculation occurs is equal to

(9-13)

Now, because Rs is greater than Rl (see Fig. 9-7), Ql is greater than Qo.
There is still one important point to note: as pOinted out earlier [8, 16-18],
recirculation always acts toward increasing the developed head. Consequently, when
we start out with a closed valve and open it gradually, recirculation eventually stops
at a certain flow rate Qo. At that flow rate, the head drops suddenly owing to the
termination of recirculation. Conversely, when we start to close the valve from a wide-
open position, the head jumps suddenly at the flow rate of Ql (see Fig. 9-8).
Of course, in practice things are rarely so simple. The onset of recirculation is
affected by a great number of additional factors, such as the geometry of the pump,
position of the inlet edge [20], speed with which a given valve is closed or opened, and
many other factors.
Moreover, the differences between Rs and Rl are not always as large as those
illustrated in Fig. 9-7. In most cases, they are significantly smaller. Consequently, the
differences between Qo and Ql are rarely as large as those shown in Fig. 9-8. Very
often, this difference is so small that the test shows only a small dip (17) (Fig. 9-9).
Pumps seldom develop curves that look like Fig. 9-8). The only cases in which I have
observed such QH curves were in tests of sewage pumps with very wide passages and
inlet edges parallel to the axis.

Figure 9-8 Hysteresis in the QH curve owing to


inlet recirculation.

Figure 9-9 A dip in the QH curve caused by inlet


recirculation.
88 Performance Factors

'I
II
II
II
II
II

,,
I ,
I ,

.
I \
I ,
--../ ~ Figure 9-10 Sudden surge in the NPSH
requirements of a pump caused by recirculation
at the suction nozzle.

Sometimes the instabilities caused by the onset and disappearance of recirculation


were not noticeable from the shape of the QH curve, but did appear as a sudden
jump [19J in the NPSH requirements of the pump (Fig. 9-10). At present. little is known
about how recirculation produces a jump in the NPSH requirements. It seems to be
related to the fact that the action of the impeller on the surrounding liquid is of a
cyclical nature, caused by the finite number of impeller blades. Occasionally, the
frequency of such an action may produce resonance with another cyclic effect
occurring in the pump, resulting in a jump. In Chapter 20, we will examine such a
case.

RECIRCULATION AT THE IMPELLER OUTLET

Relatively little is known about the causes and effects of recirculation at the impeller
outlet. but what is known can help solve serious problems. Generally, one or more of
the following kinds of recirculation may exist at the impeller outlet:
1. Recirculation from the impeller outlet into an intermediate zone between the impeller
blades
2. Recirculation from the impeller outlet into the eye of the impeller
3. Recirculation from the discharge nozzle toward the impeller outlet

We have already seen that recirculation can increase the head developed by an
impeller (Fig. 9-3). We can increase the heads developed by a pump at low flow rates
by increasing the inlet recirculation with an inducer.
A similar effect can be achieved by forcing some of the liquid that has already been
acted on by the blades and entered the casing to return to the impeller, where it is
acted on for a second time.
In practice, this is accomplished either by opening slots in suitable locations of the
impeller shrouds (Fig. 9-11) or by removing parts of the shrouds (Fig. 9-12).
Figure 9-11 demonstrates how slots opened at a proper location of the shrouds can
increase the heads developed by a pump. Such an increase occurs only when the
liquid returns to a region within the passage between the impeller blades from which
it can receive an additional amount of energy. If the size and location of these
recirculation slots are not chosen correctly, it is likely that the recirculating liquid
will return to the impeller inlet and lose the total amount of energy supplied to it
during its first passage through the impeller. Figure 9-13 illustrates such a case.
Recirculation 89

Plain impeller
200 - - - With slots S
0- - - - - 0 Calculated
190

180

00170
ID
l160
J:
-u 150
Cl!
Q)
I 140

130

120

110

100
Figure 9-11 The effect of slots Son 0 10 20 30 40 50 60 70 80 90 100 110 120
performance. Flow rate (m 3/hr)

Figure 9-12 Impellers with parts of their shrouds removed to increase the shut-off head.

A series of slots A has been opened in the shrouds in the vicinity of the leading
faces of the blades between the radii Ra and Rb. As can be seen from the illustration,
the pump has shown a significant increase in the developed heads at all flow rates.
Next, the slots were closed with epoxy, and identical slots B were opened at the
trailing faces of the blades between the same radii Ra and Rb. This time, the pump
has shown extreme reductions in head over its total working range.
The reason for this great difference in performance can be found in our discussion
of Eq. 6-19. We showed that the pressure on the trailing faces of the blades is smaller
than at their leading faces. When the recirculated liqUid returns above a certain
90 Performance Factors

------ ~

I '- /"~~
~----
-~ ,,~

~
Ra /

I Rb

-~~-- ______ ' ----------

1.3

--
1.2
~_-----.::: - _ A
C 1.1 - - - - - - - _ _ _ Plain if/) - _
:s:l1.0
Q) ---
- ... _ Pel/er ...... .......
15 ............"
8 0.9 B ............ '"
-g 0.8 ...............
"'~"
,
ID
I 0.7

0.6 "
0.5 '----_---'-_ _---'---_ _--"--_ _L--_---'-_ _--"--_ _L · _
o 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Figure 9-13 The effect of slot
Q/Qp location on performance.

radius of the leading faces of the blades, the pressure at this location inside the
impeller plus the action of the blades proper prevent the recirculated liquid from
returning to the eye of the impeller. Instead, the liquid is acted on by the blades for a
second time, thus increasing the total head developed by the pump.
The pressures are significantly lower along the trailing faces of the blades (Fig. 6-2).
Consequently, the recirculated liquid finds little resistance to its flow toward the
impeller inlet. On its way toward the impeller eye, the liquid dissipates all the energy
it acquired during its first passage through the impeller. This, in turn, causes a
significant reduction in the heads developed by the pump.

Recirculation between the Discharge Nozzle


and the Impeller Outlet
The recirculation between the discharge nozzle and the impeller outlets is
commonly associated with the droop in the QH curve. Tests performed with a single
impeller in two identical casings [15] with different discharge nozzles (Fig. 9-14) seem
to indicate that such an opinion is justified. However, the problem is more
complicated than meets the eye.
Recirculation 91

---- .....
1.1

1.0

0.9

15.
:to 0.8
J:

0.7

0.6

0.5

Figure 9-14 The effect of discharge o 0.2 0.4 0.6 0.8 1.0 1.2 1.4
nozzle geometry on performance. Q/Qopt

In Ref. 24, Guiton reports that the magnitude of the shut-off head is often
determined by factors such as the layout of the piping system or the manner in
which the flow has been stopped. According to Guiton, "The zero flow head Ho
depends on how the flow is stopped: whether there is a valve closure or not, where is
the valve, and where are the delivery pressure measurements made between the
pump and this valve."
These observations have been confirmed indirectly by my own experience as
illustrated by the following examples.
A standard 2-in. production pump, that has been on the market for over 12 years
and tested on many occasions, frequently exhibited a drooping curve (Fig. 9-15). On
many other occasions, however, identical pumps have developed steadily rising
curves (Fig. 9-16) (for a wide range of impeller cut-downs).
The results shown in Figs. 9-15 and 9-16 are from tests performed on identical
pumps with different impeller cuts over many years. Indirectly, they confirm the
claims made by Guiton.
Guiton's observations and the results of the tests performed on the 2-in. pumps
mentioned earlier are not the only reasons to look for additional causes of a drooping
curve. Here is a typical case history.
92 Performance Factors

1.2 ~__6;;.;..;;,5..;;.in;.;... ..;;;D.;.;;ia~__

ii1.0.~
::r:.° 0.8
l:
0.6 ~_-;::-:--:::::--_~
5 in. Dia
0.4

0.2

OL-__~~__~____~____- L____- L____- L___ Figure 9-15 Test results of 11


o 0.2 0.4 0.6 0.8 1.0 1.2 identical impellers with different
Q/Qopt impeller cut-downs.

6.5 in. Dia


1.2

1.0

ii 0.8
::r:.0
l:
0.6

0.4

0.2

0 Figure 9-16 Test results of differ-


0 0.2 0.4 0.6 0.8 1.0 1.2 ent units of pumps identical to the
Q/Qopt pumps represented in Fig. 9-15.

I once observed that a certain pump usually exhibited a drooping curve, whereas
another pump using the same impeller usually produced a steadily rising curve.
The casings for both pumps were designed for the same head and flow rate, but the
base circle of the volute of one pump was about 6% smaller than that of the second
pump.
Hoping to find the reason for this difference in performance, I modeled both
casings in Plexiglas® and tried to observe the flow in the volutes with the aid of a
stroboscopic light.
The results were rather disappointing. Each time either of these pumps was
stopped and started again, a different flow pattern appeared in the volute. Also, the
head developed at shut-off was different after each new restart.
Recirculation 93

To make things worse, this effect seemed to be time-dependent. Whenever one of


the pumps was allowed to operate against a closed valve for about 5 to 10 minutes,
several changes could be observed in the flow pattern. With each change, the head
that developed at zero flow also changed.
Despite such inconsistencies in the test results, the following general picture
emerged. One casing usually developed a drooping curve, whereas the second usually
produced a steadily rising curve. There also seemed to be some sort of a relationship
between the flow pattern observed in the casing and the head developed at zero flow.
In the casing with a drooping curve and also when the other casing produced a
drooping curve, the flow pattern at shut-off was similar to that shown in Fig. 9-17.
The liqUid went a short distance into the discharge nozzle and then returned to the
casing and impeller.
A steadily rising curve seemed to be associated with a flow pattern such as the one
shown schematically in Fig. 9-18. After passing the (transparent) discharge nozzle,
the liquid entered the (opaque) discharge pipe. From there it returned to the volute,
without any clear sign of returning to the impeller.
The schematic sketches presented in Figs. 9-17 and 9-18 may hold a clue as to
why one casing usually produced steadily rising curves, and the other drooping
curves. These sketches may also hold the key to an explanation of the effects
reported by Guiton [24].
A glance at Fig. 9-17 reveals that the liqUid returning from the discharge nozzle
enters the impeller near the trailing face of the blade. Against this, in Fig. 9-18, the
liquid seems to approach the impeller near the leading face of the blade. (These
observations were made under stroboscopic light, and therefore it was impossible to
determine the true nature of the back-flow. It seems, however, that the back-flow was
not continuous, but cyclical).

Oscillations
Transverse

Figure 9-17 Flow pattern associated with a


drooping curve.
94 Performance Factors

Oscillations

Transverse
circulation

\
T Figure 9-18 Flow pattern associated with a
steadily rising curve.

The possible implication of the difference in the pattern of recirculation can be


explained by the pressure distribution along a blade at shut-off (Fig. 6-2). In that
case, the pressure coefficient Cp at the outlet of the leading face is well over 0.8,
whereas at the outlet of the trailing face, it is less than 0.65.
Now assume that a particle of liquid has left the impeller with a Cp value of 0.8
and, after returning from the discharge nozzle, it approaches the impeller with a Cp
value of, say, 0.7. If that particle approaches the impeller in the vicinity of the leading
face, where the pressure coefficient is well over 0.8, it will be expelled outward. At the
same time, its energy will increase owing to the repeated action of the blade. However,
if the same particle returns to the blade in the vicinity of its trailing face, where the
pressure coefficient is less than 0.65, the particle will enter the impeller. As the
pressure along the trailing face of the blade decreases steadily toward the center of
the impeller, the particle is continuously accelerated toward the center.
Consequently, once the liquid begins to move inward, it continues to move along the
trailing face of the blade until it reaches the entrance of the impeller vanes. During
this process, it loses most of its energy, which is converted into heat. It is reasonable
to conclude that such a dissipation of energy produces lower readings on the
pressure gage, and exhibits a droop in the QH curve. Indirectly this explanation is
also supported by the results of the tests presented in Fig. 9-13.
The hypothesis presented above explains some of the phenomena reported by
Guiton [241. The location of the re-entering liqUid in relation to the outlet tip of a
blade depends on the position of the impeller when the flow of the liqUid has been
shut off completely. Consequently, the manner and timing of a complete shut-off of
the flow determines the magnitude of the shut off head.
Recirculation 95

The position of the re-entering liquid relative to the blades also depends on whether
the frequency of this cyclic recirculation is a multiple of the passing frequency of the
blades in relation to the volute tongue. This can explain the time-dependency of the
shut-off readings that I have observed.
At this writing, I know of no experimental data that would directly confirm or deny
the hypothesis presented above. Indirectly, however, it is confirmed by the test data
presented in Fig. 9-13. These data illustrate unequivocally how the peripheral
location at which the recirculated liqUid returns to the impeller affects the developed
heads.
CHAPTER 10
Axial and Radial Thrust
and Balancing

When a liquid is acted on by an impeller, its head increases by an amount equal to


H2. This head consists of the static pressure head Hp2 and the velocity head C~/29,
where C2 is the absolute velocity of the liquid at the impeller outlet.
The pressure P2 at the impeller outlet is equal to -yHp2, where is the specific weight
of the liquid. This pressure P2 of the liquid acts in all directions, including the space
between the impeller shrouds and the casing. However, due to the rotational motion
of the liquid contained within that space, the pressure acting on the shrouds varies
with the radius.
For practical purposes, we can assume that the liquid between the shrouds and
the casing rotates with half the angular speed of the impeller, and therefore the
pressure distribution on the shrouds of a closed impeller can be approximated by
the equation

(10-1)

where U2 is the peripheral speed of the impeller outlet and U=wR is the peripheral
speed at any other radius.
As a result of the effect of the rotation of the liquid contained between the shrouds
and the casing, the pressure distribution on the impeller shrouds can be represented
as shown in fig. 10-1.
On the hub-side of the impeller, the pressure of the liquid acts on an area that
extends from the outer radius R2 of the impeller down to the radius Rs of the shaft.
On the suction side, the pressure P acts on an area that extends only down to
the radius Rw of the wearing ring. As a result, there remains an unbalanced force on
the hub side of the impeller owing to the pressure the liqUid exerts on the shroud
between the radii Rw and Rs. The magnitude of this force is equal to

Rw
T = 2rt f PR dR .
JR, (10-2)

97
98 Performance Factors

Figure 10-1 Pressure distribution on


the shrouds of a closed impeller.

AXIAL THRUST ON A SEMI-OPEN IMPELLER


The axial thrust exerted on the hub side of a semiopen impeller is exactly the same
as that of a closed impeller (Fig. 10-2). In a semi-open impeller, however, this thrust
is counteracted by a significantly smaller force.
In a closed-type impeller, the axial forces acting on the outer radii of the shrouds
between the radii R2 and Rw are equal but opposite on each side of the impeller. In a
semi-open impeller, the axial forces remain unbalanced up to the outer radius R2 of
the shrouds. On the hub side of the shrouds, the pressure distribution follows
in the manner expressed by Eq. 10-1. On the suction side, however, the pressure
distribution takes the form of curve a in Fig. 10-2.
The shape of curve a depends largely on the design of the impeller waterways. In
theory, the shape could be calculated from the head-rise of the liquid within the
impeller passages using Eq. 1-7. In reality, the pressure within these passages is not
constant at a given radius, but varies along any given arc between two consecutive
blades (compare Eq. 1-9). It therefore seems more realistic to assume that the
pressure within the passages of the vanes increases either directly as the radius or as
the square of the radius. I have tried both approaches. Neither has produced a
perfect balance, although the results were adequate to prevent overload of the
bearings. (Chapter 29 presents an approach that can lead to more accurate calcul-
ations of the axial loads on semi-open impellers.)
Axial and Radial Thrust and Balancing 99

Figure 10-2 Distribution of


axial pressures on a semi-
open impeller.

BALANCING THE AXIAL THRUST


The most straightforward way of compensating for the axial thrust is to take it up by
a thrust bearing. In most cases, it is also the most efficient way. In some cases,
however, it increases the cost of the pump because of the need for a heavier bearing
frame, and so forth. In other cases, the axial thrust may turn out to be so large that
it can be balanced only with specially designed bearings. This may increase the cost
of the pump by so much that it becomes noncompetitive. Thus, alternative means of
balancing the axial thrust are frequently used.

Balancing the Axial Thrust in a


Single-Stage Pump
The most common way to balance a single-stage impeller is to provide its back
shroud with a second wearing ring W2 (Fig. 10-3) and to connect the space C inside
that wearing ring with the impeller eye by means of the balancing holes B. This
procedure is based on the assumption that the holes B will equalize the pressure
within the space C and within the impeller eye. Consequently, if we make the radius
of the wearing ring W2 equal to the radius of the wearing ring on the suction side, we
achieve a perfect balance.
100 Performance Factors

Figure 10-3 Balancing the axial thrust by means of a back


wearing ring W2 and balancing holes B.

In reality, however, this is not completely correct. Unavoidable clearances between


the stationary and the rotating surfaces of the wearing rings allow a flow of liquid
from the pressure side of these rings toward the eye of the impeller. On the suction
end of the impeller, this flow is direct. On the hub side, however, the liquid first
passes from the pressure side into the space C and from there into the eye of the
impeller, via the holes B. Within space C, the pressure must be higher than in the
impeller eye, owing to the pressure drop through holes B. However, it is usually
possible to keep this pressure difference within acceptable limits by choosing a
suitable size and number of balancing holes.
The situation with semi-open impellers is different. Here, the axial forces on each
side of the shroud are quite different. To compensate, the radius Rw of the wearing
ring (fig. 10-4) must be large enough as such that the axial force that results from the
pressure of the liquid on the remaining annular surface extending between the radii
R2 and Rw of the hub side of the shroud, is opposite and equal to the axial force
exerted on the suction side of the impeller between the radii R2 and Rl (Fig. 10-4).

Figure 10-4 Balancing a semi-open


impeller by means of a wearing ring.
Axial and Radial Thrust and Balancing 101

Figure 10-5 Balancing axial thrust by


means of back vanes B.

Here. however. arises a problem of a different kind. Pumps are usually sold with
different cut-downs of the impeller diameter. If the wearing ring of a semi-open
impeller has been designed so that the axial thrust is balanced in a pump with a full-
diameter impeller. there will be a significant axial imbalance when the outer diameter
of the impeller is cut down. For example. when the outer diameter of the impeller is
cut down to the radius Rw of the wearing ring. the axial thrust may act in the
direction away from the impeller eye. depending on the magnitude of the residual
pressure within the space inside the wearing ring of radius Rw and the pressure
distribution in the space between the blades.
One way to prevent such an occurrence is to balance the axial load on a full-
diameter impeller but not completely. leaveing some margin for the cases of cut-down
impellers.
Another way to reduce the axial thrust on a impeller is to provide it with back-
vanes (Fig. 10-5). Such a means [25. 26] is particularly popular in pumps that handle
slurries or abrasive matter because it also provides a way to keep the solids away
from the stuffing box. Also. in semi-open impellers. back-vanes make it easier to
maintain the balance of axial forces acting on both sides of the impeller over a wider
range of impeller cut-downs. Such back-vanes. however. usually increase the power
consumption of the pump, thus reducing its overall effiCiency.
Finally. the axial thrust in a Single-stage pump can be avoided by designing the
pump with a double-suction impeller (fig. 3-6). In theory. such a design provides a
perfect balance. In practice. however, the axial balance may be seriously distorted by
inaccuracies in manufacturing and improper layout of the suction line, or both.

Balancing the Axial Thrust in Multistage Pumps


One way to balance the axial thrust in a multistage pump is to balance each
impeller individually by providing individual balancing rings (Fig. 3-8). Another way
is to arrange the impellers in opposing directions, as shown in Fig. 10-6. With an
even number of impellers, such an arrangement can eliminate the axial thrust
completely. Such an arrangement. however, requires complicated passages from one
stage to another. This increases the hydraulic losses and requires very complicated
102 Performance Factors

castings, significantly increasing the cost of producing good-quality casings, when


large numbers of stages are involved.
Another drawback of such an arrangement is that it requires right-handed and left-
handed impellers that, externally, look identical. This may often lead to very costly
errors during assembly or re-assembly of a pump.

Figure 10-6 Balancing


axial thrust by means of
DISCHARGE opposed impellers.

To discharge line

Space connected to suction n-.:;:;;;..._--__a;;;;;; To suction nozzle


- - - - -.......

.. Impeller eye

Figure 10-7 Balancing


axial thrust by means of
a balancing drum.
Axial and Radial Thrust and Balancing 103

To suction nozzle

Balancing disk

Figure 10-8 Balancing axial thrust by


means of balancing disk.

Another way to balance the axial thrust is to use a balancing drum (Fig. 10-7) or a
balancing disk (Fig. 10-8). In such a case, the axial thrust is being taken up by a
single drum or disc. This device is subjected to the total pressure developed by the
pump on one of its faces. On its other face, it is subjected to the suction pressure at
the inlet of the first impeller.

BALANCING AXIAL LOAD WITH INDIVIDUAL BALANCING RINGS


VERSUSS A SINGLE BALANCING DEVICE

Adding a wearing ring to each impeller (Fig. 3-8), increases the losses due to the
leakage through the additional gap between the additional wearing rings. From
Eq. 7-7, the amount of leakage QL through such a gap can be expressed as

QL= KA[2gHdlo. 5 (10-3)


104 Performance Factors

where K = a coefficient whose magnitude depends on the ratio of the (length of


gap)/(width of gap), and so forth
A = through-flow area of the gap
Hd = difference of the pressure head on both ends of the gap

The total additional loss of power owing to this leakage is equal to n times the
additional loss due to leakage through one stage:
L=-y(nH)QL (10-4)

where n = number of stages


H = head developed by one stage

Now assume that instead of balancing each impeller individually, we are using a
single balancing drum of the same diameter and width of gap as the back wearing
rings of the individually balanced impellers. In that case, the leakage through this
balancing drum is equal to

QL t = KAI2gnH d ]o.5
(10-5)
= QLlnf5.

This means that the additional power lost from leakage through the balancing drum
is now equal to (nHJQL t, or
Lc=-y(nH)QL[n]O.5 (10-6)

This means that the loss of power caused by the leakage through the balancing drum
will be [njO.5 times greater than when we balance each impeller individually.
There are, however, certain advantages to using a single balancing device. It is
easier to access one mounted at the end of the pump than to access each impeller
individually. In the latter case, the entire pump must be dismantled to repair or
exchange even a single balancing ring. This, of course, significantly increases the cost
of maintenance and repair.

RADIAL THRUST

In addition to the axial thrust, the pumped liquid also exerts a thrust on the impeller
in the direction perpendicular to the axis of the pump. The magnitude of that thrust
depends, of course, on the size of the pump and the head it develops. Its magnitude
also depends on many additional parameters, such as the design features of the
casing and the ratio of the flow rate Q of the pumped liquid to the flow rate Qd, for
which the casing has been designed to deliver at the given head.
If the casing consists of a single volute of conventional design, the radial thrust is
smallest when the flow rate ratio Q / Qd = 1 and it increases with increased deviation
(Fig. 10-9).
At flow rate ratios of Q / Qd < 1, the direction of the force is usually somewhere
between 30 degrees and 100 degrees downstream of the volute tongue, as indicated
by the arrow A in Fig. 10-10. At flow rate ratios greater than one the force is usually
in the opposite direction.
Axial and Radial Thrust and Balancing 105

600

500
en
g400
ti
£ 300
:::!

(ij
'g 200
a:
100
Figure 10-9 Radial thrust on a
6x4 in. pump impeller at different 100 200 300 400 500 600 700 800 900 1000 1100 1200
flow rate [33]. GPM

Figure 10-10 Unmodified casing used by Worster [28] in his tests {(Xv is the angle between the volute
curve and the impeller periphery).

In practice, centrifugal pumps are seldom used for flow-rate ratios higher than
Q/Qd=l. In such cases, as can be concluded from Fig. 10-9, the radial thrust
achieves its maximum value at shutoff. As mentioned earlier, the magnitude of the
radial thrust can vary significantly with the geometry of the casing. Consequently, to
be able to study the qualitative effects of various design parameters on the magnitude
of the radial thrust, the pump industry has agreed to relate its magnitude to the
projected area (B2~) of the impeller outlet (Fig. 10-11) and to the head Ho developed
by the pump at shut-off. In the United States, the most widely accepted equation for
expressing the radial thrust is
106 Performance Factors

,-
1.8 sin U v ,-'-
0.40
,- ,- K
,-
,-
,-
0.35 ,-
/.

0.30

0.25

K 0.20

0.15

0.10
Figure 10-11 Comparison
between the k values pub-
0.05
lished by the Hydraulic
Institute in 1965 and the
500 1000 1500 2000 2500 3000 values according to Eq.
Specific speed 10-8.

(10-7)

where To = total thrust at shut-off, in pounds


B2 = outer width of impeller, in feet (Fig. 10-11)
~ = outer diameter of impeller, in feet
K = a coefficient specific for a given design
Ho = total head at shut -off, in feet
S = specific gravity of the liquid

Effects of Casing Geometry on Radial Thrust


Experience teaches us that the magnitude of the coefficient K increases with
specific speed. In other words, it increases with the ratio Qo.5jHo. 7 5. The effect of an
increase in that ratio on the geometry of a conventional volute, is an increase of the
volute angle (Xv (Fig. 10-10) and can be expressed [33] as

K=L sin (Xv (10-8)

where L is a coefficient whose magnitude depends on a series of geometrical features


of the pump.
Stepanoff [27] has compiled a graph that represents the general trend of the
manner in which the magnitude of (Xv increases with the specific speed. Using
Stepanoffs data, I have superimposed the K values calculated according to Eq. 10-8
over the experimental data published by the Hydraulic Institute in 1965. As can be
Axial and Radial Thrust and Balancing 107

seen in Fig. 10-11, there is a fairly good agreement between the calculated data and
the data recommended by the Hydraulic Institute, up to a specific speed of about
2000.
However, the problem is that the radial thrust depends on so many other geometrical
features of the casing that the data presented in Fig. 10-11 can serve only as a rough
gUide of what can be expected from a given design.
The effects of casing geometry on the radial thrust are well illustrated by the tests
performed by Worster [28]: after measuring the radial thrust on an impeller operating
in a conventional volute casing (Fig. 10-10), two rings r have been installed in the
casing to reduce the axial clearances on both sides of the impeller (Fig. 10-12). Sub-
sequent tests have demonstrated that this change increases the radial thrust three to
four times over the thrust in the unmodified casing.
Next, material was added to the tongue to reduce the distance between the volute
tongue and the impeller (Fig. 10-13). this modification increased the radial thrust by
more than 10 times.
These tests indicate that it is possible to reduced the radial thrust by simply
allowing greater clearances between the impeller and the casing tongue.
Much can be learned about the effects of casing geometry and radial clearances on
the radial thrust from the tests performed by Agostinelli and co-workers [29]. The
radial thrust developed on a given impeller at a given operating speed was tested with

Figure 10-12 Casing with reduced axial


clearance.
108 Performance Factors

Figure 10-13 Casing with reduced


radial clearance.

four different casings. One was a conventional volute, and the other three were
concentric circular casings of the same width but different ratios of casing diameter
Dc to impeller diameter Di.
Figure 10-14 shows the results of these tests. First, it confirms the conclusion
derived from Eq. 10-8, that a concentric casing (for which the volute angle is equal to
zero) produces a smaller radial thrust than a conventional volute. It also reveals a
very interesting aspect of the effects of the radial distance between the impeller rim
and the casing.
With a circular casing, when the ratio of the casing diameter Dc to the impeller
diameter Di was equal to DclDi= 1.145, the radial load at shut-off was approximately
68% of the radial load generated in a conventional volute. However, the variations of
radial thrust with changes in flow rate followed a pattern that was not much different
from the pattern in the case of the conventional volute.
When the ratio DclDi was increased to 1.290, the radial thrust at shut-off dropped
to about 25% of that developed with the conventional volute. Moreover, the curve
representing the changes of the radial thrust with flow-rate ratio has assumed a
completely different shape from that in the case of a conventional volute or of a
circular casing with a diameter ratio of 1.145. A further increase of the ratio of Del Di
from 1.290 to 1.430 has shown relatively minor changes in the magnitudes and
distribution of the radial loads. Being intimately familiar with the test equipment and
methods used in calibrating the instruments used during these tests, I have reason
to believe that the difference in the test results of the last two casings may have been
caused by inaccuracies in testing.
The results of tests with the three circular casings lead to the conclusion that the
radial clearance between the impeller and casing can affect the radial load only up to
a certain point.
Finally, here is an additional example of how the casing geometry can affect the
radial load. Tests on a newly developed pump revealed that it was developing radial
loads that were too large for the bearing frame on which the water end was destined
to be mounted. To reduce this load, the casing was remachined, as shown in
Fig. 10-15. This made about 20% of the periphery of the casing adjacent to the volute
Axial and Radial Thrust and Balancing 109

100

80

60 Concentric casing, DcID; = 1.290

Concentric casing, Del 0; = 1.430


40

20

20 40 60 80 100 120 140 160


Capacity, percent of normal

Figure 10-14 Radial thrust on the same impeller tested in four different casings [29].

Figure 10-15 Remachined casing


for reduction of radial thrust.
110 Performance Factors

tongue concentric with the impeller. Subsequent thrust tests have shown that this
change reduced the radial loads by about 60%.

Other Alternatives for Reducing Radial Thrust


A common means of avoiding excessive radial loads is to subdivide the casing into
two volute-shaped passages, with their tongues located 180 degrees apart. Each
volute is designed for half the flow rate delivered by the pump and their output is
combined somewhere inside or at the end of the discharge nozzle. This is called a
double volute.
In theory, such a casing should provide a perfect balance at all flow rates. In
practice, however, the designer must always incorporate certain unavoidable
differences into the shape of each volute or length of diffuser passages. As a result,
double volutes usually develop residual radial thrusts of between 10% and 20% of
the thrust developed by a single volute.
CHAPTER 11
Miscellaneous Factors
that Affect Pump
Performance
CAUSES OF A DROOPING CURVE

At present. two major causes of a drooping curve are recognized. One was discussed
in Chapter 9 in connection with Figs. 9-17 and 9-18 and supported by the test results
presented in Figs. 9-15 and 9-16. The second factor that can cause a drooping curve
usually appears when a pump is designed for a high head coefficient and is due to
the hydraulic losses associated with the flow of the pumped liqUid through the
waterways of the pump.
Figure 11-1 presents a typical efficiency curve. Had there been no losses. the
efficiency curve of the pump would have been 100% at all flow rates. represented by
line A-B. However. owing to a combination of different losses (shown in Fig. 11-1 by
the thin vertical lines). the effiCiency curve assumes the shape presented in Fig. 11-1.
The losses that are generating the shown effiCiency curve consist of mechanical
losses. such as friction in the bearings and stuffing box; leakage losses; losses due to
disk friction; and head losses owing to resistance within the wetted passages of the
pump. Of all the losses listed above. only the head losses are significantly affected by
the flow rate. Consequently. the shape of the efficiency curve shown in Fig. 11-1
indicates that the losses of head vary with flow rate according to a curve similar to
the one shown in Fig. 11-2. Keeping this in mind, we can now see how a high head
coefficient can generate a drooping curve.
Consider two impellers of identical outer dimensions that have been designed to
deliver the same flow rate Qd at the same speed. However, at the design flow rate Qd,
the pump designed for the high head coefficient will develop a theoretical head equal
to Hdl (Fig. 11-3) whereas the pump designed for the low head coefficient will develop
a theoretical head equal to Hd2 (Fig. 11-4).
For ease of explanation, we are making the assumption that the theoretical QH
curves of each of the impellers can be expressed by a straight line (Figs. 11-3 and
11-4). (This assumption is based on the already obsolete theory. that the outlet angle
is the major deSign-parameter that determines the magnitude of the developed head.
This topic is discussed in detail in Chapter 29.) We are also assuming that the losses
111
112 Performance Factors

100 A
- t
'Q)E
r B
Overall
e >-
(,)
Q) losses
s c::
Q)

~
>- '0
==
(,)
c::
Q) W
'0
==
w

Figure 11-1 Overall losses and the


Flow rate ---- efficiency curve.

t
'C
ctl
Q)
.s:::.
'0
C/)
Q)
C/)

~
Figure 11-2 Variations in head losses
Flow rate ---- with the flow rate.

Theoretical QH line
-- -- -- -~-~- "'1---__
HL

Figure 11-3 The effect of head losses


HL on a pump designed for a high head
1-----Qd-----+l·1 coefficient.
Miscellaneous Factors That Affect Pump Performance 113

Figure 11-4 The effect of head losses


HL on a pump designed for a low head
coefficient. 1--- - - - Q d - - - - - - . - I

of head at any given flow rate are identical in both impellers. If we subtract from the
theoretical QH lines the losses of head (which we have assumed to be identical in
both impellers for any given flow rate), we observe the following. The pump designed
for the high head coefficient produces a drooping curve, whereas the pump designed
for a low head coefficient develops a steadily rising curve. While a pump designed for
a low head coefficient is more likely to develop a steadily rising curve, it does not
guarantee that it will always produce such a curve. In our discussion related to Figs.
9-17 and 9-18, we saw that other factors can produce a droop in the QH curve. In
addition, a pump designed for a lower head coefficient is a physically larger, heavier,
and more costly unit that will also require more floor space. Very often, a pump
designed for a lower head coefficient is also less efficient because of increased losses
owing to disk friction.

CAUSES OF A DIP IN THE QH CURVE

We have already discussed one of the causes of a dip in the QH curve (Fig. 9-9). The
second case in which such a dip is frequently encountered usually happens in
pumps in which the outlet edge of the impeller blades is not parallel to the axis.
Figure 11-5 presents such an impeller schematically. The impeller can be regarded
as being composed of an infinite number of partial impellers of width flB, all of which
are operating in parallel.
There are various ways to design such an impeller. In the most popular method, all
partial impellers are designed to develop the same head Hd at the design flow rate Qd
(Fig. 11-6). Because of the differences in the outer diameter, each partial impeller
develops a different QH curve. All the curves intersect at point I, where Q= Qd.
The partial impellers that terminate at the outer radii of the discharge edge develop
steadily rising curves because of the low head coefficient, for which they have been
designed. This is shown schematically in Fig. 11-6 by the curve A-I-A. However, the
partial impellers that terminate at the inner radii of the outlet edge may develop
drooping curves because of their relatively high head coefficient. This is shown in
Fig. 11-6 by the curve B-I-B.
114 Performance Factors

Figure 11-5 Partial impellers of width £lB.

T
r I--Qd~1
A B c
Flow rate
Figure 11-6 The resultant QH curve generated
by parallel operation of two partial impellers of
different outer diameters that are designed for the
same head and flow rate.

When two partial impellers are operating in parallel, they produce the resultant
QH-curve A-I-C, in which the flow rate at any given head is equal to the sum of the
flow rates delivered by each of the individual partial impellers when operating against
the same head. This produced the dip shown in Fig. 11-6 in the vicinity of the flow
rate Qd.
PART
Problems
Encountered with
Centrifugal Pumps

The manner in which a pumping system is laid out and the qualifications of
the people active in the production, installation, testing, and maintenance
of the pumping units affect pump operation. This section outlines a series
of do's and don'ts and discusses a great variety of problems encountered in
practice. Finally, it presents a few particularly difficult case histories that I
have been confronted with in the past.
CHAPTER 12
Testing
In Chapter 2 we discussed the performance characteristics of centrifugal pumps and
their importance for selecting the proper pump for a given duty. These characteristics
are usually determined on the basis of tests performed by the manufacturer during
the development stage of a given unit. However. to be of real value to a pump user.
such tests must be carried out correctly. Otherwise. the pump user my be confronted
with very unpleasant surprises.
The example given above is not the only reason that correct test results are needed.
Pumps are frequently being tested on site to determine whether they really satisfY
the expected conditions. Serious deviations of these results from the rated perfor-
mance curves usually indicate that something is wrong either with the pump or the
pumping system. or both. Such findings usually precipitate (often very) costly and
time-consuming searches for their causes and for means of correcting any possible
malfunction.
False test results may point to problems that do not exist or may redirect the
search for solutions in the wrong direction. Finally. as will be discussed in Part V.
situations may arise in which there is a need to modifY an existing pump in order to
adapt it to new. or modified operating conditions. False tests results may indicate
that the adopted modifications are correct while. in reality. they are faulty. or vice
versa. These are only a few of the examples that demonstrate the importance of
following correct test procedures.
In theory. we can distinguish two classes of problems:

1. Problems related to the hydraulic performance of a pump. such as differences between


the rated and the tested performance curve, cavitation at higher-than-specified NPSH
reqUirements, and higher-than-specified power consumption (Le .. lower-than-specified
efficiencies)
2. Problems of a mechanical nature, such as operating with noise and/or vibrations,
overheating of bearings or other parts of the pump, and frequent breakdowns of various
pump parts

In reality. all these effects are so intimately entwined that it is often impossible to
distinguish between these two classes of problems.
Sometimes a mechanical fault initiates a chain reaction. Examples include
misalignment of the pump and the driver, pipelines imposing stress on the casing. or
lack of balance of rotating parts because solid matter is clogging one of the impeller
117
118 Problems Encountered with Centrifugal Pumps

passages. In each case, the unit vibrates and stationary and rotating parts often rub
together. Rubbing may increase the clearances of the wearing rings, resulting in
increased leakage losses. Also, rubbing causes an increase in power consumption,
thus reducing efficiency even more. In addition, rubbing parts wetted by the pumped
liquid may cause cavitation owing to higher temperatures.
Vibrations may also loosen the bolts between a pumping unit and foundations and
the bolts that hold different parts of the pump together. This, in turn, aggravates the
misalignment of different parts, giving rise to other disturbances.
An analogous chain reaction is initiated by factors that cause heating of pump
parts, such as incorrectly mounted or inadequately lubricated bearings, packing the
stuffing box, a mechanical seal that is too tight. insufficient lubricant or insufficient
cooling of stuffing box or seal, incorrect type of packing or mechanical seal, excessive
grease in the bearings, and incorrect lubricants.
Any of these faults may cause excessive heating of certain parts of the pumping
unit. This may destroy the lubricant, causing still more heating, with resultant
seizing and scorching of different parts. In most cases, this reduces the life of the
affected parts.
Excessive heating of certain pump parts also causes uneven expansion. This may
ruin the alignment of the different parts. This, in turn gives rise to vibration, uneven
wear of parts, and so on. The faults owing to improper installation, assembly and
machining are so numerous and diverse that it is impossible to make a complete list.
Often, only an in-depth knowI?rlge and understanding of the factors that affect pump
performance can help solve a particular problem. To make matters worse, a pump
may be thought to operate unsatisfactorily while, in reality, there is nothing wrong
with its performance. Such cases are more common than one might expect. These
cases are the result of problems associated with pump testing.

TESTING AN~ TEST PROCEDURES

The manner in which a centrifugal pump must be tested is established by standards


issued by several official and semi-official institutions (e.g., the Hydraulic Institute).
However, none of these standards guarantee that a test will be performed correctly.
The factors affecting the outcome of a performance test are simply too numerous to
be covered fully by standards.
Although there are differences between the test codes established by different
authorities, all are based on certain common principles.
Power consumption is also recorded Simultaneously, making it possible to draw the
Q-HP curve for the given pump, as shown in Fig. 2-l. To calculate the efficiency at a
given flow rate, the power added to the pumped liqUid is divided by the power
consumed by the pump.
The definition of power is work performed in a unit of time. According to this
definition, one (U.S.) horsepower is equivalent to lifting 550 lb to an elevation of one
foot in one second (or lifting 110 lb/sec, to a height of 5 ft, etc.). In metric units, one
(metric) horsepower is equivalent to lifting 75 kg to a height of one meter in one
second.
This definition leads to the following equation for calculating the effiCiency (in U.S.
units):
Testing 119

EFF = QxHxS (12-1)


3,960 x HP

where Q = flow rate in gallons per minute


H = head in feet
S = specific weight
HP = horsepower consumed (in U.S. units)
When using metric units, the efficiency is given by the equation

EFF = QxHxS (12-2)


270xHP

where Q = flow rate in cubic meters per hour


H = head in meters
HP = horsepower in metric units

CONSEQUENCES OF INCORRECT TESTS

An incorrect test may do more harm than good. It may incorrectly indicate that a
pump performs unsatisfactorily. This usually causes significant loss of time and
money spent in attempts to locate the source of trouble that does not exist. Even if
there are certain faults in a pump, an incorrect test may make the real cause less
evident. This again will mean a waste of time and money.
In general, a pump operates as a part of a system. A malfunction in the hydraulic
performance of a pump manifests itself as a malfunction of the whole system. The
opposite, however, is not necessarily true. A malfunction of a system does not always
mean that a pump does not perform properly. The source of the trouble may lie in
another part of the system.
The most straightforward way to determine whether a pump is the cause of a
malfunction is to perform a test. Such a test, however, is of real significance only if it
is executed properly. Incorrect tests inevitably lead to false conclusions, which often
result in enormous expenses and lost time.
While there are differences between the test codes established by different
authorities, all are based on certain common principles. First, all measurements
should be made with reliable and accurate instruments. To ensure accuracy of
measurements, these instruments should be checked and calibrated at frequent
intervals. For example, some test codes require Bourdon pressure gages to be
checked for accuracy before and after each test.
Nevertheless, problems may still arise. It is important to be able to analyze tests
and determine alternative procedures to ensure reliability.

MEASUREMENTS OF POWER CONSUMPTION

Power input is the power consumed by the pump proper. In the case of a pump
directly coupled to an electric motor, this power is equal to the power output of the
120 Problems Encountered with Centrifugal Pumps

driver. That is, it is equal to the power input into the motor multiplied by the
efficiency of the driver.
Therefore, a reliable test requires that only calibrated motors should be used for
test purposes. In the absence of a calibrated motor, the motor manufacturer's
efficiency data for an equal motor are frequently used. In this case, however, the
accuracy of the power measurements cannot be greater than 3%.
To make things worse, even a calibrated motor cannot always be relied upon.
Different codes prescribe different methods of calibration. While some of these methods
produce consistent results, there have been cases in which the same motor, when
calibrated in accordance with different test codes, has demonstrated differences in
efficiency as high as to 3% [30].
In belt drives, the power lost owing to slippage and continuous bending of the belt
must be subtracted from the output of the driver. For the power lost due to bending
of the belt, a value of 0.25% to 0.50% of the total power consumption is often
assumed.
To calculate the power lost due to slippage, first measure the operating speed of
the driving pulley. Next, calculate the theoretical speed of the driven pulley by
multiplying the speed of the driving shaft by the ratio of the effective radius of
driving pulley to the effective radius of the driven pulley. For a flat belt, the
effective radius is equal to the sum of the radius of pulley and half of the belt
thickness. For a V-belt, the effective radius is determined from the belt manu-
facturer's instructions. Now measure the actual operating speed of the driven
pulley. This speed will be lower than the calculated theoretical speed as a result of
slippage. Finally, subtract the actual measured speed from the theoretical calcul-
ated speed. The percentage of power lost owing to slippage will be equal to the
percentage of speed lost.
Determining the amount of slippage may also help prevent certain failures
associated with belt drive. For a belt to be capable of transmitting power, it must be
pretensioned. This is usually accomplished either by providing the drive with a
special tension pulley or by installing the driver on slotted rails and fixing it in a
position that will assure that the belt is properly pretensioned.
Whenever a belt has been pretensioned by an excessive amount, two things may
occur. First, it may impose excessive load on the bearings, thus reducing bearing life.
Second, it may bend the shaft of the driver, or the pump, or both. This, in turn, may
lead to a whole series of problems.
On the other hand, a belt with too much shack will not only cause excessive loss of
energy, but will also lead to frequent belt breakage, from excessive wear and over-
heating.
Under normal tension, the slip in a flat belt should be about 3%, and in V-belts it
should be about 1%. Significant deviations from these numbers are indications of
improper belt tension.
In deep-well pumps with long columns, the power consumed by the lineshaft is
usually subtracted from the power output of the motor. The power consumed by the
thrust bearing of the motor is also subtracted from the power output. These
calculations are based on thrust data furnished by the pump manufacturer and the
power consumption data furnished by the motor manufacturer. In the case of a belt
drive via a geared head, the power loss in the gearbox and in the belt are also
subtracted from the motor output.
Testing 121

All these methods imply a certain amount of inaccuracy and are used mainly
during field tests. Only the pump directly coupled to a calibrated electric motor
ensures a certain degree of accuracy and is suitable for a certified field test. However,
errors can occur. The most common errors are dirty contacts on the watt meter, one
or more wires leading to the watt meter not connected tightly to the contacts, a
broken wire between the watt meter and the main circuit, and the watt meter
incorrectly connected to the power line.
Because of these failures, the watt meter is sometimes circumvented by inserting a
torque meter between the pump and the driver. However, failures may also occur
here. The most common occur when there is dust or dirt on the contacts that connect
the pressure sensor to the readout instruments. To eliminate this problem, the power
is sometimes determined by coupling the pump directly to a dynamometer.
In practice, a dynamometer consists of an electric motor mounted on ball bearings.
Thus, the motor frame can rotate freely under the influence of the torque generated
by the electric current. The motor frame is prevented from rotating by a system of
levers and cables connected to the stator and loaded with weights. To keep the stator
at rest, the magnitude of these weights must be chosen so that they exert a torque
opposite and equal to the torque exerted by the electric current.
The magnitude of the torque is calculated by multiplying the magnitude of these
weights by the distance of the point of their suspension from the center of the motor.
The product of this torque, multiplied by the speed of the motor, gives the exact
magnitude of the power supplied to the pump.
A dynamometer is regarded as the most reliable means of measuring power. Still,
inaccurate measurements can occur. The most common causes are dirt or inaccurate
lubrication of the bearings supporting the motor frame; inaccurately aligned or
damaged bearings; binding fulcrums of levers to which the weights are being
attached; the weight of the electric wires that are supplying current to the motor
exerting additional torque to the motor frame, and the stiffness of the electric cables
preventing the frame from rotating freely.

MEASUREMENTS OF THE FLOW RATE

The most accurate and reliable method of measuring the flow rate is by diverting,
instantaneously, the flow from the discharge pipe into a container of known
dimensions. The liquid is allowed to flow into this container for a given period of time,
then the flow is instantly diverted into a bypass. After the waves generated by the
liqUid have subsided, the volume of the liquid in the container is calculated. The
volume of liqUid in the container divided by the period of time it flowed into the
container gives the exact flow rate.
A container of large dimensions must be used for accurate results, or the time
required to divert the flow may introduce a significant error into the time measure-
ments. A large container, however, is very costly and can cause space problems. For
these reasons, the flow rate is most frequently measured by a flow meter. Several
kinds are on the market.
The flow meter should be mounted in a location that guarantees that the liqUid
flowing through it has a uniform velOCity distribution. This requires a certain minimum
length of straight piping ahead and downstream of the meter. The minimum length of
122 Problems Encountered with Centrifugal Pumps

this piping, and the frequency of calibration checks are usually specified in any valid
test code.
Still, these precautions do not eliminate the occurrence of faulty flow measure-
ments. The most common cause is solid matter entering the flow meter and
obstructing the flow. Another cause that is particularly common in closed circuits is
cavitation within the flow meter.
Special precautions must be undertaken when the flow rate is measured with a
venturi meter or orifice. In this case, the flow rate is usually determined indirectly by
means of a differential manometer. The tubing that leads from the flow meter to the
manometer requires special care. Any looseness or leakage will result in incorrect
results.
Similarly, when the flow meter is located in the suction line and the pressure there
is below atmospheric pressure, the tUbing should be full of air. Also no liquid should
enter the tubing during the test.
Finally, incorrect readings are often caused by dirt entering the tUbing and clogging
passages and connections. Special problems are encountered with measurements of
very high flow rates. Most of the instruments designed to measure such flow rates are
based on the fact that the amount of liquid flowing through an area of a given size is
directly proportional to the velocity of the liquid. Very high flow rates usually imply
very large cross-sectional areas, in which it is impossible to maintain a uniform
velocity. Consequently, the results of such measurements are rarely more accurate
than 3% to 5% .

MEASUREMENTS OF HEAD

Perhaps the most vulnerable of all measurements is the head developed by a


centrifugal pump. Often, it is impossible to measure this head directly, and therefore
indirect methods must be used.
The total head developed by a pump is the sum of the heads required by the liquid
to do the following things: overcome the resistance to the flow of the liquid in both
the suction and discharge lines, overcome the pressure differential on both sides of
the pump, overcome the static head against which the pump must deliver, and
compensate for the difference in the velocity heads.
To account for the resistance in the suction line, the suction pressure is usually
measured a short distance upstream of the pump. However, for accuracy, the flow of
the liqUid must be uniform at the pressure measurement locations. Often, this can
be achieved only if this location is preceded by a long section of straight piping of
uniform diameter or by a shorter section preceded by a flow straightener.
Even then, the results can be affected by air entering the suction line or by a low-
pressure core generated within the center of the suction line. Such a core may appear
whenever the suction tank is laid out incorrectly or when the suction line consists of
several consecutive elbows or bends. Even with a correctly laid out suction tank and
line, the pressure measurements will be correct only at flow rates that are not
significantly smaller than the design flow rate of the pump.
At very low flow rates, an interchange occurs between the liqUid that has already
entered the impeller and the liqUid approaching the pump via the suction pipe
(Chapter 9). Part of the liquid that was already acted on by the impeller vanes returns
Testing 123

into the suction pipe. There, it mixes with the incoming liquid and transfers to it part
of the energy it absorbed when being acted on by the impeller vanes. This exchange
of energy sets the incoming liquid into rotational motion and increases its total
energy contents at the periphery of the pipe's inside diameter. As a result, the
measurements in the suction line will be higher than the actual ones.
The total head developed by a pump is defined as the total amount of useful energy
transferred from the impeller blades to the liquid. In the absence of prerotation, this
total head equals the total head measured at the pump discharge, minus the total
head measured at the inlet to the pump. However, when prerotation is present, the
impeller also transfers useful energy to the liquid upstream.
Because the incoming liquid has already absorbed a certain amount of energy while
it was still in the suction pipe, any pressure readings in the suction pipe will be a
combination of the true suction pressure plus the energy added to the liquid by the
rotating impeller. Under such circumstances any pressure measurements taken in
the suction pipe are eqUivalent to those taken in the pump itself.
The only correct method of determining the true suction pressure in the presence
of prerotation is to measure the pressure at the surface of the liqUid in the suction
tank at a place where it is free from vortices and significant velocities. The true
suction pressure will be equal to this pressure reading plus or minus the difference
in the levels of this surface and the centerline of the pump. At low nonzero flow rates,
the velocity head and the frictional losses in the suction pipe must be added to the
suction head. The velocity head can be calculated. The frictional losses, again, can be
extrapolated on the basis of measurements performed at higher flow rates where
prerotation is almost absent. However, as both values are usually very small at the
flow rates at which prerotation begins to develop, they are frequently ignored.
Another method of establishing the suction pressure in the presence of prerotation
is based on the fact that in the absence of prerotation the losses in the suction line
increase as the square of the flow rate increases. Consequently, if we plot the suction
pressure against the square of the flow rate, we get, in the absence of prerotation, a
straight line such as line AB in Fig. 12-1.
In the absence of prerotation, this line would have followed up to the point C at
shut-off. However, owing to prerotation the actual pressure readings near shut-off
are higher and produce the curve BD. (Compare our study related to Fig. 9-2.)
To account for the effects of prerotation, the measured suction pressures are
graphed as shown in Fig. 12-1. However, at flow rates lower than the flow rate at
point B, the actual values of the suction pressures are assumed to follow the
extension BC instead of the measured values expressed by the curve BD. (This
method also implies a certain amount of error because it disregards the frictional
losses caused by the rotation of the liqUid within the suction pipe.)
The pressure readings in the discharge line are also not immune from errors. They
may be affected by the layout or other faults in the discharge line, although these
cases are uncommon. The layout and installation of the discharge piping may affect
the pressure readings in a unpredictable manner. This is further complicated
because they are inconsistent.
In a few cases, a small projection of the gasket into the waterways at the discharge
flange A (Fig. 12-2) reduced the readings on the pressure gage by as much as 60 ft.
In other cases, however, even greater projections into the discharge line did not affect
the pressure readings at all.
124 Problems Encountered with Centrifugal Pumps

tc
.El
::l
(5
(/)
.0
<1l
-0
<1l
OJ
.r::
c
0
U
::l
C/)

Figure 12-1 The effect of recirculation


Flow rate squared - on suction head readings.

Figure 12-2 A gasket project-


ing into the discharge line.
Testing 125

h
A

-- ---- . --.:..-

Figure 12-3 The effect of the location of the pressure gage on pressure readings.

In Fig. 12-3, the pump tested was a double-suction, split-casing type. For some
inexplicable reason, the pressure readings at B were always higher than at A despite
losses in the elbows E, and the difference h in the elevation of the two pressure
gauges was almost 3 ft.
Referring to Fig. 12-4, let the level of the liquid in the tubing, over the centerline of
the pump, be A in the tubing that connects the suction gage G I , and B in the tubing
that leads to the pressure gage Ch,. In this case, the total head developed by the pump
is

(12-3)

where: P2-PI = difference in the pressure readings between the gages Ch, and G I (in
our particular case, PI is negative)
W = specific weight of the liquid
(Vi-Vr)/2g = difference in the velocity heads of the liquid that is flowing in the
discharge pipe 2 and the suction pipe 1
hab = difference between the levels of the liquid in the gage tubing
hL = losses of head in the elbows E and the short, straight, pipe lengths

In practice, however, it is impossible to measure the difference hab in the elevation


of the liqUid contained in the tUbing, whenever the latter is opaque. Even when
transparent tubing is being used, such measurements would have been very time-
consuming because the volume of the air that remains within the tubing changes
with any fluctuation in pressure.
126 Problems Encountered with Centrifugal Pumps

s--.t

'-1
J

Figure 12-4 A typical layout of an open test loop.

To eliminate these difficulties, care must be taken to ensure that the tubing
connecting the pressure gage G2 (Fig. 12-4) is always full of water. This is
accomplished by mounting a three-way air valve AV immediately upstream of the
pressure gage. Mter the pump has been started, the air in the tube that leads to the
pressure gage ~ is allowed to escape to the atmosphere. During this deaeration
process, initially air will only escape through the air valve AV. Later, intermittent
gulps of air and liquid will begin to issue through the air valve. Finally, only a steady
stream of liquid will continue to flow out of the valve. To play it safe, this stream of
liquid is allowed to flow for a few minutes. Finally, the air outlet is closed, and the
gage ~ is connected to the discharge pipe. Only now can the pressure gage ~ be
regarded as ready for use.
Whenever a pump is receiving liquid under pressure, the same procedure must be
adopted for measuring the suction pressure. However, when the pressure in the
suction pipe is less than atmospheric, such as in the case of Fig. 12-4, a different
procedure must be adopted.
The reduced pressure within the suction line will cause dissolved air to escape from
the liquid and to accumulate within the tubing. In addition to the factors cited above,
Testing 127

any differences in suction pressure will cause significant fluctuations in the level A of
the liquid.
To eliminate these problems, it is common practice to make sure that the suction
tubing is filled with air. An effective way of doing this is by installing a line trap LT at
the connection to the suction pipe (Fig. 12-4). Such a trap is provided with a
detachable transparent container C to accommodate any excessive liquid that enters
owing to increased suction pressure (which is inevitable during closure of the
discharge valve for the purpose of reducing the flow rate). A small valve Venables the
line trap LT to be shut off from the suction line and to empty the pot C without
interrupting the operation of the pump.
Whenever the suction tubing is full of air, the term hab in Eq. 12-1 must be
replaced by the term h (Fig. 12-4) that is, by the difference in the elevation between
the center of the suction line at the spot where the tube leading to the suction gage is
connected and the center of the discharge gage 0:2.
A very common error occurs when an air pocket is allowed to get trapped in the
tubing that leads from the pipelines to the gages. Such air pockets are apt to occur
whenever part of the tubing has a negative slope (as in Fig. 12-5) or is shaped like an
inverted U tube (as in Fig 12-6).
One common mistake is to leave sharp edges or burrs at the connections between
the pump piping and the tubing that leads to the pressure gauges. For correct
results, these edges should be carefully rounded (see point A in Fig. 12-7).
Another common cause of incorrect pressure readings is the nonuniform pressure
in a liquid flowing through a given cross-sectional area of a pipe. It is rarely possible
to determine the true average pressure across a given section. To reduce the error,
drill and tap several equally spaced holes around a given pipe section and connect
them in parallel to the same pressure gage. Figure 12-8 demonstrates how disregard-
ing this practice can affect the pressure measurements.

Figure 12-5 A suction


line with a negative slope.
128 Problems Encountered with Centrifugal Pumps

Figure 12-6 A pipe section


in the shape of an inverted U

/
\ I /

I
I

--~-- I

I
I

Figure 12-7 The location at which


the pressure readings may be affected
by burrs.

Another important factor is damaged pressure gauges. This is particularly likely to


happen with Bourdon gauges that have been handled roughly during installation.
False pressure readouts are sometimes caused by a dampening valve, which is
often installed upstream of a pressure gage to reduce fluctuations in the readouts.
Such a valve usually achieves its dampening effects by allowing a very small amount
of liqUid to pass between the gage and the main stream. Because there is always a
certain small clearance between the stationary and the movable parts of the valve,
some valve elements may be displaced by the fluctuations of pressure so as to
temporarily shut off the passage between the gage and the piping. This may lock-in
Testing 129

80

-*"--~
.... ~- EFF 70
40 QH .x!

60
C
F 30
Ql 50 ~
Q) Iiia.
.s 40 ;:
-0 u
c:
al 20 0---0 With 4 taps equally spaced round the pipe periphery Ql
6 I
and connected in parallel to the pressure gage 30:2
*" - -x With only 1 pressure tap
ffi
0... 4 20
I 10
III
2 10

0
0 5 10 15 20 25
Flow rate (m%r)

Figure 12-8 The effect of the location of pressure taps on pressure readings.

any intermediate pressure that existed in the gage when the passage was shut off,
resulting in misleading pressure readings.
Another often overlooked source of incorrect readings is cavitation at the point
where the measuring instruments are connected to the pumping system, or
cavitation in the instruments. Such a case may be particularly easy to overlook when
cavitation occurs in the discharge line. When the pump is part of a closed loop, the
pressure in the loop may be relatively low.
One of the most vulnerable spots, in this case, is the throat of a venturi meter or
orifice, when these are used to measure flow. At these spots there is a considerable
drop in pressure. When the pressure in the unobstructed parts of the piping is
already low, the additional pressure drop in the venturi or the orifice may easily
cause cavitation.
A second spot is the last valve or fitting that regulates the return of the liquid to the
suction tank. Downstream of such a valve, the pressure is usually very low because
this location is directly connected to the suction tank. This spot, therefore, is very
susceptible to cavitation.
Another error in determining the actual pump performance is a result of test data
processing, particularly the calculation of the velocity heads and of certain losses. It
is customary to use ready-made charts to calculate velocity heads and frictional
losses in standard pipes of given nominal diameters. However, these charts disregard
the fact, that the velocity heads and frictional losses are a function of the magnitude
of the inside diameter (ID) of the given pipe, whereas the nominal sizes of pipes refer
to the outside diameter (OD). The inside diameters may vary Significantly, according
to the working pressure for which these pipes have been designed.
Sometimes these variations are negligible compared with the ID and can be ignored.
However, when a small-diameter pipe is designed for high pressure, the ID differences
130 Problems Encountered with Centrifugal Pumps

can considerably increase the velocity head and the frictional losses. The actual pipe
diameter should be carefully measured and noted in calculations.

TESTING AT OTHER-THAN-SPECIFIED SPEEDS

Sometimes, a pump must be tested at a speed that is different from the specified
speed. Such cases occur when, for example, there are power restrictions at the time
and location where the tests are being performed, a calibrated motor of suitable size
is not available, or there is some other limitation of the testing equipment.
In such a case, the pump operation at the specified speed can be calculated with
the aid of the following equations:

(12-4)

(12-5)

NPSHR s = (Ns)2 (12-6)


NPSHR t Nt

EFFs=EFFt (12-7)

(12-8)

where Qs = flow rate at the specified speed


Qt = tested flow rate
Ns = specified speed
Nt = tested speed
Hs = head at the speCified speed
Ht tested head
=
NPSHR t required NPSH at the tested speed
=
NPSHR s = NPSHR requirements at the specified speed
EFFs = specified efficiency
EFFt = tested efficiency
HPs = power consumption at the specified speed
HPt = tested power consumption

Equations 12-4 to 12-8 are widely accepted and usually produce correct results.
However, they are not foolproof. The most common deviations occur when the tests
are performed at a lower-than-specified speed. It is not unusual to get test results
that indicate lower efficiencies and higher NPSH reqUirements than when tested at
the specified speed. Also, a pump tested at a reduced speed may indicate that the
maximum limit of the flow rate it can deliver at the specified speed is higher than its
real value at the specified speed.
Testing 131

When a pump is being tested at a higher-than-specified speed, the results may


indicate that the pump is developing heads that are lower than they really are, and
that it is operating at efficiencies that are lower than the real efficiencies.

NPSH TESTS
Classification of Test Loops
Basically, there are three classes of test loops for performing NPSH tests (Figs.
12-9-12-11). In Fig. 12-9, the pump takes liquid from a container with a constant
level. The suction head is regulated by valve A in the suction line, and the flow rate
by the valve B of the discharge line. To eliminate any disturbances in the incoming
flow of liqUid, a flow straightener is usually inserted between A and the pump inlet.
This arrangement may show that the pump requires higher NPSH than it really does
because the turbulence at A tends to accelerate the release of any gas or air that is
dissolved within the incoming liqUid.
In Fig. 12-10, the pump receives liqUid from a container in which the liqUid level
can be adjusted by a suitable control arrangement. The suction head is regulated by
the level of the liqUid in the suction tank and the flow rate by a valve in the discharge
line. Such an arrangement often uses a relatively short suction line without any
significant obstructions in the path of the incoming liqUid. The arrangement may
show lower NPSH requirements than under normal working conditions because the
suction line remains short, reducing the probability that excessive amounts of
dissolved air or gas will be released from the oncoming liqUid. Pumps with long
suction lines show higher NPSH requirements because long lines provide more time
for release of gas.
Figure 12-11 shows a closed loop arrangement. The NPSH is regulated by lowering
the pressure in tank T with a vacuum pump or by heating or cooling the pumped
liqUid by means of coils C. When a vacuum pump is used, the liqUid can be stripped
of most dissolved gas or air. Such an arrangement will show lower NPSH
requirements for a given pump than when the same pump is tested in one of the
loops in Figs. 12-9 or 12-10, because any gas entrained in a liquid increases the
NPSH requirements of a pump.

Flow straightener
7
Valve B

I ~
Constant level
Pump

- ----a

Figure 12-9 An open test loop in which the NPSH is controlled by valve A.
132 Problems Encountered with Centrifugal Pumps

Variable level

Figure 12-10 An open test loop in which the NPSH is controlled by the level in the suction tank.

In practice, all three classes of test loops are widely used. However, cases exist in
which only a closed loop can be used for testing a pump for its NPSH requirements.
This relates particularly to cases when the tested pump is designed for a low head. It
may happen that the sum of the total head developed by such a pump plus its
minimum NPSH requirements are lower than the atmospheric pressure.
Take, for example, a pump that requires an NPSH of 4 ft at flow rate Qa. Let the
head developed at Qa be equal to 20 ft. If this pump is tested in an open loop, the
discharge valve can be adjusted so that the pump will deliver Qa as long as the
available NPSH is higher than 14 ft (because the head that corresponds to the
atmospheric pressure is equal to about 34 ft). However, if the NPSH falls below 14 ft,
the pump will have to develop a head higher than 20 ft to overcome the atmospheric
pressure. As soon as the available pressure drops below 14 ft, the pump will not be
able to deliver Qa when tested in a open loop.
When tested in an open loop, the minimum NPSH required by this pump at flow
rate Qa will seem to be not less than 14 ft; in a closed loop, the same pump can
deliver the same amount of liquid even when the NPSH falls to 4 ft.

Time Effects of Air on NPSH Tests


It has been known for decades that air dissolved in a liquid may adversely affect
the performance of a pump operating under reduced NPSH conditions [31].
In my experience, the presence of free air in the liquid usually changes an H-NPSH
curve with a sharp break-off point (Fig 12-12 curve a) into a curve in which the head
To vacuum pump

Tank T

Flow Meter

td - IJ:::::
• $= - It><1J
=!-::..--_ -_. - -==--
- - - '.;;;;a- . .___ -- -- -
Perforated plate

Pressure gage

To manometer
Cooling (heating) coil
t
Flow straightener

-----------
---
~
....
""
k
Figure 12-11 A closed test loop.
.....
c,.:)
c,.:)
134 Problems Encountered with Centrifugal Pumps

decreases gradually (Fig. 12-12 curve b). Also, in the presence of air, the total
deterioration of the performance occurs at a higher available NPSH than in its
absence.
Recently, I came across test results that shed additional light on how dissolved air
affects the suction performance of a pump.
A pump designated to operate at 50% of its best efficiency point in the presence of
3 ft of available NPSH, has been tested at a constant available NPSH and variable
flow rates, and it has been found to operate perfectly. When the customer was called
in to witness the test, the pump failed completely. The pump was retested at a
constant specified flow rate and variable NPSH. The test has produced a two-level
curve, as shown in Fig 12-13 by the curve a. Both the installation and the test
procedures underwent careful scrutiny, but no clue to the source of these differences
in the test results could be found.
The pump was tested in a closed loop, in which the vacuum was maintained by a
continuously operating vacuum pump, and the magnitude of the vacuum in the
suction tank was regulated by valves. As a last resort, we decided to deaerate the
water in the system by switching on the vacuum pump prior to the test and letting it
operate for about half an hour at a vacuum of 15 in. Hg. That did the trick. The test
produced a satisfactory H-NPSH, as presented in Fig. 12-13 by curve b.
The customer was invited again to witness the test, and the pump failed again. This
time, the curve had a shape like that shown in Fig.12-14 by the curve a. A series of
additional tests has been performed, but the results were never consistent. In one
case, the curve assumed the shape of curve b in Fig. 12-14.
Mter a significant amount of trial and error, the answer to the mystery was finally
found. The observed inconsistencies in the test results were caused by the air
dissolved in the liqUid. The results were also dependent on the number of steps in
which the NPSH was lowered to 3 ft, as well as on the time lapse between one set of
readings and the next.
The initial test in the laboratory was carried out at a constant NPSH and a varying
flow rate. Prior to this test, the suction tank was evacuated to create a constant
vacuum of 27 in. Hg, and kept under this vacuum until the test engineer was ready

H a
/

Figure 12-12 The effect of air on


NPSH NPSH requirements.
Testing 135

~70

~
<ll ~------®
\a
·60
al<ll
I50

Figure 12-13 The time effect on o 2 4 6 8 10 12 14 16 18 20 22


NPSH requirements. NPSH (feet)

~70
<ll
<ll
~60
"0
ttl
<ll
I 50 a

Figure 12-14 The time effects o 2 4 6 8 10 12 14 16 18 20 22


of air. NPSH (feet)

to take the readings. Enough time passed to allow the vacuum pump to remove any
excess air, which was liberated from the water. The pump then performed as
expected.
With regard to the tests at a constant flow rate and a varying NPSH, the time effect
can be explained as follows. The initial removal of air was carried out at a vacuum of
15 in. Hg. Consequently, below this pressure, additional air was being liberated. This
air appeared in the water as a mass of tiny bubbles dispersed throughout the liquid.
This lowered the specific weight of the pumped mixture of the fluids and resulted in
the observed test measurements. The amount of air liberated each time the pressure
was lowered was, of course, dependent on the magnitude of the drop in pressure: The
greater the drop, the more air liberated.
136 Problems Encountered with Centrifugal Pumps

When the measurements were carried out in many small steps, as presented in Fig.
12-13 by curve b, only a small amount of air was liberated with each additional
reduction is suction pressure. This allowed enough time for the vacuum pump to
remove the liberated air. Consequently, the pump produced satisfactory results.
When the measurements were carried out in large steps, as shown in Fig. 12-14 by
curve a, the rate at which the air was liberated from the solution exceeded the rate at
which the air was removed by the vacuum pump. This resulted in a continuous
increase in the amount of free air in the water and caused a continuous drop in
pressure.
Sometimes, with an intermediate number of steps, a state of equilibrium set in
between the rate at which the air was liberated and the rate of its removal. In such
cases, pressure readings remained constant but were lower because of a reduction in
the specific gravity of the liquid-air mixture that was caused by the presence of air
bubbles or by a partial obstruction of the impeller inlet (see Chapter 4), or both.
Under certain circumstances, the initial amount of liberated air was greater than at
subsequent measurements. This caused an initial large decrease in pressure, followed
by a partial recovery, as shown by curve b in Fig. 12-14.

The Effects of the Suction Line


Any factor that causes an uneven velocity distribution at the impeller inlet will
increase the NPSH requirements of a pump. Such factors include a valve, bend,
increaser, reducer, T joint, or any other irregularity in the path of the liquid.
There are two ways to eliminate the danger of such an uneven distribution of the
velocity: a flow straightener should be inserted between the pump inlet and the
nearest obstruction (Fig. 12-9), or the suction line in the immediate vicinity of the
pump should have a long, straight pipe. For best results, the length L of this pipe
should be at least 20 times the diameter D of the pipe (Fig. 12-10).
Even with a correctly chosen suction line, test results may be incorrect. This will
happen whenever the layout or assembly of the suction line is not carried out
correctly. One of the most common assembly errors is that the bore of the suction
line is displaced off-center relative to the centerline of the pump inlet (Fig. 12-15).

Figure 12-15 Misalignment


between the suction nozzle
and the suction pipe.
Testing 137

Another factor that can affect the NPSH requirements and the performance of a
pump is an opening in the gasket that is too small (Fig. 12-16), or the gasket's
location is eccentric in relation to the bore of the suction pipe or of the pump (Fig.
12-17). In some cases, both performance and NPSH requirements are affected when
the gasket projects into the suction line by an amount K (Fig. 12-17), as little as 3%
of the bore diameter of the suction pipe. To safeguard against gasket failure, the
gasket bore should be at least 0.25 in. larger than the ID of the suction pipe.

The Effects of the Discharge Line


Whenever a closed loop is used to test the NPSH requirements of a pump, the size
of the discharge line can affect the test results. In particular, when testing pumps
that develop a relatively low head, the absolute pressure in the discharge line may be
quite low. A discharge line of small diameter may reduce this pressure even more
owing to the high velocities of the liquid. This, in turn, may cause liquid to cavitate
somewhere downstream of the pump. Venturi meters located within the discharge
line may be particularly vulnerable in this respect.
Figure 12-18 shows, schematically, a calibration curve of a venturi meter. If a large

fWl ,
----+
)
- ( t - - - - - - -t-+--- -

Figure 12-16 An undersized gasket in the


suction line.

. - - - . ---+-1 - ' - - -

Figure 12-17 A misaligned gasket.

j 1-1--
H2

Figure 12-18 A typical calibration curve for


i1~~~~____~~
C 0 A B
a venturi flow meter. Flow rate
138 Problems Encountered with Centrifugal Pumps

flow meter is chosen. the measured flow rates will probably fall between points A and
B (Fig. 12-18). The calibration curves between these two flows are very flat. and
therefore a small error in the readings will produce a large error in the actual flow
measurements. The size of a venturi meter is chosen so the measured flow rates lie
somewhere between C and D. Within this range. the calibration curve is relatively
steep. but the pressure inside the throat of the venturi meter will be lower than the
pressure upstream of the meter by an amount somewhere between HI and H2.
Whenever a pump develops a relatively low head (e.g .. when tested at low speed)
and particularly when the available NPSH is also low. the drop in pressure in the
throat of the venturi meter may cause cavitation at this location. although the pump
proper is not cavitating. In such a case. the test results will be similar to the results
when cavitation appears in the pump proper.
U sing an oversized flow meter whose readings lie somewhere between A and B (Fig.
12-18) can eliminate this possibility. but then a manometer filled with a liqUid of low
specific gravity (instead of mercury) must be used Simultaneously to measure the
pressure differential across the venturi meter.
In a closed loop. another vulnerable point in the discharge line at which cavitation
may develop is where the discharge pipe returns to the suction tank (Fig. 12-11).
When this point is near the water level. the static pressure there may be lower than
the NPSH available at the impeller eye. This static pressure may be reduced even
more by the turbulence and hydraulic losses caused by the re-entering liqUid. This
can result in cavitation at the end of the discharge line. This may cause a blockage of
the flow by a vapor-filled pocket. in a manner similar to the case discussed in
connection with Fig. 13-11.
A similar effect can be seen when the suction tank is preceded by a pipe of large
diameter and the rest of the discharge line is of small diameter (Fig. 12-19). In this
case. cavitation may occur where the liquid passes from the smaller pipe into the
larger. Similarly. a partially open discharge valve may also become a spot where
cavitation may occur. particularly when the piping downstream of the valve leads
directly to the suction tank with no obstruction. This may occur when the valve is
located high above the level of the liquid in the suction tank.
One remedy for such cases is to choose the discharge line of as large a diameter as
possible. Another effective remedy is to install the discharge line as far below the level
of the liqUid in the suction tank as possible (Fig. 12-19).

Discharge pipe Suction tank


\
A

// ...... ~----- ,,//

~-------(" ,
-+- -- - --,-- - -+--- - - - - -----or--
,/"
.-------, I / Figure 12-19 Different return arrange-
'- .............. IL.. _ _ _ _ _ / / /
ments of a discharge pipe in a closed loop,
Testing 139

Other Sources of Test Loop Related


Disturbances
The location where the liquid returns from the discharge line to the suction tank
may also affect the test results. This location should be completely submerged below
the surface of the liquid in the suction tank. Otherwise, the returning liquid will mix
with the surrounding air and will propagate it into the pump, resulting in false test
results.
Another possible cause of misleading NPSH tests is air leaking into the system.
This is, perhaps, responsible for more incorrect test results than all other factors
combined. While the effects of air leakage on the results of NPSH tests are well
known, locating and eliminating such leakage is not always easy.
The most reliable way to test for air leakage is to put the whole system under high
vacuum for about 30 minutes to see whether the loop loses some of its vacuum. If it
does, air is leaking into the system, and the next problem is to find where leakage is
coming from.
Sometimes, the leaks can be located by putting the entire system under pressure
and observing any liquid leakage. This however, is not foolproof. Sometimes, the
system can be leakproof against a liquid and still leak air. Also, a system occasionally
is leakproof against internal pressure, but can leak under external pressure.
If a pressure test does not reveal the source of leakage, another method is to put
the entire system under vacuum and tighten and seal all suspected places until the
system holds vacuum. Even if the system can hold vacuum for a reasonable period of
time, this does not mean that there cannot be air leakage during the tests. When
placing the system under vacuum, some of the instruments (e.g., pressure gages)
must be shut off or disconnected to avoid damage to them. Later-especially in a
closed loop system-air is likely to leak through the connections to these instru-
ments.

Misleading Test Results Caused by the


Pump Proper
Manufacturers commonly offer the same pump with different mechanical options.
For example, the same pump might be supplied with either a stuffing box or a
mechanical seal. Also, there is usually a choice of mechanical seals (Le., a pump may
be supplied with either a single-acting or a double-acting seal). This, too, may cause
incorrect test results. In particular, this is true in cases where an impeller is provided
with holes for balancing the axial thrust (Fig. 12-20), or with back-vanes (Fig. 12-21).
When an impeller is provided with balancing holes, the pressure inside the space S
(Fig. 12-20) will not be much higher than the pressure in the impeller eye. When the
available NPSH is low, the pressure in space S may be less than atmospheriC pressure.
Under such circumstances, a single seal may open up and allow air leakage.
Air may also leak through a seal when an impeller is provided with back-vanes (Fig.
12-21). In this case, a vacuum may be created near the shaft when the pump
operates at high flow rates. At high flow rates, the head generated by the back-vanes
is often significantly higher than the head generated by the impeller. This will create
a tendency for the liqUid to be thrown out from the space contained between the back
shroud and casing, creating a vacuum near the impeller hub.
140 Problems Encountered with Centrifugal Pumps

Balancing
holes

Single mechanical seal Figure 12-20 The effect of balancing holes


on a single mechanical seal.

Connection to sealing liquid


~I

Figure 12-21 Sealing of stuffing


Back vanes box.
Testing 141

In both cases, air may leak into the pump when the shaft is sealed off by a stuffing
box (Fig. 12-21). This however, can be prevented easily by sealing the lantern with a
pressurized liqUid.
A different source of incorrect NPSH results may be a result of worn wearing rings
(Fig. 7-12).
An NPSH test may also give incorrect results when there are vibrations in the pump
or in the piping. Such vibrations may arise from misalignment between the pump
and driver, and they can be due to the pump supporting the weight of the pipelines or
to the strain imposed on the pipelines by the pump casing. Often the effect is that
cavitation begins at conSiderably higher NPSH values than when the same pump is
installed properly.

Testing at Shut-off and Low Flow Rates


Special problems occur during NPSH tests when the flow rate is only a small
fraction of the normal capacity of the pump. When a pump operates at a flow rate
significantly lower than its design point, the through-flow velocity of the liqUid
becomes small. This gives the liqUid more time to liberate any dissolved gas when
passing through the zone of low pressure. This increased amount of liberated gas
affects the performance of the pump in a way analogous to vapor bubbles, that is, it
reduces both the head and efficiency. Such cases are particularly pronounced with
liquids other than water (e.g., fuels).
The solubility of air in some fuels is much higher than in water (Fig. 12-22).
Consequently, tests performed on same pump with water and with fuels show that,
at low flow rates, the NPSH requirements for fuels are much higher than for water
[32].
Another factor that affects the suction capability of a pump near shut-off is the
very low efficiency of a pump at very low flow rates. Most of the power wasted owing
to low efficiency goes into heating the pumped liqUid. This, in turn, reduces the
vacuum under which the pump can operate satisfactorily. This effect is particularly
pronounced when a pump is tested in a closed loop, where the same liquid is heated
over and over again. The best way to handle this problem is to install a cooling coil C

>. I \
0
c::
Q)
-(j
:EO
Q)
-0
c::
'"
-0 0:
'"
Q)
.r::
I
--~~~1~~~416~
.E z
-0
Q)
:t:'
-f:
a
(/)
Q)
<ii
0
(fJ
---f-------f-----j 4

L-----~----~-----L----~----~----~O
Figure 12-22 The effect of dissolved 0 100 200 300 400 500
air on NPSH requirements. (Taylor.) GPM
142 Problems Encountered with Centrifugal Pumps

into the loop which could keep the temperature of the pumped liquid constant (Fig.
12-11).
Special problems are created during NPSH tests because prerotation is induced in
the suction pipe at partial and very low flow rates [211. At very low flow rates, part of
the liquid that enters the impeller returns to the suction line. This returning liquid
causes the liquid contained within the suction line to rotate (see Chapter 9), causing
a pressure differential across the diameter of the suction pipe (see Fig. 12-23). As a
result, the vacuum meter shows only the local pressure that exists at the spot at
which the measurements are taken, and not the true, absolute pressure of the liquid
that enters the impeller.
In practice, a vacuum meter is connected to the walls of the suction pipe to
measure the pressure near its walls. Consequently, when prerotation is present, the
readings always show higher suction pressure than the pump really has.
Prerotation generally begins at low flow rates and reaches its maximum at shut-off.
Because of this, it is more correct in open loops to measure the elevation between the
centerline of the pump and the level of the liquid in the tank and add to this value
the velocity head and losses owing to liquid flow through the suction line.
The velocity head is calculated by dividing the flow rate by the cross-sectional area
of the suction pipe. Frictional losses can be determined by calibrating the suction line
at higher flow rates where prerotation is almost absent, and extrapolating the data to
lower flow rates. In most cases, however, both losses and velocity head at low flow
rates are so small that they can be ignored without introducing any significant error.
When the test is performed in a closed loop, the procedure is somewhat different.
Instead of measuring the static elevation of the pump over the level of the liquid in
the suction tank, a vacuum gauge is mounted on the tank at the same level as the

13 Peripheral velocity
component
12
11
10 \ 0
en 9 \
.E-8 \
;:. \ 0.5
'u 7
0
\ ""C
Qi 6 \ <1l
Q)
> 5 \ ..c:
\ 1.0.2
4 \ Cii
(jj
3 \
\ 1.5
2

3 in.
-1
-2
-3 ~ - suction pipe
-4 Figure 12-23 The effect of pre-
rotation on velocity and on static
head distribution.
Testing 143

centerline of the pump impeller. The readings of this gauge minus extrapolated losses
and minus the velocity head in the suction line equal the value of available NPSH.
Even with all these precautions, it may be impossible to determine the NPSH
requirements of a pump at very low flow rates.
In addition to making the pressure measurements in the suction pipe inaccurate
and unreliable, prerotation often induces such intense instabilities at low flow rates
(compare the discussion related to Fig. 13-18) and low NPSH values, that it becomes
impossible to test at all. These instabilities manifest themselves in high noise and
vibrations and fluctuations in gauge readings. In many cases, the hammering effects
of these instabilities can ruin a pump in a very short time. Because no pump can
operate under such conditions, there is little sense in determining what its NPSH
requirements might be at very low flow rates. At such flow rates, a pump should be
used only in the presence of a high available NPSH.
CHAPTER 13
Pump Performance
at Reduced NPSH
To choose the right pump for a given low-NPSH application, more information about
the suction performance of a centrifugal pump than a single NPSH value at a given
flow rate is needed. Rarely does a pump operate under constant conditions. In
practice, it usually operates over a wide spectrum of heads, flow rates, and available
NPSH values. This may impose more requirements on a pump than simply operating
satisfactorily under a given NPSH.
Take, for instance, a fuel pump for a jetliner. At takeoff and landing, as well when
flying at low altitudes, the plane consumes much power because of air resistance.
Under these conditions, NPSH is no problem. The pump must supply a huge amount
of fuel, but it is not subjected to low-NPSH problems.
At high altitudes, when flying within the rarefied air, the available NPSH is at a
premium. Against this, the air resistance is much lower than at lower altitudes.
Under these conditions, the motors reqUire less fuel, but the pump must be able to
supply these amounts even under the available NPSH.
Assume that for this application, we must choose between two pumps. Pump A has
an H-NPSH curve as in Fig. 5-2, and pump B has an NPSH characteristic as in Fig. 5-3.
(see also Fig. 13-1). At high NPSH values, both curves show the same performance.
As the NPSH is reduced, the performance of B drops off at an NPSH value at which
the performance of A is still unaffected. However, at low NPSH values, B continues to
deliver liqUid long after A has broken down completely. From this description, it is
evident that B is much more suitable for this application than is A.
Now choose between the same two pumps for water supply. In water supply
applications, the NPSH values are usually subjected to smaller fluctuations than in
jetliners. Against this, power consumption often presents 80% to 90% of all operating
expenses. Any slight reduction in efficiency (owing to reduced available NPSH) can
result in enormous financial losses. For this application, A is more suitable than B.
Figure 13-2 presents the NPSH requirements of two different pumps with
comparable noncavitating performances. One of these pumps must be used for an
application in the process industry. Assume that the pump must take liqUid from an
open tank that is refilled continuously from another source. During normal
operation, the pump's function must be interrupted periodically, but the refilling of
the suction tank continues. Each time the pump is started again, it must deliver a
145
146 Problems Encountered with Centrifugal Pumps

:r:
t a....-- ---
A

-0 ,..

I
m /

Figure 13-1 Two of the different


NPSH - - - - possible shapes of H-NPSH curves.

t
I
(/)
Il..
Z

Figure 13-2 Two different shapes


Q Flow rate _ _ of Q-NPSH curves.

large amount of liquid until the level in the suction tank returns to normal. For this
application, B is better suited than A.
Consider another application in which the pump must operate without interruption
while the liqUid in the tank is refilled intermittently. In this case, the pump must
operate satisfactorily even when the level in the tank falls, but the flow rate must be
reduced accordingly to eliminate the danger of emptying the tank completely. For this
application, A is obviously a better choice than B.
Note that for the same pump, both H-NPSH curves of Fig. 13-1 will appear at
different flow rates. In fact, the H-NPSH curves may also assume several additional
forms [34) (see Figs. 13-3-13-5 and 13-lO-13-12).
Generally, any pump can be expected to produce curves that look like Figs. 13-3-
13-5, depending on the flow rate at which it operates. Curves such as those in Figs.
13-10-13-12 are usually a result of some unusual effect in the system or the pump.
Pump Performance at Reduced NPSH 147

I
t
o
I-

Figure 13-3 An H-NPSH curve at design flow rate. NPSH - - - - , I....

I
t
o
I-

Figure 13-4 An H-NPSH curve at partial flow rates. NPSH - - -

I
t
o
I-

Figure 13-5 An H-NPSH curve at high flow rates. NPSH ---II"~


148 Problems Encountered with Centrifugal Pumps

THE SHAPE OF H-NPSH CURVES UNDER DIFFERENT OPERATING


CONDITIONS

In a conventional end-suction pump, a curve such as that in Fig. 13-3 is usually


generated when a pump operates at or near its design pOint. In this case, the liquid
enters the impeller in a direction parallel to the blade inlets (Fig. 13-6).
When the available NPSH is reduced, cavitation develops in zone A-A of the blades.
This creates a concave interface between the flowing liquid and the vapor-filled
cavity. This cavity also reduces the passage area of the liquid. As a result, the velocity
of the liquid increases, causing an additional reduction in pressure near the
interface. The cavitating zone grows larger, increasing the size of the obstruction. As
a result, once a cavitating zone develops, it grows rapidly, causing an almost sudden
breakdown in performance (Fig. 13-3).
The curve in Fig. 13-4 generally appears at flow rates that are significantly lower
than design flow rate. Here, both the head and efficiency increase at reduced NPSH
values before complete breakdown in performance occurs.
At low partial flow rates, the liquid enters the impeller at an angle smaller than the
blade angle (Fig. 13-7) This creates a dead space S in the region, where cavitation
usually begins to develop. As long as the vapor-filled area is not larger than the dead
space, performance does not suffer. In fact, when the dead space is relatively large,
cavitation even reduces the hydraulic losses inherently associated with such a dead
space when completely filled-out with liquid. As a result, cavitation increases both
head and efficiency. Performance breaks down only after the available NPSH is
lowered to the point at which the cavitating zone becomes larger than the dead space.
These effects are not merely theoretical speculations but are well-substantiated by
experimental observations. Because of their importance for certain applications in the
process industry and because they are relatively unknown, the following experiment
is presented.

.. U
Figure 13-6 Schematic of flow
between two consecutive blades
of an impeller when the pump
operates at the design point flow
rate. (A) High NPSH. (8) Low NPSH. U=Velocity of the blade. 1-1 marks
the location of the passage between two consecutive blades at their inlet.

Figure 13-7 Schematic of flow


between blades at flow rates much
lower than the design point flow rate.
Pump Performance at Reduced NPSH 149

Figure 13-8 Cavitating zones at


different flow rates and NPSH.

A centrifugal pump was tested for NPSH in a loop with a transparent suction pipe.
As the NPSH was lowered, a cavitating zone A (Fig. 13-8) about 0.25 of an inch wide
appeared at the inlet tips of the impeller blades. Neither the head nor the efficiency
were affected by this cavitating zone.
As the NPSH reduced further, the cavitating zone enlarged without any adverse
effect on performance. Only after the cavitating zone reached a width of 5/8 of an
inch, did performance break down completely.
The same test was repeated at a lower flow rate. At this rate, even when the
cavitating zone widened to 5/8 of an inch (zone B, Fig. 13-8), there was no effect on
performance. Only after the cavity grew to a width of 1.38 of an inch (zone C) did
performance break down.
This test shows two things. First, a cavitating zone of significant size may exist in a
impeller without affecting the performance of the pump. Second, the size of an
undetected cavitating zone grows at reduced flow rates.
These effects are important in cases where cavitation can be detrimental to the
consistency of the pumped liquid. Such cases may occur, for example, when a pump
handles certain organic substances or colloidal solutions. In these cases, a pump
may seriously damage the pumped liquid without exhibiting any detectable signs of
cavitation. In such cases, one must choose a pump carefully.
In Chapter 5, it was pOinted out that the NPSH required by a pump is often
determined by the value at which performance drops by a certain percentage.
However, a large cavitating zone may exist within the impeller without exhibiting any
warning Signs.
Consequently, when choosing a pump for applications where there should be
absolutely no cavitation, choose a pump that will operate at flow rates no lower than
its design point, or provide the pump with a higher margin of NPSH.
The curve of Fig. 13-5 usually appears when a pump operates at a flow rate
significantly higher than the design flow. In such a case, the liqUid enters the
impeller at an angle larger than the blade angle (see Fig. 13-9). When the liquid
enters the blades, the direction of its flow is deflected to coincide with the direction of
the passage. This deflection generates a force F on the trailing side of the blade inlets
that, in turn, increases the local pressure in the zone where cavitation is most likely
to occur.
It is known from hydraulics that the magnitude of F is directly proportional to the
relative velocity of the liqUid. Consequently, whenever a cavitating zone develops at
A-A, the reduced pressure brought about by the increase in the velocity head of the
liqUid is partially counteracted by the increased force F. As a result, the head
developed by the pump under cavitation conditions falls off more gradually than in
the other two cases.
150 Problems Encountered with Centrifugal Pumps

Figure 13-9 Schematic of flow between


blades at high flow rates. (A) High NPSH.
(8) Low NPSH.

An H-NPSH curve such as that in Fig. 13-5 can occur at design flow rates and even
at partial flow rates. Presently. not all cases have found an explanation. However.
knowledge of the existence of such cases may help the pump user to choose the
proper pump:

1. In pumps of high specific speed. the suction performance is often like the performance
shown in Fig. 13-5. even at partial flow rates. This can be attributed mainly to the
position of the inlet edges of the blades. In pumps of higher specific speeds. the position
of the inlet edge is usually significantly more inclined to the axis of rotation than in
pumps of low specific speed. This means that the peripheral speeds of the inlet edge at
the outer radii is Significantly higher than at its inner radii. This causes the outermost
parts of the blade inlets to start to develop cavitation at higher NPSH values. whereas at
the inner radii. no cavitation has yet appeared. This allows the pump to continue to
pump liquid. but at a reduced rate.
2. With special geometry. it is possible to obtain suction performance as shown in Fig. 13-5
for pumps of low specific speed (see Figs. 24-2 and 24-3).
3. Admitting small quantities of air or gas into the suction line usually changes the shape of
the H-NPSH curve from that of Fig. 13-3 to that of Fig. 13-5 (see also Fig. 13-22).
4. Excessive leakage through worn impeller rings often produces a similar effect on the
shape of the H-NPSH curve as air is admitted.

When a curve that is shaped like Fig. 13-5 is undesirable. the follOwing preventive
or corrective means could be adopted: choose a pump of low specific speed that
operates at less than the design flow. eliminate any air leakage. and check and
replace worn wearing rings.

MUTATIONS IN THE SHAPE OF THE H-NPSH CURVE NOT CAUSED BY


PERFORMANCE CONDITIONS

The curves in Figs. 13-10 to 13-12 usually indicate some uncommon condition or fault
in either the pump or the system. The curve in Fig. 13-10 indicates an effect known as
alternative vane cavitation. Some of the impeller passages become completely blocked
with vapor. while the flow in others continues unobstructed. This alternative vane
cavitation may be stationary or rotating. In the first case. the same passages are
blocked all the time. In the second. the blocks rotate continuously from one passage to
another. In both cases. the H-NPSH curve is shaped like in Fig. 13-10 [351.
Pump Performance at Reduced NPSH 151

I
t
o
I-

1 1

Figure 13-10 The effect of alternate vane cavitation. NPSH ---I.~

t
I
,. - - --------
I
I
o
I- I
I
I
I
I
I
I

Figure 13-11 The effect of an undersized discharge line. 8 A NPSH ---I.~

t
I
0 C A

0
l- 18
I
I
I
I
I
I
III II

Figure 13-12 The effects induced by a vibrating discharge d b a i


valve. NPSH-.
152 Problems Encountered with Centrifugal Pumps

Figure 13-13 A typical impeller


curve with (1) all passages open and
(2) some passages blocked.

In Fig. 13-13, let the QH curve for noncavitating performance of a pump be


represented by curve 1. When some of the impeller passages are blocked, the pump
delivers smaller quantities over its entire operating range. Consequently, the QH
curve will look like curve 2.
To see how such blockage affects the H-NPSH curve, assume that a pump is tested
for its H-NPSH requirements at flow rate QA. As long as there is no cavitation, the
pump develops a head HI. However, after some of the passages become blocked, the
head at flow QA equals H2.
Alternate vane cavitation may be caused by a faulty impeller casting or an incorrect
layout of the suction line or with splitter vanes. When an NPSH curve as in Fig. 13-10
is objectionable from an application point of view, the first step should be to check
whether the suction line was laid out and installed according to the gUidelines in this
chapter and in Chapter 14. If the line was installed correctly, open the pump and
check to see if the impeller vanes are spaced at exactly equal distances from one
another. If the line is laid out incorrectly, a flow starightener may eliminate the
problem (Fig. 14-26).
The NPSH curve in Fig. 13-11 has the following properties: below a certain value of
available NPSH, the pump loses its ability to deliver the speCified flow rate. When the
available NPSH drops below this critical value, the pump continues to deliver liqUid,
but the output moves to a lower flow rate.
This case indicates an undersized discharge line or any of the elements that
comprise this line, for example, a valve or a flow meter. An undersized element may
cause cavitation in the discharge line. This, in turn, may present such a high
resistance to the flow that the pump must deliver less liqUid.
To correct this, replace the affected element of the line with an element of larger
dimensions, or change the layout of the discharge line so that the elevation of the
element where cavitation may occur is reduced to the lowest possible value.
The NPSH curve in Fig. 13-12 seems to be caused by a faulty valve disc. The disc may
vibrate as a result of flowing liqUid causing rapid pressure fluctuations throughout
the entire pumping system. These fluctuations are believed to excite cavitation effects
within the pump. Presently, there are no adequate explanations for the mechanism of
this effect.
Pump Performance at Reduced NPSH 153

Fig. 13-14 schematically presents the loop in which such an NPSH curve was
observed. The operator started the testing procedure from point I through A (Fig.
13-12), using only the 102-mm (4.0-in.) valve. When the operator reduced the
available NPSH to b, the operator still used the 102-mm valve to adjust the flow rate.
At this point, cavitation bubbles were observed through the transparent suction pipe.
They seemed to be generated in the impeller eye and to collapse in the impeller eye
and in the suction pipe. Also, the head of the pump dropped from A to B.
Next, the operator lowered the available NPSH to a value of d. However, the
operator first opened the 102-mm valve completely and started to adjust the flow rate
with the 203-mm (8.0 in.) valve. The head immediately increased to D, and the
cavitation bubbles disappeared. Only after the NPSH was lowered still further did the
performance break down.
To study this effect more in detail, the test was repeated using the, 102-mm valve
alune. Below the NPSH value a, the available NPSH was lowered at very small
intervals and the heads were recorded at each step. The curve (Fig. 13-12) was
labeled lAC-II. Also, when the NPSH was lowered below a, cavitation developed within
the impeller eye with ever-increasing intensity.
This test was repeated with the 102-mm valve wide open. Only the 203-mm valve
was used to regulate the flow rate. In this case, no cavitation bubbles appeared in the
impeller eye above an NPSH value of b. The head remained constant down to an
NPSH value of d. It seems that the described effect has been caused by the disc of the
102-mm gate valve. When protruding into the flowing liqUid, this disc seemed to
vibrate at higher flows, causing cavitation. When wide open, the disc of that valve
cleared the path of the flowing liqUid. This stopped the vibrations.

I i
f--f~<j
II
6 m Approximately ---------1'1
I I /203mm
Butterfly valve

3m
203 mm I.D.
/ Pressure gage
connection

Suction tank

- -- . - - - --1-1-1-

Figure 13-14 The test loop for the case illustrated in Fig. 13-12. 1.0. Inner diameter.
154 Problems Encountered with Centrifugal Pumps

THE EFFECT OF PREROTATION ON PERFORMANCE AT LOW NPSH

The NPSH requirements of a pump decrease with a decrease in flow rate [36]. This
general rule, however, sometimes fails at very low, partial flow rates. There have been
cases where the NPSH requirements suddenly underwent a drastic increase [37]. This
effect seems to be due to prerotation (compare Chapter 9).
In the presence of prerotation, the pressure distribution in the suction pipe
assumes a parabolic shape (Fig. 13-15). At the outer diameter (OD) of the suction
pipe, the pressure rises above the average suction pressure, while near its center, it
drops well below this value.
The intensity of prerotation and its influence on the suction performance of a pump
depend largely on the geometry of the impeller and its operating speed. For an
impeller inlet of given diameter and speed, geometry is the main factor affecting
prerotation.
In an impeller of low specific speed, the inlet edges of the blades are relatively short
and almost parallel to the axis (Fig. 13-16). In such an impeller, the low pressure
zone created by prerotation is relatively remote from the impeller vanes and, there-

Suction pipe diameter

ct
I
+
t
Average 1

suction pressure
0-------- ---1---
Actual suction
pressure (in the Figure 13-15 Pressure distribution at the
presence of prerotation) pump inlet in the presence of pre rotation.

\ I
\ I
\ I
\ I
\ I
\ I
\ I
\ I
\ I
Figure 13-16 Impeller of low specific speed.
Pump Performance at Reduced NPSH 155

\ I
\ I
\ I
\ /
\ /
\ /
\ /
\ /
\ /
\ I
Figure 13-17 Impeller of high specific speed.

fore, has little effect on performance. On the other hand, the inlet edges of blades
with high specific speed are relatively long and more inclined relative to the axis (Fig.
13-17). Here, the low pressure zone created by prerotation contacts a large portion of
the blade inlets. Consequently, it hase a great effect on performance. It increases the
NPSH requirements of the pump by reducing the pressure along significant portions
of the blade inlets.
Thus, pumps of high specific speed may cause a lot of trouble when operating at
partial flow rates with low available NPSH. In extreme cases, the entire pumping
system may operate with high noise and vibration, as if subjected to heavy hammer
blows.
Observations through a transparent suction pipe have revealed the following. At
low flow rates and high NPSH values, intense prerotation has been observed within
the suction pipe. As the NPSH was lowered, a large single cavity appeared periodically
in the center of the suction pipe (Fig. 13-18) and collapsed vigorously, vibrating the
entire system. In a 5-in. diameter suction pipe, the central cavity reached a diameter
somewhere between 2 to 3 in. and a length of over 24 in.
When the NPSH was lowered further, prerotation stopped and the large cavity
disappeared. Instead, large vapor-filled cavities were observed on the trailing faces of
the blades, and the pump began operating quietly again, although it developed a
lower total head (Fig. 13-19).
These effects can be explained easily based on what we know about the origin of
prerotation. As pOinted out in Chapter 9, prerotation occurs when some of the liquid
that has already been acted on by the blades returns to the pump inlet. As a result of
that action, the liqUid acquires a tangential velocity component Cu. Owing to inertia,
the liqUid also retains this velocity component after it has left the impeller and
returned to the suction line.
Strong experimental evidence supported by theoretical studies suggests that the
magnitude of the C u component imparted to the liqUid by the inlet parts of the blades
156 Problems Encountered with Centrifugal Pumps

--- --

--- --
Figure 13-18 Central cavity in
prerotating liquid.

Vapor filled cavities

.. u

Figure 13-19 Elimination of the


angle of incidence caused by the
vapor-filled cavities.

depends, to a great extent, upon the magnitude of the angle of incidence, or the
difference between the angle ~*, at which the liquid enters the impeller, and the blade
angle ~l (Fig. 13-20). Below the design flow rate, the magnitude of the angle of
incidence, and with it the magnitude of C u , is smallest at ~*=~l. It increases with
reduced angle ~*, that is, with reduced flow rate [38].
When prerotation sets in, a low-pressure zone is generated at the center of the
suction pipe (Fig. 13-15). Under cavitating conditions this low pressure zone fills with
vapor, reducing the cross-sectional area of the suction pipe (Fig. 13-18). This, in
turn, increases the velocity with which the liquid enters the impeller eye. However,
such an increase in the velocity of the approaching liquid increases the magnitude of
the angle ~*, eliminating the cause of prerotation. Consequently, the pressure
throughout the suction pipe returns to its average value, causing a violent collapse of
the central cavity.
As soon as the cavity collapses, the total cross-sectional area of the suction pipe
becomes available for the liquid flow. This, in turn, reduces the velocity of the
incoming liquid and the angle ~*. Consequently, prerotation develops, starting the
cycle again.
When the available NPSH is lowered even more, the passages between the blades
become blocked, largely by vapor bubbles (Fig. 13-19). This increases the magnitude
of the velocity component em; and with it the magnitude of the angle ~*. The increased
magnitude of the approach angle eliminates the primary cause of prerotation, and
the pump operates quietly again.
Pump Performance at Reduced NPSH 157

Figure 13-20 The angle of


approaching liquid versus the
blade angle.

THE EFFECT OF A SUCTION LINE ON NPSH REQUIREMENTS

When a suction line includes several bends that follow in the same direction (e.g., in
the direction of a right-handed spiral), vortices may be induced in the liquid entering
the impeller.
One property of such vortices is that the pressure at their center is lower than the
average pressure in the suction pipe. Consequently, a sequence of unidirectional
bends increases the NPSH requirements of the pump. Figure 13-21 presents test

5 I
Double-bend I
I
ahead of I
~4 I
Q) I
a3 I
-S /Straight
I 3 I end suction
(j)
a.. / .In Iet
z /
/

2 /
/

"""
" ""
OL-____~______~______~____~___
Figure 13-21 The influence of a suction line on o 25 50 75 100
NPSH. Flow rate (m3/hr)
158 Problems Encountered with Centrifugal Pumps

results of a pump with two bends in the suction line. The same pump was later
tested with a straight line (Fig. 13-21).
The suction performance of a pump is also affected by mismatches between the
suction pipe and pump inlet (Fig. 12-15) and by a gasket forming an obstruction to
the flow of the incoming liquid (Figs. 12-16 and 12-17).

THE EFFECT OF AIR OR GAS ENTRAINED IN THE LIQUID

When the suction line is not airtight and the inside pressure is below atmospheric
pressure, air enters the pump. The rate at which this occurs increases with the
reduced available NPSH. Increased air leakage, in turn, increases the drop i1H of the
head developed by the pump (Fig. 13-22). Consequently, the shape of the H-NPSH
curve changes from a to b.
Air in the liquid entering the impeller will also completely block the passages earlier
(at higher available NPSH) than in an airtight suction line.
Although it always reduces the head developed by a pump, the presence of air can
have certain beneficial effects in the presence of cavitation. When a pump operates
under cavitation, small amounts of air often act as a cushioning medium, which
reduces the mechanical impact of the collapsing bubbles. This reduces the damage
caused to pump parts by cavitation. It also reduces noise and vibration. Therefore,
controlled amounts of air are sometimes introduced intentionally into a pump that
must work under cavitating conditions. This, however, is done only in extreme cases
when there is no alternative.
Similar effects will occur when the pumped liqUid contains significant amount of
dissolved air. The solubility of any gas in a liqUid decreases with reduced pressure.
Low available NPSH, therefore, liberates a major part of the gases dissolved in the
liqUid into the low-pressure zones of the suction line and impeller. The effects are
identical to air leakage.

t
::r:
a !::.H

/b
,-
,., ---
-0
CIl
Ql
I

,
I I
I

Figure 13·22 The effect of air on an


NPSH - - - H-NPSH curve.
Pump Performance at Reduced NPSH 159

THE EFFECT OF TIME ON SUCTION PERFORMANCE

The suction performance of a pump is also affected by time. As a pump operates


during a prolonged period, the smoothness of the waterways undergoes gradual
changes. Even the shape may change owing to erosion and corrosion. All of these
factors affect the suction performance.
Leakage owing to worn out wearing rings may be particularly damaging. Figure 13-23
shows how the change in wearing ring clearances has affected the suction perfor-
mance of a pump. Within the working range of the pump, the NPSH requirements for
a given flow rate moved toward lower flow rates by the amount of leakage q. In
addition to the above, increased leakage often changes the shape of the H-NPSH
curve from that in Fig. 13-3 to that in Fig. 13-5.

THE EFFECTS OF TEMPERATURE ON SUCTION PERFORMANCE

Equation 4-1 can be rewritten as

(13-1)

For a given pump, flow rate, and speed, the inlet velocity C is constant. It is also
reasonable to assume that the NPSH requirements remain constant. Equation 13-1
can, therefore, be rewritten as

(13-2)

where M =
c2
NPSH - - = constant.
29

Equation 13-2 states that the absolute pressure Ps required to suppress cavitation
increases directly with the vapor pressure Pv . Indirectly, the equation states that Ps

~4
"""
"
CD
'iii
E
~3 ""
..,"'0.12 mm
(f) clearance
a.
z2

Figure 13-23 The influence of


wearing ring clearance on NPSH 25 50 75
requirements. Flow rate (m 3/hr)
160 Problems Encountered with Centrifugal Pumps

increases with the temperature by the same amount as Po. In reality. however. this is
not always the case.
The increase in the required manometric suction head at higher temperatures is
usually remarkably lower than the increase in vapor pressure. as if the NPSH
requirements of a pump are lower at increased temperatures. This is because the
behavior of liquid at elevated temperatures depends on its thermodynamic properties.
Consider. for example. a certain volume Va of saturated steam at two different
temperatures. 50°F and 200°F. The specific volume of saturated steam is 1703 ft3flb
at 50°F and 33.6 ft3flb at 200°F. Consequently. it is necessary to evaporate 50 times
more water to fill the given volume Va at 200°F than at 50°F.
The latent heat of steam at 200°F is about 90% of the latent heat at 50°F. This
means that, at 200°F. it is necessary to transfer 45 times more heat to the liquid for
generating the same volume Va of steam than at 50°F.
The liquid that flows through the impeller usually remains in the low-pressure zone
for a short time. Consequently. a smaller volume of vapor is liberated during this
time period at 200°F than at 50°F. This explains the reduced intensity of cavitation
under the given available NPSH at elevated temperatures.
The damage caused by the collapse of vapor-filled bubbles may also be lesser at
higher temperatures than at lower ones. The time necessary to remove the latent heat
from a given volume of steam increases with the amount of heat liberated. Here. a
word of caution is in order: while the impact of vapor bubbles collapsing at an elevated
temperature is usually less severe than at lower temperatures. the damage may still be
significant because the strength of many materials is also reduced at higher
temperatures.

THE EFFECT OF ENTRAINED IMPURITIES


Cavitation can occur only if impurities (nuclei) are present in the liquid [39]. The
cohesive forces between the molecules of an absolutely pure liquid are so great that,
without such nuclei. a liquid can withstand even significant negative pressures
(tension) without cavitating. In the laboratory. scientists have succeeded in creating
tensile stresses up to 20 atm before cavitation occurs [40].
In practice. any liquid has enough impurities to allow cavitation at vapor pressure.
However. excessive amounts of impurities may lead to higher NPSH requirements.
This is mainly because liquids that contain a great amount of impurities entrain
many tiny bubbles. These expand under reduced pressure. leading to the same kind
of disturbances as air leakage.
Such cases are particularly common with liquids containing organic matter (e.g ..
paper pulp). In this case the liquid may also contain significant amounts of other
gases that were created by. say. fermentation.

SUCTION PERFORMANCE WHEN HANDLING DIFFERENT LIQUIDS


The thermodynamic properties of a liquid may have a pronounced effect on the NPSH
requirements of a pump. Because different liquids usually possess different thermo-
dynamic properties. the NPSH requirements of a pump can be expected to vary from
one liquid to another (Fig. 12-22).
Pump Performance at Reduced NPSH 161

However, the thermodynamic properties are not the only parameters that make the
NPSH requirements different for different liquids. Surface tension and viscosity also
come into play (compare Chapter 8).
Because of the multitude of parameters, it is almost impossible to accurately
predict the NPSH requirements of a pump when pumping other than the liqUid with
which it has been tested. Consequently, when an exact knowledge of such
requirements is crucial, the only foolproof method is to carry out actual NPSH tests
with the liquid to be pumped.

THE EFFECT OF A SHARP EDGE ON CAVITATION

It has been known for decades that a sharp projection in the path of a flowing liquid
can adversely affect the performance of a centrifugal pump. Until recently, however,
very little has been known about the causes and the mechanism of such an effect.
On one occasion, I came across a case history that sheds light on the nature of that
effect. I also found a theoretical explanation for the cause of that effect. A detailed
account of that case history is given in Ref. 41. Here, we present only a few highlights
of that case history and its implications.
An unconventional inducer was tested in a loop provided with a transparent
suction pipe. This allowed the action of the inducer to be observed with the aid of a
stroboscopic light. During these tests, small bubbles began to appear at each sharp
edge of the inducer blades and were carried away with the flowing liqUid. It was not
known whether these were air-filled or vapor-filled bubbles. The total loop was put
under high vacuum, and care was taken to remove any air that might have been
dissolved in the water. After this was accomplished, the pump was started again.
However, the bubbles again began to appear at the same locations as during the
initial tests. Now it was certain that these were vapor bubbles. To eliminate the
bubbles, the suction pressure was gradually increased to 20 m above the atmos-
pheric pressure, but to no avail. The vapor-filled bubbles continued to be generated
at each sharp edge of the blades and carried away with the flowing liquid.
Recently, I found out that what was observed during these tests was not a
coincidence. It has been proven that when a liqUid flows along a solid boundary, it
must separate at all the points at which the curve representing this boundary has a
discontinuity in its derivative. A cusp, or sharp edge, forms such a Singular point.
This explains the observed effect.
When a liqUid that has been in direct contact with a solid wall flows past a sharp
edge, it separates from the wall, creating an empty space (vacuum) between the
flowing liqUid and the solid wall. This causes some of the liqUid to evaporate into that
empty space and to be carried away in the form of vapor-filled bubbles. When these
bubbles enter the zone of higher pressure they collapse vigorously, causing typical
cavitation damage.
CHAPTER 14
Pumping System
Layout
The size and geometry of a sump, tank, or container from which liquid enters the
suction line has enormous effects on pump performance. In particular, the choice of
these parameters may be critical when the dimensions of the source where the pump
gets its liquid is compared to the handled flow rate.
As a rule, if a pump can handle the volume of liquid contained in the suction sump
in less than 2 min, the sump design must be subjected to close scrutiny.

EFFECTS OF PUMP INTAKE AND PIPING LAYOUT

Even in large sumps, there may be problems due to insufficient submergence of the
suction inlet or incorrect relative positioning of several suction lines. When
centrifugal pumps were used mainly for irrigation and water supply, sumps created
problems almost exclusively in large installations. With the widespread adaptation of
centrifugal pumps by almost all industries and other applications, troubles caused by
improper design of pump intake has become much more common.
Often pumps must be installed in locations where space is at a premium. Also, the
cost of any pump intake grows astronomically with its size. Thus, pump intake size
should be kept as small as possible. Under these circumstances, an incorrectly
designed suction sump may have detrimental effects on pump performance.
Another application in which an incorrect suction sump design can have catas-
trophic consequences is in power stations, where vast amounts of water for cooling
are used. Again, the design and construction of a suction sump are of utmost
importance. They clearly show the need for in-depth knowledge and understanding of
the phenomena occurring at the pump intake.

Vortices in the Suction Sump


In Fig. 14-1, liqUid enters through duct A into a circular sump S and leaves it
through B. Both ducts are tangential to the walls of the sump, but lie at different
levels. Owing to the eccentric position of duct A, the liqUid that enters the sump S
moves along circular paths, creating a vortex.
163
164 Problems Encountered with Centrifugal Pumps

L
--- - - - 1\ I -- - - - -

-----~T---­
ill
--- ----\-t--t----- - - - -

-------ii j--- --=:::----.- B =---=- --


--~ ==~- 4+-~-~~- ----x
-rrr-
- - - - -------t--l j - - - - - -

o
l

Figure 14-1 Development of


an air funnel in a vortex.

Assuming that the losses in the sump are negligible. the motion of the liquid is
subject to the law of conservation of energy. This law is expressed by Bernoulli's
equation:

P V2
- +- +H = constant (14-1)
y 29

where P = pressure in the liquid


'Y = specific weight of the liquid
V = absolute velocity of the liquid
H = static elevation over a certain datum X-X (Fig. 14-1)
Whenever a liquid moves in concentric circles. Eq. 14-1 is fulfilled only if the
tangential velocity component of the liquid fulfills the condition
Pump System Layout 165

CXR=CoXR,,=constant (14-2)

where: Co = tangential velocity of the liquid at a certain reference radius Ro


C = tangential velocity component at any other radius R

These equations have important implications as far as the properties of a vortex


and its effect on pump performance are concerned. For simplicity, assume that the
sump is open to the atmosphere and that the downward velocity of the liquid is so
small that it can be neglected. In this case, replace V 2 in Eq. 14-1 with CZ. Keep in
mind that the pressure on the liquid surface is constant (atmospheric). Equations.
14-1 and 14-2 lead to the following equation for the shape of the vortex surface:

(14-3)

where Ho = elevation of the free surface of a liquid over the datum X-X at a distance
Ro from the center
Ro = reference radius from Eq. 14-2 equals the distance of the centerline of
channel A from the center of the vortex.

Equation 14-3 shows that elevation H of the liquid at any radius R decreases with a
reduction in R. Consequently, the center of the vortex assumes the shape of an air-
filled funnel.
In practice, this is not always the case. Whenever outlet B from the sump is located
deep below the surface of the channel A, the funnel is usually filled with a core of
liquid that rotates as does a solid body.
For subsequent discussions, we use the submergence L=(Ho-H) (see Fig. 14-1)
instead of H. Thus, Eq. 14-3 can be rewritten as

(14-4)

Equation 14-4 states that at a large submergence L, the radius of the funnel
becomes very small. However, a small value of R results in increased surface tension.
This effect, combined with other properties of a real liqUid (viscosity, friction along the
solid boundaries, and so forth), causes the funnel to close at a greater depth. This, in
turn, triggers an interaction between the particles that are already in the funnel with
particles of the surrounding liqUid and causes the funnel to fill with a rotating core.
Now assume that, instead of duct B, the sump is provided with duct D located
directly below the core of the vortex. As a result, the liqUid is continuously drained
from the core, allowing it to remain full of air. The presence of D creates a tendency to
keep the funnel full of air, however, the factors that act toward filling it with liquid
also persist. One factor is surface tension.
Surface tension increases inversely to the radius of the core. This radius decreases
with increased submergence (Eq. 14-4). Consequently, when the submergence of Dis
great enough to generate adequate surface tension, the funnel remains full of liquid.
Conversely, when the submergence of D is small, an air funnel appears (the "bathtub"
effect).
166 Problems Encountered with Centrifugal Pumps

70
v
./ VV ......... I--"
....

60 ,1--'" ,," / V
..,-
V
~ ~ .........
I--"

/"V V
fo"'"
e / / /' ~
~\~ /'
,9' .~~e I--'" V
.~
/'
,~
.~.' 'Y,\'Y,
,9' .~~e / '
/I co\~·
CO ~
/ ,95'1/
,/ \().~
~ L ./
V
10 / /
o Figure 14·2 Minimum sub-
100 150 200 300 400 600 800 1000 2000 3000 4000 6000 mergence of suction intake.
Flow rate (GPM) I.D., Inner diameter.

This analysis can be related directly to the suction line of a centrifugal pump.
When the entrance to such a suction line is not adequately submerged below the
level of the liquid, air may enter the pump due to vortices present in the suction
sump.
Submergence is not the only factor that determines whether an air funnel will be
created at the inlet to the suction line. However, certain general gUidelines can be
used to judge whether a given submergence is adequate (Fig. 14-2). However, these
guidelines apply only to correctly designed sumps.
In practice, the required minimum submergence may vary significantly from these
guidelines, depending on the position of the suction inlet in relation to the vortex, the
velocity of the liquid in the suction pipe, the velocity of the liquid as it enters the
sump, and the combined shape of the inlet duct and the suction sump.

Position of the Suction Inlet in Relation to Vortex


Assume that the inlet to the suction pipe is located near the center of the vortex
(Fig. 14-3). Often, the liquid flow into the suction pipe drags the lower end of the
vortex into the pump. Consequently, the pump drains the liquid from the core of the
vortex. The rate of drainage depends on the velocity of the flow in the suction pipe.
The higher the velocity, the greater the danger that the core will become filled with
air. Extensive experiments with suction pipes have demonstrated that the velocity of
the liquid in the suction pipe is perhaps the most decisive factor determining funnel
creation in a correctly designed sump [42).
The likelihood that the vortex will be drawn into the suction pipe also depends on
the distance of the pipe from the core. The smaller the distance, the greater the
danger that an air funnel will be formed. Indirectly, this implies that a suction sump
should be large enough to allow suction -line installation as far from any vortex as
possible.
In a well-designed suction sump, certain general gUidelines of location and
dimensions are established by the Hydraulic Institute (Figs. 14-2 and 14-5).
Pump System Layout 167

----~

Figure 14-3 Entrance of air into the


suction pipe.

Velocity of Liquid in the Suction Pipe


Velocity of flow in the suction pipe has a decisive effect on the creation of air
funnels. The higher the velocity, the greater the danger of air funnels and the greater
the submergence required to suppress this danger. Fig. 14-2 indirectly presents this
relationship between velocity and required submergence.
While the rule is that the danger of generating an air funnel increases with the
velocity of the liquid in the suction pipe, sometimes the opposite is also true.

Figure 14-4 Air funnel generated by


prerotation.

C B S H Y

-
...... V ;.V
150,000
100,000 1 / V I I I lJ-t-H A
80,000
60,000 I
I
II V '" V/.V ./
I-'"
J-...1-

...... ......
V I-'"
40,000 II J ~
30,000 I / Ip ...... i-"'"
~
c..
20,000 II 1/ / I
I I t%
~ 15,000 1 A L -~--~~-------~~~---
-----t - - - - - - - - - - - - ~
- - t%
II / //' V ,/ --- -r---- ---- ---- --- -r---cc-- t%
~ 10,000 I Y I B ..... ~
8000 'I
~
~
0
u::: 6000 I /, V : H
I f/ / :
Lc ~ ~~
4000
3000
I I /
I II
I
I I ~
2000
1500 II II / /////)//////////////~//////////&
1000 IJ I I I I I I I I I I I I I I I I I I I

100 200 300 400 500

Figure 14-5 Recommended sump dimensions.


168 Problems Encountered with Centrifugal Pumps

It was pointed out in Chapter 9, that recirculation takes place only at low inlet
velocities in the suction nozzle. At such velocities, a vortex is generated within the
suction pipe, by prerotation. Under certain circumstances, this vortex propagates to
the surface of the suction sump, thus creating the well-known air funnel (Fig. 14-4).
In such cases, much greater submergences may be needed than shown in Figs. 14-2
and 14-5.

Velocity of Liquid Entering the Sump


In Fig. 14-1, Ro is the distance of the centerline of the channel of the approach A
from the center of the core of the sump. In such a case, Co can be the average velocity
of the liqUid in A.
Radius R of the funnel, found from Eq. 14-4 is equal to

(14-5)

For a given submergence L and a given distance Ro of the inlet duct from the
center, R increases with the velocity of approach Co. A large value of R means a small
value of surface tension. This reduces the forces that tend to fill the central funnel
with a liqUid core and increase the probability of creating an air funnel.

Combined Shape of the Inlet Duct


and Suction Sump
Figures 14-6 and 14-7 illustrate the effect of the combined inlet duct and suction
sump shape on the formation of vortices.
In Fig. 14-6, the liquid enters the sump from a channel of the same width as the
sump itself. As a further precaution, a screen is prOvided at a distance Y upstream of
the pump. Also, distances B, C, and H are established according to the gUidelines in
Fig. 14-5. Under these circumstances, the liqUid can be expected to enter the suction
line without any major disturbances.
Fig. 14-7 presents another extreme case. The small dimensions of the inlet duct
result in high inlet velocities. They also result in the appearance of two vortices. To
complicate things, the pumps are located at the center of these vortices. This
arrangement usually results in very unsatisfactory pump performance.
The manner in which a sump vortex affects the performance of a centrifugal pump
depends largely on its location in relation to the suction pipe. When the center of the
vortex lies outside the suction pipe (Fig. 14-3), the vortex draws air into the pump

Figure 14-6 A recommended sump layout.


Pump System Layout 169

-------~------

Figure 14-7 An unsatisfactory sump layout.

with all its adverse effects. When the center of the vortex is located inside the suction
pipe (Fig. 14-7), no air is drawn into the pump. However, this may lead to very
unstable operation.
There is a basic difference between a vortex whose core is open to the atmosphere
and one whose core is completely shut off from the air. In the first case, the pressure
can never be less than the atmospheric pressure. In the second, the pressure inside
the core can drop well below the vapor pressure of the pumped liquid. Consequently,
when the core of a vortex is totally contained within the suction pipe, cavitation may
develop within the pump.
Another kind of disturbance sometimes occurs when the core of the vortex is
outside the suction pipe and the submergence is so great that no air funnel is
created. In this case, the pressure in the core cannot fall below the pressure of the
surrounding liquid, and air does not enter the suction pipe. Consequently, when
such a core is drawn into the suction pipe, it causes neither air leakage nor
cavitation, but it does change the velocity distribution of the liquid entering the
impeller. This, in turn, may reduce the output or cause noisy operation.

PRINCIPLES OF SUMP DESIGN

A characteristic feature of Fig. 14-5 is that dimension B is not the minimum recom-
mended value, but the maximum recommended distance. This can be explained
using Figs. 14-1 and 14-7, and Eq. 14-2.
Figures 14-1 and 14-7 indicate that the cores of the vortices are most likely to
appear somewhere in the center of the sump. On the other hand, the pump intake
must be kept away from such cores. This means that the suction inlet should be
located near the walls. This is the main reason an upper limit is imposed on distance
B.
However, too small a distance may allow all the liquid to enter from one side of the
suction inlet (Fig. 14-8), causing an uneven distribution of the velocity of liquid
entering the impeller. This is especially likely to happen when the suction pipe is
equipped with a suction bell of large diameter that touches the wall of the sump. To
eliminate this disturbance, increase B above the value recommended in Fig. 14-5.
170 Problems Encountered with Centrifugal Pumps

Figure 14-8 The effect of close proximity of a wall


downstream of the intake.

Generally, keep distance X (Fig. ·14-8) between the rim of the suction bell and the
nearest wall somewhere between X=O.25xD and X=O.50xD.
It is best to make the width of the approach channel equal to the width of the
suction sump. However, this is not always possible, especially in industrial applic-
ations where the suction tank gets its liquid from a pipe. In such cases, the next best
thing to do is to make a tapered approach (the dotted lines in Fig. 14-7). For such an
approach, the Hydraulic Institute recommends that the angle a should be between 45
degrees and 75 degrees. As a further improvement for such a case, the Hydraulic
Institute also recommends placing a flow straightener at the end of the conical
section (in the form of a bar screen) and a certain minimum value for the distance Y.
Although, in many cases, the shape in Fig. 14-6 usually gives satisfactory results,
some designers prefer to fill out the corners of the pump intake (Fig. 14-9). Others
recommend providing the sump with two semicircular walls (Fig. 14-10). These
recommendations are believed to be capable of eliminating vortices that may appear

Figure 14-9 One of the recommended shapes of the


suction sump.

WO
-------------------------------------~~ 0 Figure 14-10 An alternate recommended sump layout.
Pump System Layout 171

in the corners of the sump. These designs may perform satisfactorily, but practice
has not confirmed their superiority. In fact, vortices that could be expected to be
generated in square corners could well be replaced by vortices whose centers are
located near the suction pipe, in centers 0 of the semicircles (Fig. 14-10).
The shape of the plan of the sump is not the only factor that may lead to vortices.
The shape of the vertical section may be important. In Fig. 14-11 the sharp corner at
A may easily generate vortices. This occurs when the distance Y is relatively small
(Fig. 14-11). To prevent a vortex from forming, allow any changes in the elevation of
the sump and duct to follow along an inclined slope (Fig. 14-12).
When liquid enters the sump through a pipe P (Fig. 14-13), the sharp edge E may
produce vortices. To reduce this danger, the edges should be rounded off as much as
possible.

I
~---------i-

.. -=-=--=---=----=--=--~
----=-=--~=-=-
-===- --- ---- ---- ----
I
r
f==
---

I
Figure 14·11 A vortex generated by
I
I+----y----+jl
a sharp corner in a suction sump.

---
----
----
----
----
----
----
----
----
----
----
---- ---

Figure 14·12 Sump design that is


intended to prevent the vortex
shown in Fig. 14·11.

Figure 14-13 Vortices generated by a sharp inlet


edge of the supply pipe.
172 Problems Encountered with Centrifugal Pumps

Extremely harmful effects may be caused by an inlet pipe positioned above the
liquid level in the suction tank (Fig. 14-14). The resulting waterfall usually draws air
into the sump, and from there it usually propagates into the suction pipe.
Special precautions are necessary when more than one suction pipe is installed in
a common sump. No pump should be installed upstream of another (Figs. 14-15 and
14-16). When the liquid flows past the suction pipes, vortices are shed (Fig. 14-17).
The vortices shed beyond the upstream pipe propagate into the downstream pump,
causing disturbances.
A satisfactory arrangement of several pumps in one common sump is shown in Fig.
14-18. Here, the pumps are arranged in a row perpendicular to the direction of the
flow of the liquid. No pump is upstream of the other. However, even such an arrange-
ment will function satisfactorily only if the distance S between the particular pumps
and all other dimensions are chosen according to the recommendations given earlier.
In certain cases, even an arrangement as in Fig. 14-18 causes problems. In this
case, partitions between neighboring pumps may be helpful. One of the special
features is that a certain distance should be left between these partitions and the
back wall of the suction sump in the event the flow is not Uniformly distributed along
the entire width of the sump. In such cases the partitions often become the centers of
the resulting vortices (Fig. 14-19). This, in turn, keeps the centers of these vortices at
the greatest distance from the pump intakes.
When a partition extends to the back of the wall, an uneven velocity distribution in
the suction sump usually causes a vortex in each of the compartments. The centers
of such vortices usually lie near or even coincide with the centers of the suction pipes
(Fig. 14-20).

Figure 14-15 Harmful Figure 14-16 The worst


arrangement of two in- arrangement of two in-
takes within a common takes within a common
suction sump. suction sump.
Pump System Layout 173

Figure 14-17 Vortices shed behind


a cylindrical object. Flow

Figure 14-18 Recommended


arrangement of two intakes in a
common sump.

+ ~ l ! !! )

l ! !!)
Figure 14-19 The effect of
a space left behind a baffle
separating two pumps.
+ ~

Figure 14-20 The effect of


extending partitions down to
the back wall.
174 Problems Encountered with Centrifugal Pumps

REMEDIAL MEANS FOR IMPROVING SUMP PERFORMANCE

It is not always possible to build a suction sump that conforms to all requirements. In
some cases, there are space and cost limitations; in others, a pump must be installed
over an existing tank. There are also instances in which well-designed sumps cause
problems for unknown reasons. In such cases, remedial means may solve them.
When air funnels are generated in the vicinity of a suction pipe, one remedy is to
cover the surface of the liquid with an inert floating material. For example, when the
liquid is water, pieces of wood can be spread over the surface. These floats usually
break up the existing air funnels and prevent others from forming.
When the core of the vortex lies within the suction pipe, such floats are useless. In
this case, some kind of flow straightener-either baffles or screens-can be installed
in the sump. For instance, allow the water to pass consecutively under overhung and
oversubmerged baffles (Fig. 14-21). When using this remedy, the following pOints
must be observed:

1. The distance between the overhung baffles and the bottom of the tank should be great
enough to allow liquid to pass under the baffles with a relatively low velocity.
2. The submerged baffles should terminate below the surface of the liquid so that the liquid
can flow over with a relatively low velocity.
3. In no case should the submerged baffles be terminated so high as to generate a waterfall.
This may draw air into the liquid and propagate it into the pump.

Whenever the liquid enters the sump above the water level, as shown in Fig. 14-14,
the entrance of air into the pump can be prevented by a somewhat different
arrangement of the baffles, such as the one shown in Fig. 14-22.
Here, the first baffle imparts an upward velocity to the entering bubbles, thus
assisting in their escape from the liquid. Any remaining bubbles can then be removed

--- Figure 14-21 Breaking up vortices


with baffles.

Figure 14-22 Separating air drawn


in by the incoming liquid.
Pump System Layout 175

by redirecting the flow by means of an overhanging baffle, and then by a second


submerged baffle, and so on.
Other remedial means that can eliminate the adverse effects of an incorrectly
designed sump are a wide suction bell at the inlet opening of the suction pipe ( Fig.
14-8), a cone beneath the suction inlet (Fig. 14-23), and a baffle behind the suction
pipe (Fig. 14-24). In some cases, these have proven to be effective, in others, they are
completely useless. Only trial and error can determine if they will help in any given
case.

PREDICTING SUMP PERFORMANCE FROM MODEL TESTS

Notwithstanding what is already known about pump sumps and their effects on
pump performance, many unknow aspects relating to the effects of sump design
remain. This is particularly true in cases when location and other factors require
new, untried sump designs.
When designing large and costly installations, it is necessary to construct a
reduced-scale model of the final outfit and carry out any necessary modifications on
the model before finalizing the design.
For these model tests to produce significant results, the experiments must be
carried out according to certain model laws. It is believed that the most significant
condition is that Froude's number should be identical for the model and prototype.
Froude's number is defined as

V
F=-- (14-6)
(gh)05

Figure 14-23 A conical guide under the


suction bell.

Figure 14-24 A baffle installed behind the pump intake.


176 Problems Encountered with Centrifugal Pumps

where V = velocity of the liquid in the sump


h = depth of the liquid at some significant location

If all dimensions of the model are made T times smaller than those of the prototype,
the velocities V m in the model must be increased according to the equation

(14-7)

where Vp is the velocity of liquid in the prototype.


Recent studies have shown that the actual performance of a sump also depends on
the Reynolds number, which is defined as

Vh
Re=- (14-8)
v

where v is the kinematic viscosity of the pumped liquid.


It is impossible to build a model that simultaneously satisfies both Eqs. 14-7 and
14-8, except for the case when the dimensions of the model are identical to the
dimensions of the prototype. Also, because the effect of Froude's number F has
proven to have a much greater influence on sump performance than the Reynolds
number, therefore, in practice, only Froude's number is usually the same for both the
model and the prototype.
In critical cases, it is possible to test a model at identical values of F and of Re, even
when the dimensions of the model and the prototype are not identical. This can be
accomplished by using for the model a liqUid with a viscosity that is different from
that of the liqUid handled by the prototype.

THE EFFECT OF THE SUCTION LINE LAYOUT

When a suction line contains many bends, each of which turns the liqUid in the same
orientation, vortices may occur in the liqUid. Figure 14-25 presents results of tests
carried out on a given pump. Curves marked I represent the performance of a 3-in.
end-suction pump run at 1770 RPM.
During the test, the suction line had four bends, each turning the liqUid so it
moved like a right-handed screw. Mter the test was completed, the suction line was
rearranged so that two of the bends continued to turn the liqUid in the same
direction while the other two turned the liqUid in the opposite direction. This resulted
in curve II.
To avoid disturbances of this kind, eliminate any bends from the suction line. When
this is impossible or impractical, insert a flow straightener between the last bend and
the pump (Fig. 14-26).
Many types and kinds of flow straighteners exist. The choice of a given straightener
for a given application is usually governed by the available space, the available NPSH,
and the readiness with which the straightener can be purchased. A small straightener
usually poses significant resistance to flow, thus requiring a higher available NPSH.
Larger straighteners, on the other hand, may be prohibitive due to space limitations.
Pump System Layout 177

200
190
~ 180
"0
gl170
I
160 70

150 60

50 ~
~
40 B
30 ~
c::
Q)

20~
w
10

Figure 14-25 The effect of suction line o 100 200 300 400
layout on pump performance. Flow rate (GPM)

----~/~
-----------------~------

~""
7---T---7--~_

_'""-- _ _ ~
~
___
""/
~
"
_ _ _ J-
""
~~---

/
-
--
Figure 14-26 A typical
flow straightener. Strainer of perforated metal 7

Although a well-designed flow straightener eliminates many problems caused by


improper layout of suction line, it can sometimes be a source of problems, partic-
ularly when the pumped liquid contains solid matter of a stringy consistency. In such
a case, the flow straightener may easily become clogged with the solid matter, thus
reducing, or sometimes even stopping completely, the flow of the pumped liqUid.

CAUSES AND EFFECTS OF WATER HAMMER


The term water hammer is a carryover from the period when pumps were primarily
used to handle water. It refers to a group of phenomena closely associated with the
inertia of the liqUid within the pumping system.
When a pump is started, the liqUid within the pipeline must be accelerated until it
reaches its full velocity. Similarly, when the pump is stopped, the liqUid in the
pipeline must be decelerated until it comes to rest.
Similar effects also occur when the pumped liqUid is redirected from one line to
another. In this case, both effects may occur Simultaneously: deceleration in one line
and acceleration in the other. These changes in the state of motion of the pumped
liquid can generate forces that can cause a pump casing to burst. In other cases,
they may cause severe cavitation.
178 Problems Encountered with Centrifugal Pumps

Water hammer can be particularly damaging where a process requires a rapid


switchover of the pumped liquid from one pipeline to another or frequent and rapid
starts and stops. In such cases, pump damage may even occur more frequently than
in long pipelines used for water supply or fuel transport.
In long pipelines, starts and stops are relatively infrequent. Consequently, only
pressures that exceed the strength of the casing may rupture it. In a case of frequent
and rapid starts and stops, a casing may burst under a smaller pressure surge as a
result of fatigue. Water hammer may also cause pump damage by inducing cavit-
ation. In this case, the frequency at which a pump is put out of order is directly
proportional to the frequency at which cavitation occurs.

Principles of Water Hammer


According to Newton's laws of motion, a body will change its state of motion only
when subjected to a force. In the absence of such a force, any mass that moves with
a given velOCity will continue to move with the same velocity, and a body at rest will
remain motionless. The force required to change the state of motion of any body is
proportional to its mass, to the magnitude of change in its velocity, and to the rate at
which this change takes place.
Closing a valve located near the discharge end of a long pipeline brings the total
volume of the flowing liqUid to a sudden stop. Such a change in the state of motion,
however, can occur only under the action of a force. In this case, the force is a
sudden increase in the local pressure. This pressure surge is generated where the
flow was stopped. There, it causes a local compression of the liqUid and an elastic
expansion of the metal surrounding this high pressure zone. The elastic local
deformation of the liqUid and of the surrounding metal creates a pressure wave that
moves throughout the pumping system.
The magnitude of the pressure buildup caused by the closure of the valve is pro-
portional to the mass of the liqUid contained within the pumping system and the
speed with which the valve is shut down. In long pipelines or ones of significant
diameter, rapid valve shut-down may generate a pressure wave of intensity great
enough to burst a pump casing.

Preventing Damage from Pressure Surges


The first immediate remedy to prevent a pump casing from bursting as a result of a
pressure surge is simply to shut down the valve very slowly. In practice, the
attendant should close it slowly over a specified period of time, or an automatic shut-
off device that performs the same duty can be installed.
In industry, such arrangements may be undesirable because of the necessity to
shut-down or redirect the flow quickly. In such cases, alternatives must be adopted.
One method is to install surge tanks in strategiC locations. These are open tanks that
are partially filled with pumped liqUid and connected to the main lines so that
whenever the pressure wave reaches the point of connection, the liquid rises. This
dissipates the wave's energy, eliminating the danger of damage.
A second method is to install air vessels in strategic locations. These are
hermetically sealed vessels connected to the main pipelines and filled with air or inert
Pump System Layout 179

gas. When the pressure wave reaches the vessel, the gas within compresses and part
of the energy is converted into heat, which is dissipated through the vessel walls.
In a third method, a pressure relief valve installed between the source of the pressure
wave and the pump opens under the influence of the pressure wave, allowing part of
the liquid to escape through a bypass. This is perhaps the most effective way of
reducing the intensity of a pressure surge, but it can be used only if the escape of
liqUid is acceptable.
Deciding on which means to use depends on cost, properties of the pumped liquid,
size and layout, special requirements, topographical layout, resistance to flow, special
problems, and the interaction among all these factors. In many cases, the problems
caused by water hammer are so complicated that they are best handled only by
highly qualified specialists in this field. In other cases, however, they may be
alleviated by one of the methods listed above.

Cavitation Caused by Water Hammer


and Its Prevention
Pump damage caused by inertia of the liqUid is not limited to buildup of pressure.
It may be caused by a transient reduction in pressure. In this case, the pressure at a
given location in the pump may fall below vapor pressure and cause cavitation.
In Fig. 14-27, consider a simple pumping system consisting of a short suction line
with a foot valve F, a pump P, and a long discharge line D which delivers the pumped

s--.f

Figure 14-27 A pumping system in which water hammer may cause cavitation.
180 Problems Encountered with Centrifugal Pumps

liquid to a storage tank S. In this system, the discharge valve V is often kept open all
the time and is used only when a pump must be dismantled for repairs or some other
emergency.
When the pump is shut down, its rotating parts continue to rotate because of
inertia. During this transient period, the speed of the rotating parts decreases
continuously until they come to a complete stop. At the same time, the liquid in the
pipeline flows under the influence of inertia, and new liquid enters the suction line
under the influence of the available NPSH, to replace the liquid discharged into the
storage tank. Both effects continue until all come to a standstill.
In Fig. 14-28, the solid line schematically presents a head capacity curve of a pump
operating at a constant speed. The most significant feature is that the head developed
by the impeller decreases with increasing flow rate Q. At flow rate Q= Qo, the pump
ceases to develop any head at all. At still higher flow rates, the head becomes
negative. This means that the pump, instead of adding energy to the pumped liquid,
actually removes some of its energy owing to the available NPSH. This, in turn,
means that reduced NPSH is available to suppress cavitation at the impeller eye.
When pump speed is reduced, the relationship between its head and flow rate is
altered continuously (the dotted curves in Fig. 14-28). Each speed reduction results
in a lower head-capacity curve.
Now assume a pump that originally operated at flow rate Qa against head Ha has
stopped operating. After a time interval T, its speed drops to N=NT and its head
capacity curve is represented by curve T (Fig. 14-28). At the same time, owing to
inertia, the liquid in the system continues to flow, although at a lower rate QT. Under
certain circumstances, this flow rate may still be large enough to cause the pump to
generate a negative head. This, in turn, generates a very low pressure at the
discharge side of the impeller. Under certain conditions, the pressure can cause
severe cavitation and damage at the outlet tips of the impeller blades.

---- ...
--- ..... ......
---... ........ ....., .....

-
t I+-----Qr----I
~--------Qa--------~

~----------Qo----------~
Figure 14-28 Illustration of how cavitation can
develop when a pump is shut off.
Pump System Layout 181

One of the simplest and least expensive ways of preventing cavitation is to install
an air valve at the pump discharge. The valve is adjusted to open when the pressure
at the impeller outlet falls below a predetermined pressure. This breaks up the
vacuum and prevents cavitation. Unfortunately. this remedy cannot always be
adopted. Sometimes. air cannot be admitted into the pump because of subsequent
problems during starting. In other cases. this is undesirable because of the nature of
the pumped liquid. and one of the following remedies can be adopted:

1. Increase the available NPSH. For example. install the pump at a lower elevation. and.
respectively. pressurize the suction tank.
2. Provide the rotating element of the pump driver combination with a flywheel that allows
the speed of the pump to decrease more gradually.
3. For an electric drive. replace the valve V (Fig. 14-27) with a solenoid-operated valve that
will shut down as soon as the power supply to the motor is stopped.
4. Replace valve V by a pressure-operated valve that will begin to close as soon as the
pressure at the discharge drops below a certain minimum.

A pump may be also subjected to cavitation during the transient conditions that
exist during starting. When a pump is started. the liquid in the suction and discharge
lines must be accelerated. The force that accelerates the liquid in the suction line is
generated by the difference in the total head at the inlet to the suction pipe and the
total head at the impeller inlet. The maximum value of this force is limited because
head difference cannot exceed the difference between the NPSH available at the inlet
to the suction line and the NPSH required at the impeller eye. The force generated by
this difference must be able to accelerate the liqUid in the suction line to its full
velocity of flow within a limited period of time. This time limit equals the period
required by the pump driver combination to attain its full operating speed. When the
difference between the available and the required NPSH is insufficient to accomplish
this. cavitation will occur at the impeller eye.
Generally. the following factors determine whether the inertia forces will generate
cavitation at the impeller eye:

1. The length and diameter of the suction line. The greater the mass of the liqUid contained
within the suction line. the greater the force required to bring it to full flow within the
required time period.
2. The difference between the NPSH available at the inlet to the suction line and the NPSH
reqUired at the impeller eye. The greater the difference. the greater the force available for
accelerating the liquid.
3. The length of the discharge line. The shorter the line. the smaller the time interval
required to accelerate the liquid to its full velocity. When the liqUid in the discharge line
attains full flow before the liqUid in the suction line does. cavitation occurs in the
impeller.
4. The time interval during which the rotating parts of the pump accelerate to full speed.
The smaller this time interval. the shorter the time available to accelerate the liqUid in
the suction line without causing cavitation.

In addition to these four factors. the magnitude of the time interval required to
accelerate the liqUid to its full velocity of flow is also significantly affected by the
resistance to flow of the different sections of the suction line.
182 Problems Encountered with Centrifugal Pumps

There are several ways to prevent the inertia of the liquid from causing cavitation at
the impeller inlet. First, the suction line should be as short as possible, should have
as few bends and valves as possible, and should be large enough to reduce resistance
to flow. Second, the available NPSH should be as high as possible. For example, the
pump should be located as low as practical or the suction tank pressurized.
Third, the pump should be started against a partially closed discharge valve. It
should be opened slowly, allowing enough time to accelerate the liqUid in the suction
line. Certain pumps, however develop great instabilities and high power requirements
at the partially closed valves (e.g., propeller pumps). This makes it necessary to start
them against an open valve.
Fourth, certain automatic or manual devices that buildup pump speed may gradually
be incorporated. This allows enough time to accelerate the liqUid in the suction line.
A fifth possibility is to install an auxiliary tank filled with liqUid in the vicinity of
the pump inlet and connect it to the suction line via a pressure operated valve (Fig.
14-29). This valve is preset to open whenever the pressure in the pump inlet falls
below a certain minimum, thus allowing some liqUid from the tank to enter the low-
pressure zone. With proper location, tank proportioning, and adjustment of the valve,
cavitation during startup can be prevented.

Additional Inertia Effects


Inertia effects during startup can affect different pumps in various ways. Some
pumps are subject to high stresses when operating against a closed valve. For example,

Auxiliary tank

Figure 14-29 Auxiliary tank for preventing


cavitation during startup.
Pump System Layout 183

propeller pumps, when operating near shut-off, develop high instabilities accom-
panied by intense noise and vibrations. These vibrations can ruin a pump in a
relatively short time.
To prevent this kind of damage, such pumps are always started against an open
valve. However, with long discharge lines the inertia of the liquid causes additional
resistance to flow. This is equivalent to starting the pump against a partially closed
valve.
This kind of damage can be prevented by providing a bypass between the pump
outlet and the check valve of the main discharge line (Fig. 14-30). In this installation,
the pump is started with the bypass valve wide open. The valve is gradually closed,
redirecting the flow toward the main discharge line, allowing the liquid in the main
pipeline to accelerate gradually without any serious inertia effects.
Another adverse effect is "slam pressure." This effect is characterized by the fact
that it can occur in pipelines as short as 10 to 15 yd.
Figure 14-31 presents a pumping system consisting of a suction line S provided
with a foot valve F, a pump P, and a discharge line D that terminates with tank T.
When the pump is stopped, the liquid in the system continues to flow for a short
time. During this period, the foot valve remains open and new liquid enters the
system, replacing the liquid that was discharged into the tank. At the end of this
period, the liquid stops and reverses its direction.
A foot valve usually consists of a simple flat disc. When the backflowing liquid
acquires a significant velocity, this builds up stagnation pressure on the top of the

Check valve

I
I

Bypass

Pump

Figure 14-30 Bypass


for preventing pump
damage during startup.
184 Problems Encountered with Centrifugal Pumps

Figure 14-31 Pre-


loaded suction valve.

valve disc and reduces the pressure on its underside. Consequently, the disc valve is
slammed against its seat, shutting off the flow suddenly. The slamming action
creates an intense pressure wave even in a relatively short pipeline. Such an intense
pressure wave can easily fracture the pump casing.
One of the popular ways to prevent the buildup of excessive slam pressure is to
preload the disc of the foot valve with a dead weight Wor with a spring (Fig. 14-31).
The force exerted by preloading should be great enough to shut the foot valve before
the returning liquid has acquired a significant velocity. However, the foot valve can be
preloaded only when there is adequate reserve of available NPSH to overcome the
increased resistance of the foot valve to prevent cavitation.
Another remedy is to install a relief valve in the suction line. Such a valve opens
when the pressure in the suction pipe exceeds a certain value. This allows the liquid
to bypass back into the suction sump or tank.
Damage caused by slam pressure can also occur in the absence of a foot valve,
whenever a nonreturn valve is installed near the pump discharge (Fig. 14-32). In this
case, the liquid between the nonreturn valve and the suction sump continues to flow
toward the suction inlet, under the influence of inertia, after the valve has closed.
This can generate a vacuum in the pump, leading to cavitation.
To prevent damage, the disc of the nonreturn valve should be preloaded so that it
closes immediately after the liquid stops and begins to reverse its flow. This is done
by mounting a valve disc D on a long shaft S, which extends outside the valve casing
Pump System Layout 185

Figure 14-32 Preloading


the check valve at the
pump exit.

through a stuffing box. Lever L is fixed to the external end of the shaft and loaded
with a weight Wor a spring.
Another remedy is to install an automatic air valve near the pump outlet. This valve
is adjusted to open when the local pressure between the pump and the nonreturn
valve drops below a certain level. This allows air to enter the line and to break the
vacuum.

OPERATION IN SERIES
Whenever several pumps operate in series. one or more may fail if an upstream pump
is shut down or broken. Upstream pump stoppage reduces the pressure at the inlet
of the next pump, causing cavitation. If there are additional pumps farther
downstream, these may also fail. The first downstream pump is unable to develop the
required head because of the cavitation in the upstream pump.
Another kind of failure may occur in a system where a Significant portion of the
required head is due to friction. In this case, the liqUid may continue to flow although
at a reduced flow rate, even if one pump has broken down and the combined head
developed by the pumps is reduced. This may happen because the resistance of the
system is lower at reduced flow rates. Because the pumps are arranged in series, the
pumped liqUid continues to flow through the idle pump, rotating the impeller. This,
in turn, may loosen the nuts that hold the sleeves and impeller in proper axial
position, seriously damaging the rotating parts of the pump.

PARALLEL OPERATION
Whenever the suction nozzles of several pumps are connected to one manifold, air
leaking into one pump may also enter other pumps. This affects the performance of
all pumps.
186 Problems Encountered with Centrifugal Pumps

Such a leak may occur in the suction line of one pump in operation. It may also
occur at the discharge side of one or more pumps. In the latter case. however. such
air leakage is harmful only when the pump through which the leak occurs is not in
operation.
Another kind of trouble may occur when there is a breakdown in the discharge
piping of one of the pumps. This decreases the discharge pressure and significantly
increases the flow rate at which the affected pump is operating. The first direct result
of such a breakdown may be cavitation in the affected pump. Indirectly. however.
such an increase in flow rate also reduces the NPSH available for other pumps of the
system. Consequently. such a breakdown may also produce cavitation in the rest of
the pumps.
Different kinds of problems with pumps operating in parallel may occur in the case
of a drooping head capacity curve. however. they happen infrequently. Figure 14-33
presents the resistance curve Hr. of a system in which two identical pumps are
installed to work in parallel. The QH curve of each is represented by curve A. When
one pump is in operation. the pump delivers the flow rate QI against the head HI.
When the pumps operate in parallel, two alternate QH curves. Band C. are created.
Two different curves result because. between the shut-off head Ho and the
maximum head Hmax each pump may deliver two different flow rates against the
same head. Consider. for example. the performance of the two pumps when operating
in parallel against the head H 2 • which lies between Ho and Hmax. (Fig. 14-33).
When both pumps operate in parallel. they can deliver three different flow rates.
One equals the sum of Qb from one pump plus Qb from the second pump. or Q2=2 Qb.
The second Q4 equals flow rate Qa from one pump. plus flow rate Qa from the other.
or Q4=2Qa. The third is the sum of Qa from one pump. plus Qb from the other. or
Q3=Qa+Qb.
By adding. in this way. the flow rates of each pump at each given head. two
different curves (B and C. Fig. 14-33) are obtained when two pumps with drooping QH
curves operate in parallel. There is. however. one important difference between curves
Band C.
Curve B extends from Q=O to a flow rate where H=O. Curve C. however. extends
only from flow rate Qm. which is equal to twice the flow rate that each pump delivers
at head Hmax. down to a flow rate at which the head is equal to the shut-off head Ho.
Below this head. the combined output of both pumps follows only curve B.

Figure 14-33 Parallel operation of


two pumps with drooping curves.
Pump System Layout 187

As long as the resistance curve of the system intersects only curve B, the combined
output of both pumps is determined unequivocally. Therefore, the drooping shape of
the QH curves poses no problem. Only when the resistance curve intersects both B
and C (Fig. 14-34) is there trouble because the resistance curve of the system is now
in equilibrium with both QH curves, which were generated due to parallel operation
of both pumps.
In this case, the flow rate may fluctuate rapidly from one value to the other,
resulting in noise and vibration. In practice, however, this is rare. Curve C covers
only a small range of flow rates (Figs. 14-33 and 14-34). This alone reduces the
probability that the resistance curve of the system will intersect both branches of the
combined QH curve.
Even when the resistance curve Hr intersects Band C, noise and vibration do not
necessarily appear. These are usually generated when the fluctuations in flow rate
occur at a relatively high frequency. However, such rapid changes in the flow rate
meet high resistance owing to the inertia of the liqUid in the pipeline. This resistance
is low only when the pipeline is short.
Another reason pumps with a drooping curve rarely cause problems even when
operating in parallel is that when the pumps operate in one of two flows Qb or Qc,
they are in a state of stable equilibrium. In stable equilibrium, the easiest and
shortest way in which equilibrium can be restored is by returning to the original flow
rate, when the flow rate changes by a very small amount.
The equilibrium is unstable only when the difference between Qc and Qb is very
small. In this case, a small deviation of flow rate Qc toward the right brings it near
flow rate Qb. Similarly, a small deviation of flow rate Qb toward the left will bring it
into the vicinity of Qc. In practice, such cases occur when the resistance curve is very
steep. Steep resistance curves, however, are rarely favored because they are always
generated by a high amount of friction in the pipeline. This, in turn, means large and
costly energy losses.
When pumps with drooping QH characteristics operate in parallel, the shut-off
head Ho of each pump may be lower than the head Hi against which one single pump
is operating (Fig. 14-35). In such a case, one pump might be unable to begin
delivering liquid when the second pump is already operating. In modern practice,

Figure 14-34 A special case in


which two pumps with drooping
curves may cause instabilities
when operating in parallel.
188 Problems Encountered with Centrifugal Pumps

System resistance

u'-----_ Figure 14-35 Parallel operation of


two pumps with drooping curves
and with shut-off heads lower than
their operating heads.

however, this rarely happens. The shut-off head on a QH curve of a pump is usually
the average value of a range of heads through which a pump fluctuates at shut-off
(compare Chapter 9). In reality, the actual head at shut-off fluctuates between a
value lower than the nominal Ho to a value higher than Ro. In most cases, the value
at the peak is even higher than the maximum head Hmax (Fig. 14-33). Such peaks at
shutoff allow a pump to start delivering liqUid even when the other pump is operating
and developing a head equal to Hi. However, such possibilities can be different for
different individual cases.
Only pumps with extremely large droops may cause problems (the dotted lines in
Fig. 14-35). Such curves were common in older designs, but are extremely rare in
modern pumps.

STRESSES IMPOSED BY PIPELINES

Three basic factors can cause a pipeline to impose a serious stress on a pump casing.
One factor is poor workmanship during installation. A pipelines should always be
installed so that its ends match the pump suction and discharge nozzle exactly, both
in length and concentricity, before they are connected to the pump. Pipe flanges
should not be forced into position by tightening the bolts. This imposes stresses on
the pump casing and often leads to trouble.
A second factor is changes in the length of piping caused by temperature fluctu-
ations that occur when ambient temperatures altered or when hot liquids are
pumped. The magnitude of the changes is directly proportional to the length of the
pipeline and the change in temperature. To avoid problems from stresses by such
changes, the pipeline requires freedom of movement relative to the pump. This can be
provided by expansion joints, flexible piping, or an omega-shaped section of piping
inserted between the pump and the remaining pipeline. Another alternative is to lay
out the pipeline so it will turn 90 degrees at a short distance from the pump and slide
freely on its support. (Such an arrangement, however, may be prohibitive because of
its effect on the hydraulic performance of certain types of pumps.) Another solution is
to mount the pump and motor on a specially constructed bed plate. The bedplate
Pump System Layout 189

must be designed for extra rigidity and must rest freely on its foundations without
being bolted down. The pumping unit can slide freely. and the rigid bed plate keeps it
aligned with the driver.
The settings of foundations and of pipe-supporting structures are also sources of
stress. Any foundation or structure continuously changes its position because the
ground under it changes. particularly during the first year after the construction
work has been completed. During this period. the earth fills out any loose spots and
voids created by excavations and landfill. Even later. climate and other forces
influence the process.
Often. the situation may be corrected by periodically loosening the bolts connecting
the pump and driver to the bedplate and realigning the unit. In more severe cases.
new holes may need to be drilled in the bed plate. Sometimes the only remedy is to
exchange entire sections of pipeline.
As pointed out earlier. to reduce the danger of stresses being imposed on the pump
by the pipelines. it is advisable to insert an expansion joint between the pump and
the pipelines. However. under certain circumstances. this too is likely to cause
problems. if not carried out properly. Figure 14-36 illustrates such a case.
A large vertical pump was installed over a sump. as shown in Fig. 14-36. To isolate
the pump from the thermal expansions and contractions of the pipes. an expansion
joint E was inserted between the pump and the pipelines. During operation. this
pump developed intense vibration and noise. The cause turned out to be the
expansion jOint. Although it protected the pump against fluctuations in pipeline

-------

Pump

Figure 14-36 Preventing vibration caused by forces generated within the pump head due to change in
the direction of flow.
190 Problems Encountered with Centrifugal Pumps

length, it left the pump exposed to a huge force resulting from the change in direction
of the liquid. This force occurred in the discharge head of the pump, as a result of the
change in the direction of flow of the liquid. (This particular pump happened to
handle huge flow rates.)
Mter the cause of the problem was determined, the system was provided with tie
rods that took up the forces generated by the turning of the pumped liquid, and
transferred them to a special support. This arrangement is shown in Fig. 14-36 by
the dashed lines. This arrangement solved the problem.

CONSEQUENCES OF PIPELINE STRESSES

The direct consequence of pipeline stress is an elastic distortion of the pump casing,
which can lead to metallic contact between the casing ring and the impeller ring,
increasing the rate of wear on the sealing rings. Increased sealing clearances, leakage
losses, and reduced output result. Distortion may also cause early mechanical
destruction of the rings and cavitation from heat generated by friction between the
metallic surfaces. It may also lead to eccentricity of the shaft in relation to the
stuffing box and, respectively, the mechanical seal seat: both will leak and may be
destroyed. Elastic distortion may also result in misalignment between pump and
driver. This may cause overheated bearings, noise, and vibration. The vibration, in
turn, may cause cavitation and fractured casings. Finally, elastic distortion can
result in excessive power consumption owing to metallic contact between the rotating
and stationary parts. This may overload the driver and impose a prohibitively high
torque on the shaft.

EFFECT OF SUCTION LINE LAYOUT ON PERFORMANCE

In Chapter 13 we demonstrated how a double bend upstream of the pump inlet


affects the NPSH requirements of a pump (Fig. 13-21). This is not the only case in
which the layout of the suction line affects pump performance.
As yet, the direct link between the suction line layout and pump performance
remains yet mostly unexplored, however, it is known that when a suction line
contains several bends that are turning the liquid in the same direction (Fig. 14-37),
a vortex that is likely to propagate into the impeller can be generated. Such a vortex
always causes a drop in pressure within certain regions of the incoming liqUid,
causing an increase in the NPSH requirements of the affected pump. This has been
demonstrated in the test results given in Fig. 13-21.

Figure 14-37 Four consecutive bends,


all turning the flow in the same direction.
Pump System Layout 191

190
~180 ,
2 I 'Q..,
:;- 170
ell
Q)
1770 RPM " "\J
I160 \,-
II __ 70

150 60
50E
Q)
~
40 Q)
-S
30 G'
c:::
Q)
20:2
:t::
UJ
10

Figure 14-38 The effects of changes in o 100 200 300 400


suction line layout on performance. GPMQ

Figure 14-39 Four consecutive bends,


two of which turn the flow in one direction
and the other two in the opposite direction.

The magnitude of the effects caused by the presence of bends in the suction line
depends, in part, on their spacing in relation to one another and on their relative
orientations. To the best of my knowledge, little research has been done in this area.
Consequently, the follOwing case history may serve as a useful illustration.

A 3 x4 in. end-suction pump persistently produces lower-than-rated heads and efficiencies (Fig. 14-38, curve 1).
A careful inspection of the pump, installation, and testing procedure revealed no reasonable cause for such a
poor performance. However, a careful inspection of the suction line revealed that it included four bends, all with
the same orientation, as shown in Fig. 14-37.
The suction line was of uniform 4-in. diameter, and the spacings between the individual bends were fairly
large, ranging /rom 40 to 70 in. Still, the fact that all four bends turned the liquid in the same direction proved
to be enough to cause the observed deficiencies in performance.
Because of the existing local conditions, the use of the bends could not be avoided. However, by changing the
layout of the suction line in a manner that caused two bends to turn the flow counterclockwise and the other two
to turn the flow clockwise, as shown in Fig. 14-39, it was possible to restore the operation of the pump to its
original rated performance. This performance is shown in Fig. 14-38 by the curves marked II.
CHAPTER 15
Installation, Handling,
and Operation of Pumps
and Pumping Systems
THE EFFECTS OF AIR OR GAS

When a mixture of gas and liquid is acted on by a rotating impeller, the heavier liquid
is thrown outward by centrifugal force, displacing the gas toward the center.
Depending on the amount of gas contained in the liquid, this displaced gas fully or
partially blocks the inlet to the impeller vanes. In the first case, the pump is not able
to deliver any liquid. In the second case, the pump may start to deliver a reduced
amount of liquid, provided that it can develop the head required to overcome the
resistance of the system. In this case, one of the following may occur: The flow rate
increases gradually until it is restored to the expected value; the flow rate decreases
gradually until the pump ceases to deliver any liquid; or the pump operates at the
same, reduced, flow rate.
These differences in pump behavior are intimately related to the manner in which
pumping occurs and to the nature of the interaction between a flowing liqUid and a
gas.
When an impeller operates, it throws the contained liqUid outward, causing a
vacuum at the inlet. This vacuum, however, never occurs in practice (except under
cavitation conditions). This is because the displaced liqUid is continuously replaced
by new liqUid that enters through the suction nozzle of the pump.
The force that makes the new liqUid replace that contained in the impeller is due to
the difference between the pressure at the inlet to the suction line and the pressure
existing at the eye of the impeller. LiqUid will enter the impeller only if the pressure at
its eye is less than the pressure upstream.
Almost any gas dissolves in a liqUid under the action of pressure, As the pressure
increases, more gas dissolves in a given volume of liquid, and vice versa.
Consequently, when the liqUid enters the eye of the impeller, some of the dissolved
gas is liberated due to the reduced pressure. This increases the volume of the
entrapped gas, causing the gas blockage to grow continuously until the flow of the
liquid becomes totally interrupted.
Against this, a flowing liqUid can carry away some of the gas it comes into contact
with owing to the action of adhesive forces. This phenomenon tends to reduce the
193
194 Problems Encountered with Centrifugal Pumps

amount of gas entrapped in the impeller eye until it vanishes completely. In such a
case, the pump resumes its full, rated flow.
In a pump with a partially blocked impeller, both effects occur simultaneously.
Depending on the ratio of the amount of gas liberated as a result of the reduction of
pressure and the amount of gas carried away by the liquid owing to adhesive forces,
one of the aforementioned effects occurs. Therefore, decreasing the size of the gas
blockage can often be achieved by opening a nearby bypass, thus reducing the
discharge head. This increases the flow rate and, with it, the amount of gas carried
away by adhesion.
Still, it is always advisable to be safe and prevent a gas blockage. As the same
considerations apply both to air and to other gases, the term air blockage is used for
both cases.
Basically, two sources of air entering a centrifugal pump exist: air pockets and air
leakage. Air pockets may affect performance and may occur in the suction line, the
pump casing, or the discharge line.

AIR POCKETS IN THE SUCTION LINE

Air pockets may occur in the suction line when a portion of it has a negative slope
(Fig. 12-5). They may also occur when there is some obstruction in the upper part of
the pipeline. Such cases occur, for example, when an upstream section of the pipeline
is located eccentrically in relation to a downstream portion in such a way that a step
S is formed at the upper part of the suction line (Fig. 15-1). A gasket with an
undersized bore (Fig. 12-16) and a gasket projecting from the top of the pipe joint
(Fig. 12-17) have a similar effect. Air pockets may also occur when the suction line is
larger in diameter than the pump inlet (Fig. 15-2). Such an arrangement is very
common in systems with long suction lines and limited NPSH availability. In such
cases, the suction pipe must be made as large as practical to reduce hydraulic losses.
Air pockets created when a portion of the suction pipe has a negative slope can be
avoided with careful layout and installation of the suction line. The line must always
be installed to exclude any part from having a negative slope. Theoretically, a
perfectly level pipeline fulfills this condition. In practice, however, some positive slope

-. - - . - - . t++t-H-H ++-+-+- .......+--+- -+l1---I+++I-+-

Figure 15-1 Air pocket created by a step in a pipejoint.


Installation, Handling, and Operation of Pumps and Pumping Systems 195

--- --=.-:....=-.- -===-=:-" ::--


.~-~-

Figure 15-2 An oversize suction pipe may


enable an air pocket to form.

Figure 15-3 Steadily


rising suction line of uni-
form diameter eliminates
air pockets.

should be provided (Fig. 15-3). Generally, the suction line should have a continuously
rising slope of not less than 1%.
Air pockets created by a mismatch of the pipe centerline can be avoided only by
careful installation; whereas air pockets caused by a gasket that projects into the
pipeline can be avoided by making the bore of the gasket larger than the bore of the
pipes (Fig. 15-4).
Finally, air pockets that may occur when the diameter of the suction pipe is larger
than the bore of the pump inlet can be avoided by providing the suction line with an
eccentric reducer (Fig. 15-5) However, when the diameter of the suction line is more
than a size or two larger than the bore of the suction nozzle, a section of straight pipe
of the same diameter as the pump inlet should be installed between the reducer and
the pump. This is necessary because the eccentric reducer may deflect the liquid
stream toward one side of the impeller eye, reducing output and efficiency.
196 Problems Encountered with Centrifugal Pumps

.- -p- ,.
---- ~
,~

B
'--

I -_. \
)
(

~~.
-~~.-

Figure 15-4 Gasket opening B should be larger than the pipe diameter.

Figure 15-5 An eccentric reducer may prevent an air pocket.

When it is impractical to follow all steps required to avoid air pockets, the second
best remedy is to install air vents in all affected places. These vents should be kept
open during priming, thus allowing the air to escape.
This practice, however, should be applied only in extreme cases when there is no
other choice because the solubility of gases in liquids decreases with pressure.
Consequently, at the reduced pressure that usually exists in the suction lines, gas
may be liberated from the pumped liqUid and may collect in the spaces where air
pockets occur.
In practice, pumps sometimes operate successfully even when air pockets are
present in the suction line. This creates confused notions among pump users, and
often induces them to neglect the precautions required to eliminate air pockets. They
simply believe that air pockets are harmless.
It is important to remember that air pockets in a suction line are harmless as long
as the air remains stationary. The trouble begins when air moves and enters the
Installation, Handling, and Operation of Pumps and Pumping Systems 197

pump. The rotating impeller expels the heavier liquid outward, thus locking the air in
the impeller eye. Sometimes, this blocks the flow of liquid through the pump, causing
a complete failure. In other cases, the blockage may be only partial and liquid
continues to flow through the pump at a reduced rate. This reduced flow has a
twofold effect. First, the absolute pressure in the liquid passing through the air lock
is less than at the entrance to the suction pipe because of higher velocities through
the partially blocked areas and as a result of losses in suction pipe. This liberates
some of the air dissolved in the liqUid, increasing the volume of locked-in air. Second,
the liquid passing through the air lock carries away some of the entrained air,
reducing its volume.
Depending on the ratio between the volume of air added and the volume carried away
from the air pocket, this pocket may either increase in volume or remain unchanged.
Occasionally, it decreases until it vanishes.
In the first two cases, the air pocket significantly impairs the performance of the
pump and often stop the flow completely. In the third case, however, full performance
will be restored in a relatively short time, creating the impression that air pockets can
be tolerated. At present, it is impossible to predict which cases will occur. Thus, it is
mandatory to be safe and eliminate any air pockets from the pump.
The manner in which an air pocket can affect pump performance is illustrated by
the following example. In Fig. 12-6, a tank located at a high elevation supplies liqUid
to a pump. The pipes between the tank and the pump create room for an air pocket
at A. To prime the pump, valves 1 and 3 are opened until all the air is expelled from
the pump and the suction line, leaving only the air pocket at A. Mter this is
completed, valve 3 is closed, valve 2 is opened, and the pump is started.
During priming, the following may occur. First, the suction pipe has a generous
diameter and the flow rate delivered by the pump is relatively small, resulting in a
low velocity in space B. Consequently, the flowing liqUid will be unable to carry any
air away from the air pocket into the pump, and the pump will operate satisfactOrily.
Second, when the diameter of the suction pipe is relatively small and the flow rate
is high, the flowing liqUid carries air from the air pocket A into the pump. However,
when the velocities of the liqUid inside the pump are also high, some of the air is
removed through valve 3 as long as the pump is not put into operation. If enough
time is allowed, all air may leave the pumping system, and the pump will again
operate trouble-free.
The only time an air pocket adversely affects performance is when the velocity of
the liqUid during priming or operation is relatively high and the pump started
before all the air escaped from space A. In this case, the air that was brought into
the pump from A, after the impeller was put into rotation becomes locked in the
impeller eye.
A similar effect may be caused by air pockets generated by incorrect pipe layout
(Figs. 12-S, IS-I, and IS-2) or a gasket that projects into the suction line (Fig. 12-16).
Whenever an air pocket is located above the level of the liqUid in the suction sump
(Fig. 12-S), or at a significant distance from the pipe intake, the effect of the air
pocket is also affected by the reduction in pressure caused by its location or the
frictional losses in the suction line, or both. This reduction in pressure causes some
of the air that is dissolved in the liqUid to be released into the air pocket, thus
increasing its size. In such a case, the danger of pump failure is significantly higher
than in the case illustrated in Fig. 12-6.
198 Problems Encountered with Centrifugal Pumps

AIR POCKETS IN THE PUMP PROPER


In single-stage volute pumps, air pockets usually occur in the highest zone Z of the
casing (Fig. 15-6). During priming, the entering liqUid gradually fills out the casing,
allowing air to escape through the discharge nozzle N. However, when the level of the
liqUid reaches the tongue T of the volute, the escape route of the air is cut off and an
air pocket is created.
Analogous to the effect of an air pocket in the suction line, an air pocket in the
casing is not always harmful. Its effect depends mainly on its size. An air pocket in
the casing is subjected to two classes of force. The liqUid that flows through the
casing tends to carry the air into the discharge nozzle, deaerating the pump
completely. On the other hand, the centrifugal forces generated by the rotating
impeller tend to throw the liqUid outward and displace the air into the suction eye.
Depending on the size of the air pocket and the pump geometry, one or the other
effect dominates. This determines whether an air pocket is harmful.
To eliminate such an air pocket, an air valve V may be installed at the highest
point of the casing and may be kept open during priming (Fig. 15-7). In certain pump
geometries, it is also possible to prevent the formation of an air pocket by connecting
the affected zone with the discharge nozzle by means of a hole H. However, this
is usually effective only when the discharge nozzle is directed vertically upward
(Fig. 15-6). It is completely useless when the discharge nozzle is horizontal (Fig. 15-7).

.-.- _·_·-t·_·_-_·-
i
1
i
1
-1------
i

Figure 15-6 Hole for deaerating a volute.


Installation, Handling, and Operation of Pumps and Pumping Systems 199

~
-
I -==
-r- -
~----..~

-+---- ~ -
.~--.:==.-i-""H

Figure 15-7 Valve for deaerating


a volute.

Figure 15-8 Air pockets at the


top of the suction nozzle.
200 Problems Encountered with Centrifugal Pumps

Figure 15-9 Air pockets in a multistage pump.

In a side-suction pump, (Fig. 15-8) an air pocket may also occur in the suction nozzle
above the impeller eye (see B in Fig. 15-8). In multistage "donut" pumps, an air pocket
at the top A of the casing is usually separated from the rotating impeller I by a
stationruy diffusor D (Fig. 15-9). The liquid issuing from such a diffusor is, therefore,
not subjected to the centrifugal forces generated by the rotating impeller. Consequently,
the liquid does not displace the air from the pocket back into the impeller eye, but it
carries the air into the eye E of the next impeller, with damaging results.
Mter each stage, the pressure in the liquid increases significantly, increasing the
solubility of the air in the liqUid. Under high pressure, the flowing liqUid gradually
absorbs all the air from the air pocket. In most cases, the pressure after the second
stage can accomplish this. Thus, it is rarely necessruy to deaerate more than the first
two stages of a multistage pump.
These are the most common air pockets in the pump casings. Sometimes, however,
an air pocket may occur in an unexpected location because of a protruding fin or
bulge caused by a fault in the casting. A pocket can also occur behind a piece of
trapped solid matter.

AIR POCKETS IN DISCHARGE LINE

Air pockets in the discharge line may affect the performance of a centrifugal pump in
two cases. First, they may occur when the discharge line is lower than the centerline
of the pump (Fig. 15-10), and the pump was primed with the discharge valve V
closed. In this case, air may remain trapped between the discharge valve V and the
Installation, Handling, and Operation of Pumps and Pumping Systems 201

C
I

Figure 15-10 Air pockets in


a discharge line.

check valve C. When the valve V is opened before the pump is started, the air trapped
between V and C enters the pump casing. This, in turn, affects the performance.
Air pockets can also form when the discharge line is higher than the pump (Fig.
15-10). In this case, air trapped in the discharge line may cause a different kind of
trouble. When the pump is started, a sudden blow may be heard as if the whole
pumping system were struck by a blow from a heavy hammer. The blow may be
particularly intense when there is considerable back pressure on the disc of the
check valve.
In Fig. 15-11, D is the check valve disc, S is the seat, Al is the outer area of the
seat face, PI is the pressure downstream of disc D, and A2 is the area of the opening

Figure 15-11 Section through a typical


check valve.
202 Problems Encountered with Centrifugal Pumps

sealed off by the disc D. When the check valve is in the closed position. disc D is
subjected to a total closing force approximately equal to FI =AIXPI. To open this disc,
it is necessary to build a force F2=P2xA2 on the upstream side that is greater than
force Fl. However. area Al is greater than area A2. Consequently. pressure P2 must
become greater than pressure PI in order to build up a force F2 greater than force Fl.
As the pump is started. the pumped liquid starts to compress the air contained
upstream of the valve disc D. This continues until the pressure P2 becomes large
enough to overcome force Fl.
As soon as D begins to open. P2 spreads over Al causing the valve disc to fling wide
open under the action of the expanding air. At the same time. the excess pressure
energy stored in the compressed air is converted into kinetic energy. causing disc D
to hit the nearest stop at a very high velocity.
To eliminate this effect, a float -actuated air vent should be installed upstream of D.
This vent remains open as long as air is in the pipeline. allowing the entrapped air to
escape freely. As soon as the line fills with liqUid. though. the float closes the vent
and prevents any leakage of the pumped liqUid.

AIR LEAKAGE

Air may leak into a pump from several sources. If the water level of the source from
which the liqUid is taken falls below the inlet of the suction pipe after the pump
commences to work, air can enter through the holes of the strainer and fill out the
pump. Air can also leak through any loose joint or hole in the suction line or pump
casing.
To detect this second type of leak. one of the following methods may be applied.
l. If the suction line has a foot valve. fill the pump and suction line with water when the
pump is at rest and apply some pressure. For example. fit a standpipe to the suction line
and fill it with water or use the discharge line to serve as a standpipe. Any leak in the
suction line will manifest itself as water leakage.
2. If a pump operates in an nonflammable environment and it handles a nonflammable
liquid. an open flame (e.g .. a burning candle) may be brought near all pOints where air
leakage is suspected. A flickering flame indicates the source of the leak.
3. Pour water (or some other liquid) on the suspected points of the suction line while the
pump is operating. and observe the readings on the vacuum and pressure gauges. Leaks
are indicated by a corresponding change in the vacuum and pressure readings when the
water is being poured on the suspected spot.

Air leakage is most likely to occur in three places:


l. Flanges, bushings, nipples, manholes. drain plugs, and so forth, in the suction line (Le ..
at any jOint)
2. The suction side of the pump casing (from holes or porosities)
3. Through the stuffing box when it is on the suction side of the pump or on the pressure
side of the impeller, but connected to the suction eye through balancing holes (Fig. 12-20).
To prevent this, stuffing boxes are generally provided with a seal arrangement (Fig. 12-21).

Such a seal arrangement, however, is vulnerable to failures. For instance, the pipe
delivering the seal water can become clogged, or the water seal lantern may not be
Installation, Handling, and Operation of Pumps and Pumping Systems 203

located where the seal water enters the stuffing box. To ensure that the stuffing box
is properly water sealed, the gland should not be tightened too much. Water leaking
from the gland indicates that the water seal arrangement functions properly. In small
pumps, a water seal is often not provided and the gland must be tightened properly.
Another cause of air entering the pump through the stuffing box may be that the
part of the shaft that passes through the stuffing box is covered with a sleeve that is
eccentric in relation to the shaft center or is badly worn. Other causes may be that
the shaft itself is bent or runs eccentrically because of excessive lateral forces or
vibrates excessively as a result of critical speed or excessive strain on the bearing
housing. Air leakage through a stuffing box occurs mostly on the suction side of the
pump. However, when the impeller is provided with balancing holes and a seal ring,
(Fig. 12-20) or with back-vanes (Fig. 12-21), there may be a vacuum near the gland
on the pressure side. This requires the same precautions as when the stuffing box is
located on the suction side.
A special case of air leakage may occur when the impeller shaft is covered with a
sleeve for protection (Fig. 15-12). Here, S is the shaft, I is the impeller, SL is the
sleeve, N is the nut that presses the sleeve against the impeller, and G is the gland. In

Figure 15-12 An air leakage through the shaft sleeves.


204 Problems Encountered with Centrifugal Pumps

this case, air may enter between Nand S at a, pass between the sleeve and shaft, and
enter the eye of the impeller at b. To prevent this, such sleeves are often designed to
accommodate an 0 ring O. In practice, if the attendant forgets to reinstall the 0 ring
during re-assembly, air leakage can occur.
Air leakage into the pump may also result from the aforementioned vortices in the
suction sump.

AIR HANDLING CAPABILITY OF IMPELLERS OF DIFFERENT


GEOMETRIES

Experimental evidence indicates that an axial flow impeller can tolerate a larger
percentage of air in the liquid than an impeller of a lower specific speed. Although
there are only a few case histories that support this conclusion, there seems to be a
logical reason to this claim. In an impeller of low-specific speed, a partial blockage of
the vane passages often blocks the flow completely (Fig. 15-13). A corresponding
blockage in an axial-flow impeller will still leave an open passage for the pumped
liquid (Fig. 15-14).

Figure 15-13 An air pocket in a radial impeller.

Figure 15-14 An air pocket in an axial impeller.


Installation, Handling, and Operation of Pumps and Pumping Systems 205

IMPROPER HANDLING AND OPERATION

Faulty workmanship often leads to mechanical problems: immediate breakdown of


the pump, reduced pump life, noise pollution, or pollution from leaks. In many cases,
mechanical problems can also impair performance. For example, vibration of a pump
element or an element in the pipeline may induce cavitation in the pump proper.
Conversely, hydraulic disturbances can lead to mechanical problems. For example,
cavitation from insufficient NPSH may destroy some parts of a pump in a short time
and can be a source of noise pollution. Similar effects also occur from operating at
low flow rates.
In addition, a pump operating at very low flow rates may heat the pumping unit,
which, in turn, will cause cavitation. Heating of the pumping unit may also be caused
by a bent shaft, faulty bearings, improper lubrication, uneven tightening of the bolts
connecting various pump parts, and improper alignment between the pump and
driver. Noise, vibration, and, occasionally, increased leakage through the stuffing box
may also occur.

Figure 15-15 Section through a


vertical inline pump.
206 Problems Encountered with Centrifugal Pumps

Often, the axial load on an impeller is balanced by a wearing ring W and balancing
holes B (Fig. 15-15). In this case, the pressure in region R is often little higher than in
the suction eye. In many cases this means that the absolute pressure at R is lower
than atmospheric pressure. A bent shaft may enable air to enter this space through
the stuffing box and from there into the impeller eye.

The Effects of Careless Starting


A pump may fail shortly after starting for the following reasons:

1. It has not been adequately primed. Parts may run dry, resulting in overheating or
seizure. If some parts are overheated more than others, uneven expansion may occur.
This can distort alignment, which, in turn, can bring about a chain reaction of failures.
Seizure of rotating parts can cause fracture of the shaft or coupling. In the mildest case,
it will blowout the fuses.
2. Whenever certain parts of the pump reqUire cooling, care should be taken to ensure that
an adequate amount of coolant flows through the cooling jacket before the pump is
started. Otherwise, the affected parts may overheat. The amount of the cooling liqUid
also affects pump operation. There are no fixed rules regarding what this amount should
be. Even with the same pump, it varies with the operating conditions and the environ-
ment. It may be as low as 2 GPH or as high as several GPM. A good indication is the
temperature of the cooling liqUid at the exit from the cooling jacket. As far as possible,
this temperature should be kept below about 140°F. However, when the cooled parts
require lubrication, the final temperature of the coolant should not be kept too low to
avoid increasing the viscosity of the lubricant.
3. Some liquids are so viscous at room temperature that they cannot be handled by any
centrifugal pump. To overcome this problem, these liqUids are usually heated before they
enter the pump (viscosity decreases with an increase in temperature). Failure to preheat
the pumped liquid to the proper temperature before it enters the pump may cause some
of the impeller blades or the shaft to fracture.
4. Special precautions are required when starting a pump for the first time after it has been
installed on site. The first step is to check the alignment between the pump and the
driver. Even when these two elements have been aligned by the manufacturer on a
common baseplate prior to shipment, they usually shift out of alignment during
transportation or during installation. The next step is to try turning the unit by hand. A
piece of solid matter (e.g., bolts, wrenches, or some other tool) can fall inside the pump
during installation and jam the impeller against the casing. Finally, the direction of
rotation must be checked. This is best done with the pump uncoupled from the driver to
prevent shaft or impeller nuts from coming loose, which usually causes damage to all
rotating parts.

FAULTS IN WORKMANSHIP AND THEIR EFFECTS

The Effect of a Blow against a Pump Part


When a smooth surface is subjected to a blow with a hard object, a dent is made
(Fig. 15-16). This dent causes some material to be displaced and rise above the
affected surface. If a part with a flat surface is tightened against this spot, these two
surfaces will assume a nonparallel position (Fig. 15-16).
Installation, Handling, and Operation of Pumps and Pumping Systems 207

Shaft (will bend


when sleeves
are tightened)

~ Extrusions

~ Dent

Figure 15-16 Effect of a blow against a flat face.

In a pump, such cases are not rare. Dents on the face of the impeller hub or on the
shoulder against which it bears will prevent the impeller from running square to the
axis. This, in turn, will cause rubbing of the wearing rings, resulting in excessive
power consumption.
In a long shaft covered with sleeves, a dent on the face of a sleeve may bend the
shaft when the sleeves are tightened against one another (Fig. 15-16). If the shoulder
against which a ball bearing bears is damaged, the bearing may overheat. Similar
effects occur when some dirt is clamped between two flat surfaces and when surfaces
are not machined exactly perpendicular to the axis of the shaft.
When a pump part falls on a hard surface, there is always a danger that it will be
damaged in a manner that will adversely affect pump performance. A shaft may be
bent and thus heat the bearings, cause wear on the wearing rings, create noise and
vibrations, or cause increased power consumption. A blow against the impeller may
bend the shrouds inward, reducing the width of the impeller waterways, resulting in
a reduced output of the pump. The most straightforward way to restore the
performance of a pump that has been damaged in this way is to straighten the
shrouds. In certain cases, the effects of the reduction in the width of the impeller
waterways can be compensated for by enlarging the throat area of the volute.

Misalignment between the Pump and Driver


Whenever a pump is coupled directly to the driver, any misalignment can cause the
pump to operate with noise and vibration and the bearings to overheat. In severe
cases, this may lead to more serious effects. For example, vibrations may loosen
bolts, increasing misalignment and further aggravating the effects. In extreme cases,
208 Problems Encountered with Centrifugal Pumps

Figure 15-17 Axial mismatch between the impeller


and the casing.

this can lead to early breakdown of the pump or some of its parts. Overheating the
bearings can also lead to their early destruction even without loosening any bolts.
In a belt-driven unit, improper alignment between driving and driven pulleys can
cause the belt to fall off or break frequently. It may also overheat the bearings.

Axial Position of the Impeller


In Fig. 15-12 the axial position of the impeller is determined by two nuts N on the
shaft. When assembling such a pump, care should be taken that the centerline of the
impeller discharge should coincide with the centerline of the casing. Any mismatch
between these centerlines may adversely affect performance (Fig. 15-17). This is
particularly apt to occur when the width B of the casing is only a little greater than
the width b of the impeller.
CHAPTER 16
Problems \Nith Bearings
Bearings support the rotating elements of the pump and keep them in proper
position relative to other parts and the driver. Failure of a bearing may change the
position of the rotating elements, which will cause interference between the
stationary and moving parts of the pump, and may lead to fractures or seizing of
parts.
Certain bearing failures may cause binding of the shaft and can result in its
fracture. In other cases, faulty bearings may cause local overheating of pump parts.
resulting in a distortion of their alignment with other, cooler parts of the pump. This
again may lead to a great number of problems, as discussed throughout this book.
Generally. there are two distinct types of bearings: sliding and rolling. While their
failures may have similar effects on a pump and its performance, each type requires
different handling and care.

SLIDING BEARINGS

A sliding bearing is used to support and keep the rotating elements in place in both
the axial and radial directions. Radial support usually consists of a cylindrical shell
of suitable material and dimensions, mounted in a rigid housing. Axial support
usually consists of rigid rings that are fixed to the bearing housing and that bear
against rotating collars that are firmly attached to the rotating element. Sometimes.
these collars and rings are conical or spherical to provide support both radially and
axially.
A special form of sliding bearings are self-adjusting pads that provide better
lubrication, lower friction, and higher loads. However, they are usually much more
expensive than bearings with solid shells.

Advantages and Disadvantages


Compared with roller bearings, sliding bearings have the following advantages:

1. Whenever the shaft is subjected to steady forces, the loaded parts of the bearings are
subjected to a constant stress, which reduces the danger of failure caused by fatigue.
2. When made of suitable materials and properly proportioned and supported, sliding
bearings are often capable of operating within the pumped liquid, which both lubricates
and cools them.
209
210 Problems Encountered with Centrifugal Pumps

3. With proper cooling and lubrication, at high speeds sliding bearings can support much
higher loads than rolling bearings.

Sliding bearings have the following disadvantages:


1. Their coefficient of friction is often 10 to 15 times greater than that of rolling bearings.
This results not only in costly waste of energy, but increases the rate of wear of the
mating surfaces.
2. The higher friction coefficient often increases the temperature of the lubricant to the
extent that it is mandatory to install elaborate and costly cooling systems.
3. Sliding bearings usually occupy much more space than rolling bearings designed for the
same duties. This makes them, in most cases, more expensive.
4. Even with proper cooling, the heat generated at the mating surfaces of the bearing is
greater than that at rolling bearings. This accelerates the oxidation of the lubricant and
contributes to the loss of many of its lubricating properties. The relatively high coefficient
of friction also accelerates the contamination of the lubricant with particles worn away
from the mating surfaces.

These factors make proper lubrication and frequent oil changes more critical. The
preceding discussion is only one example of how improper care may lead to bearing
failure.

Problems with Sliding Bearings


Generally, sliding bearings are apt to fail for one or more of the following reasons:
1. Improper lubrication. This includes the quality of the lubricant used, as well as the
frequency of oil changes.
2. Inadequate cooling oj the lubricant (whenever applicable). This may be caused by the
failure of the cooling system or the attendant's failure to open the valve that delivers the
cooling medium before starting the pump.
3. Bearings running dry. This is due to the attendant's failure to ensure that there is
enough lubricant in the bearing housing before the pump is started.
4. Contamination oj the lubricant. This is due to inadequate attention to details (e.g.,
ignoring the need to store the lubricant in a clean closed container; using a dirty refilling
funnel, not rinsing the oil reservoir during each oil change, and not making sure that all
oil seals are in good shape).
5. Misalignment. Misalignment may occur during reassembly after a pump has been opened
up for repair or other maintenance. Dirt may get clamped between the bearing bracket
and the pump casing, owing to an extrusion caused by a blow (Fig. 15-16) or to uneven
tightening of the bolts. Misalignment may result in excessive loads on the bearings,
bending of the shaft, or metallic contact between the stationary and the rotating parts,
which often results in excessive wear or seizure. Externally, misalignment often makes
itself evident by excessive heating of the bearing housing, which, in turn, accelerates the
oxidation of the lubricant and may cause bearing failure owing to improper lubrication.
6. Loose bolts, Another source of problems caused by bearings appears when the bolts that
keep the bearing bracket in position are not evenly and adequately tightened or become
loose during pump operation (e.g., from vibration or omission of the necessary locking
devices). In such a case, the bearings may move from their original position by so much
that all loads will bear against the impeller rings or packing. This may ruin the impeller
or destroy the packing in a very short time.
Problems with Bearings 211

ROLLING BEARINGS

Rolling bearings basically consist of two rings or races and a set of balls or rollers
that fit snugly between the races. The balls or rollers are kept apart by a cage made
of sheet metal, brass, plastic, or any other suitable material.
The main advantages of rolling bearings are the following:

1. The intial cost is low.


2. They can run unattended for prolonged periods
3. They usually require smaller and less expensive housings than sliding bearings used for
the same duties
4. Replacements are readily available from many sources.
5. They are energy efficient. Their low coefficient of friction requires changing the lubricant
less frequently than for sliding bearings. In fact, many rolling bearings are supplied with
sealed-in lubricant that lasts for their entire life. The low coefficient of friction also makes
the need for cooling less frequent than for sliding bearings.

The main disadvantages of rolling bearings are the following:

1. Races and all rolling elements are subjected to rapid, intermittent stresses, making them
susceptible to fatigue failure.
2. Most classes of rolling bearings require special precautions and care during assembly
and disassembly.
3. Special care is required in using the exact amount of lubricant (neither too much nor too
little) .

Lubrication of Rolling Bearings


Inadequate lubrication causes bearings to deteriorate very rapidly. Similarly, exces-
sive lubrication can also shorten the life of a rolling bearing. Too much lubricant
causes the bearings to run hot, thus accelerating the rate of oxidation of the
lubricant. This will ultimately result in an early failure of the bearings.
Failures from improper lubrication manifest themselves in one or more of the
follOwing ways:

1. Absence of lubricant in the bearing housing.


2. Water in the lubricant and the bearing housing.
3. Discoloration of races and balls (rollers).
4. Spalling (Le., a series of isolated and irregular indentations distributed over the periphery
of the races, Fig. 16-1).
5. Smearing (i.e., rows of curved clusters of hairlike marks on the races, Fig. 16-2).
6. In extreme cases, the heat generated owing to the absence of lubrication may cause some
of the balls (or rollers) to become welded to the races.

Most pump manufacturers offer a choice of grease or oil lubrication. It is up to the


pump user to decide what kind of lubrication is best for a particular application. An
improper choice of lubricant can lead to early bearing deterioration.
212 Problems Encountered with Centrifugal Pumps

Figure 16-2
Smearing in
Figure 16-1 Spalling a ball thrust
caused by edge loading. bearing.

Grease. The advantages of grease lubrication are the following:

1. Grease can be retained without elaborate enclosures, even in vertical or inclined shafts.
2. Some calcium-based greases can seal out moisture. a feature that may be very desirable
when a pump must operate in a humid environment.
3. Some lithium-based greases can protect the bearings against corrosion.
4. Stiff greases seal out contaminants.
5. Filtered greases, with high-viscosity oil mixed in, allow bearings to operate relatively
quietly.
6. Grease requires less frequent refills and relubrication than does oil. In general, the
frequency of relubrication varies widely. Shielded ball bearings prelubricated with sealed
in grease have lasted up to 20 years. In other cases, however, frequent relubrication has
proven to be necessary. In a promotional pamphlet entitled "A Guide to Better Bearing
Lubrication," issued in 1979 by SKF Industries Inc., the relubrication periods shown in
Fig. 16-3 are suggested. These periods are recommended for moderately loaded bearings
at 70°F (21°C). For each rise in temperature by an additional 25°F (14°C), these time
intervals have to be halved. The data presented in Fig. 16-3 should serve only as a
general guide. The optimal time intervals for a given application can be determined only
on the basis of actual observations carried out over a prolonged period of time.

The main disadvantages of grease lubrication are the following:

1. Grease-lubricated bearings are very difficult to cool effectively. This precludes the use of
grease for very high speeds and high loads, where much heat is being generated owing to
friction.
Problems with Bearings 213

b
20000
15000
a
20000 10000
a; 15000 7500
i::c;
~ B 10000 5000
.- III
.§ ~
1ii 0 5000 ?5OO
.go
.J:JUl
..2 S 3000 1500
~~ 2000 1000
1500 7~

1000 500

500 250
~Oo 1( 00 2(ioo 10000
Relubrication interval n r/min
a Radical ball bearings
b Cylindrical roller bearings, needle roller bearings
c Spherical roller bearings, taper roller bearings, thrust ball bearings

Figure 16-3 Time interval for relubrication (T/), as recommended by SKF Industries (1979).

2. The viscosity requirements of greases for different applications varies significantly with
temperature. For low temperatures, a low-viscosity grease must be used. Similarly, for
high temperatures, high-viscosity greases are recommended. Consequently greases are
not suitable for environments with significant temperature fluctuations.
3. A grease suitable for the operating temperature of a bearing may be too stiff during
starting. Particularly when a pump is located outdoors and must be started on a cold
winter morning. Such an excessive stiffness of the lubricant may easily ruin the bearing.
4. It is very difficult to determine the actual amount of grease in the bearings. This can
easily lead to under- or overlubrication.
5. Changing the lubricant also poses certain problems. In some cases, bearings may have
to be dismantled, cleaned of the old grease, reassembled, and relubricated. However,
whenever the housing has a drain plug, the old grease can often be flushed out by means
of a light oil, or transformer oil, preheated to between 200°F and 240°F.

Oil. The main advantages of oil lubrtcation are the following:


1. The oil level can be determined easily and kept constant.
2. Oil can be cooled easily. In fact, the lubricating oil by itself is often used as a coolant for
the bearings, making it particularly suitable for high speeds and loads, and for operating
at high ambient temperatures.
3. Certain oils have a very high viscosity index (i.e., their viscosity changes very little with
temperature). This makes oil particularly suitable for use in environments in which
significant temperature fluctuations are likely to occur.
4. Oil change is usually significantly easier to carry out than grease change.
5. Some oils have a lower coefficient of friction than grease. This makes them more suitable
for extreme loads and speeds.
214 Problems Encountered with Centrifugal Pumps

The disadvantages of oil lubrication are the following:


1. It is more costly.
2. It requires more frequent changes of lubricant than does grease.
3. For vertical shafts, it requires an elaborate and costly design of the bearing housing.
4. It is less suitable for humid and corrosive environments than is grease.

Lubrication Procedures
In relation to cleanliness, proper choice of lubricant, provisions against running
dry, and so forth, the same rules apply to rolling and sliding bearings. Rolling
bearings, however, require two additional provisions:
The first relates to cleanliness. Dirt has many more places to hide in a rolling
bearing than in a sliding bearing. This makes it mandatory to exert much more care
in cleaning. The second provision concerns the amount of lubricant in the bearings.
All rolling bearings run hot when full of lubricant. Because an increase of about 25°F
in temperature doubles the rate of oxidation of the lubricant, hot running should be
avoided as much as possible. As a general rule, rolling bearings should not be filled
with lubricant to more than one third to one half of their capacity.

Assembly, Dismantling, and Storage


One of the most important precautions required during assembly and dissaembly
of rolling bearings is cleanliness. Both the bearing and the housing must be carefully
inspected for any traces of dirt or other inclusions. If they have to be dismantled for
more than several hours, or for a shorter period of time in a dusty environment, they
should be protected by wax paper or lintless rags.
Before reinstallation, the cleanliness of a bearing should be checked both visually
and by rotating it slowly by hand. A clean bearing in good condition should tum
evenly, without any trace of change in the smoothness of its motion. It should also
show no signs of discoloration or rust.
Before installation, rolling bearings should be inspected and checked for scratches,
signs of erosion, or excessive clearances between the races and the balls. In most
cases, it is less expensive to discard a suspicious bearing before installation than to
have to dismantle a pump later because of a bearing failure.
A rolling bearing usually has an interference fit with either the shaft or the
housing. When the interference fit is on the shaft, the bearing should never be forced
in place by applying pressure to its outer race (Fig. 16-4). Similarly, when the
interference fit is in the housing, only the outer race should be subjected to force
during mounting (Fig. 16-5). During assembly, care should be taken to ensure that
surfaces of the bearing races should move parallel to the mating surfaces of the shaft
or the housing. Any cocking of a bearing while it is being driven into position may
damage both the races and the rolling elements.
When outside the pump, a rolling bearing should never be spun with compressed
air. If this is done while there is still dirt in the bearing, it will certainly cause
damage. However, even after a bearing has been cleaned of all dirt, the high speed
imparted to a race by the compressed air may easily cause damage from lack of
proper lubrication.
Problems with Bearings 215

Incorrect Correct
I I

Figure 16-4 Mounting of a bearing


with inner interference fit.

Incorrect Correct
I

Figure 16-5 Mounting a bearing


with outer interference fit.
216 Problems Encountered with Centrifugal Pumps

Some rolling bearings have an asymmetrical design. They will successfully resist
axial loads applied in one direction, but will not sustain any loads in the opposite
direction. When installing such a bearing, care should be taken to ensure that its
orientation conforms with the pump manufacturer's recommendations.
If a pump must remain idle for a prolonged period of time (e.g., 3 years), the
bearing frame should be dismantled and the bearings cleaned and relubricated before
starting .. Also, whenever a pump must remain idle for a significant period of time
within a humid or a corrosive environment, the bearing should be covered with a
rust-preventive compound. Failure to take these precautions may lead to premature
bearing failure, which, in turn, will initiate a series of additional pump problems.

Failures of Rolling Bearings


Some causes of failure were discussed earlier. Other causes include the following:

1. Inadequate or improper lubrication invariably causes damage to the mating surfaces.


The first sign of inadequate lubrication may be discoloration of the running surfaces.
When allowed to run without adequate lubrication, damage progresses rapidly to
failures that are not much different from fatigue failure. In more advanced stages,
spalling may occur, as shown in Fig. 16-6.
2. Improper sealing of the bearing housing may allow dirt or water to mix with the
lubricant or may enable the lubricant to leak out of the housing.
3. Fatigue may be due to radial or axial loads, or both. In unidirectional radial loads, the
race of the rotating ring will show signs of wear at the center, along the entire
periphery. The stationary ring, however, will exhibit signs of wear only over a small
portion of the periphery, at the location that has been subjected to the load. In a pure
axial load, the signs of wear appear on the opposite face of each ring and extend over
the total area of the races. In a pure axial load, the width of the wear marks are
constant. In combined axial and radial loads, the width of the wear marks on the
stationary ring are widest in the direction of the radial load component.
In centrifugal pumps, the most common causes of premature bearing failure due to
fatigue are excessive operating speed, excessive belt tension, excessive suction
pressure, and excessive total head.
4. Faulty mounting practice is often recognizable by a bearing that runs hot or operates
noisily at the time of installation. Internally, such a bearing often shows isolated signs
of damage, while the rest of it is in perfect condition.

Figure 16-6 Spalling caused by inadequate lubrication.


Problems with Bearings 217

5. A too-tight fit is often recognizable externally because it is usually much more difficult
to turn the shaft by hand, than it is with a normal fit. A too-tight fit results in the
bearings running hot and also brings about premature fatigue failure. Internally, this
fequently results in an axial crack in one (or both) of the races. It may also show signs
of smearing (Fig. 16-7), generated during installation.
6. Too-loose fits are usually indicated by wear or fretting signs (Fig. 16-8) on the outside
of the outer ring or on the inside of the inner ring. Very often, a too-loose fit also
causes discoloration and wear of the bearing seats.
7. Defective seals on the shaft or housing are often indicated by excessive localized wear
or cracks in the bearing races.
8. The effects of misalignment are often similar to the results of excessive bearing
loading.
9. Vibrations transmitted to the bearings from other sources while the pump is not in
operation can also lead to failure. During standstill, each ball of the bearing is in
continuous contact with the same spot of its races. When such a bearing is subjected

Figure 16-7 Smearing caused by installing


in a housing with a too-tight interference fit.

Figure 16-8 Fretting corrosion.


218 Problems Encountered with Centrifugal Pumps

Figure 16-9 False brinnelling.

to significant vibrations for a prolonged period of time, an effect known as "false


brinnelling" (Fig. 16-9) may occur. This means that circular indentations are generated
in the races at the spots where they are in contact with the balls.
10. Passage of electric current may cause pitting of the races or the rolling elements, or
both. Such cases are known to have resulted from electrical repair (e.g., welding)
performed on the pump on site.

Figure 16-10 Vertical shaft supported by an angular contact


bearing.
Problems with Bearings 219

Less Evident Causes of Rolling Bearing Failure


Figure 16-10 is a typical arrangement of a vertical wet pit pump. Impeller I is
mounted on shaft S, which is suspended from an angular-contact thrust bearing T
and supported by a sliding-line bearing L. Sometimes the impeller is provided with
balancing holes B, to reduce the axial thrust. Such an arrangement runs satis-
factorily with relatively long and heavy shafting. However, when the shaft is short and
of small diameter, its weight is not heavy enough to keep the races of the bearing Tin
permanent contact with the balls, and the pump vibrates and operates noisily. In
extreme cases, this can ruin the pump in a relatively short time. One of the simplest
remedies for such a case is to plug the balancing holes B, thus increasing the axial
load on the bearing T.
Whenever an impeller is subjected to high axial loads, the thrust bearing T should
be located on the outboard side of the bearing frame (Fig. IS-IS) and the line bearing
L on the inboard side.
For bearings to operate satisfactorily, their seats should be machined accurately,
allowing no radial play between the bearings and the casing. If the fit between the
line bearing L and its seat is too tight and the fit between the thrust bearing T and its
seat is less tight, the axial load may be taken up by the nearer line bearing instead of
by the thrust bearing. This can ruin the line bearing in a very short time. A similar
effect can occur even if the casing is machined to the correct tolerances, but some
dirt has been wedged between the outer race of the line bearing and the bearing seat.
To prevent this from happening, the line bearing L should be able to move freely in
the axial direction. But while it is important for the bearing to be able to move axially,
it should not be so loose that its outer race can rotate against its seat. In practice,
however, such cases are common. The result is always a breakdown of the bearings
in a short period of time.
CHAPTER 17
Sealing Rotating Parts
PACKED STUFFING BOXES

One of the first problems that arises with packed stuffing boxes is the choice of proper
packing material. Until recently, the most popular kinds of packing were asbestos-
filled, Even today, these packings are still regarded by many as the most versatile
and economical. However, because the use of these packings might constitute a health
hazard under certain circumstances, substitutes are being used with increasing
frequency.
The most promising asbestos substitutes are filled with either graphite yarn,
aramid fibers, or fiberglass. Unfortunately, both aramid-filled and graphite yarns are
very expensive. Graphite yarns also tend to oxidize under normal atmospheric
conditions, and aramid fibers are not suitable for use in temperatures above 260°F.
Fiberglass-filled packings seem to be economically competitive, but ordinary
fiberglass is rather abrasive and has a tendency to soften at higher temperatures. To
compensate for these deficiencies, mixtures of the fibers named above are sometimes
used, and attempts to develop better fibers are ongoing.
Even when an asbestos substitute is perfectly suitable for a given application, it
often requires special handling. For example, packing containing aramid fibers or
fibers made of other thermoplastic material usually reqUire more cooling than
asbestos-filled materials. They may also require special running-in periods, during
which a pump must be started and stopped at frequent intervals until the packing
starts to perform properly.

Handling, Installation, and Troubleshooting


Packings are sold in two forms: cut-to-size and continuous braided rope. The pre-
cautions required during installation of precut packings are as follows: they should
be amply lubricated before insertion into the stuffing box, the butt joints of each ring
should be offset by apprOximately 90 degrees relative to the neighboring rings, and
the lantern ring (Fig. 12-21) should be located in its proper position [Le., opposite the
inlet of the cooling (sealing) liquid).
When the packing rings are cut from a continuous rope, they must be cut to the
exact length required. The ends of each ring should just touch when wrapped around
the shaft. Segments that are too long or too short will cause excessive leakage and
may lead to early deterioration of the packing.
221
222 Problems Encountered with Centrifugal Pumps

Special Problems
A special problem was encountered with a pump in which bore B of the stuffing
box was eccentric relative to the centerline of the shafts by as little as 0.004 in. (Fig.
17-1). As the space between the shaft and the bore was filled out with a soft packing,
it was thought that such an eccentricity would have little adverse effect on pump
operation. In particular, that shaft (2 in. in diameter) operated at only 1470 RPM.
In reality, however, the eccentricity made the shaft act as a grinding machine,
grinding away the packings at the rate of about 3 rings in 4 hr and converting them
into a paste. This paste, in turn, squeezed out through the clearance between the
gland and the stuffing box, leaving the stuffing box empty and ineffective.
After the stuffing box cover was replaced by a concentric one, the grinding stopped.
One possible explanation for this mystery is that the eccentricity caused the packing
to be more compressed on one side of the shaft than on the other. This, in turn, may
have resulted in a sort of a plastic flow of the packing under the influence of the
rotating shaft.

MECHANICAL SEALS

The most common problems with mechanical seals are: short life, leakage,
overheating, high power consumption, and allowing air to enter the pump. Seals are
designed in a great variety of shapes and sizes. Some problems relate only to specific
types or brands of seal. It is therefore best to consult the pump manufacturer when a
particular seal fails repeatedly, but some rules apply to all seals.

I,

I, s
I,

II
'I

Figure 17-1 Cause of packing grinding.


Sealing Rotating Parts 223

Dirt is one of the greatest enemies of mechanical seals. To seal effectively, the faces
of a mechanical seal have to be flatter than 0.00002 in. Any trace of dirt trapped
between the sealing faces will not only result in leakage but will lead to early
deterioration of the seals.
Other problems occur when the faces are not exactly perpendicular to the axis of
rotation. During operation, this may lead to rapid separation of the sealing faces on
one side of the seal, and a return to normal on its opposite side. The effect is twofold:
some liqUid leaks, and wears away particles from the seal faces, leading to their early
deterioration; dirt may enter between the sealing faces.
The lack of perpendicularity of the sealing faces may be caused by a number of
faults in other parts of the pump, including a bent shaft, worn-out bearings, dirt
clamped between mating faces of pump parts, unevenly tightened bolts, strain
imposed by the pipe lines, and so forth.
Effects similar to those caused by a lack of squareness may be caused by vibration
that, in turn, may be caused by any number of factors, such as loss of balance owing
to a solid object becoming wedged between two impeller vanes, misalignment between
pump and driver, resonance between the operating speed of the pump and the
natural frequency of the bedplate or of the foundations, and so forth.
Yet another source of seal trouble is excessive pressure. To prevent leakage, the
rotating part of the seal is kept in contact with its stationary part by means of a
spring or a series of springs. These springs have to be compressed by a pre-
determined amount. This is achieved by installing the shaft collar at a predetermined
distance from the sealing plane. Whenever this collar is mounted too close, the spring
exerts excessive pressure on the sealing faces. This, in turn, may cause increased
power consumption, which leads to overheating. Increased temperature can convert
the thin lubricating film of liqUid between the sealing faces into vapor, which destroys
the lubricating properties of the liqUid film and results in seal failure. Overheating
can also make rubber parts of the seal brittle. This, too, leads to premature seal
failure.
Another kind of problem may arise when a single seal that is meant to seal ag8.inst
inner pressure operates under a vacuum. This may happen when an impeller is
provided with balancing holes (Fig. 12-20) or back-vanes (Fig. 12-21). When the
vacuum within the space in which the seal is operating exceeds about 17 in. Hg, the
seal is likely to open and allow air to enter the pump.
SpeCial problems arise when the pumped liqUid contains abrasive particles or
dissolved solids which may crystallize during standstill, particularly when it is a hot,
saturated solution. In the latter case, the dissolved solids crystallize when the liqUid
cools off and cause serious damage when the pump is restarted. In both cases, a
clean external liqUid should be supplied at a pressure higher than the pressure
existing in the stuffing box. In this way, it may be possible to keep the pumped liqUid
away from the sealing faces. If the access of the pumped liqUid to the sealing faces
cannot be eliminated, special attention should be given to the choice of the
mechanical seal. This is best accomplished by consulting the seal manufacturer. As a
further precaution, the stuffing box (and the rest of the pump as well) should be
thoroughly flushed with a clean liqUid immediately after operation of the pump has
been interrupted. Finally, whenever a seal requires cooling, it may fail as a result of a
fault in the cooling system. For example, the attendant may fail to turn the cooling
water on prior to starting the pump, or scale may build up on the cooling surfaces.
224 Problems Encountered with Centrifugal Pumps

Diagnosing Causes of Seal Failures


Very often. the wear pattern and other features of the sealing faces may be helpful
in determining the cause of a seal failure.
A mechanical seal consists essentially of two rings (Fig. 17 -2), A and B, operating
against each other. The sealing face of the so-called primary (rotating) ring A (which
is usually made of a harder material) is generally narrower than that of the secondary
ring B. In addition to the sealing rings, a mechanical seal is equipped with two
secondary seals: seal D seals the rotating element against the shaft, and seal E seals
the stationary ring against its seat. A secondary seal may consist of an 0 ring,
bellows, elastometric cup or ring, or any other suitable device.
During operation, the matching faces of the sealing rings may develop wear. When
a seal fails, the location and shape of the wear pattern can often reveal the source of
the failure.
Under normal operating conditions, the primary ring A generates a uniform wear
pattern on the secondary ring B of the same width as the sealing face of the primary
ring A. If a leaking seal shows such a regular wear pattern the leak is usually not
through the engaging seal faces but is due to the failure of either or both the
secondary seals D and E. Such failures are usually due to faulty installation or
incompatibility of the secondary seal material with the chemical composition or
temperature of the pumped liquid.
When subjected to excessive pressures, the primary seal A may distort, as shown
(very exaggerated) in Fig. 17-3. This reduces the contact between the sealing faces to
their outer diameters alone, which in turn, causes heavy wear on the outside of the

Figure 17-2 Principal components


of a mechanical seal.

Figure 17-3 Distortion of a primary ring from overpressurization.


Sealing Rotating Parts 225

secondary ring B, and no contact on the inside (Fig. 17-4). It may also cause the
outer rim of the sealing face of the primary ring to chip. During operation, such a
seal may show little or no leakage at high pressures, but it will leak steadily at low
pressures.
When a seal is not cooled adequately, the primary ring may be subjected to thermal
distortion, as shown ( in a greatly exaggerated form) in Fig. 17-5. This produces
identical marks on both sealing rings as in the case of excessive pressure, but on the
inside of the sealing faces. During operation, a seal subjected to such thermal
distortion usually leaks steadily when the shaft is rotating, but it shows no leakage
when the shaft is stationary.
Whenever the wear pattern on the secondary ring is wider than the sealing face of
the primary ring, there is serious misalignment within the pump proper. Such a
misalignment may be due to faulty bearings, dirt clamped between two mating faces,
or an extrusion caused by a blow against a machined surface (Fig. 15-16).
Misalignment can also be caused by uneven tightening of bolts.
If a seal leaks steadily when the shaft is stationary or rotating, mechanical
distortion of the secondary ring may be the cause. Sometimes two large contact spots
will appear on the sealing face of the secondary ring (Fig. 17-6), and the faces of the

Figure 17-4 Mark on a secondary ring


caused by overpressurization.

Figure 17-5 Thermal distortion of a primary ring.


226 Problems Encountered with Centrifugal Pumps

Figure 17-6 Two high spots


in a secondary ring.

Figure 17-7 Contact over more than 180


degrees of the secondary ring.

primary ring may also be damaged. Mechanical distortion can also cause other
patterns on the face of the secondary ring. Sometimes it shows contact over more
than 180 degrees of the face of the secondary ring (Fig. 17-7) and no contact on the
remaining part of the sealing face. In other cases, it shows more than one or two
contact spots on the secondary ring (Fig. 17-8).
Some likely causes of mechanical distortion are the following:

1. Gland plate distortion owing to uneven or excessive torque applied to the bolts of the
gland plate
Sealing Rotating Parts 227

Figure 17-8 Multiple contact spots on a stationary


Wear
ring.

Figure 17-9 Heat checking cracks on a stationary


ring. Heat checking

2. The gland surface that is in contact with the secondary ring is not flat
3. Nicks and burrs on the gland surface that is in contact with the secondary ring

When a leaking seal shows no sign of wear on the sealing faces, the rotating seal
element may not be secured adequately to the shaft [e.g., the screw S (Fig. 17-2) was
not locked). This allows the seal to remain stationary while the shaft rotates. The
compression spring C may be ineffective. This may occur whenever the rotating
element is not fixed in the correct pOSition on the shaft or when the spring is clogged
with solid matter.
228 Problems Encountered with Centrifugal Pumps

Figure 17-10 Heat checking marks on


isolated spots.

When wear is only on one side of a ring. the face of the ring is not square with the
axis of rotation.
Chipped edges of sealing faces are often caused when the liquid entrapped between
these faces flashes into vapor. This causes a significant separation of the faces and
an increase of pressure within the separation zone. This increase in pressure causes
the vapor to condense violently. As a result. the ring faces slam against each other.
causing the sealing faces of the primary ring to chip. Very often this also causes
surface cracks on the secondary ring (Fig. 17-9); this is usually referred to as heat
checking. The most common symptoms of this effect are the following: the seal leaks
steadily. whether or not the shaft is rotating. and sounds occur from flashing or from
face popping. To eliminate the second problem. the suction pressure of the incoming
liqUid should be increased to prevent vapor flashing. An alternative remedy is to
increase the cooling rate of the seal faces.
Chipping may also be caused by vibration. that. in turn. may be caused by cavit-
ation. misalignment, or something else. Chipping also often results from mishandling
during assembly.
Very high wear on the sealing faces may be due to one or both of the following:
1. Poor lubrication of the mating faces caused by excessive pressure or overheating. This.
again. may be the result of improper location of the rotating seal on the shaft. improper
cooling. or a poor choice of seal.
2. Abrasive particles embedded in one (usually the softer) face of the rings.
In some cases. seal failure may result from several factors acting simultaneously.
For example. when ring distortion and a thermal distressed surface occur
simultaneously. heat-checked areas occur only at certain isolated spots (Fig. 17-10).
CHAPTER 18
Miscellaneous Studies
NOISE POLLUTION

Noise is a sequence of rapid variations in the pressure of the air. When these
pulsations reach the human ear, we encounter it as sound or noise. The source of
these pulsations is vibrations of some object or medium that is in direct or indirect
contact with the atmosphere.
When an object vibrates, it causes rapid compressions and expansions of the layer
of air with which it is in contact. These rapid changes in air pressure are propagated
through the atmosphere in the form of pressure waves that reach the human ear.
This alone is not enough to make someone hear a sound. For the human ear to
percieve a sound, both the intensities and frequencies of these pulsations must lie
within certain limits. The normal human ear hears frequencies between 30 and
20,000 cycles per second (cps). The lowest intensity lies at 0.000000003 psi. Theor-
etically, there is no upper limit for the intensity of pressure pulsations that can be
heard by the human ear. However, an intensities higher than 0.0003 psi to 0.003 psi
cause temporary deafness; intensities around 0.03 psi or higher cause permanent
deafness.
In practice, it is customary to use the decibel as a measure of sound intensity.
Figure 18-1 shows the relation between this unit and the pressure fluctuations in psi.
Because vibrations are the direct source of sound, it is inconvenient to discuss the
causes of each of these effects separately. Therefore, this discussion makes no
distinction between noise and vibration.

Hydraulic Sources
A spectrum of noise-related effects are associated with the hydraulic performance
ofa pump.
Pressure Pulsations. In a Single-stage volute pump, the blades of an impeller
move past the volute tongue with a frequency equal to the number of impeller vanes
multiplied by the rotating speed of the pump. This causes pressure pulsations in the
liquid, whose frequency F is equal to
F=Nz (18-1)

where z = number of blades


N = speed of impeller, in revolutions per second
229
230 Problems Encountered with Centrifugal Pumps

160

140

120
.0
0100
:i-
·00 80
c
Q)
C
60

40

20

0.3 psi

Intensity, Db Sound level


20 Whisper
40 Quiet residential area
60 Small office
80 Public office
100 Heavy traffic
120 Rock orchestra Figure 18-1 Intensity of sound in decibels
140 Air drill versus psi.

Depending on the intensity and frequency of these pulsations, the human ear may
perceive a sensation ranging from an almost inaudible whisper to an unbearable
noise.
In a multistage volute pump, both the intensity and frquency of these pressure
pulsations depend largely on the relative angular positions of the impeller vanes and
the volute tongues. When the impellers and volutes are located so that the blades of
all impellers pass the tongues of all volutes simultaneously, the highest intensities
are produced. On the other hand, when they are arranged so that only one impeller
blade at a time passes one volute tongue. this produces the highest frequency.
If a pump is provided instead of a volute with diffusor blades. frequencies Fl of the
pressure fluctuations at each individual stage are equal to

Nnz
F[ =-- (18-2)
m

where n = number of vanes in one diffusor


m = largest common factor of z of impeller vanes and n of diffusor blades.

The relative angular position of the individual impellers and diffusors are subject to
the same rules as are multistage volute pumps. In a multistage diffusor pump. the
relative angular position of the diffusor rings can be changed. With certain construc-
Miscellaneous Studies 231

tions, the relative angular position of the individual impellers can be changed. This
can be effective in reducing excessive noise.
Cavitation. When the vapor-filled bubbles of a cavitating liquid reach a zone of
higher pressure, they collapse vigorously, giving rise to noise. It is not always
possible to distinguish between noise created by cavitation and that caused by other
factors, but a dry, cracking noise from the suction end of the pump is likely to be due
to cavitation. This is especially true when the pump operates at low net positive
suction head (NPSH) values or at very low flow rates.
Prerotation. The primary characteristic of this source of noise is a sound such as
a blow from a very heavy hammer that occurs at a relatively low frequency (compare
the discusion related to Figs. 13-16-13-18).

Mechanical Sources
One source of noise is misalignment of the pump and driver. Vibration caused by
such misalignment usually display a frequency equal to twice the rotating frequency
of the pumping unit. Whenever power is transmitted by a coupling with pins or teeth,
misalignment can also induce vibrations at frequencies that are multiples of the
product of the number of pins (or teeth), multiplied by the operating frequency of the
unit.
Finally, vibration can be induced by a loose coupling or by unbalanced rotating
parts. This usually manifests itself in a frequency equal to the number of revolutions
per second. Vibration also occurs at this frequency for a bent shaft or an eccentrically
machined impeller. A third source is plain journal bearings, where vibrations at half
the operating frequency often indicate an oil whirl.
In pumps with ball bearings, faulty races may cause vibrations whose frequency
equal the rotating speed of the pumping unit multiplied by the number of balls in the
bearing with the faulty races.
Other mechanical sources of noise are loose bolts, unequally tightened bolts, and
stresses imposed by piping. The frequencies of these noises are usually not directly
related to the operating speed. Frequently, they are related to the natural frequencies
of the vibrating parts. These parts may simply respond to any other periodic exciting
force that happens to occur within or near the pump house. (In one case,the source
of a noise was a valve disc that had the same natural frequency as the speed of a
machine on the second floor of the same building.)
Noise can also be caused when rotating parts rub against stationary pump elements.
This source produces frequent, short, periodic squeaks or one long, continuous squeak.
Finally, vibration may be induced by a loose copuling or by a coupling with an
eccentrically machined bore. Such vibrations usually occur at frequencies that are
multiples of the operating speed.

Combined Sources
Vibrations caused by mechanical sources may produce cavitation in the liquid.
This, in turn, forms a secondary source of noise and vibration that can be more
intense than the primary source. Radial forces in a pump with a single volute can
also generate noise. These forces may become high at partial flow rates (compare
232 Problems Encountered with Centrifugal Pumps

Chapter 10). This bends the shaft with each revolution. These radial forces are also
subject to fluctuations at the frequency at which blades move past the volute tongue.
In certain unfavorable instances, this may cause the entire pumping unit to
vibrate. The simplest, most effective way to eliminate the source of vibration
originated by radial forces is to operate the pump only at the flow rates recommended
by the manufacturer.
A special kind of vibration is often encountered in double-suction pumps (Fig. 18-2).
Because of unavoidable inaccuracies in casting, there are always some differences in
the flow pattern of the liquid entering each side of the impeller. Sometimes these
differences become so intense that the entire rotating unit is forced toward one side
of the casing.
Assume that the impeller was forced toward the left. This displacement reduces the
gap XL, thus decreasing the leakage through the left wearing ring. This, in turn,
increases the axial pressure PL on the left shroud of the impeller, pushing the entire
rotating unit to the right, which in turn reduces the gap XR. Consequently, the
pressure PR on the right shroud becmes greater than the pressure on the left shroud,
causing the rotating unit to move to the left again, thus starting another cycle of axial
movements.
Although this may sound unbelievable, such periodic axial movements of the rotat-
ing unit are observed on a high percentage of double-suction pumps in operation.
(See also the discussion related to Figs. 19-10-19-12.)
A special source of noise and vibration is often generated by resonance. This
happens when the natural frequency of the rotating unit, pump casing, driver, base

PL--+-I~ ....-+-~+-- PR

Figure 18-2 Source


of axial vibrations in a
double-suction pump.
Miscellaneous Studies 233

plate, aggregate, foundations, or any structural element of the pump house coincides
with the rotating speed of the pump or with the lower harmonics of this rotating
speed.
Resonance also occurs when the frequencies of one or more of these elements
coincide with the frequency with which the impeler blades move past the volute
tongue or with the frequency of any other periodic effect that may arise during the
pump's performance.
The primary characteristic of resonance is that it produces much more intense
vibrations than the exciting source. It also begins with a relatively low intensity that
increases rapidly within a short period of time. Excessive vibration caused by
resonance is particularly apt to appear in a variable speed drive. With such a drive,
vibration occurs only within a limited range of speeds.
An effective remedy for the last case is to provide the drive with some sort of
automatic means to prevent the unit from operating near the exciting speed. Another
option is to alter the natural frequency of the resonating element by changing its
shape or weight.

Special Cases
Sometimes excessive noise can be generated by a combination of circumstances. In
one case, a pump developed a high noise level at the passing frequency of the blades
at a flow rate at which it previously operated quietly. Inspection revealed that the
foundation bolts were improperly tightened. After retightening, the noise returned to
its expected level.

DAMAGE TO PARTS

Pump parts may be damaged in many different ways. The three principal classes of
damage are corrosion and erosion; abrasion and wear; and fractures, cracks, extru-
sions, and dents.
In practice, the nature and causes of damage are often much more numerous and
complex. In particular, many forms of damage result from the combined simul-
taneous action of several factors or from a certain factor acting as a catalyst that
triggers another damaging process.

Fractures, Cracks, Dents, and Extrusions


Fractures, cracks, dents, and extrusions usually result from excessive stress.
Fractures Caused by Excessive Tension Forces. Fractures caused by
excessive tension forces are recognizable by the granular structure of the surfaces
along which the part failed. Depending on the physical properties of the affected
material, fractures caused by tension may be accompanied by cracks in the direction
of the applied stress and by elongated, spearlike projections in the same direction.
When such a fracture occurs in a ductile material, it may be detected by observing
that when the fractured parts are put together, their combined length is always
longer than the length of the original, unbroken part. Another trademark of such a
234 Problems Encountered with Centrifugal Pumps

fracture is that the cross-section of the part in the failure zone undergoes a reduc-
tion. In brittle materials, surfaces may still exhibit a granular appearance, but there
is no increase in the combined length of the broken parts and no reduction of the
cross-section of the failure zone of the broken part.
Fracture from Compression. Depending on the physical properties of the affected
material, fracture from compression appears as a pulverization of the affected part or
as a large lateral expansion. The latter is usually accompanied by many cracks at the
outer rim of the affected part.
Bending. In many metals, bending fractures are recognizable by a granular struc-
ture of the fractures surface on one side of that surface, and a smooth, squeezed-in
structure on the opposite side. In a ductile material, such a fracture is also usually
accompanied by a plastic distortion of the material in the vicinity of the edge of the
fracture's surface.
Shear. Shear usually manifests itself by smooth surfaces of the fracture, often
accompanied by lines that indicate the direction of the acting force.
Fatigue. With forces that vary periodically, failures caused by other factors
occur at lower stresses than the ultimate strength of the affected material. This
property is known as fatigue.
Fatigue usually originates at locations of concentrations of stresses. To these
belong a sharp transition from a thick section to a thin section, as well as minute
cracks and other surface imperfections. Sometimes, a tiny tool mark can initiate a
failure owing to fatigue. Such failures may also be caused by frequent expansions
and contractions caused by fluctuations in temperature. Failures from fatigue usually
occur only after a certain number of cyclic changes in the magnitude or direction of
the stress.
Often, fatigue starts at or as a small crack in the material. This crack grows
gradually under the influence of the fluctuating forces until fracture occurs. This
effect is often marked by the fact that a portion of the fracture surface shows signs of
rust or scale, while the remainder has the color and structure of a fresh fracture.
Dents and Extrusions. Dents and extrusions are usually the result of a blow.
Their causes and effects have already been discussed in connection with Fig. 15-16.
They can also appear when excessive pressure is applied to a very small area, such
as when a granule of hard material is clamped between two softer surfaces or when
two sharp edges press against each other.

Erosion and Corrosion


Whenever a liqUid is in contact with a solid for a prolonged period, the wetted
surface of that solid may deteriorate as a result of one or a combination of the
following: chemical reactions, electrolytic action and crevice corrosion.
Chemical Reactions. Certain liqUids may react chemically with the material of
the wetted pump parts. To avoid damage, the pump can be built of special materials.
Appendix B presents a comprehensive list of the chemical compatibility of different
materials. lt should be emphasized, however, that this list should serve only as a
gUide, and neither the author, the publisher, nor the company that compiled the list
Miscellaneous Studies 235

can guarantee that all data are correct. In particular, as will be explained later in this
chapter, even chemically compatible materials may fail because of electrolytic action,
the presence of a factor that may act as a catalyst that triggers a chemical reaction,
or some other unforseen cause. Consequently, before using any of the recommended
materials, check with the manufacturer(s) of the given product.
Electrolytic Action. Even when a pump is made of materials that are compat-
ible with the pumped liqUid, it may fail when the wetted parts are made of different
materials. When two different materials are immersed in a liquid that conducts
electricity (an electrolyte), they may form a voltaic cell. Electrolytic action can destroy
the part made of a less "noble" material within a relatively short time. (In common
parlance, the term noble refers to the chemical reactivity of a given material,
especially metals. The more chemically inert a material is, the more noble it is
considered to be.)
When damage caused by electrolytic action is suspected or anticipated, all wetted
parts should be made of one single material. If this is impossible or impractical, use
materials with nearly the same electrochemical potentials.
Sometimes a material that is more chemically active is covered with a more inert,
protective layer. For example, a shaft made of mild steel is frequently protected by a
bronze sleeve. In this case, all contact faces between the two materials should be
sealed off from the pumped liqUid or electrolytic action that will ultimately cause
serious damage.
One methods of reducing corrosion caused by electrolytic action is called cathodic
protection. This method consists of providing a direct electrical current in the opposite
direction of the current generated by the wetted materials. There are two ways of
accomplishing this. One is to supply the proper direct current from an external
source, (e.g., from municipal utilities and a rectifier, from a battery.). The second is to
install in the liquid a material that is less noble than any of the materials of which
the pump is made. The most popular material used for this purpose is a zinc anode
(Le., a bulk of zinc suspended at the proper location within the pumped liquid).
Crevice Corrosion. Some materials are attacked chemically only in the presence
of another substance (catalyst). For example, mild steel, when half submerged in
water, rusts qUickly because of its simultaneous contact with air. However, when
totally submerged in distilled water, it remains completely inert.
In particular, air and gases often act as such catalysts. They are likely to be
entrapped in small cracks and crevices. This often starts crevice corrosion. As its
name implies, this kind of corrosion begins in any zone where two wetted surfaces
touch one another.
In some cases, crevice corrosion may occur for exactly the opposite reason. The
corrosion resistance of certain materials is often due to a protective layer of oxide.
Whenever a crevice is wide enough to permit liqUid to enter and form a stagnant
zone, the liqUid loses some of its oxygen to the liquid that flows by, thus becoming
oxygen deficient. The oxygen deficiency of the stagnant liqUid is replenished by the
oxygen contained within the protective layer, thus making it vulnerable to corrosion.
Effects of corrosion may manifest themselves in different ways, depending on the
physical properties of the material and the circumstances under which it occurs.
When the corrosion rate is low and the velocities of the flOwing liqUid high, the affected
areas may be polished (this is the basis of the process of chemical machining).
236 Problems Encountered with Centrifugal Pumps

In other cases, it produces an irregular surface with many holes of different sizes
that overlap and look as if the entire surface has been eaten away by insects. In still
other cases, especially in composite, nonhomogeneous materials (e.g., cast iron
consists of iron and trapped particles of graphite that are finely dispersed throughout
the material) corrosion may eat away the more active element, causing the rest to
disintegrate.

The Effect of Velocity on Corrosion


Regardless of whether the deterioration of the surface is caused by chemical or
galvanic action, or even by corrosion, the rate at which it takes place may be
Significantly affected by the velocity of the attacking liqUid.
In a perfectly still liqUid, conditions are not uniform. Some parts of the wetted
surfaces may remain unaffected by the liqUid, whereas others undergo very intense
corrosion. In the presence of a slow-moving liqUid, the corrosive activity may spread
over a wider area of the wetted surface. This may make corrosion more uniform, but
less intense.
Higher velocities may set up vortices that cause local pitting. Still higher velocities
may carry away the corroded material at a higher rate, thus exposing new material to
the attack of the liqUid. On the other hand, it may have a protective effect in the case
of crevice corrosion by replenishing oxygen to the liqUid contained within these
crevices. Against this, however, high velocities may have an abrasive effect on the
affected surfaces and may remove the protective film that covers these surfaces.
In general, although the velocity of the liqUid often has a very pronounced effect on
the rate of corrosion, the nature and intensity of that effect may be different in
different situations.

The Effect of Elevated Temperatures


on Corrosion
Some materials are chemically compatible with certain liquids at low temperatures,
but are attacked by the same liqUid at higher temperatures. Elevated temperatures
may also liberate gases dissolved in the liqUid, that, in turn, may act as catalysts
triggering a corrosion process.
CorrOSion, in general, is a very complicated process controlled by many factors in
addition to their mutual interaction. For example, cast iron will operate in water even
at elevated temperatures, if it contains a certain amount of neutral salts. In some
installations, neutral salts are added continuously to the pumped liqUid to keep
corrosion to a minimum. Failure to add neutral salts (popularly called stabilizers)
may increase the rate of corrosion manifold.
The minimum salt concentration required to prevent corrosion of cast iron in water
depends primarily on the following factors: the pH value of the pumped water; its
temperature, and the concentration of oxygen as a result of air being dissolved in the
pumped water.
For example, at 200°F and at a very low oxygen concentration, a minimum of 10
parts per million of stabiliSing salts is capable of preventing corrosion in water with a
pH value of 7.0 (Le., neutral). In water with a pH of 9.0 (Le., slightly alkaline), the
same concentration of stabilizer prevents corrosion even at 300°F. However, if the
Miscellaneous Studies 237

water contains more than 0.03 cclL of oxygen (Le., 30 parts per million by volume),
no stabilizing salt can prevent the corrosion of cast iron. Alternatively, with a low
oxygen content and a pH value of 7.0, the water requires at least 15 parts per millon
of stabilizer to prevent corrosion at 300°F.
General gUidelines for the requirements of stabilizer concentration for different pH
values and temperatures at very low oxygen concentrations are given in Table 18-1.

Table 18-1 Minimum concentration of stabilizers in parts per million.

pH values

Temperature, OF 7.0 7.5 8.0 8.5 9.0

200 10.0 7.5 5.0


250 12.5 10.0 7.5 5.0
300 15.0 12.5 10.0 7.5
350 15.0 12.5 10.0 7.5
400 15.0 12.5 10.0

Other Factors that Affect Corrosion


Intermittent Operation. Pumps operating continuously in a given medium
often exhibit a smaller rate of corrosion than pumps operating intermittently. During
stoppage, air may more readily accumulate inside the pump, accelerating corrosion.
Cavitation. Cavitation usually damages pump parts through a combined action
of erosion and corrosion. Rough surfaces of any material are more readily attacked
by chemical agents than are smooth surfaces. One reason is that rough surfaces
attract and retain small nuclei of air and gas which act as catalysts.
Erosion caused by cavitation can convert any smooth surface into a rough surface,
thus triggering and accelerating corrosion. Another reason cavitation often triggers
corrosion is that it converts some of the pumped liquid into vapor. In some instances,
a chemical may be more active as a vapor than as a liquid.
Reduced NPSH. Corrosion caused by reduction in available NPSH may occur
even in the absence of cavitation. This may occur when a gas that is dissolved in the
pumped liquid has a chemical affinity for the material that makes up the pump.
Under reduced available NPSH, some of the gas becomes liberated from the solution
and attacks the pump material.
This problem can be dealt with in one of two ways: either increase the available
NPSH, or replace the unit with a pump constructed of a different material.

Abrasion and Wear


Solids in Pumped Liquid. Unless the wetted parts are made of highly abrasion-
resistant materials, solids will cause rapid reduction in the cross-sections and wall-
thicknesses of the affected parts. Sealing surfaces may be worn away, increasing
leakage. Reduced cross-sections and wall-thicknesses also reduce the mechanical
strength of the affected parts and ultimately lead to fracture.
238 Problems Encountered with Centrifugal Pumps

A typical symptom of this damage is mirror smooth surfaces on the rotating parts.
The stationary parts, in contrast, exhibit oblong dents and elevations covered with
secondary grooves and scratches. Often, the grooves, dents, and elevations are in
wavy patterns.
Vortices, Secondary Flows, and Cavitation. Secondary flows may arise in the
pumped liquid, leading to abrasion in some zones of the pump. Such effects usually
occur when a pump operates near shut-off. The most vulnerable zones are near the
volute tongue and the impeller eye, at the blades and shrouds on the impeller outlet,
on the high-pressure side of the wearing rings in the casing, and near the balancing
holes.
Abrasion caused by these factors sometimes produces the same kind of surface as
the one damaged by solids in the pumped liquid. In most cases, however, all the
effects produce cavitation which usually generates an irregular surface covered with
many interfering holes of different depths.
Mechanical Wear. In the liquid end of a pump, mechanical wear usually occurs
in the wearing rings and the stuffing box. In the stuffing box wear is a result of the
stationary packing wearing away particles of metal from the rotating shaft. Wear can
be particulary severe when the gland is too tight or when there is inadequate liqUid
flow between the shaft and the packing.
Mechanical wear is manifested by scratches on the periphery of the shaft, often
accompanied by dark discoloration from overheating. In extreme cases, the shaft may
be so badly eaten away by the packing, that it fractures.
To eliminate defects caused by wear, only the correct grade of packing should be
used (preferably the brand recommended by the pump manufacturer), and the gland
should never be tightened enough to prevent all leakage from the stuffing box. In
most cases, a leak of 100 drops/min considerably reduces the wear on the shaft.
In the wearing rings, wear may result from high velocities of liquid flowing between
the sealing faces, combined with the high peripheral velocity of the sealing rings, or it
may result from metallic contact between the stationary and rotating parts. Wear
from the combined action of liquid flow and relative velocity of the rotating parts
demonstrates itself in the form of enlarged clearances. These, in turn, may be
detected in a manner discussed in connection with Fig. 7-11.
Wear caused by metallic contact manifests itself in scratches of the surfaces of the
wearing rings that extend in the direction of the relative motion of the matching
parts. As in wear of the stuffing box, such marks are often accompanied by dark
discolorations caused by overheating.

Galling
When some materials (especially metals) rub against another surface, some of the
material is worn off one surface and fused to the other. In more serious cases, this
can lead to seizure resulting from the welding together of the two surfaces. Such
cases are particularly apt to occur when both surfaces are made of same material
and possess the same hardness. Particularly vulnerable to this effect are series 400
soft stainless steels. On the other hand, stellite is known to offer excellent galling
resistance. Among the less expensive materials, most varieties of cast iron have
shown excellent resistance to galling.
Miscellaneous Studies 239

To reduce the danger of galling, one of the following things may be done:
1. Make each of the mating parts of a different material. (Appendix B-1)
2. Heat treat the mating parts so as to generate a hardness differential of at least 80 points
on the Brinell Hardness Scale.
3. If the moving part is cylindrical, machine peripheral grooves onto the face of the rotating
cylinder.

OTHER CAUSES OF DAMAGE

Human Factors
These effects include problems caused by carelessness, negligence, human error,
and lack of proper knowledge and understanding of the do's and don'ts related to
centrifugal pumps. These problems include, among others, the following:
1. Inaccurate alignement between the pump and driver.
2. Forcefully connecting mismatched flanges of the pump and the pipelines by means of
the connecting bolts.
3. Uneven tightening of the bolts of a flange or a cover.
4. Starting the pump without lubricant in the bearings.
5. Starting the pump without liqUid in the casing.
6. Starting the pump without checking the direction of driver rotation.
7. Allowing the pump to operate for a prolonged time with noise and vibration without
checking the source of these abnormalities.
8. Dropping solid objects into the suction tank.
9. Allowing the pumping unit to operate in a unventilated location (this may overheat the
driver and bearings).
10. Applying a blow or excessive force on the pumping unit after it is installed on site.
11. Allowing water to enter the lubricating oil.
12. Subjecting the pump to thermal shock.
13. Failing to preheat liquids that the pump can handle only when they are hot.
14. Starting a pump with a closed valve in the suction line.
15. Leaving a nut, bolt, or tool in the pipeline during assembly. These objects may enter the
pump, causing irreparable damage.

Case Histories
In practice, there is no end to the number and versatility of problems that can be caused by human error. The
following case histories illustrate the range and types ofguesses one must consider to find the cause of a problem
encountered with a given pump.
A pipefitter installed a blind gasket between a pair offlanges in the suction line. This closed the passage of the
liquid from the suction well into the pump, thus preventing the pump from delivering any liquid.
To find out what was wrong, the pump was dismantled, all parts were thoroughly checked, and any detected
minor inaccuracies were corrected. The pump was then carefully reassembled and reinstalled, but it still failed
to deliver any liquid. A second team of mechanics tried again to find the cause of the malfunction, but they also
failed. After four or five unsuccessful trials, someone jokingly remarked, "Perhaps there is a blind gasket
somewhere in the suction line." The suction line was dismantled, and it turned out that this was not just a joke.
240 Problems Encountered with Centrifugal Pumps

An experienced pump mechanic spent several days working on a pump that failed to deliver any liquid, until
he accidentally discovered that the discharge line was equipped with left-handed valves that opened clockwise.
The pump was intended to be started against completely open valves, and therefore the mechanic turned them
counterclockwise before switching on the power: However, because these were left-handed valves, he was closing
instead of opening them, preventing the liquid flom flowing.
The following problem is one whose source took some time to detect. A pump was installed on a test stand and
showed erratic readings of power consumption. It was dismantled, and excessive rubbing was evident on the
wearing rings. It was then reassembled, and great care was taken to prevent the piping from imposing any strain
on the pump. Still, the power readings remained erratic. The alignement between the pump and driver was
checked and corrected for accurate concentricity and squareness, but nothing that was done eliminated the
pump's erratic behaviour:
After several weeks of trial and error, it was finally discovered that the mechanic who was working on the
assembly of the pump did not know that the water end needed to be firmly connected to the bearing frame. He
simply did not adequately tighten the bolts that connected the pump casing to the bearing frame.
Another hard to solve case occurred when the level in the suction tank and the available NPSH were high,
and the pump developed intense cavitation. As the level decreased, the intesity of cavitation decreased. Finally,
at very low levels, cavitation disappeared completely. The only known explanation was that when the level in the
suction tank rose, the pump was delivering very high flow rates (compare the discussion related to Fig. 6-10).
The problem, however, was that all flow measurements showed that the pump was operating within its normal
range offlow rates, even when the level of liquid in the suction tank was high.
A search was conducted for a possible leak in the discharge line and the posibility that some liquid was being
diverted to a bypass before reaching the flow meter: Still, there seemed to be no evidence that the pump operated
far out on its QH curve.
Then came the idea that even if there were no external signs that the pump was handling excessive flow rates,
there might exist an excessive flow through the impeller: This can happen when there is a large gap between the
wearing rings that seal off the passage of the liquid from the impeller periphery to the impeler eye. A suspicion
arose that the mechanic who assembled the pump forgot to install the wearing ring in the casing. The pump was
opened, and this turned out to be the case. The large gap between the impeller ring and the casing allowed a
significant amount of liquid to return from the impeller outlet to its inlet (compare Chapter 7).
A study revealed that the clearance between the impeller and casing allowed 30% of Qmax to be recirculated
into the suction eye. When the suction tank was lowest, the pump delivered 65% of the maximum recommended
flow rate. Even with the addition of 30% internal leakage, the pump operated satisfactorily. When the level in the
suction tank was highest, the flow delivered by the pump was only slightly lower than the maximum
recommended flow rate Qmax. However, the actual flow rate handled by the impeller was almost 30% higher: This
excessive flow rate was responsible for the observed problem.

The case histories presented here illustrate how difficult it can be to detect the real
reason for a malfunction in a centrifugal pump when it is caused by human error.
Fortunately. they happen infrequently. Nevertheless. such cases can be solved with a
thorough knowledge and understanding of the factors that cause pump failure.

Time-Related Causes of Part Failure


The same time-related factors that cause deterioration of performance may also
lead to part failure. These include erosion. corrosion, and wear. An additional time-
related cause is the aforementioned effect of fatigue of materials. In addition to the
above, the following time-related factors may also lead to the failure of parts.
Settling of Foundations or Other Structural Elements. Most construction
work requires that certain preparations be carried out on site. These may include
Miscellaneous Studies 241

excavations or landfills, or both. The prepared site is later subjected to the combined
weight of the structures and their contents, as well as to the action of the elements
(e.g., rainwater seeping under the foundations). Depending on the nature of the
preparatory work, and the quality of that work, and the nature of the ground, the
entire structure may set evenly, tilt to one side, or dislocate in some other manner.
This may cause misalignment between the pump and driver or between the pump
and pipelines, thus imposing excessive stresses on the pump. In severe cases, these
stresses may easily lead to fracture of some of the pump's parts.
One of the most popular ways to prevent this kind of damage to pumps is to
connect them to their pipelines by way of a flexible unit. Such units are readily
available.
Changes in Operating Conditions. When the total head against which a
centrifugal pump is operating changes, the flow rate delivered by the pump is usually
altered. This may result from an accumulation of scale or debris in the pipelines
(causing an increase in head), or erosion of the pipelines, which may cause a
reduction of the velocities and a subsequent reduction in head.
Depending on design, some centrifugal pumps exhibit excellent performance near
their design flow rates, but they may develop very intense instabilities at off-design
flows. In some cases, these instabilities may become intense enough to fracture some
of the pump's parts within a relatively short period of time. Such instabilities are
likely to occur in pumps of high specific speed and may be particularly vigorous in
larger units.
The most effective way to reduce the danger of such a failure is to replace the pump
with a better design. Alternatively, whenever such instabilities occur at reduced flow
rates, the problems they are causing can be eliminated by providing a bypass
equipped with suitable sensors that will cause the bypass valve to open whenever the
flow rate falls below a certain point. When such instabilities occur at high flow rates,
specially designed pressure-operated valves, which will automatically reduce the flow
through the pipelines as soon as the discharge pressure in the pipelines falls below a
certain point, can be used.
CHAPTER 19
Problems Related to
Specific Circumstances
PUMPS WITH SEMI-OPEN IMPELLERS

A semi-open impeller has only one shroud. The second shroud is replaced by a
stationary wear face cast integrally with the casing or by a stationary wear plate
firmly attached to the casing. Experience has shown that the clearance C (Fig. 19-1)
between the impeller vanes and the wear face has a profound influence on the
performance of the pump.
Up to a value of about C=0.016 in. the effect of the clearance is rather
insignificant. but when it is increased the performance of the pump deteriorates
rapidly with any additional increase (compare Fig. 7-13). Because the performance is
sensitive to the magnitude of the clearance, all pumps with semi-open impellers have
some arrangement that permits the magnitude to be adjusted. In most modern
pumps, this can be done without dismantling them. In horizontal pumps, it is
usually done by mounting the thrust bearing T in a cartridge be and adjusting the
axial position of the cartridge by means of shims S (Fig. 19-3).
With all shims removed, the cartridge and the rotating elements are pushed
forward until the impeller vanes touch the wear face. The thickness of the shims
required to maintain zero clearance can now be measured with a filler gauge. By
adding between 0.010 and 0.015 in. to this thickness, the correct setting is obtained.

Figure 19-1 Gap between the wear plate and the blades of a
semi-open impeller.
243
244 Problems Encountered with Centrifugal Pumps

Although this is often the only way to determine the clearance between the impeller
vanes and the casing, it is not necessarily foolproof. A bent shaft, uneven tightening
of bolts, or uneven wear may cause the faces of the impeller vanes to be nonparallel
to the wearing face of the casing (Fig. 19-2). In such a case, the actual clearance will
be too large on one side of the casing and will adversely affect performance.
When there is reason to believe that such a case exists, the simplest way to verity it
is to glue a small amount of a plastic material (e.g., modeling clay) to several places
on the blade's surfaces. The wear face of the casing should be covered with an
antistick compound (e.g., silicon grease), and the pump assembled. The excess
plastic is squeezed from the clearance between the vanes and the casing, leaving only
enough to fill out clearances. The pump is then be opened and the thickness of the
fillers at different locations on the vanes measured to verity whether the vane faces
are parallel to the wear face.
Another method is to remove the casing from the pump and leave the impeller
mounted on the frame. The distance of each vane from the plane against which the
casing was mounted are measured (Fig. 19-3, plane Pl. This is repeated for each vane
while the impeller is in a fixed angular pOSition.
The same procedure is repeated after the impeller is turned 90 degrees and 120
degrees. Similarly, measurements are taken to determine the distances of several
pOints of the wear face from this plane of the casing which is mounted against plane
P. Now it is possible to determine the order of magnitude and the variations in the
sizes of the clearances between the vanes and the wear face with simple calculations.
In practice, neither method is easy to apply. When the face of the vanes forms a
plane surface, it is easier and simpler to determine the clearances by measurements.
However, when these faces are mostly curved, it is simpler and more reliable to use
the first method. Sometimes, particularly in deep-well pumps, semi-open impellers
are used in mUltistage arrays. Here, the likelihood of problems is compounded by the
possibility of improperly locating some of the impellers on the shaft. In Fig. 19-4. the
first impeller is in contact with its wear face A. while all others are at different
distances from their respective casings, as at B. When the shaft is moved to create
the correct clearance at A. the clearances of all other impellers become excessive.
Closely related to this problem is the case in which any sections of the blades are
not parallel to the axis of the shaft. In Fig. 19-5 the impellers are of the mixed-flow
type and the section through the blades at A is inclined to the shroud at an angle 'Y.

Figure 19-2 The effect of a nonparallel wear plate on the


gap C.
Problems Related to Specific Circumstances 245

Figure 19-3 Adjusting the axial position of an impeller by means of shims S.

A standard procedure of assembling such a pump is as follows. First, impeller 1 is


fixed to the shaft in its proper position, and the shaft with the impeller on it is placed
into the first stage and locked into position by the nut N. Next, the casing of the
second stage is assembled and impeller 2 is fixed to the shaft so that the vanes bear
against their wearing face. This procedure is repeated for all subsequent stages.
Sometimes one of the impellers is pressed against its seat with an excessive force.
This may cause the vanes of the particular impeller to deflect, as shown at B.
After the assembly is completed, the nut N is removed to enable the shaft and
impellers to rotate. However, as soon as the nut is removed, the vanes of the impeller
that were deflected flex back into their normal position. This moves the shaft axially
and lifts all other impellers from their seats. During operation, only the impeller
whose vanes were originally deflected operates with the proper clearance; all the
others are located too far from their seats.
246 Problems Encountered with Centrifugal Pumps

en
Figure 19-4 Possibility of uneven distance between semi-op
A impellers and the casing in a deep-well pump.

distorted
Figure 19-5 How the vanes of a semi-open impeller can be
during assembly.
Problems Related to Specific Circumstances 247

A different kind of problem is encountered in pumps with semi-open impellers


when the pumped liquid contains stringy or fibrous material. This material can
wedge between the impeller blades and the wearing face, breaking the affected blades
or shaft, or both.

VERTICAL DEEP-WELL PUMPS

Deep-well pumps, often called borehole pumps, were originally developed to pump
water from deep wells of small diameter. They have now gained widespread use in all
branches of indUStry.
There are several reasons for this acceptance. First, deep-well pumps require
considerably less floor space than conventional horizontal pumps used for the same
duties. This difference can be particularly great for pumps with relatively high flow
rates and high heads. (Compare, for example, two pumps for 8000 GPM against 1000
feet at 1770 RPM.). Second, because the pumping end is submerged in the pumped
liqUid, deep-well pumps do not need to be reprimed each time they are restarted.
Third, when the inlet to the first impeller must be located deep below the level of the
liquid in the suction tank, a deep-well pump costs significantly less than a horizontal
pump because of smaller floor space requirements.
There are three principal parts of a deep-well pump: the pump or turbine, the
column, and the head. In Figs. 19-6 and 19-7, pump P is essentially a multistage
pump constructed Similarly to a donut pump but with a smaller outer diameter. The
column C is a line of piping that keeps the pump suspended in position and serves as
a discharge line that brings the liqUid from the well up to above-ground level. The
column also encloses the line shaft, which transfers the power from the driving unit
to the turbine. The head H is a bracket from which the column and pump are
suspended. In addition, the head also serves as a support for the driving unit (electric
motor, gear head, or pulley drive).
Deep-well pumps are subdivided into a number of classes. They are classified by
impeller type (closed or semi-open) and column type (open shaft with water-lubricated
bearings or enclosed shaft with oil-lubricated bearings).
Deep-well pumps share many of the same problems of other pumps. They also have
special problems because of their unique construction features and operating
conditions. Some problems are common to all classes of deep-well pumps, whereas
others relate to only one type. Of course, any problem is aggravated because the
pump operates deep below the earth's surface. Anyone who is familiar with the
problems encountered with borehole pumps that operate in deep wells should be able
to solve problems in smaller settings more easily.

Problems Common to All Deep-Well Pumps


Extension of Lineshaft. When a pump is in operation, axial thrust usually
builds up on the impellers. This thrust acts toward the suction sides of all impellers.
Consequently, each time a pump is started, its lineshaft undergoes an extension.
With long columns, this extension may be so large that the lower parts of the
impellers begin to rub against the casings, damaging or destroying the impellers and
casings, or it may cause a breakdown of the lineshaft. Excessive lineshaft extension
248 Problems Encountered with Centrifugal Pumps

SR

RB

Figure 19-6 Section through a deep-well pump. (Worthington.)

usually manifests itself in great noise and vibration, and fierce fluctuations in power
consumption readings.
Most deep-well pumps have some means of regulating the axial position of their
impellers. For an electric motor or a right-angle gear drive with a hollow shaft, this
device is an integral part of the driving unit. In other cases, this device is usually part
of the head.
To avoid problems caused by too much extension, the pump should be started with
its impellers in the highest position. If no rubbing occurs, the pump should be
stopped, the impellers slightly lowered, and the pump restarted. The procedure
should be repeated until the first signs of rubbing appear. The impellers are then
lifted slightly, just enough to avoid rubbing.
Variations in Water Level of the Well. When the pump is not operating, the
level of the water in the well is exactly the same as in the surrounding strata. When
the pump is started, the level of the well starts to fall until there is a state of
eqUilibrium between the rate at which water is removed from the well and the rate at
which water is seeping into the well from the surrounding strata.
The level at which this state of equilibrium is reached is subject to seasonal and
weather conditions. Sometimes, the water level in the well falls so much that air
Problems Related to Specific Circumstances 249

Figure 19-7 Section through an oil-lubricated,


deep-well pump.

enters the pump. This reduces the flow immediately and allows the water level to rise.
This shift in water level increases the flow rate of the pumped liquid, causing the
water level to fall again. As a result, the pump delivers water in periodic bursts
instead of a continuous stream.
As a temporary remedy, the discharge valve can be partially shut down, thus
reducing the amount of liquid delivered by the pump. The only permanent remedy
known is to add at least one column to the pump. In some cases, this may reqUire
deepening the well.
The effects of Sand in Water. The first direct result of sand in the water supplied
to the pump is the excessive wear of parts. This quickly leads to deterioration in
performance and may also cause mechanical failure. The only possible remedies are
to make the parts that are subjected to wear from wear-resistant materials or to
replace the worn parts at shorter intervals.
The most significant amounts of sand appear in newly drilled wells. After a period
of time, the amount of sand usually decreases to a tolerable level. However, during
the initial period, when the amount of sand contained in the liqUid is excessive,
special precautions must be taken every time the pump is stopped. This is because in
such cases, all the sand in the liqUid that remains in the column will settle and clog
the impellers and caSings.
To avoid this, the pump should never be stopped suddenly. The amount of sand
carried by the water depends on the velocity of the water in the column. The
250 Problems Encountered with Centrifugal Pumps

discharge valve should therefore be shut down very gradually, so as to reduce the
amount of sand carried with the flowing liquid-and thus amount contained in the
column-to a tolerable level before the pump is stopped completely.
Air in the Column. To prevent water in the pipelines from flowing back into the
well, a deep-well pump is usually provided with a nonreturn valve close to the
discharge. This however, cannot prevent the liquid in the column from flowing back
into the well. When the column extends for more than about 34 ft (10.33 m) above
the water level, a vacuum is created in the column above the water level every time
the pump is stopped. This vacuum extends up to the nonreturn valve. When the
pump is restarted, cavitation shocks may occur. In addition, in a pump provided with
an open shaft, this vacuum may drain all the lubricant from the stuffing box, causing
damage to the shaft when the pump is restarted. To prevent this, a deep-well pump
with a long column should always have a vacuum-breaking valve.
Another problem arises when air enters the column during stoppage but is unable
to escape when the pump is started. This may cause an extremely severe blow in the
nonreturn valve, as explained in connection with Fig. 15-11. In deep-well pumps,
these effects can be more severe than in horizontal pumps, because long columns are
capable of entrapping huge amounts of air.
Faulty Installation or Faulty Well A deep-well pump is connected to a head
that rests on a solid foundation (Figs. 19-6 and 19-7). The column and pump are
suspended from this head. Because of the length of the column, any lateral force can
easily bend it, causing misalignment of the bearings. This, in turn, usually results in
noise and vibration and may often lead to early destruction of the pump.
To eliminate such a possibility, the head and motor must be installed with their
centerlines in perfect alignment with the centerlines of the column and pump. In
addition, care must be taken that the casing of the well does not exert any lateral
forces on the pump assembly.
When the well is straight and vertical, this is achieved by installing the head
assembly with its centerline perfectly vertical and coinciding with the centerline of
the well. When the well is straight but not vertical, this is often achieved by simply
mounting the head assembly in a slightly inclined position, so that its centerline
forms an extension of the centerline of the well. When the well is not straight, the
only remedy is to install a pump of a smaller outer diameter.
Special attention is required when installing a deep well pump on a moving vehicle
(e.g., a large ship). The orientation of the pump base will change with each change in
the vehicle's position. To prevent any bending movements on the column because of
this effect. the head and pump must be supported on both ends by one common,
rigid structure.
Improper Column Assembly. A deep-well pump will not operate satisfactorily if
some of the line bearings and the shaft are not properly aligned. To ensure
alignment, each shaft section and column is machined for perfect concentricity of the
threads and perfect squareness of the faces. Both the concentricity and squareness
of these parts may be easily ruined by allowing a column shaft to fall or bend,
damaging a face by a blow, or allowing dirt to be trapped between the faces of
consecutive column elements. Each factor can cause noise, vibration, and early
destruction of the pump.
Problems Related to Specific Circumstances 251

Pumps with Oil-Lubricated Columns


In addition to problems common to all deep-well pumps, the following are related
specifically to pumps with oil-lubricated columns (Fig. 19-7).
Inadequate Lubrication. Before starting the pump, all line bearings must be
properly lubricated. In pumps with long columns, it may take several minutes for the
oil to reach the bottom bearings. When the pump is started before this happens, the
unlubricated bearings are ruined. This, in turn, causes the shaft to vibrate, which
may ruin the remaining bearings.
Improper Assembly of the Oil Piping. Improper assembly of the oil piping may
be another source of poblems. The column bearings are kept in line by the oil piping
o (Fig. 19-7). The piping, in turn, is kept straight and rigid by the tension imposed by
the nut N. This nut should be tightened firmly enough to keep the oil pipe straight
under all working conditions.

Pumps with Water Lubricated Columns


Prelubrication. Water-lubricated columns are fitted with rubber bearings. When
they run dry, these bearings possess a high coefficient of friction, and therefore their
prelubrication is even more critical than is that for oil-lubricated bearings. Even a few
seconds of running dry may ruin them. To make sure this never occurs, the bearings
should be adequately prelubricated before starting, and this should continue at least
until the pumped water completely fills the column.
Swelling of Bearings. Most kinds of rubber swell when exposed to water
containing certain chemicalf" particularly in water containing sulfides or sulfates.
When this occurs, the bearings may bind the shafts and cause a breakdown.
Swelling of the bearings requires some time. In certain cases the rubber swells by
only a certain amount and then it stops. A good remedy is to allow the pump to run
without interruption (or with very short interruptions) for several weeks. The rotating
shaft wears away the excess rubber generated by the swelling process. Mter the
process of swelling has terminated, the pump is then able to operate trouble free.
Another remedy is to use bearings with extra-wide clearances. Finally, if time
allows. submerge different kinds of rubber in the water taken from the well, and test
which kind is most resistant.
The most accepted practice is to make the shafts SH (Fig. 19-6) of carbon steel and
provide them with stainless steel or bronze sleeves SL where the shafts are supported
by the bearings. To make sure each sleeve falls within its bearings, the length of each
shaft is made equal to the length of column pipe plus spider ring SR. During
installation, all column elements are tightened so that each face of the element bears
firmly against the opposite face of the next element.
When assembling the column shafts, air may become trapped inside the coupling,
preventing metallic contact between the faces. To prevent this, a vent hole V is
provided in each coupling. This hole must be located exactly where the faces of both
shafts meet, or air can become trapped inside the coupling and prevent metallic
contact of the shafts.
252 Problems Encountered with Centrifugal Pumps

Replacing Oil-Lubricated with Water-Lubricated Columns. This should be


avoided whenever possible because a well with an oil-lubricated pump always contains
a certain amount of oil. When this oil comes into contact with the rubber bearings, the
rubber swells. The result is usually broken shafts and early deterioration of the
bearings.
Sometimes, however, such a changeover from one type of pump to another is
unavoidable. In this case, remove as much oil from the well as possible by lowering
an absorbent material (e.g., a ball of cotton attached to a rope) into the well. This
material should touch only the upper, oil-containing layer of the liqUid and should
not come into contact with the water located below this layer.
After the material has absorbed a certain amount of oil, it is pulled up and the oil
squeezed out. This process is repeated until water begins to appear in the squeezed-
out oil. This is all that can be done in this case. There is no way to remove the last
traces of oil from the well; a thin layer usually remains (Fig. 19-8).
When a pump is lowered into the well, some of this oil enters the suction pipe and,
from there, the bearings. A primitive, but effective, way of reducing the probability of
such a case is to wrap the strainer that is located upstream of the pump inlet with
wet, soft paper (e.g., old newspapers). When done properly, the pump inlet can be
lowered below the layer of oil before any oil enters the suction pipe (Fig. 19-9). When
the pump is started, the wet paper diSintegrates and the pump continues to operate
satisfactorily. As a further precaution, the suction inlet to the pump must be
submerged so deeply into the well as to ensure that the water level never falls so low
that some oil from its surface is able to enter the pump.

Figure 19-8 Remains of oil in a well formerly equipped with an oil-


lubricated, deep-well pump.
Problems Related to Specific Circumstances 253

Oil

~....iiIHli:;;---Water

~--±'~- Paper-covered
strainer
Figure 19-9 Installing a water-lubricated
pump in a well in which an oil-lubricated
pump had been used previously.

Deep-Well Pumps with Semi-Open Impellers


In addition to the problems encountered with all other pumps with open impellers,
special problems arise with semi-open impellers as a result of elongation of the
shafts.
To achieve and maintain optimum performance, it is important to maintain proper
clearances between the impeller vanes and the matching wear faces. In multistage
pumps, it is important to maintain the same clearance between all impellers and
their corresponding wear faces. In deep well pumps, this is complicated by the fact
that the preset relative locations of the impellers change not only as a result of the
extension of the lineshaft but also from the extension of the pump shaft proper.
When the pump is started, additional axial thrust is generated onto the impellers,
which often leads to a very substantial extension of the shafts. To provide for this
extension, the initial clearance should be set large enough so as to prevent metal
contact between the impeller blades and the wear faces while the pump is in
operation. At the same time, the clearances should be small enough to ensure
efficient operation of the unit.
For correct initial setting, the impellers should be set at their highest position. The
pump is then started. If there is no rubbing (the latter is usually accompanied by
noise and vibration) the performance at one flow rate is taken. If the performance is
20% to 30% of that expected, the impellers may be lowered by about 0.10 in. and the
pump restarted. However, if the difference in performance is 10% or less, the
254 Problems Encountered with Centrifugal Pumps

impellers should be lowered not more than 0.008 in. at a time. This procedure is
repeated until the first sign of rubbing appears. The impellers are then lifted by about
0.012 in., and this should be the final setting.
A different problem associated with shaft extension relates to the pump proper. In
turbines with 15 stages or more, the shaft of the pump proper may extend so much
that the bottom impeller touches the casing, while the clearances of the other
impellers are still excessive. To accommodate this, the pump must be assembled with
varying impeller clearances.
First, the first impeller is held firmly against its seat by the nut and the sub-
sequent, say, 10 impellers are assembled so that each touches its seat. Now the nut
is released slightly to allow an axial movement of the shaft offrom 0.004 to 0.008 in.,
depending on the shaft diameter and the calculated axial thrust.
The eleventh impeller is now clamped against its seat, lifting all other impellers by
the distance allowed owing to the previous slackening of the nut. This procedure is
repeated for each tenth stage, allowing for the extension of the turbine shaft.

SPLIT-CASE, DOUBLE-SUCTION PUMPS


When a bend in the suction line is located asymmetrically relative to the suction
nozzles of the pump (Fig. 19-10), distribution of the flow rate between the two halves
of the impeller may be uneven. This occurs when a uniform flow passes through a
single bend, because it exits with a nonuniform velocity distribution [43). as shown in
Fig. 19-11. As a result of that distortion of the uniformity of flow, each suction inlet
of the impeller receives a different flow rate.

Figure 19-10 Asymmetrically mounted


elbow upstream of a double-suction
pump.
Problems Related to Specific Circumstances 255

A double-suction pump can be considered to consist of a combination of two


identical, single-suction impellers that operate in parallel and discharge into a
common casing. Let curve A in Fig. 19-12 represent the QH characteristics of each of
the single-suction impellers. Because of their parallel operation, the pump delivers a
resultant flow rate equal to the sum of all the flow rates that each of these impellers
deliver against the given head. Consequently, the actual QH curve of the pump shows
a flow rate equal to twice the flow rate delivered by each of the given impellers against
the given head. This is shown in Fig. 19-12 as curve B. In Fig. 19-12, we also have
entered a third curve C. This curve represents the resistance of the pumping system

I--- 50
'I
-----
1--1.20--+1 V/Vmean V/Vmean

Flow

Figure 19-11 Distortion of a


uniform flow due to a simple
bend [43]. Vmean = Flow rate RID = 1
divided by area of pipe cross- -0
I

t
section. V =Actual velocity of
the liquid.

I
I Curve A
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I

I~ 1---,1Q
,10-1 -
Figure 19-12 Effect of an uneven flow ----Oa-
distribution within a double-suction pump.
I..-----Ob ( = 2 0a)---I
256 Problems Encountered with Centrifugal Pumps

to a given flow rate. The actual flow rate against which the pump operates is at the
intersection M of curves Band C. It is equal to the flow rate Qb=2Qa. For the given
pumping system, this is the actual flow rate delivered by the pump, regardless of
whether both partial impellers are delivering exactly the same flow rate Qa, or one is
delivering more (e.g., Qa+aQ). and the second less (Le., Qa-aQ). When operating
within the given pumping system, their combined output is always equal to the flow
rate Qb, operating against the head Hm.
The test results presented in Fig. 19-11 lead to the conclusion that a bend
upstream of the suction inlet of a double-suction pump causes an uneven
distribution of the flows entering each side of the impeller. This means that one side
of the partial impeller delivers a flow rate equal to Qa-aQ (see Fig. 19-12) against
head HI. The other side of the impeller delivers a flow rate Qa+aQ against head H2.
At the outlet of the impeller, both flows encounter the same resistance Hm which has
been determined by the intersection of curves B and C. This results in the following.
At the impeller outlet, the flow rate Qa-aQ encounters a sudden drop in pressure,
from HI to Hm. This effect is similar to that of a sudden breakdown in the discharge
line. A sudden acceleration of the flow occurs, resulting in cavitation upstream of the
corresponding partial impeller. The flow rate Qa+aQ, however, encounters a sudden
increase in the resistance to flow, from H2 to Hm. This has an effect eqUivalent to a
sudden shut-off of a discharge valve, and violent water hammer may result. In
practice, both effects occur Simultaneously and cause the pump to operate with
intense vibrations, accompanied by cavitation noises.
Of course, this is a Simplified presentation of the actual effect. The flow is
continuously subjected to modulations owing to the periodic passage of the impeller
blades past the volute tongue, and to a vast number of additional effects. Also, the
separation between the two flows that enter from each side of the impeller is often
being terminated, not at the impeller outlet but at an upstream radius, complicating
the situation even more. In general, however, the study presented furnishes a
reasonable explanation of the effects of an asymmetrical bend being installed
upstream of the suction nozzle of a double-suction pump.
In the past, the general opinion was that such an occurrence could only be
eliminated by removing the bend and replacing it with a straight section of piping,
not shorter than 20 diameters of the diameter of the suction pipe. However, recent
experience has demonstrated that it is possible to eliminate the adverse effect of an
elbow upstream of a double-suction pump by inserting a SUitably designed flow
straightener between the pump and the elbow. (Compare also the discussion related
to Fig. 18-2).
Another problem frequently seen in double-suction pumps occurs because the
impeller is symmetrical relative to the plane 0-0 (Fig. 19-13). As a result, each half of
the impeller is a mirror image of the other half. This creates the possibility that an
impeller may be mounted on the shaft in such a manner, that its left side is located
on the right side of the plane, and vice versa. This causes the blades to rotate in the
opposite direction than was originally intended. Such an error drastically reduces
efficiency and creates noise and vibration. It also produces a QH curve that is
completely different from the rated curve.
A third problem can arise when the axial position of the impeller is determined by
the nuts N. (Fig. 19-13). These nuts should be adjusted so that the centerline 0-0 of
the impeller coincides with the center of the volute. Significant deviations from this
Problems Related to Specific Circumstances 257

o R

o
Figure 19-13 Section through a double-suction, split-case pump. L =Left side of double-section
impeller. R =Right side of double section impeller.

Figure 19-14 Mismatch between


the two halves of a split-case pump.
258 Problems Encountered with Centrifugal Pumps

rule may cause shrouds of the impeller to rub against the inside of the casing,
resulting in noise, vibration, and excessive power consumption. In addition, incorrect
impeller positioning may cause excessive axial loads (compare the discussion related
to Fig. 18-2) and reduced performance, even when the impeller shrouds do not touch
the casing.
Another special problem with this kind of pump sometimes occurs when a hole H
connects the volute with the suction (Fig. 19-13). Such holes may develop as a result
of the erosive or corrosive action of the pumped liquid. Often they are too small to be
easily detected (especially because of their location), but they are large enough to
cause a significant reduction in performance and efficiency.
Sometimes, after a pump has undergone extensive repair, a misalignment between
the two casing halves may occur (Fig. 19-14). Such a misalignment M often impairs
performance. When there is no other way to correct the misalignment, the next best
remedy is to grind off the edges that protrude into the waterways. Another problem
unique to split-case pumps arises when the gasket that seals the joint between both
halves of the casing protrudes into the waterways.

MULTISTAGE DONUT PUMPS

Many of the previously described problems also occur in multistage pumps. There are
also other problems. A common one arises when the impeller outlets A (Fig. 19-15) do
not fall exactly opposite the diffuser inlets B. This usually reduces both efficiency and
output.
This may be caused by one of several factors. First, when the axial position of the
impellers on the shaft is determined by a pair of nuts and the position of the shaft is
determined by a thrust bearing, the impeller outlets and diffuser inlets may not be

Figure 19-15 Section through a multistage donut pump.


Problems Related to Specific Circumstances 259

opposite one another. This may happen simply because the impellers were not
located in the correct position. Second, the length of each impeller hub may not be
exactly equal to the length of each stage casing. This is likely to happen when the
faces of the impeller hubs have been remachined, but equal lengths have not been
removed from the casing faces. This defiCiency can sometimes be corrected by
inserting a few shims between the impeller hubs. (When the differences are small,
one shim can be inserted between every third or fourth impeller.)
A related problem arises when the pump is provided with a common hydraulic
balancing device (Fig. 19-16). In such a device, high-pressure liqUid from the last
stage passes through a cylindrical gap G into a space formed by the recessed faces of
discs Bl and B2. From there, it passes through the throttling annulus T into cavity C
and exits through drainpipe D, back to the suction line. The distance between the
balancing discs at their throttling faces is kept small to build up enough pressure to
prevent them from coming into metallic contact.
The balancing device in Fig. 19-16 determines the axial pOSition of the impellers.
Its failure may alter the axial pOSition of the impellers, affecting performance and
efficiency, or it may cause a mechanical breakdown or other problems. The balancing
device may fail because of a hole or porosity in discs Bl or B 2 , which reduces the
pressure of the liqUid passing between them, causing metal-to-metal contact and
wear. Also, it may fail because drain D is clogged, increasing the pressure in space C

Figure 19-16 Section through a disc-


type balancing device of a multistage
donut pump. N = Nut which locks the
balancing disk in proper position on
shaft.
260 Problems Encountered with Centrifugal Pumps

and causing metallic contact between discs B1 and &. It might also fail from clogged
passage G. This might be caused by scale buildup in a pump left filled with liquid for
a prolonged time without being operated.
A different problem arises when passage G is very worn. This increases the
pressure in the liquid that flows between discs B1 and B2. thus enlarging their
distance. This. in turn. allows more liquid to leak through the balancing device.
causing less liquid to enter the discharge line. Consequently. at a given total head.
the pump delivers less liquid.
In a multistage donut pump. the liquid that leaves an impeller passes through a
diffuser D (Fig. 19-15) to the back-flow b. From there. it flows into the next impeller.
Both the diffuser and back-flow are located within the casing. The assembly. consist-
ing of the diffuser. the casing. and the back-flow may be designed in different ways.
Sometimes the diffuser and back-flow are cast in one piece and located in the casing.
In other instances. the back-flow and casing are designed as one integral piece. and
the diffuser is cast separately. While designs vary. the diffuser is usually designed to
be easy to replace.
To keep efficiency high. the diffuser blades are generally made very thin at their
inlet tips. However. the liqUid enters these blades at a relatively high velocity. which
makes them more susceptible to abrasion than any other part of the waterway. The
diffuser. therefore. may need to be replaced more often than any other pump part.
When a diffuser is replaced. it should not be permitted to rotate when acted on by
the flowing liqUid. This is usually done by inserting a pin in a matching hole or by a
cast-in stop. When a stop breaks or the pin is omitted. the diffuser rotates. reducing
both the head and efficiency.
In some donut pumps with removable diffusers. the waterways of the back-flow
form a continuation of the diffuser passages. When reinstalling a diffuser. great care
must be taken to ensure that the diffuser is fixed relative to the back-flow so that
there is a smooth passage from the diffuser waterways to the back-flow waterways.

PROBLEMS RELATED TO THE NATURE OF THE SITE

Climatic Conditions
Only a part of the power supplied from a source to the pumping unit is converted
into water horsepower. The remainder is used to overcome various mechanical and
hydraulic losses. Ultimately. these losses are converted into heat and cause an
increase in temperature of the pumping unit and of the liqUid. Consequently. when a
pump must operate in a hot climate. it should be installed in an adequately
ventilated (sometimes even in a cooled) pumphouse to prevent the ambient
temperature from rising above a tolerable limit. Because high ambient temperatures
are likely to reduce lubricating qualities. special heat-resistant greases are sometimes
recommended.
In pumps installed in areas frequented by sandstorms. sand should be prevented
from coming into direct contact with the rotating elements of the pumping unit. In
many cases. a lithium-based grease is recommended for use with the bearings. When
a pump operates in a very humid environment calcium-based greases are often
recommended as lubricant.
Problems Related to Specific Circumstances 261

Effluent-Handling Pumps Installed in a Dry Pit


It is common practice to allow effluent (stormwater, sewage, and so on) to flow
under its own weight to a collecting Cistern via special ducts or channels. From there
it is pumped to its final destination. Such runoff water usually carries stones, pieces
of wood, and other solid materials, which often penetrate the grates installed at the
inlet to the collecting cistern and enter the suction line, or even the pump proper.
This may cause intense cavitation, owing to blockage in the suction line, or may jam
the pump, when a solid becomes wedged between the impeller and casing. Because
such occurrences are not uncommon, it is a good idea to arrange the pump in a dry
pit (Fig. 19-17), separated from the cistern C by a waterproof wall W. This gives the
pump attendant easy access to handholes H whenever an obstruction needs to be
removed. The wall W is prOVided with a stuffing box(es) to give the suction line
access to both of its sides.
Because such a collecting cistern is usually located at the lowest elevation of the
surrounding terrain, the pumphouse is often prone to flooding during rainstorms. To
prevent the electric motor and other electric equipment from coming into contact with
floodwater, they are all placed at a higher level (as shown in Fig. 19-17) and the
pump is connected to the motor by a relatively long shaftline.
The problem with such an arrangement is that it is impossible to maintain the
pump and motor in satisfactory alignment. No matter how carefully the two parts of
the pumping unit are aligned initially, they will soon become misaligned as a result of
settling of the building and the differences in the coefficients of thermal expansion of
steel and masonry. To prevent these problems, the shafting should be connected to

w~

.::->. . c
., r----------------------
~u

Figure 19-17
Drainage pump
installed in a
dry pit.
262 Problems Encountered with Centrifugal Pumps

both the motor and the pump by universal joints U (Fig. 19-17), and the shaftline
should be provided with splined joints at suitable locations to compensate for the
differences in length caused by the differences in thermal expansion.

Pumps Installed In Floating Vessels


When a vessel is floating in water, it is always subjected to a certain amount of
rolling and pitching. The term rolling is used to describe the side-to-side rocking of
the vessel; pitching refers to the fore and aft rocking. During motion, these effects are
even more intense, because of the dynamic interaction between the vessel and the
surrounding liquid.
Whenever the head of a deep-well pump or of a wet-pit pump is mounted on a
higher deck and the water end is immersed in a sump located at a lower position, the
rolling and pitching of the vessel will continuously cause the centerline of the column
to deviate from its vertical position, thus imposing bending moments on the entire
unit. This, in turn, will result in serious problems during operation. To eliminate
problems caused by the rocking of a vessel, both ends of the pump should be firmly
connected to a single common rigid frame.
CHAPTER 20
Special Cases that
have Proven Very
Difficult to Solve

INTENSE CAVITATION AT INCREASED NPSH

All discussions about cavitation in centrifugal pumps lead to one firm conclusion:
cavitation is a direct result of an available net positive suction head (NPSH) that is
too low. This implies that when the available NPSH is lowered, the intensity of
cavitation increases and that an increase in the available NPSH reduces the intensity
of cavitation. However, there are cases in which the opposite appears to be true. In
these cases, the pump developed intense cavitation when the level of the liqUid in the
suction tank was very high. As that level was lowered, the noise was reduced. Finally,
when the liqUid neared its lowest level, cavitation disappeared completely.
The pump shown in Fig. 20-1 operated with unmistakable signs of cavitation when
the level of the liqUid in the suction tank was at ~. As liqUid was removed from the
tank and its level lowered, cavitation became less intense. Finally, when the level
approached its lowest position, Lt, cavitation disappeared completely.
This "unnatural" behavior was intimately related to the fact that the NPSH
requirements of a pump increase drastically at very high flow rates. The head
capacity curve for the pump involved and the resistance curve of the system are
shown schematically in Fig. 20-2. In this illustration, Qrnax is the maximum flow rate
at which the pump is designed to operate.
When the liquid in the tank was at its lowest level, the pump operated against a
static head HSTl. The total resistance curve of the system was represented by curve
Hrl, and the pump operated without any problem. When the level rose to its highest
position, the static head dropped to HST2. The total resistance curve is represented by
curve Ha (Fig. 20-2).
The intersection of this resistance curve with the QH curve of the pump occurred at
flow rate Q2, which is Significantly higher than Qrnax. At this flow rate, the NPSH
requirements of the pump became so high that even the increase in the available
NPSH (by the difference HSTt - H ST2) was not sufficient to suppress cavitation.
263
264 Problems Encountered with Centrifugal Pumps

. - - - . - - - - - ---------

------1-----.~

L2
- - - - - - - - - - t = -........

Figure 20-1 The effect of increased net


positive suction head (NPSH) on total head.

Hr2

I
lHr~__________~~~~__
Hsn

Figure 20-2 The effect of change of


static head on the flow rate.

DIFFERING PERFORMANCE OF IDENTICAL PUMPS


UNDER IDENTICAL CIRCUMSTANCES

In a chemical processing plant, two identical pumps were installed side by side to
transfer liquid from same source into the same pressurized container. Each pump was
provided with a separate suction pipe and discharge line, and the two pumps never
were used simultaneously. While one was running, the other served as a standby.
Everything about these two systems seemed identical, except for one detail: one
Special Cases That Have Proven Very Difficult To Solve 265

performed perfectly, and the other operated with great noise and vibration. The
troublesome pumping system was dismantled and inspected several times, but no
problem was found.
I realized that the successful pump's 2 in. diameter discharge was connected to a
2x 1.5 in. reducer that was connected directly to a 1.5 in. diameter pipeline, which
led to the pressurized container. The troublesome loop, however, had a 6-ft long pipe
of 2 in. diameter connected directly to the pump outlet, which led to the same
container. When the 2 inch diameter pipe section was replaced by a reducer and a
1.5 in. diameter pipe, the problem-causing pump began to operate satisfactorily.
The difference in performance was caused by the same problem encountered in the
previous case. Up to the theoretical flow rate Qc (Fig. 20-3) the NPSH available at the
inlet of these pumps was adequate to suppress cavitation. However, when a 2 in.
diameter pipe was installed in the discharge line of the pump, the total head against
which the pump was operating became Significantly lower than the head He, which
corresponds to flow rate Qe. This, in turn, increased the flow rate delivered by the
pump to well above flow rate Qc. This produced a demand for an NPSH much higher
than the NPSH available on site. This was the reason that the troublesome pump was
operating with high noise and vibration.

THE EFFECT OF THE LENGTH OF BALANCING


HOLES ON CAVITATION

A pump that had been in production for over 10 years and had been tested many
times and found capable of operating satisfactorily at the NPSH values shown in Fig.
20-4 by curve A, was tested and showed a satisfactory performance when operating
at an available NPSH of 5 ft and delivering 40 GPM. However, when the flow rate was
reduced to below about 15 GPM, the pump began to exhibit a significant drop in
head, even with the available NPSH of 5 ft.
The pump's performance at the available NPSH of 5 ft is shown in Fig. 20-4 by
curve B. As mentioned earlier, identical pumps were tested many times in the past
and were found capable of operating at these flow rates, well below the available
NPSH of 5 ft.

'0
C1l
Q)
I
I
en
c..
z

Figure 20-3 The effect of flow rate on a pump's Qc


NPSH requirements. Flow rate ----
266 Problems Encountered with Centrifugal Pumps

fE 22
I 20
I 18
I
160
I 16
140 14 ~
... ······C·····o......... -o ~
120 12 a:
Q) \ 8 I
~ 100 10 ~
""0
C1l 80
\ 8
z
Q)
..c \
~ 60 \ 6
0
I-
40 "- ' - 4
20 2
Figure 20-4 The effect of the length
0 0
0 10 20 30 40 50 of the balancing holes on the suction
Flow rate (GPM) performance of a centrifugal pump.

To determine the cause of that discrepancy, the impeller was replaced with an
identical impeller from another foundry. This time the pump produced satisfactory
results, as shown in Fig. 20-4 by curve C.
A scrupulous comparison of the two impellers revealed that the impeller that failed
had a recessed hub, as shown in Fig. 20-5 at A, whereas the second impeller had a
solid hub. The same series of tests was repeated on a second pump, with identical
results.
To verity whether the presence of the recess in the hub was the real cause of the
problem with the first impeller, the recess of one of the tested impellers was filled
with epoxy and redrilled, as shown in Fig. 20-5. This immediately restored the full
suction capability of the pump, as evidenced by curve D in Fig. 20-4.

Figure 20-5 Hub modification of a tested impeller.


Special Cases That Have Proven Very Difficult To Solve 267

At one point during the tests on the recessed impeller, the test engineer decided to
determine the magnitude of the available NPSH at which the head developed by the
pump at 11 GPM would return to its rated value, This happened only after the
available NPSH was increased to about 23 ft, as shown in Fig, 20-4 by point E. This
indicates that a sudden surge in the NPSH requirements might have occurred, which
might have a shape similar to the dashed curve that connects point E with curve A.
The first explanation that came to mind was that the observed effect might have
been caused by the differences between the resistance of the balancing holes, to the
flow of the liquid that comes from the back wearing ring. However, preliminary
calculations revealed that this is very unlikely.
During other previous performance tests, a pressure gage was connected to the
space behind the balancing holes, at location A (Fig. 20-6). The purpose of this gage
was to get an idea about the order of magnitude of the back pressure behind the
impeller hub.
As could be expected, the readings were different for different units of identical
pumps, owing to the differences in the clearances between the wearing rings, and so
forth. However, all test results have indicated that the difference of pressure that
exists between both sides of the balancing holes usually varies from somewhere
between 4.5 psi (0.3 atm) and 7.0 psi (0.45 atm).
The diameter of each balancing hole was 0.25 in. (6 mm). The length of each hole

Tap for
pressure
gage

Figure 20-6 Location of a tap for


measuring the pressure upstream of
the balancing holes.
268 Problems Encountered with Centrifugal Pumps

was about 0.75 in. (19 mm) before the recess was filled with epoxy, and 1.25 in. (32
mm) after the recess was filled with epoxy.
For a pressure differential of between 4.5 and 7.0 psi (0.3 to 0.45 atm), the
maximum velocity of the liquid flowing through the balancing holes could not have
been more than 10 m/sec (30.5 ft/sec). Consequently, an increase of 0.5 in. in the
length of the balancing holes could not have increased the resistance to flow by more
than 0.001 atm. The flow rate through a duct varies as the square root of the
pressure difference between both its ends. Consequently, the change in length of the
balancing holes could hardly have had any effect on the amount of liqUid flowing
through the balancing holes. There are few explanations for this effect.
However, one possible explanation could be that the observed effect was caused by
periodic shedding of vortices that may have occurred at the inlet of the impeller
vanes, when the flow rates were very low. Such a case has been discussed indirectly
by Levi in Ref. (44]. When a free stream of liquid is flowing past a zone of separation,
it induces vortices within that zone. These vortices appear and disappear at regular
time intervals. In his analysis, Levi shows that frequency f, with which these vortices
appear and disappear is a function of Strouhal's number S, which is defined by the
equation:

(20-1)

where d is the width of the zone of separation in the direction perpendicular to the
velocity U of the undisturbed stream.
According to Levi, this frequency flies in the vicinity of Strouhal's number, which is
equal to "IT /2 or one of its harmonics. According to Levi's findings, the same law
applies equally to surface waves and to the periodic shedding of vortices within a
totally submerged zone of separation. To cover both cases, Levi relates his findings to
flows of liquids that have one free surface. However, there is nothing in Levi's theory
that would have tied the appearance of vortices within a totally submerged zone of
separation to the presence of a free surface. This leads to the conclusion that we can
apply Levi's findings to centrifugal pumps. In fact, when observing the flow of liquids
at the impeller eye through a transparent suction pipe, I have often witnessed the
appearance and disappearance of vortices. Levi's theory might well explain the
experiences encountered with the tested pump.
The NPSH requirements of a pump decrease parabolically with a reduction in flow
rate (see Ref. (4] and compare Eq. 6-8). In a few cases, however, there is a sudden
jump in the NPSH requirements of a pump over a very narrow range of flow rates, as
shown schematically in Fig. 20-7 by the dashed line. Any attempt to attribute this
effect to a stall brought about by an excessive angle of incidence or by prerotation
cannot withstand criticism. Both the angle of incidence and the intensity of
prerotation increase continuously from the point of inception down to complete shut-
off. Had one of these factors been the cause of the sudden jump in the NPSH
requirements, the effect would have continued to increase all the way, with any
additional reduction in flow, down to shut-off. Against this, as can be seen in Fig. 20-7,
the jump in the NPSH requirements extends over a very narrow band of flow rates.
The shape of the dashed lines in Fig. 20-7 seems to indicate some sort of eigenvalue
effect.
Special Cases That Have Proven Very Difficult To Solve 269

,,,,
,,,,
I ,
I \
I \
/ ....
Figure 20-7 Evidence of eigenvalue
effects in the suction performance of a
centrifugal pump. Flow rate

Such an eigenvalue effect becomes possible when the frequency of appearance and
disappearance of vortices coincides with the natural frequency of some other element
that is present in the pump.
In our case, each of the balancing holes contains a column of liquid. The natural
frequency of such a column of liquid varies inversely with the square root of its
length. Consequently, the length of the balancing holes in the impeller could have
been the source of the observed effect.
Of course, there might have been other causes. However, it is important to keep in
mind that the performance of a pump can be affected by unexpected factors such as
the length of the balancing holes.

THE UNEXPECTED EFFECT OF THE PRESENCE


OF STRAIGHTENING VANES IN A SUCTION BELL

An 8 in. propeller pump, designed for 400 m 3 /hr (1760 GPM) against an head of 3.25
m (10.6 ft) at 1460 RPM, has been tested and found to operate satisfactorily at the
design flow rate. However, as soon as the flow rate is reduced below 380 cm3 /hr, the
pump begins to operate with high noise and intense vibration.
For 2 days, the staff tried in vain to eliminate the noises and vibration. Then, on
day 3, after operating for several hours, the pump suddenly qUieted down and
continued to operate smoothly. When the pump was pulled out of the well it turned
out that it had lost its suction bell.
As it turned out, the intense vibrations loosened the bolts that connected the
suction bell to the pump, causing it to fall into the well. The suction bell was re-
attached to the pump, and the noise and vibrations reappeared. Following widely
accepted practice, the suction bell had been equipped with straight, axial vanes. This
is generally regarded as an effective way to assure a uniform approach of the pumped
liqUid into the impeller eye. In despair, I decided to remove the straightening vanes
from the suction bell. The pump consequently began to operate smoothly down to
about 20% of its design flow rate.
There is no valid theoretical explanation for this problem. However, this case
history teaches us about the infinite diversity of the unexpected problems one may
encounter with centrifugal and other rotodynamic pumps.
PART

Solving Pump
Problems

This part presents an outline of a general procedure which is recommended


to follow when handling a problem with centrifugal pumps. It points out the
different steps which should be undertaken, before visiting the site. It also
suggests in what order the activities of an on-site inspection should be
carried out.
While the actual procedures must, necessarily, vary from case to case,
following the presented outline is capable to reduce significantly the amount
of time and efforts spent on solving a given problem.
CHAPTER 21
Solving Problems Prior
to Visiting the Site
PREPARING TO VISIT THE SITE

Upon learning that a pump does not operate properly, the first step is to ask as many
questions as possible. The source of many problems can often be located before an
on-site inspection is undertaken. Several kinds of questions should be asked.

The Site, Equipment, and Its Use


1. What is the pump being used for? Is it being used for water supply, effluents, industrial
purposes, or some other purposes?
2. What are the properties of the pumped liquid? Is it pure liquid, or does it contain solids?
What is the liquid's temperature? What are its chemical properties?
3. What are the type and size of the pump? If possible, obtain a cross-sectional drawing of
the pump(s) in use. Sometimes, studying these data can reveal that the pump is not
suitable for a given application, or that the operating conditions have changed since the
pump was installed, and so forth.
4. Ask about the site. Is there more than one pump? If so, do they operate in parallel, or in
series, or is there a standby pump? If possible, obtain a plan of the site, showing the
layout of the suction sump, piping, and so forth.

The answers to these questions can reveal faults in the layout of the suction sump,
the relative pOSitioning of the pumps, and the layout of the piping.

The Complaint?
If the complaint is that the pump does not perform according to the specifications,
the cause of the problem can sometimes be detected simply by comparing data of the
specified and the actual performance. A few examples follow.
QH Curve Correct. but Low Efficiency (Fig. 21-1). Low efficiency may be a
mechanical problem caused by rotating parts rubbing against stationary parts, tight
packing rings or tight mechanical seal, excessive hydraulic pressure on the seal or
packing (in a pressurized pumping system), or faulty bearings in the pump or motor.
273
274 Solving Pump Problems

- - - Rated
- - - - Tested

EFF

Figure 21-1 The QH curve is correct, but the


Q efficiency is low.

Another cause may be rubbing due to misalignment of the pump and driver, from
running near critical speed, or resuling from a casing deformed by pipeline or
bed plate stresses. Another cause may be a bent shaft.
The list is almost infinite. Generally speaking, such a curve points to mechanical
faults in the pump or the pump-driver assembly. For example, the bore of the impeller
hub may be eccentric in relation to the wearing ring diameter, or it may not be square
with its face. Either case may cause rubbing between the mating wearing rings,
resulting in excessive power consumption. If the face of the hub is not square with
the bore, this may bend the shaft when the impeller is tightened against the shaft
shoulder. The bent shaft may cause excessive power consumption in the stuffing box
as well as overheating of the bearings.
Another cause may be improper bearing installation. This often results in bearings
running hot and, thus, consuming excessive power. In a pump with an uncooled
bearing housing, this can easily be detected by touching the housing. However, when
the housing is provided with a cooled mantle, the problem can remain undetected
until the bearing is destroyed completely.
Another source of mechanical damage is resonance between the operating speed of
the pump and the natural frequency of its foundations. This vibrates the pumping
unit while the piping remains relatively stable. Consequently, the piping imposes
periodic deformations on the vibrating casing, which in turn may cause rubbing
between the rotating and stationary parts of the pump.
Yet another source of excessive power consumption is excessive clearance at the
bottom of the stuffing box. Some of the packing may extrude into the clearance,
increasing frictional losses.
An analogous case occurs when the cylindrical space where the packing is located
is eccentric relative to the shaft (compare the discussion related to Fig. 17-1). This
causes the packing to be more compressed on one side of the shaft than on the other.
When the shaft rotates, this difference in pressure may cause the packing to flow and
be eaten away by the shaft, which in turn increases power consumption.
Solving Problems Prior to Visiting the Site 275

Excessive power consumption can also result when the bearing housing is
eccentric relative to the water end, or from a damaged or cracked bearing housing.
This problem may also occur when the balancing device fails (wherever applicable;
e.g., clogged balancing holes and excessive suction pressure). In both cases the
bearing may be subjected to excessive axial loads, which in turn causes excessive
power consumption.
Lower Overall Output with Unaffected Efficiency (Fig. 21-2). Lower overall
output with normal efficiency indicates either that the pump is running at lower-
than-specified speeds or that the impeller's outer diameter was reduced below the
specified magnitude. In some cases this is due to a faulty or distorted impeller. For
example, the width of the impeller waterways may be narrower than normal as a
result of faulty foundry procedures, or because the shrouds of the impeller are bent
inward as a result of a blow against a hard surface.
Low Output, Efficiency, and Power Consumption (Fig. 21-3). Low output,
efficiency and power consumption indicate a partial blockage of the impeller eye, a
solid obstruction upstream of the impeller, or rough waterways. These things may also
be caused by improper suction sump layout or large clearances between the wearing
rings of the impeller and casing. For a pump with an open impeller, these faults
indicates that excessive clearance exists between the impeller vanes and the wearplate.
Difference between the Rated and the Actual Flow Rate Constant over a
Range of Capacities (Fig. 7-11). These differences usually indicate moderately
excessive clearances between the wearing rings of a closed impeller or a moderately
excessive clearance between the wearplate and the vanes of a semi-open impeller. The
difference q between the rated and actual flow rates is a result of the excessive
leakage owing to these clearances. This difference is equal to the increase in leakage
through the wearing faces.
Although such a change in the head capacity curve indicates excessive clearances,
the reverse is not always true. Even in Fig. 7-11, q is constant only at higher flow
rates. Near shut-off, this difference becomes much greater.

-- -..... ..... ,
, .....
.....
Rated .....
- - - - Tested ..... ,

EFF

Figure 21-2 The overall output is lower, but the


efficiency is unaffected. Q
276 Solving Pump Problems

H r - -__

- - - Rated
- - - - Tested

EFF
,""- ------ .............
.....
""
~

~
.,," HP

Figure 21-3 Low output, efficiency, and power


Q consumption.

The same is true for large clearances, such as when a wearing ring or wearplate is
missing. This is because at very low flow rates and at very large clearances, the
amount of liquid that leaks back into the impeller inlet is of the same order of
magnitude as the flow rate delivered by the pump. When that amount of liquid leaks
into the impeller entrance, it significantly affects the flow of the rest of the liquid,
resulting in lower heads throughout the affected flow rates. In such a case, changes
in performance as in Fig. 21-3 are observed.
Head Remains Unchanged at Shut-off, but Decreases Rapidly with Increased
Flow Rate (Fig. 21-4). This problem is indicative of an obstruction to the flow of
liqUid somewhere downstream of the impeller. The obstruction's resistance increases
with the square of the flow rate. At shut-off, the flow rate is zero. Consequently, the

H
--
I""O":_ _~

- ' , .....
""
"'"
- - - Rated '" '",,
- - - - Tested ,

Figure 21-4 The head is unaffected at shut-off, but it


Q decreases rapidly with increasing in flow rate.
Solving Problems Prior to Visiting the Site 277

head remains unchanged. At high flows. however. the head deteriorates rapidly
because of increased resistance to the flow. This may also be caused by an
undersized throat area of the volute or a reduced area of the passages between the
guide vanes of a diffuser.
Reduced Output and E.f/iciency with Unchanged Power Requirements (Fig.
21-5). Reduced output and efficiency with unchanged power requirements usually
indicate an obstruction to flow or excessively rough waterways somewhere between
the impeller and the point at which the pressure measurements were taken. In this
case. the impeller performs normally. consuming the expected amount of power. The
reduction of head and efficiency occurs after the liqUid has left the impeller.
Pressure Increases with Flow Rate (Fig. 21-6). An increase in pressure
corresponding with an increase in flow rate is common in pumps with double-suction

----- ...
- - - Rated
-- ........
..............
..... .....
- - - - Tested .....
..... ,

EFF

Figure 21-5 There is reduced output and efficiency.


but the power is unaffected. Q

Figure 21-6 The head increases with the flow rate. Q


278 Solving Pump Problems

impellers. It usually indicates that the impeller is mounted inversely on the shaft.
Sometimes this problem occurs when the pump rotates in the reverse direction.
However, in that case, the shut-off head is much lower than specified.
This problem also occurs in multistage pumps that simultaneously use right-
handed and left-handed impellers of otherwise identical construction and size (to
balance the axial thrust). In this case, the left-handed impellers may have been
mounted in stages that require right-handed impellers, and vice versa.

QH Curve Breaks Earlier than Expected (Fig. 21-7). A QH curve that breaks
earlier than expected is due to low available NPSH caused by an unsuitable pump
choice, a lower leve of liquid in the suction tank, a decrease in the level of liquid in
the tank during operation, obstructed flow in the suction line, a clogged suction
strainer, or a jammed or clogged foot valve. All cause loss of head of the liqUid entering
the pump, thus lowering the available NPSH.
Other possible causes include incorrect geometry or size of the suction tank,
incorrect positioning of the inlet to the suction line within the tank, transient condi-
tions during starting and stopping, and vibrations from a loose valve disk that excites
cavitation.
Other factors can cause disturbances in performance.

Pump Starts Satisfactorily but Output Decreases Gradually. A gradual


decrease in output is an indication of air leakage into the suction line or into the
pump, or the pumped liqUid contains a large amount of dissolved air or gas. When
the liquid enters the impeller eye where the pressure is lower than in the suction
tank, some of the dissolved gas escapes and becomes trapped in the pump inlet,
generating an obstruction to the flow of liqUid, which results in reduced output.

Pump Starts Satisfactorily but Output Decreases then Returns to Normal.


When output decreases then returns to normal, this usually indicates that a small

- - - Rated
- - - - Tested

Figure 21-7 The QH curve breaks earlier than


Q expected.
Solving Problems Prior to Visiting the Site 279

air pocket is in the suction line. When the pump is put into operation, the air from
this pocket is carried away by the flowing liquid and brought into the eye of the
impeller. Provided the amount of air in the original air pocket was not too great, the
impeller inlet will only be partially blocked. This allows a certain amount of liquid to
flow through the pump, but at a reduced rate. In time, the flowing liquid may carry
away the entrapped air, and full performance will be restored.
Output Periodically Decreases then Returns to Normal. When the output
decreases periodically then returns to normal, the inlet of the pump impeller may be
located below the liquid level in the suction tank, and the pump delivers liquid at a
higher flow rate than the rate at which the liquid enters the suction tank. Under
these circumstances, the liquid level in the suction tank drops continuously until it
is below the inlet to the pump intake. This allows air to enter the impeller eye, thus
reducing or stopping the output of the pump.
When the output of the pump drops below the rate at which the liquid enters the
suction tank, the level rises until it floods the impeller eye again. The pump then
returns to its normal operation, and the cycle repeats itself.
Dry Cracking Noises. Dry cracking noises usually indicate cavitation. If the
noises come from the area of the pump inlet, the pump may be located too high
above the liquid level in the suction tank, there may be an obstruction in the suction
line, the pump may not be capable of handling the available NPSH, and so on.
When the noises come from the discharge end of the pump, this usually means
that the pump is handling a prohibitively large flow rate, causing cavitation in the
volute throat or at the inlets to the diffuser vanes. This may be the result of a drastic
drop of pressure in the discharge line caused by a pipe rupture or a similar event.
Vibration. Vibration poses questions about whether the pump and driver were
realigned after on-site installation. Additional questions include the following. Is there
any overheating of bearings? Does the vibration come from the pump alone or from
both the pump and driver? The latter case suggests misalignment between the driver
and pump.
If there is noise coming from the pipeline, a loose valve disc or resonance may be
suspected. If any structural elements or the entire pumphouse vibrates, resonance is
likely. Alternatively, Vibrating pipelines and structural parts may indicate stresses
imposed by the pipes on the pump or structure.
Malfunction Occuring Occasionally. Signs of cavitation that occur only
occasionally lead to questions about variations in the liquid level in the suction tank,
the possibility of weeds and debris entering the suction line, and the possibility that
the pump occasionally handles liquid at an elevated temperature.
If the pumping plant has more than one pump and the problem occurs only when
several pumps are operating Simultaneously, the next question should be whether
they take the liquid from a common suction sump. In this case, the trouble may well
be caused by improper sump layout or improper location of the pump intakes within
the suction sump. This leads to additional questions on the geometry of the suction
sump and how the liquid enters the sump.
If the problem occurs when one or more of the pumps are not operating, and all
pumps get their liquid via a common manifold, there may be air leakage into the
manifold through one of the pumps that is not in operation.
280 Solving Pump Problems

OTHER USEFUL INFORMATION BEFORE VISITING THE SITE

How, when, and under what circumstances was the malfunction detected? If the
complaint is about the pump not meeting the specified performance level, a detailed
description of the manner in which this conclusion was reached should be requested.
This may reveal that the source of the complaint is simply a faulty or unreliable test.
This also raises questions about the instruments used, the method of calibration, the
location of the taps to which the pressure gages have been connected, and so on.
Have any corrective steps been taken? The results of these actions may narrow
down the possible causes of the reported malfunction.
What instruments, tools, machine tools, material-handling devices, and skilled
personnel are available on the site? The answers may help determine what pre-
paratory work should be undertaken before the on-site inspection. These prepar-
ations may include adding taps to connect the pressure gages, providing special
means for measuring the flow rate, measuring the power consumption, and checking
the accuracy of the instruments.
All of this information and any other additional particulars that may come up
during the collection of that information help determine what actions should be
carried out before the on-site inspection. The information collected during this
remote approach to problem solving may prove sufficient to determine the cause of a
given malfunction, enabling the pump specialist to make certain suggestions about
the remedial steps to be taken to eliminate the problem. Even if the source of the
problem is not evident from the reported details, an analysis of the collected data
may enable the pump specialist to suggest additional tests and trials that may solve
the problem.
When this remote approach is unsuccessful, an on-site inspection is the next step.
However, before the inspection, check the information that has been collected against
the checklist in Appendix A. This will help determine the most probable cause of the
malfunction and will reduce the number of possibilities that need to be examined.
CHAPTER 22
Conclusions Dra\Nn
from Visual Inspection
of Failed Parts
Sometimes. damaged or broken parts provide clues to the cause of a pump's failure.
Small parts. can be shipped to the pump specialist. thus allowing him to solve the
problem without an on-site inspection. When the failed parts are too bulky or cannot
be removed from the site for any other reason. an on-site inspection may reveal the
source of the failure at an early stage of the inspection.
The pump parts most likely to fail are rotating parts that come into contact with
the stationary parts of the pump or with the liqUid. and all stationary parts that come
into contact with the liqUid or with the rotating parts.

BEARINGS

Chapter 16 includes a detailed discussion of the factors that can produce bearing
failures. It also presents various cases in which the cause of the failure can be
detected by means of a visual inspection of the failed parts.

SEALS AND PACKINGS

An extensive study of seals and packings is presented in Chapter 17. including a


discussion of the connection between the visual appearance of a failed seal and the
cause of that failure.
In a packed stuffing box. the most common failure is a badly worn shaft or shaft
sleeve where these parts come into contact with the packing. Sometimes. the wear
may be extensive enough to cause the shaft to fracture.
Such failures are usually caused by inadequate lubrication or cooling of the
stuffing box. This is rather common and is usually caused by the access of the
cooling liquid to the lantern ring being blocked (Fig. 12-21). This. in turn. may be
caused by installing the lantern ring in an incorrect location (not opposite the inlet of
the cooling liquid) or by failing to install it.
281
282 Solving Pump Problems

Another cause of such damage is an improper choice of packing material or


lubricant for the packings (see Chapter 17 and Appendix B-3). Human error is also a
factor. Failing to provide coolant to the cooling system of the stuffing box, for
example, can lead to failure.
Excessive shaft wear in the area of the stuffing box can be caused by abrasives in
the pumped liqUid. A common remedy in such cases is to introduce a clean liqUid
under pressure into the stuffing box lantern. In less severe cases, the situation can
be alleviated by introducing a suitable "stiff' grease into the stuffing box. Failure to
follow the proper procedure is the most frequent cause of this kind of failure.

FRACTURED OR DAMAGED SHAFT

Some causes of shaft damage have been discussed in conjunction with the problems
encountered with stuffing box packings. Here, we shall list a few additional causes.
In Chapter 18, we discussed the shape of various fractures. Sometimes, visual
inspection of a fracture can indicate its source. When a shaft shows signs of failure
from shear, for example, excessive resistance to rotation is usually indicated. This
may be due to a bent shaft, a solid object becoming jammed between the impeller and
the casing, a galling action between the rotating and stationary wearing rings (see
Chapter 18), or excessive power requirements caused by the high specific weight of
the pumped liqUid. In general, shear fracture may be caused by any factor capable of
increasing the torque on the shaft. These include binding of the bearings owing to
poor lubrication, a stuffing box that is too tight, or any other similar cause. Another
cause of shear fracture of a shaft may, of course, be due to poor shaft material or
stress concentrations caused by machining marks.
Tensile failure (Chapter 18) may be caused by clogged balancing holes or failure of
any other means of balancing (or reducing) the axial thrust. It may also have been
caused by excessive pressure caused by the high specific weight of the handled
liqUid, or by shaft material of poor quality.
When the shaft fracture exhibits signs of bending failure, it is usually the result of
excessive lateral forces caused by misalignment between the stationary and the
rotating parts. Such a failure can also be caused by a solid object becoming wedged
between the impeller vanes. This may produce serious lateral centrifugal forces on
the rotating pump element. OccaSionally, when a pump with a single volute casing is
handling a liquid of high specific gravity, shaft failure caused by bending may be a
result of excessive radial forces (compare Chapter 10).
Sometimes, a fracture shows rust or discoloration on part of the fractured area.
This often indicates that the shaft material may have had thin cracks prior to its use
for the given part, or that the fracture is due to fatigue caused, for example, by an
unbalanced rotating element.

DAMAGED WEARING RINGS

Discoloration or wear on one side of the stationary ring indicates that the casing bore
in which the ring has been installed is not concentric with the centerline of the shaft,
or that the bore of the wearing ring is not concentric with its external cylindrical
surface.
Conclusions Drawn from Visual Inspection of Failed Parts 283

Galling between the stationary and the rotating wearing rings indicates that the
ring materials have a strong galling affinity or that the pump has been running dry.
without having been primed.
When the wearing rings erode within a short time. this generally is an indication of
the presence of abrasives in the pumped liquids. One of the more popular methods of
slowing down the erosion caused by abrasive particles is to use special abrasion-
resistant materials for the rotating and the stationary rings.
I have had good experience with shrinking a stainless steel ring onto the impeller
and vulcanizing a rubber lining onto the inside of the stationary ring. In more severe
cases. stellite is sometimes used for hardfacing both the rotating and the stationary
rings. Also. whenever the pump is provided with straight, cylindrical wearing rings.
as shown in Fig. 22-1 at A. it is advisable to redesign the wearing ring into an
L-shape. as shown in Fig. 22-1 at B.
Another popular way to handle liqUids that contain solid material is to choose face
F. which is perpendicular to the axis (see Fig. 22-2), for the sealing surface instead of
the cylindrical part of the wearing ring. Such a design is usually used in conjunction
with a bearing frame in which the axial distance between the impeller and the casing
can be adjusted externally. without haVing to open the pump (compare Fig. 19-3).
Such an arrangement allows excessive leakage caused by the wear of the rings to be
easily reduced. by simply readjusting the distance between the sealing faces. For
additional protection against wear. an L-shaped ring made of an abrasion-resistant
material is shrunk on the impeller. as shown in Fig. 22-2 at A. Sometimes. a second
ring B of suitable material is pressed into the casing. as shown in Fig. 22-2 at B.
A very common error is to make the rotating ring flat and attach it to the impeller
face with countersunk screws. as shown in Fig. 22-2 at C. Such an approach has
been tried frequently in the past, for reasons of economy. In practice, however, it
invariably results in fast erosion of the face of the ring downstream of the counter-
sunk holes, as shown schematically in Fig. 22-3.
Whenever a pump with a vertical shaft is used to handle storm runoff or similar
liqUids. particularly heavy erosion can be caused by the negligence of the pump
attendant. Such liquids usually contain a vast amount of sand. small pebbles. and
similar material. As a rule. the pumps used for such duties are designed with ample
waterways that allow the free passage of these solids through the pump. Mter the
pump has been stopped. however. most of the solids that remain in the pump and in
parts of the pipelines settle on the bottom of the casing and become trapped in the
space S (Fig. 22-4). When the pump is restarted. the accumulated solids are forced

Figure 22-1 Cylindrical and L-shaped


wearing rings.
284 Solving Pump Problems

Figure 22-2 Sealing off impeller leakage on a face


perpendicular to the axis.

into the gap between the stationary and the rotating rings owing to the pressure
generated by the impeller. The only way these solids can move is to grind their way
through the sealing gap between the impeller and the casing. This generates such a
powerful grinding action that the wearing rings may be destroyed within a very short
time.
To avoid such a rapid destruction of the wearing rings, such pumps are usually
provided with handholes that enable the pump attendant to remove the debris that
Conclusions Drawn from Visual Inspection of Failed Parts 285

Figure 22-3 Erosion of the wearing


ring downstream of the countersunk
holes.

accumulates in the space S (Fig. 22-4) of the pump every time the pump is stopped.
Unfortunately. not all pump attendants are aware of the importance of following this
procedure-often with disastrous results.

DAMAGED IMPELLERS

When the wearing ring of an impeller shows wear or discoloration on only one side.
either the bore of the hub is not concentric with the cylindrical surface of the wearing
ring. the matching part of the shaft is not concentric with the centerline of the
bearings. or the shaft is bent. Another cause can be an extrusion on the shoulder of
the shaft or the face of the impeller hub. as explained in connection with Fig. 15-16.
Broken shrouds or blades frequently mean that a solid object has entered the
pump and has hit the area causing damage. Rough. irregular surfaces full of small
holes and indentations (fig. 22-5) may indicate damage from cavitation or chemical
erosion or both. It is often very difficult to distinguish visually between these two
kinds of damage. However. when the pump is handling a clean. chemically and
electrochemically inert liquid. the damage is obviously due to cavitation. Also,
cavitiation is mostly confined to certain areas of the wetted surfaces (in zones of low
pressure). whereas chemical erosion is usually spread over a wider area.

DAMAGED CASING

A broken flange or pump pedestal is often an indication that the pipelines are
imposing excessive stresses on the pump. Such a failure can also be caused by poor
workmanship during installation. this is, uneven tightening of the bolts.
286 Solving Pump Problems

Figure 22-4 Space in which


debris can become trapped.
Conclusions Drawn from Visual Inspection of Failed Parts 287

Figure 22-5 Impeller damaged by chemical


erosion or cavitation, or both.

A cracked casing may be caused by an external blow from a solid object or by water
hammer. It can also be caused by a wide spectrum of other transient causes, such as
thermal shock. Finally, a casing may crack under internal pressure owing to the poor
quality of the material of which it has been made or to residual thermal stresses that
may be present within the casing material as a result of inadequate heat treatment.
Finally, a casing can be damaged or even ruined completely by chemical erosion,
cavitation, or by the abrasive action of solid particles present in the pumped liquid.
CHAPTER 23
On-Site Inspection
and Testing
In most cases, complaints about unsatisfactory pump performance come from its
user, who is usually familiar with what the malfunction of the pump does to its
performance. However, he may know very little about the pump proper. Because of
this, reports about unsatisfactory pump performance are often distorted and are
intermingled with the personal opinions of the user. This, in turn, inaccurately
represent the problem.
In addition, the person who reports the problem is often asked questions whose
significance he neither realizes nor understands. Again, this may lead to incorrect
data. Therefore, the first step of an on-site investigation is to verifY the data obtained
prior to the visit.
Having accomplished this, the original plan of action must usually be modified
before troubleshooting can begin. It starts with a search for the causes of the
malfunction, using the simplest, least time-consuming procedures. If these do not
reveal the cause of the problem, one must resort to more elaborate procedures.

PROCEDURES THAT DO NOT REQUIRE PUMP


OR PIPELINE DISMANTLING

This includes a study of where the pumped liquid comes from, layout of the suction
sump and suction line, available net positive suction head (NPSH), whether suction
conditions are variable or constant, and whether there are periodic changes of the
liquid level in suction sump. One may also study details of the physical and chemical
properties of the pumped liqUid, its temperature, and so on.
Examine the suction line and its connections for entrapped air that may form air
pockets and for air leakage. If the suction pipe has a foot valve, check for air leakage.
When the static head against which the pump operates is Significant, the entire
system is simply filled with liqUid and the pump stopped. This subjects the suction
line to the total static pressure of the system. Any liqUid leaking from the suction line
indicates air leaks during operation.
When the static head is very low, proceed as follows. If, in addition to a foot valve,
there is a valve downstream of the pump, close this valve and subject the entire
289
290 Solving Pump Problems

system to pressure from an outside source. Here, again, a leak in the suction line
indicates air leakage during operation. (see also Chapter 15).
When the pump gets water from a natural source such as a river or well,
information about possible seasonal changes in water levels should be collected.
Also, when the pump draws liquid from an open basin or river, inspect the hygienic
status of the source. The presence of debris or solid matter indicates that the suction
line or pump may be partially or totally blocked.
The pumping unit(s) must also be inspected. All nuts and bolts should be checked
for tightness and all electrical connections l should be checked for tightness and
correctness.
When any of these inspections reveal faults that can be corrected easily, eliminate
the faults on the spot and check the pump for correct operation. However, if the
correction requires a significant amount of time and labor, continue searching for
additional causes of malfunction. Many different factors may cause a pump to
function improperly. Often, several problems occur simultaneously. Therefore, it is
usually less expensive and less time consuming to check all possibilities during one
visit to the site than it is to find each one during separate visits.

Inspect Pumping Unit While Not in Operation


This should be done only after making foolproof provisions that the pump will not
be aCCidentally restarted during the inspection. For electric drive pumps, remove the
main fuses and post a note on the fuse box, warning against reinstalling them before
the end of the inspection. If possible, put a lock on the fuse box or lock the room in
which the fuse box is located.
In small and medium-sized pumps, the first step is to turn the shaft by hand. High
resistance indicates that something within the pumping unit is binding due to an
improperly installed mechanical seal, solid matter wedged between a stationary and a
rotating element, seizure of parts operating within small clearances, or the bore is
eccentric relative to the shaft axis.
When the shaft turns easily during part of one revolution and requires great effort
during the rest of it, a combination of eccentricity of a stationary pump element and
a bent shaft is indicated. This happens when both a rotating part and a matching
stationary bore are eccentric relative to the pump's centerline.
A bent or eccentrically running shaft can be detected by putting a dial indicator
against its exposed surface and turning the shaft by hand. If a dial indicator is not
available, any pointed object can be fixed to some stationary part of the pumping unit
in such a way that there remains only a small gap between the shaft and the pointed
end of the object. Turn the shaft slowly by hand and observe the width of the gap.
Any eccentricity will demonstrate itself as a variation in the width of this gap.
While turning the shaft by hand, pay attention to any sensation of rubbing or
clicks. A click may indicate a granule of dirt in the bearings, a faulty baU(s), or faulty
races of the ball bearings.

IThe inspection of electrical contacts should be performed only after the supply of electrical power has
been absolutely and completely disconnected, and after all the necessary precautions have been taken to
ensure that the power supply will not be aCCidentally reconnected during the inspection.
On-Site Inspection and Testing 291

Next, check the shafts of the pump and the driver for axial and radial play. In small
pumps, try to move the shaft by hand. In larger units, a lever may be needed to do
this. The lever can be made from any suitable object available on site. The lever can
also help to check the rigidity of the connections between the pump and base plate,
driver and base plate, and base plate and foundations. Of course, when using such a
lever, care should be taken not to exert an excessive force so as to damage the tested
parts.
One of the most important checks is the alignment between the pump and the
driver. However, this test has significance only when preliminary inspections show
that the shaft of both the pump and the driver are running correctly.
Basically, there are two ways in which misalignment between a pump and its driver
may occur. First, the centerline of both shafts meet at one common point C (Fig. 23-1),
and form a broken line. Second, the centerlines of both are parallel and never meet
(Fig. 23-2). Sometimes the shafts are both not parallel and displaced (Fig. 23-4).
The first misalignment can be checked in the following ways (Fig. 23-3). First, if a
filler gage is available, the gap a between both coupling halves is measured at several

c
\
Figure 23-1 Centerlines of pump and
motor are not parallel, but meet in one
point.

Figure 23-2 Use of a straight-edge for


checking shaft alignment.

Figure 23-3 Some of the dimensions to


be verified during shaft alignment checks.
292 Solving Pump Problems

------ - ------
-----

Figure 23-4 Use of a dial indicator for


checking shaft alignment.

points around the periphery. In a well-aligned aggregate, all measurements show the
same dimension. Second, when a vernier gage or micrometer is available, distance b
of the outer rims should be measured along the periphery of the coupling halves. Any
misalignment will be evident in different dimensions at different points of the
periphery.
Misalignments belonging to the second group are usually checked by putting a
straight-edge S on half of the coupling and checking whether it touches the periphery
of the other (Fig. 23-2). Again, several spots along the periphery should be checked.
Sometimes the outer diameter of one coupling is slightly larger than that of the
other. This generates a gap G between the straight-edge and the smaller coupling
half. In this case, measure the gap's magnitude with a filler gage at several distant
locations of the periphery. In a well-aligned aggregate, all measurements should show
the same gap dimensions.
An alternative method of checking shaft alignment is to fix to one coupling half a
dial indicator (Fig. 23-4), after the couplings have been disengaged. The point of the
indicator is set against the other coupling half (or against an exposed part of its
shaft). The coupling half with the indicator is turned slowly around the other shaft.
Any misalignment appears on the dial as a displacement of the indicator needle.
If no dial indicator is available at the site, a needle can be used instead. In that
case, the point of the needle should be set to allow a small gap between its point and
the other half of the coupling. Any misalignment manifests itself by variations in the
width of the gap.
Always turn the shaft to which the indicator is fixed (A in Fig 23-4), not the other
shaft B. When shaft B is turned and A remains stationary, no signs of misalignment
can be discerned. Also, this method will not disclose any misalignment when point X
of the intersection of the centerlines of the shaft lies near the plane of the dial
indicator.

Inspecting the Pump during Operation


These observations include measurements of operating speed and the direction of
rotation. Observation of the direction of rotation can be made only when the speed is
relatively low. To do this, the pump is started and stopped one or more times and the
direction of its rotation is observed during the slow-down.
The next step is to listen to any excessive noises and check for vibration. To
determine the cause of noise or vibration, first determine the location of the zone
On-Site Inspection and Testing 293

from where they come. This is best accomplished with the aid of a stethoscope. In
absence of a stethoscope. a primitive method can be tried. which consists of grabbing
a pencil in the teeth. plugging the ears with the fingers. and touching the other end of
the pencil to different points of the pump.
Another way to determine the source of noise is with a narrow band analyzer. This
instrument shows the intensity of noise at different frequencies. This. in turn. may
indicate from where the noise is coming.
When using a narrow band analyzer. keep in mind that a high amplitude at a
certain frequency does not necessarily mean that the given phenomenon occurs with
an excessive intensity. Sometimes. a high amplitude may result when the frequency
of the given phenomenon coincides with the natural frequency of a certain structural
element of the pumping unit or of the pumphouse. For example, if the narrow band
analyzer indicates a high amplitude at a frequency at which the impeller blades move
past the volute tongue, it does not necessarily mean that the resulting pressure
fluctuations of the liqUid are excessive. It may simply mean that this frequency
equals the natural frequency of the pump casing or the bedplate.
A primitive. but useful means of determining the natural frequency of such a
structural member is to strike it with a hammer and measure the frequency of the
excited vibrations. The frequency of these vibrations represents the natural frequency
of the struck structural element.
Additional observations that must be made while the pump is running are related
to the temperatures of the bearings and the stuffing box. The highest allowable
temperature is determined by the properties of the lubricant or the pumped liqUid,
and by the physical properties that the affected materials possess at elevated
temperatures. However. such heating often indicates other defects, such as faulty
bearings, a bent shaft, or an excessively tight seal.
When no other data are available. a bearing should never become so hot that one
cannot place a hand on the housing and keep it there for at least one minute.
Finally. much can be learned during a running test about whether the pump is
really suitable for the duties for which it has been selected. When ordering a pump
for a new application. the characteristics of the system are, at best, only estimated.
This is usually accomplished on the basis of extensive calculations that. in turn. are
carried out after making a number of simplifYing assumptions. In real life. however.
the actual conditions may vary considerably from the assumptions made during
these calculations.
After the pump is installed, it is often very simple to determine the exact
characteristics of the system. This is particularly easy when the suction line is
provided with a foot valve. or there is a valve immediately downstream of the pump.
The only things required are a pressure gauge immediately downstream of the pump
and a means of measuring the flow rate.
The pump is allowed to run until liqUid fills the entire system. When a foot valve is
present in the suction line the pump is stopped. and the pressure reading near the
pump is taken. This pressure and the static elevation between the centerline of the
pressure gage and the level in the suction tank are the total static head of the
system. The pump is restarted and the discharge valve (if any) is opened completely.
The flow rate and pressure are measured again. The new pressure readings minus
the initial reading gives the total frictional resistance Hfa of the system at the given
flow rate Qa.
294 Solving Pump Problems

The frictional resistance of a system varies almost exactly as the square of the flow
rate. Thus, the shape of the curve of the frictional losses of the system (Fig. 5-12).
can be determined. This follows from the relation

(23-1)

where Hf is the sum of all frictional losses in the system for any given flow rate Q.
Such a method often makes it easy to determine the characteristics of a system
with a high degree of accuracy. However, measuring instruments on the site are often
nonexistent. For pressure readings, it is usually simple and inexpensive to install a
Bourdon gauge wherever it is needed. A flow meter, however, may pose more serious
problems. Some improvisations can help. When the liquid can be directed to a basin
or tank, the measurements of the container can be taken and the time required to
raise the liquid level in the tank by a certain amount can be measured. If no
container is available, an open channel can sometimes be blocked at two locations,
converting that section of the channel into a container. Flow can also be measured by
using an improvised weir in an open channel or an orifice at the end of a discharge
pipe.

PROCEDURES THAT REQUIRE PUMP OR PIPELINE DISMANTLING

In some installations it is easier and simpler to dismantle the pump than the pipe
connections. This is particularly true of horizontally split casing pumps and with
back pull-out, end-suction pumps. In other cases, it is more convenient to start with
the pipeline.
Removing pipe sections adjacent to the pump allows a limited look at the inside of
the water end. This allows checking for foreign objects in the waterways and for
broken impeller vanes.
When dismantling sections of the discharge and suction piping, first loosen the
bolts connecting the pipeline with the pump. Any motion of the flanges indicates
stresses imposed by the pipeline on the pump.
Before opening the pump proper, study the available sketches or drawings of the
cross-section of the pump and any instructions supplied by the pump manufacturer.
Carry out the dismantling process step by step, checking the instructions frequently.
Figure 19-3 presents the cross-section through an end-suction pump with a semi-
open impeller. The clearance between the impeller and the pump casing is set with
the aid of the shims S.
Even before opening the water end, check the front clearance between the impeller
and casing by removing the shims and pushing the rotating element forward until the
impeller comes in contact with the wear face of the casing. By measuring the width of
the gap between the cartridge flange ej of the bearing cartridge be and the bearing
hOUSing bh, and by subtracting this dimension from the thickness of the removed
shims, the clearance is readily obtained.
It is possible to push the impeller toward the pump casing by tightening the
cartridge bolts eb, until the shaft cannot be turned by hand. However, before taking
any measurements of the remaining gap, the bolts should be loosened again. This
On-Site Inspection and Testing 295

must be done to prevent a bent cartridge flange that might cause inaccurate
measurements.
These gap measurements should be done before dismantling. If the clearance is
excessive, remove several shims to restore the required clearance. (Depending on the
type and size of the pump, the magnitude of the clearance should be between 0.01 in.
and 0.02 in.) Afterward, test the pump for performance. The main cause of
unsatisfactory performance is often excessive clearance. In such a case, this
immediately shows up in the results of the new test.
After removing the shims and pushing the cartridge until its flange touches the
bearing housing, it may happen that the shaft still continues to turn easily by hand.
This indicates that the clearance is greater than the thickness of the shims. In this
case, tighten the cartridge bolts firmly and test the pump for performance. If the test
results are still unsatisfactory, open the water end.
First, remove the casing from the rest of the pump. Inspect all parts for excessive
wear and damage. Badly damaged parts should be repaired or replaced.
Next, measure the actual clearance between the impeller vanes and the casing.
This clearance is equal to the thickness of the shims plus any additional clearance
that remained after their removal. To determine the magnitude of the additional
clearance, measure the distance of the impeller face from the reference plane P, and
compare it with the distance of the mating face of the casing from the same plane
(compare Chapter 19). Take these measurements at several locations on both the
impeller and casing to determine whether both faces are parallel to the plane P.
Next, fix a dial indicator (or a needle) to the impeller and check whether it runs
truly in relation to the plane P and the rabbet R. If not, dismantle the impeller and
check the periphery of the shaft and its shoulder for true running relative to the plain
P and the rabbet R. Also, check the impeller hub and the matching shaft shoulder for
extrusions and dirt (compare Fig. 15-16).
Having completed this stage, dismantle the pump completely and inspect all parts.
Finally, repair or exchange all faulty parts, reassemble the pump, and test its
performance.
PART· V

Eliminating Pump
Problems
and Modifying
Performance

This part discusses different means that are available for moditying the
performance of an existing pump. The need for such modifications often
arises because of temporary or permanent changes in the existing operating
conditions or because, at the time of pump selection, no adequate data
were available that would have made it possible to make a correct choice.
CHAPTER 24
Remedial Methods
It is often very costly and time-consuming to determine the cause of a malfunction
directly. Often, it is faster and less expensive to try a simple remedy and see whether
it works. In other cases, performance of a pump needs to be modified for one or more
of the following reasons:
1. The actual resistance to flow of the pumping system is different from the calculated one.
(e.g., owing to differences in the actual smoothness of the wetted surfaces, and so forth).
2. Time-related changes such as buildup of scale or rust in pipelines or seasonal (or
permanent) changes in water levels of the well, the river, and so forth, from which the
pump draws its supply of liquid.
3. Temporary or permanent changes in demand, such as changes in the production
schedule of a processing plant, in the popUlation, or in the manufacturing facilities
located within a certain municipality.
4. Temporary or anticipated changes in the availability of an adequate power supply, or an
anticipated increase in its availability. Similarly, a pump may have to operate under
certain conditions for a limited time only. and later these conditions are expected to
change.

Although the best practice is to use the right pumping unit for the given purpose,
such a policy may sometimes require prohibitive changes in the layout of the piping,
the foundations for the bedplate of the pumping unit. or even in the building in
which the given unit operates. Consequently, knowledge of available remedial means
and procedures may be of great help.

MISCELLANEOUS REMEDIES

As has been pOinted out earlier, it is sometimes faster and less expensive to attempt
to eliminate a problem by trying certain remedial means than by trying to locate the
exact cause of the malfunction. For example, when air leakage between two flanges is
suspected (compare Chapter 15), wrap the suspected joint tightly with a plastic or
rubber sheet. When the guess is correct and the wrapping is done carefully.
performance immediately increases.
Another class of remedial means is required in emergency cases in which a pump
must continue to operate temporarily until the basic cause of the malfunction is
corrected.
299
300 Eliminating Pump Problems and Modifying Performance

Still another class of cases arises when a direct elimination of the cause would be
prohibitive owing to cost or other points of view. Such a case arises, for example,
when a large suction sump is designed incorrectly. To demolish the sump and rebuilt
it would be time-consuming, costly, and sometimes even impossible because of space
limitations.
In such a case, the next best possibility is to try to straighten the flow into the
pump intakes by means of baffles located within the suction sump. Again, this
requires much experimenting and involves time and money. To reduce such
expenditures, it is sometimes better to build a scale model of the sump and the pump
intakes and to perform all the experiments on the model.
When it is suspected that the rotating unit is operating at or near its critical speed,
the simplest remedy is to fix a concentric ring or bushing of considerable mass to the
shaft. The additional weight changes the natural frequency of the rotating unit and
eliminates the primary cause of the malfunction.
When the layout of the suction line causes disturbances in the flow of the liquid
entering the impeller, installing a suitable flow straightener verifies the assumption
and remedies the problem.
When the problem is resonance between the pump and pipeline or between the
pump and the disc of a valve installed in the pipeline, the simplest procedure may be
to remove a section of the pipe directly connected to the pump and replace it with a
suitable rubber hose.
For a problem during sudden transient conditions, install a suitable remedial
means, as recommended in Chapter 14 for water hammer.
In many pumps the intensity of noise is rather low at the design flow rate, but it
may become intolerably high at very low flow rates. Sometimes, a change in operating
conditions may necessitate operating a pump at lower flow rates than originally
anticipated.
The best solution is to remove the pump and replace it with a more suitable one.
When this is impossible, the noise level can sometimes be reduced by cutting off part
of the volute tongue or the gUide vanes, thus increasing their distance from the
impeller blades. Another solution is to allow the pump to deliver a higher flow rate
and to bypass the surplus liqUid back to the suction.
For cavitation, raise the level in the suction tank or transfer the pump to a lower
location. When the liqUid enters into the suction line from a closed tank, increase the
pressure in the tank. Another remedy is to partially close the discharge valve. This
reduces the flow rate along with the net positive suction head (NPSH) requirements of
the pump. However, at flow rates that are too low, this may cause noisy operation
and produce other instabilities.
An auxiliary pump can also be installed that is specifically designed to handle
liquids at a low NPSH upstream of the main pump. In some modern pumps, an
inducer can be installed in the suction nozzle (Fig. 24-1). This may solve the problem.
For an impeller with an even number of vanes, removing inlet portions of alternate
vanes enables the pump to operate at somewhat reduced NPSH values (Fig. 24-2).
Figure 24-3 shows the effects of such a change in a 3-in. pump running at 3560 RPM.
However, when shortening the inlet tips of the alternate vanes, remember that this
procedure somewhat reduces the total developed head. In many cases, such a reduc-
tion can be counteracted by installing an impeller of larger outer diameter or by
underfiling the outlet tips of the vanes of the existing impeller. The reduction LlH,
Remedial Methods 301

Figure 24-1 An inducer installed


upstream of an impeller.

Original impeller Modified impeller


I

Figure 24-2 An impeller modified for improved suction performance.


302 Eliminating Pump Problems and Modifying Performance

__X" ~--------------------~71
260 66

351 ~
~240 I ::2
Ql
6 a.
~ I ~
J: - - Plain impeller
I
-g 220 I 0- - 0 Modified impeller

I
x_"....----'""" 506 ~
Q)
I
'r Figure 24-3 The effect of
I 500
200 I impeller modification on net
I positive suction head (NPSH)
I requirements. X =NPSH below
180 I which the head starts to
decrease with any reduction in
0 5 10 15 20 25 30 NPSH. Y = NPSH at which the
NPSH head breaks down completely.

caused by the removal of portions of alternate blade tips can (usually) be calculated
by

(24-1)

where H is the head developed by the impeller with full blades, ilH is the difference
between the original head and the head developed after alternate blades are
shortened, and Rb, RI, and R2 are as defined in Fig. 24-2.
Sometimes, a pump must operate under cavitating conditions. In such cases,
special steps must be taken to prevent rapid mechanical deterioration of the pump.
The only known means of doing this is to make the affected parts from special
materials or to protect them with special coatings or liners.

6.25 mm
90 -11--

~
70
D1J 90

r
260
()) 80
E
i' 50 70 ~
(.)
300 60 Q;
Q.
250 50;:
(.)
[L 200 40 a5
I
III 150 30~
w
100 20
50 10
0
0 250 500 750 1000 1250 Figure 24-4 The effect of removing
Flow rate (m 3/hr) surplus metal from the blades.
Remedial Methods 303

A wide and ever-growing spectrum of materials and processes plays a role in this
method. As a general rule, the cavitation resistance of a material increases with its
wear resistance (compare Chapter 7). However, in many cases, cavitation often
increases the vulnerability of a given surface to chemical corrosion. Consequently, in
severe cases, consult the manufacturer about the recommended material before
using it to reduce damage caused by cavitation.
Sometimes a simple visual inspection of the waterways can reveal the cause of a
problem. Visual inspection may also indicate the manner in which the given problem
can be eliminated. An inspection may, for example, reveal a sharp edge protruding in
the path of the liquid (some of the effects of a sharp edge are discussed in Ref. 45), or
cast on lumps of metal and so forth. For example, in one case, a pump deSignated to
deliver 3100 GPM against a head of 211 ft was unable to develop more than 200 ft of
head at that flow rate. Also, the efficiency at that flow rate was lower than expected. A
visual inspection of the impeller revealed that the width of the outlet edges of the
blades was unusually wide at their center whereas, according to the drawing, it was
supposed to be equal to only 6.25 mm (see Fig. 24-4). Mter regrinding the vanes to
conform to the drawing, the expected performance of the pump has been restored.
Results of a similar order of magnitude can be achieved by polishing a particularly
rough spot in the waterways or by coating the casing waterways with a smooth
enamel (compare Chapter 7).
CHAPTER 25
The Effects of Speed
and Impeller Outer
Diameter on Pump
Performance
One of the most common ways to alter the performance of a centrifugal pump is to let
it run at a different speed or to machine the impeller rim to a different diameter. In
fact. it has become a standard of the pump industry to publish the performance of
any pump at several different speeds (usually the standard speeds of electrical
motors) and for different impeller cut-downs (see Fig. 25-1).

VARIATION OF PUMP PERFORMANCE WITH


VARIATIONS IN SPEED

In theory. when the speed of a pump is altered from Nl to N2 . its flow rate changes
from Q= QI. following the equation

(25-1)

The head H2 at the new speed N2 and the new flow rate ~ increase from the head
HI. which the pump has developed at N=Nl and Q=QI.tO the new value according to

(25-2)

With regard to efficiency. this is generally expected to remain the same for the two
corresponding flow rates. With regard to the net positive suction head (NPSH) require-
ments of a given pump. the concept of suction specific speed implies that it should
also follow the relations expressed by Eq. 25-2.
Here a serious question arises. In Fig. 25-1 we see that both the flow rates and the
heads developed by a given pump vary within very wide limits. Consequently. how
305
306 Eliminating Pump Problems and Modifying Performance

3550 RPM 1740 RPM


25 203 Dia
~100
t!!
.sJ::"*
90
80
.s"* 20
70
-cco 60 J::
Q)
I
175 Dia -g 15
Q)
175 Dia
50 I

r
,/

r
10

3 Q)
E

'i
2~
I 4
1~
30 OZ 3 OZ

[L 20 ~2
III
I
III 10
0
20 40 60 80 0 10 20 30 40
Flow rate (m 3/hr) Flow rate (m 3/hr)

Figure 25-1 Performance of the same pump at two different speeds.

can we know which point of the performance chart, that has been drawn for one speed
corresponds to a given point of the chart that represents the pump performance at
the second speed?
A comparison between the data calculated form the tests performed at 1740 RPM
and the actual data obtained from the test at 3550 RPM leads to the following
conclusions:

1. As far as the QH curves are concerned, the data calculated from tests at one speed seem
to be reasonably accurate for the other speed. However, at the lower speed, the QH curve
seems to be able to go out to a greater QIQd ration than at the higher speed, possibly
owing to cavittion in the casing throat.
2. The efficiency curves show certain differences, but they rarely seem to exceed ±30/0.
3. The NPSH requirements at 3550 RPM, calculated from the results of tests performed at
1740 RPM, vary considerably from the NPSH requirements determined from the tests
performed at 3550 RPM. In terms of suction specific speed, the pump seems to have a
different suction specific speed at each of the two tested operating speeds.

In the past, I made a study of the NPSH requirements of various pumps at different
operating speed that have been published by 10 of the then largest and most
prestigious pump companies in the world. In each case, I found equal or even larger
discrepancies between the suction specific speed of any given pump tested at
different operating speeds [46], than the differences appearing in Fig. 25-1. Similar
results have also been reported by Hammitt [47) and Jekat [48).
However, at the time Refs. [46)-[48] were published, and even when the pump
represented in Fig. 25-1 was tested, many of the problems discussed in chapters 12
and 20 were still unknown or unrecognized. All the pumps that I have designed and
tested since then (after about 1970) have exhibited a good conformity with the
The Effects of Speed and Impeller Outer Diameter on Pump Performance 307

80 40 1175 RPM
400 Dia
400 Dia 68
~ ~
$Q) 70
"* 30
.s .s
l: 60 l:
-ci -ci 20
ro ro
:!! 50 I
Q) 335 Dia
335 Dia
10
40

5Ui'
42Q)

3.s
300 2I
Cf)
1 a..
100 With inducer z
a.. 200 BHP 0
I
co a..
100 I 50
co
0
0 200 400 600 800 1000 1200 200 400 600 800
Flow rate (m 3/hr) Flow rate (m 3/hr)

Figure 25-2 Performance of the same pump at two different speeds.

similarity laws implied by Eqs. 25-1 and 25-2, when used to calculate the NPSH
requirements of a given pump at different operating speeds.
Figure 25-2 presents the results of such a test. In Fig. 25-2, we also see a greater
similarity between the efficiencies of the same pump when measured at the two
different speeds, than those shown in Fig. 25-1.
I have not had the opportunity to verify whether the differences in the suction
specific speed published by the companies that served as a basis of Ref. [46] were
really correct or whether they were all the result of flaws in the way the tests were
carried out. (For me, personally, it is hard to believe that all NPSH tests carried out in
the past by 10 of the most prestigious pump companies were incorrect.)

VARIATIONS IN PUMP PERFORMANCE WITH


CUT-DOWN OF THE 00 OF THE IMPELLER

Outlet Edge Parallel to the Pump Axis


The peripheral velocity of the outlet edge is directly proportional to the magnitude
of its diameter. This leads to the belief that the performance of a centrifugal pump
changes with the cut-down of the outer diameter (00) in the same manner as with
the change in the operating speed. At a given operating speed, the flow rate at a given
percentage of the best effiCiency point (bep) is assumed to vary as the OD of the
impeller varies. Following the same logic, the head changes as the square of this
diameter changes. In fact, in the past, many performance charts for a given pump
with different impeller cut-downs were not based on actual tests but were calculated
from a test with the maximum diameter. These calculations were carried out by
308 Eliminating Pump Problems and Modifying Performance

replacing, in Eqs. 25-1 and 25-2, the ratio of the OD's of the full size and of the cut-
down impellers with the ratio of the operating speeds N2/Nl. This means that

(25-3)

and
(25-4)

where the subscript f refers to the impeller with the full diameter.
In practice, however, this is sometimes very far from reality. Figure 25-3 shows a
comparison between the actual performance of an impeller tested with full diameter
and with three impeller cut-downs and the performances of the cut-down impellers
calculated with the aid of Eqs. 25-3 and 25-4. Although the first calculated QH curve
seems to follow the actual test curve very closely, the calculated data for the other
two cut-downs are very far off. This apparent inconsistency results from the manner
in which the performance varies with impeller diameter. This is directly affected by
the blade geometry and the increasing mismatch with the casing volute. Part VI
presents a study that explains how the theoretical head increases with impeller
diameter.

60L-_165
_- Dia-_ _
3525 RPM

55

40

35

_Tested
- - ... Calculated from test with
165 Dia impeller

o 5 10 15 20 25 Figure 25-3 Performance with cut-down


Flow rate (m3/hr) impellers: tested versus calculated.
The Effects of Speed and Impeller Outer Diameter on Pump Performance 309

The study presented in Part VI also shows that, in most cases, the differences
between the tested heads and the heads calculated according to the method
presented in Part VI, are very similar for different geometries.
This can provide us with a convenient tool for predicting the pump performance
with a cut-down impeller whenever the actual full-diameter performance and the
geometry of the impeller are known.
The procedure is simply to calculate the theoretical heads developed by the given
impeller with full diameter and to calculate the ratios between these heads and the
actual tested heads. Next, calculate the theoretical heads for the impeller with cut-
down diameter. The ratios between the actual heads developed by the cut -down
impeller and the calculated heads for same cut -down can be expected to be similar to
the ratios at the corresponding flow rates for the impeller tested with its full diameter.
Although this procedure is suitable for use by the pump manufacturer, the data
required for such calculations are usually not available to pump users.
Consequently, pump users must rely on test data published by the pump manu-
facturer. The problem with these data is that the actual required performance often
falls between two of the diameters published by the manufacturer.
There is no absolutely correct way to calculate the required diameter for such
cases. As a general rule, when the operating point is near the next lower diameter
(such as the position of point A in Fig. 25-4), it is safe to calculate the new diameter
by means of the modified Eqs. 25-1 and 25-2, using the lower diameter (220 mm) as
a basis. On the other hand, when the operating point B is in the vicinity of the larger
diameter shown in Fig. 25-4 (230 mm), then the upper diameter should be taken into
consideration. In the latter case, however, we can expect that the actual heads will be
somewhat lower than calculated. To counteract this possibility, it is good to increase
the calculated diameter by some 0.25% to 1.00%.
If the operating point lies somewhere between two diameters, such as point C in
Fig. 25-4, then the safest procedure is to calculate the new diameter twice: once

250 Dia
30 1-:::2~50~-'--~ 1770 RPM
28
en26 ~2,-4..;;.0....;;D;..;.ia",,--__
2~22~;";;;";;'''';;'';';;''''''-~\
24 230 Dia
~20 220 Dia
~ 18 ~2~1O:;"';D~i';:;'a--
~ 16
~ 14
12 12 en
10 102
8 ~
6 ~
NPSH, Impeller only 4 V5
2 n.
With inducer o Z

Figure 25-4 Interpolation of impeller o 100 200 300 400 500 600 700
cut-downs from published test data. Flow rate (m3/hr)
310 Eliminating Pump Problems and Modifying Performance

taking the lower diameter as a basis, and then by using the upper diameter. The
actual diameter will be somewhere between these two diameters.
At present, this is the most common method of interpolating the magnitude of a
diameter cut-down from the published test data. Sometimes, however, there are
situations for which no test data are available for impeller cut-downs. In such cases,
the pump manufacturer can sometimes make use of another approach.
In the pump industry, it is very common to use a single pump design to develop
complete pumplines. This is done by scaling this single design up or down. In such

Figure 25-5 Different methods for cutting down


the outer diameter of a mixed-flow impeller.

,
26 "" 1465 RPM
"
" ';:>6'....
" ".?';:>o
"'+
"
"
265 x 338 Dia
24 '
',0;:0 ,
'
22
,,
~
, ..... .....
5 blades\-
.....

20 , ........ -
18 "
D
338Di~' I
1265 Di~--"'--'-
~--===::J'
I

-g 12
Q)
:r:
10

2
O~_L-_L-_L-_L-_L-_L-_L-_L-_ Figure 25-6 The effect of impeller cut-
o 100 200 300 400 500 600 700 800 down in two identical designs, but with a
Flow rate (m3/hr) different number of blades.
The Effects of Speed and Impeller Outer Diameter on Pump Performance 311

cases, the pump manufacturer can predict the performance of a given pump with a
cut-down impeller by studying the effects of an analogous cut-down on a pump of a
different size that was developed using the same prototype as the pump in question.

Outlet Edge Not Parallel to the Axis


In pumps of high specific speeds, the impeller is often designed with its outlet edge
located on a conical surface (see Fig. 25-5), instead of being parallel to the axis. In

25

1790 RPM
20

.s*
~

15
:r:
-0
m10
I

Figure 25-7 The effect of reducing


the diameter of a mixed-flow impeller OL-____~----~~----~----~~--~~----~
o 100 200 300 400 500 600
parallel to the outlet edge. Flow rate (m3/hr)

1790 RPM
- - 0; = 0.87 Oa
DC ---0;=0.790a
a = ons!. ......... 0; = 0.74 Oa
35 - - - 0; = 0.68 Oa

30

90
25
80
~
Ql 70
Qi 20
.sJ: 60
2
~
-015 50;
m
I 40 g-
Ql
10
30~
w
20
5
10

Figure 25-8 The effect of altering 100 200 300 400 500
the inclination of the outlet edge. Flow rate (m 3/hr)
312 Eliminating Pump Problems and Modifying Performance

such cases, the industry uses three different approaches to cutting down the outer
diameter of an impeller. One, as shown in Fig. 25-5 at A, is parallel to the existing
outlet edge. The second, as shown in Fig. 25-5 at B, is by changing the inclination of
the outlet edge in such a manner that the outermost diameter remains unchanged.
The third method is to leave the innermost diameter of the outlet edge constant and
to reduce the outermost diameter (Fig. 25-6).
Whenever the impeller cut-downs are made parallel to the outer rim, the resultant
QH curves follow in a nearly parallel direction to the full-diameter impeller (Fig. 25-7).
When the outermost diameter of the outlet edge is kept constant, as in Fig. 25-5 at
B, the QH curve gets steeper with each cut-down, as shown in Fig. 25-8 (because the
shut-off head is primarily determined by the peripheral speed of the outermost
diameter).
When the innermost diameter is kept constant (Fig. 25-6) the QH curves tend to get
flatter with each consecutive cut-down (owing to the reduction of the outermost
diameter). It should be pOinted out, however, that the test curves shown in Fig. 25-6
are the only ones I have ever come across. Consequently, this matter requires
additional experimental verification.
CHAPTER 26
The Effects of Reducing
the Impeller Width
One of the means used in industry for altering the performance of a pump, is to
reduce the width of the impeller. In the case of a semi-open impeller, the procedure is
simple. Machine the open face of the blades until the impeller width is reduced to the
desired magnitude.
With a closed impeller, the procedure is somewhat more complicated. A common
practice is to prepare a disk that could be used as a replacement for one of the
impeller shrouds. The given shroud is then machined off and the blade width
reduced to the desired dimension. Mter this, the spare shroud is reattached (with
small screws, soldering, or using any other suitable method).
When the outlet diameter of the closed impeller is not parallel to the axis, the
preferred practice is to replace the shroud with the smaller diameter (back
shroud/hub) (see Fig. 26-1). However, when such a shroud is destined to transfer
power from the shaft to the blades, this may cause mechanical problems. In such
cases, it is always preferable to replace the suction side shroud (front). The latter
choice is also generally accepted in impellers with a cylindrical outlet surface. The
direct effect of reducing the impeller width is to make the QH curve steeper, i.e. to fall
off at a faster rate.
In pumps with a parallel outlet diameter, the shut-off heads remain almost
unchanged. However, at low partial flows the behavior of the pump may vary
Significantly with the changes in width, particularly in pumps of different specific
speeds. Figures 26-1 and 26-2 illustrate such cases.
In the pump with high specific speed shown in Fig. 26-1, the heads developed at
partial flow rates decrease continuously with each reduction of impeller width.
Against this, in the pump with low specific speed shown in Fig. 26-2, the heads
developed at the partial flow rates seem to remain completely unaffected by the
impeller width.
One of the causes of the difference in the effect of changing the impeller width in
pumps of different specific speed is as follows. It has been demonstrated in Chapter
1, that the magnitude of the eu component of the liqUid that is being directly acted on
by the blades can be expressed by the equation

(26-1)

313
314 Eliminating Pump Problems and Modifying Performance

3500 RPM

100

90

~ 80
~ 70
:t
-0 60
:ll
J: 50

40

30

20

10

40 80 120 160 200 240 GPM Figure 26-1 The effect on performance
Q of varying the impeller width.

50 1490 RPM

w40 ----~~~~~~~~~~~EFF
80
~ ""0,/......
.5.30
//"0 ..........'0..... QH
60~
:t
Intense cavitation noises CI". °EFF
~
-0
'tlQH
8.
gJ 20 40~
J:
- Impeller width 65 mm c:
Q)

10 0- - ° Impeller width 50 mm 20~


w
OL-__L -_ _L -_ _ ~_ _~_ _L -_ _L -_ _L -_ _L -_ _~
Figure 26-2 Two impellers of
o 20 40 60 80 100 120 140 160 180 different blade width tested in the
Flow rate (m 3/hr) same casing (low specific speed).

On the one hand, when the magnitude of the em component is significant


compared with the magnitude of the blade velocity U, any change in the magnitude of
em significantly affects the magnitude of the developed head. On the other hand,
when the magnitude of em is relatively small, any change in its value has almost no
effect on the developed head.
On the one hand, in the pump represented in Fig. 26-1, the meridional outlet
velocity of the liquid at the design flow rate was approximately equal to 3.5 m/sec,
whereas the peripheral velocity of the outermost diameter of the outlet edge was
23.5 m/sec.
On the other hand, in the pump shown in Fig. 26-2, the meridional outlet velocity
at the design flow rate was approximately equal to 0.48 m/sec, whereas the
peripheral velocity of the outlet edge was equal to 26.5 m/sec. This explains why
changes in the meridional velocity component of the pump with high specific speed
Effects of Reducing the Impeller Width 315

shown in Fig. 26-1 have produced significantly greater changes in the performance of
the pump than in the case of the pump with low specific speed shown in Fig. 26-2.
A change in the impeller width can affect the developed heads. the location of the
best efficiency pOint. and the suction capabilities of an impeller. The location of the
best efficiency point will be discussed later in conjunction with the effects of the
casing on performance and in conjunction with the interaction between the impeller
and the casing. Here we shall restrict our study to the effects of impeller width on the
suction capabilities of a pump.
A glance at Fig. 26-2 shows that at flow rates exceeding about 105 m 3 /hr. the
heads of the pump with the narrower impeller begin to falloff very rapidly. At
approximately that flow rate. cavitation noises seemed to have begun to develop.
These became quite intense at moderately higher flows. The reason for this rapid fall
off of the heads at higher flow rates seems to be due to the fact that a narrower
through-flow area at the impeller inlet reduced the suction capability of the pump.
This fact has been also corroborated on other occasions. as reported in Ref. [49].
CHAPTER 27
Modifying the Casing
Geometry
By the term casing we understand two, seemingly different configurations: the volute
casing (Fig, 27-1) and the diffusor prOVided with gUide vanes (Fig. 27-2). These two
configurations, while seemingly quite different, are based on the same design
prinCiples and serve the same purpose. Their task is to guide liquid from the impeller
into the discharge nozzle of the pump. At the same time, they are usually designed so
as to convert a part of the kinetic energy of the liquid that is exiting the impeller into
pressure energy.
The principles of casing design were best formulated by C. Pfleiderer [31], in the
early 1930s. I have had an opportunity to compare Pfleiderer's recommendations with
the actual performance of well over 500 different designs. In most cases, the
agreement between theory and practice has been excellent. However, even in cases in
which significant discrepancies between theory and practice seemed to exist, I have
always been able to determine that the discrepancies were due to some extraneous
factor that was completely unrelated to the theory developed by Pfleiderer.

Volute
throat

Volute
tongue

Volute curve
Figure 27-1 Nomenclature of certain volute
parameters.
317
318 Eliminating Pump Problems and Modifying Performance

/Diffusor
Impeller

Figure 27-2 Parts of


a multistage diffuser
pump.

THE EFFECT OF THE VOLUTE THROAT AREA


ON PUMP PERFORMANCE

Experience (as well as theory) teaches us that for a given peripheral speed U2 of the
impeller rim, and a given QH curve generated by the impeller, there is a certain flow
rate at which the losses in a casing of a given throat area are the smallest (31). This
combination of head and flow rate for a given throat area of the casing is often
defined as the design point of the casing. In the pump represented in Fig. 27-3, the
design point of the casing is shown on the performance chart as point A.
Theory teaches that if we install in a casing an impeller that is developing a
different head, the best efficiency point of the casing will move to a different
head-capacity combination. This combination moves along a straight line that passes
through point A and the point at which Q and H are both equal to zero [50, 51).
Figure 27-3 presents actual test results of a pump tested in the same casing with
four different impeller cut-downs. The inscribed numbers denote the efficiencies
determined during these tests, and the points marked X show the maximum
efficiency pOints at the given diameter. We see here an almost ideal conformity
between theory and practice.
However, in many other cases, no such conformity can be found. Look, for
example, at the test results presented in Figs. 25-6 and 26-1. In Fig. 25-6, we see
that the maximum effiCiency of the pump always lies at nearly the same flow rate,
regardless of the head developed by the impeller. In Fig. 26-1, the best efficiency
points lie not on the casing line OV, but along curve AB, which is completely different
from the casing line. A more detailed study of the test results presented in Fig 26-1
reveals that all tested best effiCiency points lie at a flow rate, at which the meridional
velocity of the liqUid entering the impeller blades is constant, and equal to 3.9 m/sec.
The reason for that difference in behavior between the pump represented in Fig.
27 -3 and the pumps shown in Figs. 25-6 and 26-1 lies generally in the difference of
their specific speeds for which they have been designed. This difference mandates
inclusion of different features in their design. The throat area of the casing is not the
only parameter that determines the location of the best efficiency point of a given
pump. Under certain circumstances, the geometry of the impeller may have an equal,
Modifying the Casing Geometry 319

150 252 Dia

3560 RPM
Vi
Qi 125
Qi
S
1:
en
"0

I
III 100
............_______

75

x = Best efficiency point

Figure 27-3 The effect of volute design on 25 50 75 100


performance of a pump with low specific speed. Flow rate (m3/hr)

or even a significantly larger, effect on pump performance than the throat area of the
casing.
In the majority of cases encountered in centrifugal pumps, the losses of energy
encountered by a flowing liquid increase as the square of the velocity it possess
relatively to the wetted surfaces of its passages increases. In a conventional impeller,
this relative velocity is highest at the inlet to the blades. Consequently, the
magnitude of the effect of impeller geometry on the location of the best efficiency
point depends very much on the design of the impeller inlet.
Here is where the specific speed for which a given impeller has been designed
comes into play. The pump represented in Fig. 27-3 is of a rather low specific speed.
In such a pump, the velocity head from the relative velocity of the liquid at the blade
inlets constitutes only a very small percentage of the total head developed by the
pump. A loss of several percentage points of that velocity head will hardly have any
effect on the overall efficiency of the pump. Against this, the velocity head of the
liquid flowing through the volute throat may easily constitute 25% to 50% of the total
head. In such a pump, any losses of that velocity head will immediately appear as a
significant loss in the overall efficiency of the pump. Consequently, in such pumps,
the throat area of the casing is often the most decisive factor in determining the
location of the best efficiency point. This is vividly demonstrated by the test results
320 Eliminating Pump Problems and Modifying Performance

35
.... -- _ - - - - - - - __ -QH
.... 1760 RPM

30

- - Throat area = 14.5 cm 2


Throat area = 10 cm2
EFF 80 p-
:r::
...... c:
-g 15 ...... ......
(I)

70 ~
(I)
a.
I ......

,. , "
60 ;::
o
c:
10 - 50.!!1
o
/
I 40 iii
5

OL_____ L -_ _ _ _~_ _ _ _~_ _ _ __ L_ _ _ __ L_ _ _ _~_ _ _ _~_ _ _ _ _ _L_~

o 10 20 30 40 50 60 70 80
Flow rate (m3/hr)

Figure 27-4 The effect of the throat area on the performance of a pump with low specific
speed.

shown in Fig. 27-4. An increase in the throat area from 10 cm2 to 14.5 cm2 has
moved the location of the best efficiency point from 50 m 3 /hr to 70 m 3 /hr (Le.,
approximately 50x(14.5jlO) m 3 /hr).
In pumps of high specific speed, the velocity head from the relative velocity of the
liquid entering the impeller blades may easily be of the same order of magnitude or
even exceed the velocity head of the liqUid entering the casing. This, combined with
other sensitive features of the impeller inlet, can easily overweigh the effects of the
casing. This is what has really happened in the pumps represented in figs. 25-6 and
26-1.

THE EFFECT OF VOLUTE TONGUE GEOMETRY


ON PERFORMANCE

Contrary to widespread opinion, the geometry of the volute tongue (Fig. 27-1) can
have a huge effect on the performance of a centrifugal pump. On one occasion, I
decided to try out a special tongue geometry. Unsure of what its effect on the pump's
performance would be, I designed the discharge nozzle with a constant throat area
for a length of about 1.5 in. downstream of the tongue tip. The idea was to be able to
remove up to 1.5 in. of the length of the volute tongue without affecting the area of
the volute throat.
Figure 27-5 shows how the tongue geometry affected the performance of that
particular pump. I have come across case histories in which minor changes in tongue
geometry have caused differences in pump efficiencies ranging from 10 to 16
percentage points.
Modifying the Casing Geometry 321

35
1770 RPM

30

- 80

-- -.....
--+<-~~
25 ............. --~
"-
~ -~-
Q)
Q5
-S 20
:r:
-0
co
Q)
I 15

10 - 30
c-::J Volute with original tongue
>E- - -x Major part of the tongue removed

o~ ______ ~ ______ ~ ________ ~ ______ ~ ________ ~ ______ ~~

o 50 100 150 200 250 300


Flow rate (m3/hr)

Figure 27-5 The effect of removing part of a specially designed volute tongue (without altering the
throat area).

At present, I know of no theory that would adequately explain the reason for such a
huge effect of the tongue geometry on performance. In this respect, it is up to the
pump specialist, using his experience, to choose the correct shape. There exists,
however, one consolatory fact: In most cases, whenever a problem is due to a faulty
tongue geometry, it can be easily eliminated by simply cutting away and removing a
part of the volute tongue.

THE EFFECT OF THE VOLUTE CURVE'S SHAPE

I have had the opportunity to study the effects of a wide range of volute geometries on
the performance of centrifugal pumps. To my greatest surprise, significant variations
in the shape of the volute curve (Fig. 27-1) have produced rather small differences in
the output and efficiencies of a centrifugal pump. This is rather a lucky coincidence,
because it furnishes the pump speCialist with a very simple and effective tool for
dealing with problems related to radial loads.
Whenever a volute pump is subjected to excessive radial loads (e.g. owing to the
high specific gravity of the pumped liquid), it is possible to reduce this excessive force
322 Eliminating Pump Problems and Modifying Performance

Rounded
off edge
Metal
removed
--N--- Concentric
arc

Figure 27-6 A volute remachined for reducing


radial thrust.

by modifYing the shape of the volute curve. This can be achieved simply by
remachining the part of the volute adjacent to the volute tongue to make part of the
casing volute concentric with the impeller (see Fig. 27-6). Depending on the thickness
of the casing walls. this procedure usually allows us to make about 5% to 20% of the
volute periphery concentric with the impeller axis. Such a change usually produces a
significant reduction in the radial loads. While such a change in the volute usually
reduces the efficiency. this reduction is rarely more than 0.25% to 1.5%.
CHAPTER 28
Various Methods of
Altering Pump
Performance
UNDERFILING THE OUTLET TIPS OF THE IMPELLER BLADES

The Effects on Head


One common means of increasing the heads developed by a centrifugal pump is to
underfile the outlet tips of the impeller blades, as shown in Figs. 28-1 and 28-2.
However, as can be concluded from the performance charts presented in these
figures, the effects of underfiling vary in pumps of different geometries. Even in the
same pump, the percentage of increase in head achieved by underfiling varies with
flow rate.
One reason for these differences can be traced to Eq. 26-1: the magnitude of C u ,
and with it the magnitude of the developed head, increase when the magnitude of the
product Cmcot [3 is reduced. Underfiling the outlet tips means increasing the blade
angles in the vicinity of the impeller outlet. This reduces the magnitude of the term
cot [3, resulting in an increase in the developed head. When the magnitude of Cm is
very small in comparison to the magnitude of the blade velocity U, the total effect of
the increase of the term (Cmxcot [3) on the developed head is relatively small. The
magnitude of this term, however, increases as the magnitude of Cm increases (I.e.,
with an increase in flow rate). This has been confirmed by the test results shown in
Figs. 28-1 and 28-2.
A comparison of the percentages of head gains in both pumps shows that the gains
of head in the pump with high specific speed shown in Fig. 28-2 are much higher
than the head gains in the pump with low specific speed shown in Fig. 28-1. The
reasons for that difference are due to the unavoidable differences in the design
features of both impellers. These differences are mandated by the differences in their
specific speeds.
The specific speed of the pump represented in Fig. 28-1 is about 15.5 in metric
units (800 in U.S. units). In such an impeller, the blades are relatively very long, and
underfiling increases the angles of only a very small part of the blade. Also, in such
pumps, the meridional velocity components of the liqUid are relatively small
compared with the blade velocities U in the vicinity of the outlet tips. As a result of
323
324 Eliminating Pump Problems and Modifying Performance

QH 1465 RPM
40
00-
CD 35
Q)
E
~ 30
l:
"0 - - Original blade tip
m25 - - - Underfiled tips
I
20
c
Q)
EFF
70 ~
60.e,
>.
50 g
40·~
m
Figure 28-1 Increases
o 20 40 60 80 100 120 in head achieved by
Flow rate (m 3/hr) underfiling the blade tips
of an impeller with low
specific speed.

16 - - Original outlet tips

r
15 - - - Underfiled tips

14 "
,, 2950 RPM

13 ,,
12 ',14%,t ,
11 ,
~
Q)
10 " 14%
'~,
Q)
.sl: 9

8 \ 80
-0 \
t1l
Q) 7 70
I
6 60~
Q)

5 50 Q)
~
Q.

4 40 ;::
u
c
3 30.91
u
2 20iTI
10
Figure 28-2 Increases in head achieved by
underfiling the outlet tips of an impeller with
high specific speed.
Various Methods of Altering Pump Performance 325

both factors, the head gains caused by underfiling such a pump are relatively small.
The specific speed of the pump represented in Fig. 2S-2 is approximately SO in
metric units (4570 in U.S. units). In such an impeller, the blades are relatively short,
and therefore underfiling increases the angles of a significantly greater part of the
blades than in a pump with low specific speed. Also, the meridional velocity
components of the liqUid are significantly higher than in an impeller with low specific
speed. This explains the differences in the effect of underfiling the outer tips of the
blades of impellers of different specific speeds.

The Effects of Underfiling on Efficiency


The effects of underfiling on effiCiency exhibit a somewhat erratic pattern. In the
test data shown in Fig. 2S-1, there is a reduction in the efficiencies at lower flow
rates and a slight increase at the design flow and at higher flow rates. In the pump
represented in Fig. 29-2, no changes in efficiencies at lower flow rates are seen, but
there are very Significant reductions in the efficiencies over the practical operating
range of the pump. It seems that the main reason for this drop in efficiencies is due
to the fact that the outlet edge in such an impeller is not parallel to the axis. In such
an impeller, underfiling may easily disturb the state of equilibrium between the
neighboring stream lines of the liquid that is exiting the impeller. This seems to be
the main cause of the significant drop in efficiencies observed in Fig. 2S-2.

OVERFILING THE OUTLET TIPS OF IMPELLER BLADES

The discussion relating to the effect of underfiling the outlet tips of the blades leads
to the conclusion that overfiling (Fig. 2S-3(C)) reduces the head developed by a
centrifugal pump. Surprisingly, I have never observed such an effect. On the
contrary, in every case in which I have overfiled the blade tips, I have always gotten
an increase of 0% to 2% in both head and efficiency.
With respect to the increase in efficiency, the explanation is rather simple:
experience teaches us that the liqUid leaves the impeller at an angle that is smaller
than the blade angle at the outlet tip. The liqUid exiting the leading face of the blade
must move past the sharp edge A in Fig. 2S-3). This, as has been demonstrated in
Ref. [45], inevitably leads to separation of the liqUid from the blade. In turn, such a
separation, results in loss of energy. By overfiling the blade tips in the manner shown
in Fig. 2S-3(C), this cause of hydraulic losses is eliminated, producing effiCiency
increases ranging from 0% to 2%.
The reason overfiling does not result in a reduction of the developed head is less
obvious and requires further study. However, certain indications of the cause of that
effect can be found in the results of tests performed by Copley and Worster [521.
Figure 2S-3 presents some of the results of these tests.
Figure 2S-3(A) presents the pressure distribution in the liquid in the vicinity of the
outlet tip of a blade that terminates with a blunt edge. We see here, that the pressure
differences between both faces of the blade are reversed.
For a blade to be able to transmit power from the shaft to the liquid, the pressure
on the leading face of the blade must be higher than on its trailing face. This has also
been confirmed experimentally, as shown in Fig. 6-2. A reversal of that relationship
326 Eliminating Pump Problems and Modifying Performance

c
OJ
c
OJ
'(3 '(3
:E :E
OJ OJ
o o Blunt tip
u u
~ Trailing face ~
::J ::J
rn rn
rn rn
OJ
~
0: a..

[ [

I I

~
-A/
(A) Blunt blade tip (8) Underfiled blade tip

(C) Overfiled blade tip

Figure 28-3 The effect of underfiling the outlet tips on the pressure distribution in
the vicinty of the blade outlets: (A) blunt blade tip; (B) underfiled blade tip; (C)
overfiled blade tip.

means that the liquid is returning some of its energy to the blades. Consequently, the
reversal of the relationship between the pressures on both sides of the blade near its
outlet tip indicates that the liquid is losing some of its acquired head. In practice, this
reversal of the distribution of pressure near the outlet tip constitutes one of the
reasons the actual head developed by an impeller is always lower than the calculated
one.
Figure 28-3(B) demonstrates that this reversal of pressures can be eliminated by
providing the blades with sharp edges. Overfiling of the outlet tips also provides the
blades with sharp edges, which leads to the conclusion that overfiling the outlet tips
can eliminate the reversal of pressure that occurs in impellers with blunt edges.
Consequently, by over filing the outlet tips of the blades, we are simultaneously
creating two counteracting effects. By reducing the angles of the outermost blade
Various Methods of Altering Pump Performance 327

elements, we are reducing the power transmitted from the impeller to the liquid, and
by sharpening the outlet tips, we are reducing the amount of power returned by the
liquid to the impeller blades. Depending on the relative magnitude of these two
effects, we may get a reduction, an increase, or no change in the head developed by
the given pump. In every case I have encountered, overfiling has always produced an
increase in head ranging from 0% to 2.0%. Of course, to arrive at a final conclusion,
more experimental corroboration is needed.

CHANGING THE CLEARANCE OF A SEMI-OPEN IMPELLER

Increasing the Axial Clearance between


the Blades and their Casing
Figure 7-13 demonstrates how an increase in the axial gap between the blades and
the casing affects the performance of a pump. Here we see a significant reduction in
the developed heads and in the efficiencies at any given flow rate. However,
experience teaches us that reductions in head caused by an increase in the axial
clearance are usually significantly greater than reductions in efficiency. The power
consumption of a given pump decreases with the increase in the axial clearance.
A case in which this effect has been put to practical use concerns a well that was
drilled in a remote, developing area and found to be capable of delivering enough
water to meet all future needs. Because of the remoteness of the location, it was
decided that the most economical approach would be to install a pump capable of
using the total output of the well. However, when the pump was scheduled for
installation, the area had a very limited supply of electric power. While the utility
companies were working on increasing that supply, it was important to make
immediate use of the well to be able to begin the development work.
The solution was to install a pump with semi-open impellers and a smaller electric
motor, but to increase the gap B (Fig. 19-4) between the blades and the casing to
avoid overloading the motor. Experiments carried out at the pump manufacturer's
plant demonstrated that the solution was to lift the impellers about 9 mm above their
normal position.
Mter the new power line was completed, the impellers were lowered by the extra
9 mm originally set, and the motor were replaced by a more powerful one. This made
it possible to use the total output of the well without the need for costly replacement
of equipment.

IncreaSing the Radial Clearance between


the Inducer and its Casing
The discussion presented in Chapter 7 leads to the conclusion that an increase in
the radial clearance a (Fig. 28-4) will impair the performance of an inducer. However,
numerous tests with constant pitch inducers of high solidity and outermost blade
angles ranging from 18 degrees to 23 degrees indicate that up to a clearance of
a=0.005 R2 (Fig. 28-4), the magnitude of the clearance a has no effect on the head,
effiCiency, or suction capability of an inducer. Only when the width of the gap a is
increased beyond 0.005 R2, will the performance begin to deteriorate with any
additional increase in clearance.
328 Eliminating Pump Problems and Modifying Performance

Figure 28-4 Certain design para-


meters of an inducer.

UNDERFILING THE INLET TIPS OF AN INDUCER

Figure 28-4 presents the development of the blades of a constant pitch inducer on a
flat plane. For such a blade to be effective, the blade angle f3 of the inducer must be
larger than the angle at which the liquid approaches the blade by angle A (incidence).
This means that the liquid that is approaching the trailing surface of the inducer
blades must flow past the sharp edge A. This, as has been demonstrated in Ref. [45],
will lead to separation of the flow at that point. To eliminate any possible negative
effects of that separation, I have tried on several occasions to round off that edge, as
shown in Fig. 28-4 at B. The result was always a reduction of about 6% to 10% in the
NPSH requirements of the inducer.

THE EFFECT OF ATTACHING BACK-VANES TO


THE BACK SHROUD

Occasionally, pumps with closed or semi-open impellers are provided with vanes on
their back shrouds (Fig. 28-5). Back-vanes are usually completely out of the way of
the pumped liqUid, and they serve to keep solids contained in the liqUid from
entering the stuffing box or mechanical seal. A secondary purpose of such vanes is to
reduce the pressure of the pumped liquid on the shaft seal. Finally, in semi-open
impellers, back-vanes are capable of reducing the hydraulic unbalance (axial loads),
which occurs from lack of a front shroud (see Chapter 10). Although back-vanes are
an inexpensive means of achieving these purposes, they may adversely affect other
aspects of the pump's performance. First, if not properly designed, they may reduce
the pressure at their inner radii enough to cause cavitation or to allow air to leak into
the pump via the sealing device of the shaft. Second, they usually reduce the overall
effiCiency of the pump owing to their power consumption. Against this, back-vanes
usually increase the total head developed by a pump.
The magnitude of the effects caused by fitting back-vanes depends on, among other
factors, the following parameters:

1. The head coefficient of the liqUid issuing from the impeller (compare Chapter 7)
2. Whether the shape of the back-vanes is straight and radial or the same as the shape of
the impeller vanes.
Various Methods of Altering Pump Performance 329

Figure 28-5 Some of the parameters


that affect the performance of back-vanes.

3550 RPM

90
~
OJ
- - No back vanes
24 ID 80
E.
22 :r:
-070
<Il
20 OJ 70
I
18 60 60
~ 16 50 ~
OJ 5 mm ~
14 50
4.5mm 40 3>-
s= 4.5 mm
12
8 2 = 11 mm
30 g
OJ
10 R2 = 203 mm 20~
Ra= 203 mm
8 10

Figure 28-6 The effect of back-vanes o 10 20 30 40 50


on performance. Flow rate (m 3/hr)

3. The ratio of the height h of the back-vanes to the width B2 of the impeller blades (Fig. 28-5).
4. The ratio of the height h of the back-vanes to their distance d from the back cover of the
casing
5. The ratio between the outer radius Ra of the back-vanes to the outer radius R2 of the
impeller
330 Eliminating Pump Problems and Modifying Performance

Extensive studies of the effect of various design parameters on the pump's


performance and on the axial thrust were carried out in 1962 by Zanker [25, 261.
Unfortunately, these studies did not take into account the effects of the head
coefficient of the pumped liquid (compare the discussion in Chapter 7). Nevertheless,
the matter presented in these publications offers very important qualitative details
about the action of repelling vanes.
As has been pOinted out previously, the back-vanes is sometimes shaped straight
and radial; in other cases the back-vanes are chosen to be of exactly the same shape
as the main impeller blades. To determine the difference in the effects of these two
shapes, I tested an impeller of 8-in. outer diameter and six-vanes, first without back-
vanes and then with six radial back-vanes. Finally, I removed the radial back-vanes
and replaced them with back-vanes of exactly the same shape as the main impeller
blades. The results are shown in Fig. 28-6.
I observed the following three results. In both cases, adding the back-vanes
increased the heads developed by the pump but reduced the efficiencies. The pump
with radial back-vanes developed higher heads at given flow rates than the curved
back-vanes. Both shapes of the back-vanes produced equal reductions in efficiencies.
PART
A Glimpse of
the Future

Everything in this book is based on prior experience and knowledge


acquired in the past. The checklists, discussions, and hints presented in
this book cannot provide the pump user with all the know-how needed to
choose, use, and troubleshoot for centrifugal pumps. To be able to handle
unusual and hard-to-solve cases, the user also needs in-depth knowledge
and understanding of the principles of operation of these pumps and the
various factors that affect their operation.
However, in engineering, as in all other aspects of life, the best solution of
today often is second best tomorrow. New findings are continuously being
added to our body of knowledge, and old finding revised. These findings
often shed new light on old problems, and may be able to help solve new
problems that will undoubtedly occur in the future.
The study presented in this section can be described as one that has the
potential of delivering this kind of information for future use. A few
examples that demonstrate some of the practical applications of the study
are also presented.
CHAPTER 29
Ne\N Findings
Concerning the Mode
of Operation of
Rotodynamic Impellers
The contents of this chapter are primarily of interest to the pump designer. However,
as will be demonstrated later, the knowledge of these findings may help the pump
user to solve problems that, without that information, were beyond the user's reach.
As with all new findings, it may take some time before the full potential of the
contents of this chapter are recognized and explored. However, the few problems that
will be discussed later and the discussions presented in Refs. 53 and 54 indicate that
the presented study has the future potential of greatly assisting to the pump user.
It has been demonstrated theoretically [55-57) and experimentally [57,58), that the
outlet-angle of a blade cannot serve as a significant design parameter for calculating
the head developed by a rotodynamic impeller. It has also been demonstrated that the
classical approach presently used for calculating the developed head with the aid of
the aerofoil theory leads to contradictory conclusions [59,60). Here we present an
approach that is free of these faults.
The theory presented here relates strictly to the theoretical head developed by the
impeller, without regard to factors such as leakage, head losses in the casing, and so
forth. As such, it can be regarded as a replacement of the traditional slip theories
that contradict the results of many performance tests, as will be demonstrated later.
The head developed by a rotodynamic impeller results from the combined action of
different factors. The most pronounced is the effect of the blade geometry, but the
developed head is also significantly affected by a wide spectrum of other factors. The
most important are hydraulic losses, leakage, disc friction, and recirculation.
Hydraulic losses tend to reduce the magnitude of the developed head, but as has
been demonstrated in Ref. 7, the relationship between the hydraulic losses and the
magnitude of the developed head is much more complex than was originally
presumed [7,61,62). Leakage moves the head capacity curve to the left, which is
expressed by lower head readings at a given, measured flow rate.
Disc friction, although consuming power, usually increases the magnitude of the
head developed by a rotodynamic pump [5-7,11,12,63). Recirculation is also capable
of significantly increasing the magnitude of the tested head [8,9,19).
333
334 A Glimpse of the Future

In addition, many other factors have proven to be capable of affecting the


performance of a pump [25]. Unfortunately, the nature and effects of these factors
have not yet been adequately explored.
Finally, the manner in which a given pump has been tested, can also lead to
differences in the magnitudes of the measured heads [64,65].
The discussion presented here is limited to the effects of the blade geometry.
Consequently, the calculated heads are expected to differ from actual test the results.
However, comparison of the calculated heads with the actual test results indicates
that the approach presented here produces very realistic results.
To simplity our discussions, we shall carry out our studies on a blade of an axial-
flow impeller. Later, we shall modity our findings to make them suitable for use with
a centrifugal impeller.

THE PHYSICAL MEANING OF EULER'S EQUATION

According to Euler, the energy transmitted by a cylindrical section of a blade of an


axial flow impeller is given by

E=(pV)X(Cu2-C u d . u. (29-1)

Writing U equals distance divided by time and regrouping, we get

E = (p V) x [C U2. - CUI] . distance. (29-2)


tIme

The nomenclature used herein is defined at the end of this chapter.


The first expression on the right side is mass, the second defines acceleration.
Consequently, the product of these two terms defines a force. The direction of Cu is
identical to the direction of U, therefore, Eq. 29-2 tells us that the energy added to
the liqUid by a blade expresses the dot product of force multiplied by distance.
The presented interpretation of Euler's equation leads to the conclusion that for a
given blade velocity U and for CUI =0, the magnitude of C u at a given distance Ad from
the inlet edge (Fig. 29-1) expresses the amount of energy added to the liquid by this
part of the blade, which extends between the impeller inlet and the axial distance ~.
This feature of Cu provides us with a very useful analytical tool. Consequently, in all
studies that follow, we assume that CUI =0.

THE INTEGRAL THEOREM OF THE BOUNDARY

This theorem can be formulated as follows. There are cases in which the magnitude oj
certain dynamic Jeature oj the boundary oj a volume is equal to the sum oj the identical
dynamiC Jeatures oj all partial volumes that are contained within that volume.
At least three cases are known to comply with this theorem. The first is the well-
known theorem of Green (or Stokes). It states that the circulation on the boundary of
a volume of liqUid is equal to the sum of all circulations of the liqUid elements
contained within that volume. The second can be defined as follows. The momentum
New Findings Concerning the Mode of Operation of Rotodynamic Impellers 335

..
A ..

Figure 29-1 Symbols used in equations.

of a volume of matter contained within a given boundary, is equal to the sum of the
moments of all mass elements contained within that boundary, or

U xM = fff u * dm. (29-3)

In Eq. 29-3, U and dm signifY the velocities and masses of the individual volume
elements, respectively, and U and M signifY the velocity and mass of the total volume
of matter contained within that boundary, respectively.
Equation 29-3 can be substantiated as follows. Assume that two bodies A and B of
equal mass M are moving with the same velocity U. Let A be a solid volume of a
homogeneous matter, and let B be a thin shell filled with a compressed gas. All
elements of the solid body A are moving, of course, with the same velocity U= U.
Consequently, the total momentum of A is obviously equal to Mx U. In the case of
body B, the variations in the velocities of the individual gas particles are infinite, but
relative to a stationary observer, the momentum of body B is the same as that of body
A (Le., equal to Mx U).
The third special case relates to the effect of energy transfer from a blade to a given
volume of liquid. When a blade element acts on a given volume of liquid, the energy
contents of that volume increases by an amount equal to the amount added to the
liquid by that blade element, regardless of the magnitudes and distribution of the
velocities of the discrete particles of liquid of which that volume is composed.
Consequently, in conformance to our conclusions drawn from Eq. 29-2, we are
entitled to express this gain in energy by means of an equivalent value of C u , equal to
CE (for CUI =0 and a given blade velocity U).

TRANSFERRING POWER FROM A BLADE TO THE LIQUID

Assume that a blade is completely immersed in a liquid and that both the liquid and
the blade are at rest. In that case, there will not and cannot be any dynamic transfer
of energy between the blade and the liquid.
Now assume that an observer is moving with a uniform velocity - U. Relative to
that observer, both the liquid and the blade are moving with a velocity U. It is
obvious, that this motion of the blade and of the liquid in relation to the observer will
not and cannot produce any dynamic transfer of power between the blade and the
liquid.
336 A Glimpse of the Future

A dynamic transfer of power can take place only when the motion of the blade
causes a change in the motion of the liquid. This can occur only when the liquid has
a velocity component Cu that is different from the velocity U of the blade (Fig. 29-1).
In that case the moving blade displaces a certain volume Qv of the liquid per unit of
time. The displaced liquid interacts with the particles of the surrounding liquid and
forces them to enter and fill the space left behind the moving blade. Only when Qv is
not equal to zero can an exchange of power between the blade and the liquid occur.
(Note. Relative to the flowing liquid, the blade moves in the direction of U with a
relative velocity equal to U- Cu .)
Referring to Fig. 29-1 we see at a glance that the volume Qv of the liquid displaced
by the blade (in a unit of time) is equal to

(29-4)

where B is the blade width perpendicular to A and U.


Equation 29-4 is based on the silent assumption that the velocity component Cu is
uniform and constant throughout the liquid. In reality, however, the value of e u
varies not only with the distance A from the inlet edge, but also with the distance X
from the blade surface (Fig. 29-2). This raises the question of what magnitude should
be chosen for Cu, for calculating Qv by means of Eq. 29-4. Here, the integral theorem
of the boundary has proven to lead to very encouraging results.
Referring to Fig. 29-2, let us consider a very narrow blade strip Ma. Because
earlier we assumed that CUI =0, we can use the inlet tips as a datum from which the
magnitude of Cu can serve as a measure of the energy added by the blade to the
pumped liquid.
When the liquid passes through the distance LlAa, the blade element contained
within Ma interacts with a very small volume LlQva of the liquid. This interaction
imparts to the volume LlQva a certain velocity component C u , in the direction of U. As
the volume LlQva has been acted on directly by the blade element contained within the
axial distance Ma the displaced liquid had no choice but to follow the direction of the
blade surface. Consequently, the magnitude of the Cu component imparted to the
flow rate LlQva can be expressed by the equation

(29-5)

However, this value of Cu will be added to only a very small flow rate LlQva. The Cu
value of the remaining flow rate QI-LlQva will still remain unaffected by the action of
the blade element Ma and will remain equal to zero. The blade element Ma has
changed only the Cu value of LlQv; therefore, for small values of LlQva, it is valid to use

..
A
Figure 29-2 Symbols used in the
study of interaction of an axial flow
blade and the pumped liquid.
New Findings Concerning the Mode of Operation of Rotodynamic Impellers 337

Cu=Cul =0 for the purpose of calculating aQua with the aid of Eq. 29-4. This means
that

As a result of the power being added to the partial flow aQua, the total flow rate QI
exits the distance Ma with a total power content of

(29-6)

where Cua is the Cu value determined from within the distance Ma on the basis of
Eq.29-5.
The velocity component C ua that has been imparted to the particles of liquid of the
volume aQua causes these particles to collide with other particles of the volume QI
that, originally, have not been acted on directly by the blade strip Ma. These
collisions cause exchanges of momentum. Consequently, a part of the remaining
particles of QI also begin to move in the direction of Cu.
We do not know the actual distribution of Cu within the flow rate QI, when it is
leaving the axial distance Ma. However, as follows from our earlier discussion, for a
given blade velocity U and for Cul =0, it is possible to express the total amount of
power transmitted to the flow QI by the blade strip Ma by the magnitude of the Cu
component of the liquid. Consequently, we are entitled to express the amount of
power transmitted by the blade strip Ma to the flow QI by an equivalent value of C u ,
which is equal to CE, where

CE = C lla X f).Qva . (29-7)


a QI

In that case

Wa=pxQlxUxCEa. (29-8)

In Eq. 28-8, CEa expresses the amount of power transmitted to the pumped liquid
by the blade strip M a , regardless of the magnitude and the actual distribution of the
Cu components of the discrete particles of liquid that make up the flow QI.
In relation to a stationary observer, the amount of power that is added to the flow
rate QI, as it exits the axial distance Ma can be expressed by the Cu component CEa.
Consequently, applying the same reasoning that led us to use the magnitude zero for
the Cu component of the liquid that is entering the interval M a , we can use CEa for
the Cu component of the liquid entering the interval M b . This means that

In relation to the inlet edge, the total amount of power that is added to the flow QI
as it leaves the axial distance Mb is equal to
338 A Glimpse of the Future

As a result, the flow rate QI enters the distance Me with the equivalent C u value of

In a similar manner, we find that the value LlQvn, for any axial distance M n, is given
by
(29-9)

Also, the equivalent value of Cu at the outlet of any axial interval


Mn (Fig. 29-2) is given by

(29-10)

By repeating the same procedure up to the outlet edge of the blade, we can find the
equivalent value of Cu at the impeller outlet. To find the head added to the liquid
owing to the action of the blade, we simply multiply C& by U2/ g.
Let us compare our findings with the performance of a constant pitch inducer of
high solidity. In such an inducer the value of C u , as given by Eq. 29-5, remains
constant. The value of CE, however, grows continuously with the distance A from the
blade inlet. This growth, however, can continue only until CEn- 1 = Cu, In that case,
Cun=Cu and Eq. 29-10 reduce to

CEn= CEn-l =constant . (29-11)

This means that the head increases with the distance A from the inlet edge only up
to a certain axial distance Ad (Fig. 29-1 ). After exceeding that distance, the head
remains constant. This really has been confirmed by numerous tests [7,56,59,66) as
shown schematically in Fig. 29-3.

APPLICATION TO RADIAL FLOW IMPELLERS


The approach developed for axial flow impellers can also be used to calculate the
head developed by a radial flow blade. Referring to Fig. 29-4, divide the distance

-c
ell
Q)
..c
-c
Q)
0..
a
(i)
>
Q)
o Figure 29-3 Increase in head
along the blades of a constant pitch,
Axial distance from inlet ---lI"~ axial flow inducer of high solidity.
New Findings Concerning the Mode of Operation of Rotodynamic Impellers 339

Figure 29-4 Interaction between a


radial blade and the liquid.

R2 - Rl into many annular elements 1lR. Starting with the innermost annular element
IlRl. we find the value of CEa at Rl +dR. Now. using this value of CEa• we calculate
Qvb and the value of CEb. at Rl + 2dR. We continue in the same manner to calculate
the CE values at ever-increasing radii. until we arrive at the last annular interval
IlR2. Having found C&.. we are able to calculate the head HL developed by the blade
at the radius R2-0.5IlR by multiplying C&' by [W(R2-0.5* IlRUg).
For small values of dR, we can now find H2 developed by the blade from the relation

(29-12)

However, when applying this approach to a radial flow blade, we must modify
Eqs. 29-9 and 29-10 so they account for the fact that, in a radial flow impeller, the
peripheral velOCity U of the blade varies with the radius R.
This can be accomplished as follows. Figure 29-5 presents a magnified image of the
interval IlRn, which is shown in Fig. 29-4. Let us denote the radius of the center of
IlR n by R n, and let the CE value at the entrance to this blade element IlRn, be given by
CEn-l. The total head of the flow rate QI entering IlR n is equal to

w x (Rn - O.5LlRn) x CEn_l


9

The head of the flow rate QI -Qvn did not change when passing from Rn-O.5dRn to
Rn. Consequently, the equivalent Cu component of the flow rate QI-IlQnv becomes
340 A Glimpse of the Future

Figure 29-5 Detail from Fig. 30-4.

equal to [CEn-lX(Rn-O.5LlRrJ/Rnl. at R=Rn. This means that the magnitude of Qvn is


now given by

(29-13)

Similarly. at the outlet of the annular element LlRn• the value of CE (which has been
calculated for R=Rnl has to be multiplied by [Rn/(Rn+O.5LlRnll. This leads to the
following equation for the CE value at the outlet of any annular element LlRn of a
radial flow impeller:

CEn = [AQvn x C un + (QI - AQvn) x CEn_d x Rn


(29-14)
QI x (Rn + O.5ARn)

Table 29-1 Tested pump heads vs. calculated impeller heads

Pump RPM Eff Flow H test Hca1c Ratio


(%) (m3 /hr) (m) (m) (%)

End-suction pumps
2x2.5 2950 71 27 27.5 32.7 84
4x6 1470 69 75 35 38.2 91.5
4x6-A 1760 81 198 45 50.5 89
4x6-B 1760 78 198 40.5 49.9 81
4x6-C 1760 77 198 34.9 43.1 81
Multistage donut pumps
7x9 1470 79 250 31.1 34 91.4
Double-suction. split-casing pumps
4x4 1470 76 185 37 42.2 86.6
8xl0 1760 87 590 41 46.7 87.7
Sewage-disposal pumpst
5x6 1470 89 250 22.9 25.6 89
6x6 1470 82 360 23 26.3 87.5

tBecause of the thickness of the blades. each face of the sewage impellers has a different shape and produces a different
head. The heads shown here are the average values of the heads calculated for each face separately.
New Findings Concerning the Mode of Operation of Rotodynamic Impellers 341

Eqations 29-13 and 29-14. in combination with Eq. 29-12. can be used to calculate
the head developed by a blade of a radial flow impeller. Of course. such an approach
can produce only approximate results. However. as will be demonstrated below. these
results can serve as a powerful tool for studying the effects of blade geometry on the
performance of a centrifugal pump.

COMPARISON WITH TEST RESULTS

Table 29-1 presents a comparison between the heads developed by 10 pumps of


unrelated geometries and the heads calculated for their impellers on the basis of
Eqs. 29-12-29-14.
As can be expected from the introduction to this chapter and from the discussion
presented in Ref. 57. the actual heads developed by the tested pumps turned out to
be lower than the heads calculated for the impeller blades alone. A study performed
on some 500 dissimilar pumps indicates that, on average. the actual head developed
by a pump is about 85% of the head calculated for the impeller alone [54). This
difference usually varies from pump to pump. within a range of 5% to 8% (with the
exception of a few extraordinary cases). The data presented in Table 29-1 fall well
within the expected range.

NOMENCLATURE

A = axial distance from the leading m = mass of the matter contained


edge within a partial volume that is
located within the defined
Ad = a certain defined distance from boundary
the inlet edge
m3 / sec = cubic meters per second
B = blade width
Ns = specific speed: (m3 / seclo. 5
CE = eqUivalent velocity component of x (RPM) /(m)o. 75
the liquid in the direction of U
Q = flow rate of the pumped liquid
Cm = meridional velocity component of
the liqUid QI = flow rate between two consecutive
blades
Cu = velocity component of the liqUid in
Qv = volume of liqUid displaced per unit
the direction of U
of time by a moving blade
9 = acceleration due to gravity
R = radius
H = total head U = peripheral velocity of a blade; also.
K = a coefficient velocity of the boundary of a
volume of a given mass of matter
L = blade length
u = instantaneous velocity of any
M = mass of the total volume of the individual particle of matter
matter contained within a defined located within the boundary of a
boundary given volume V
342 A Glimpse of the Future

W = power transmitted by the blades a = blockage owing to final blade


to the liquid thickness
X = distance from the blade's surface
measured in the direction of U.
Subscripts
Z = number of blades
~ = blade angle 1 refers to the leading edge of a blade
p = density 2 refers to the outlet tip of a blade
w = angular speed Rad/sec n refers to the nth element of a blade
TJ = efficiency n-l refers to the preceding blade element
CHAPTER 30
Some Future
Applications of the
Presented Theory
This chapter discusses a few of the future applications of the theory presented in
Chapter 29. It is expected that, with time, the pump user will be able to make use of
it in many additional ways.

PREDICTING THE HEAD-CAPACITY CURVE OF A PUMP


WITH A CUT-DOWN IMPELLER DIAMETER

The approach that led to Eq. 29-14 makes it possible to directly calculate the
theoretical heads developed by an impeller at different diameters (compare Ref. 53).
When we calculate the theoretical heads developed by an impeller with full
diameter and compare them with the actual test results, we find the ratios S of the
tested heads divided by the calculated heads for all flow rates of the given pump.
When we calculate the theoretical heads of the same impeller and the cut-down
impellers and multiply these data by the S values obtained earlier, we get a
significantly better conformance between the predicted and the tested heads than
when we use equations 25-3 and 25-4 (see Fig. 30-1).

CALCULATING THE AXIAL LOADS ON


SEMI-OPEN IMPELLERS

It has been pOinted out on page 000 that, for calculating the axial load of a semi-open
impeller, it is customary to assume that the pressure within the passages of the
blades varies either linearly with the radius, or as the square of the radius. Both
approaches are empirical substitutions of the actual variations of the pressure within
the impeller. A much more correct approach is to calculate the total heads at
different diameters-as we discussed for the case of cut-down impellers-and
subtract the velocity heads. The approach presented in Chapter 29 makes such an
approach possible.
343
344 A Glimpse of the Future

150 3560 RPM 150 3560 RPM


0,= 251 0,= 251

---- .... -
0=238
125

- - - Calculated
- - Tested
according to
QIQ,= DID,
*' - ~ Calculated from 0,.
using Eq. 14 and
HIH,= (010,)2
75 75 S--values for 0,

o 25 50 75 o 25 50 75
Flow rate (m 3/hr) Flow rate (m 3/hr)

Figure 30-1 Testing performance of the same pump with three cut-downs of impeller 00 versus data
calculated from a test with a full-diameter impeller versus data calculated by means of Eqs. 25-3 and
25-4.

INCREASING THE PUMP HEADS TOWARD SHUT-OFF

It has been demonstrated in Chapter 9, that it is possible to increase the pump heads
at reduced flow rates, by either removing parts of the shrouds (Fig. 9-12) or by
opening slots in the shrouds (Fig. 9-11). It also has been demonstrated that the
location of these slots has a crucial effect on the pump's performance (Fig. 9-13). The
theory presented in Chapter 29 makes it possible to determine with higher accuracy
the optimum location of these slots.

CONCLUSION

What we have discussed here can be regarded only as the start of a list of future
applications of the theory presented in Chapter 29. To make this theory useful to the
pump user, however, there is still a need to introduce certain changes in the
purchasing practices when buying centrifugal pumps.
Some Future Applications of the Presented Theory 345

The present practice is to ship a centrifugal pump accompanied by its sectional


drawing and performance chart. In the future, each pump will be accompanied by
either a tape or a floppy disc containing data, in addition to the above-mentioned
data, required for accomplishing different modifications in the performance, and to
solve many other problems that the customer may encounter during use of the
pump.
APPENDIX A
Checklist of Problems
vvith Centrifugal Pumps
and their Causes
Centrifugal pumps are one of the world's most commonly used devices. Moreover.
their field of application is continuously expanding. However. new applications often
bring about new problems. Consequently. it is impossible to foresee the problems
that may turn up in the future. The only alternative is an indepth knowledge and
understanding of how different factors may affect pump performance.
However. even with this knowledge. it is not easy to determine which of the over
120 known causes of trouble with centrifugal pumps is the most likely source of a
given malfunction. Unique cases may also occur. Therefore. one of the first steps in
diagnosing the source or sources of a given problem is to reduce the number of
factors to be checked. This can be best accomplished by studying ready-made
checklists that enumerate the most probable causes of a given problem.
The following is a list of problems that I have encountered during my practice.
Following this list are checklists of causes that might have generated the given
problems.

LIST OF PROBLEMS WITH CENTRIFUGAL PUMPS

1. Pump does not develop any head. nor does it deliver liquid
2. Pump develops some pressure. but delivers no liquid
3. Pump delivers less liquid than expected
4. Pump does not develop enough pressure
5. Shape of head-capacity curve differs from rated curve
6. Pump consumes too much power
7. Pump does not perform satisfactorily, although nothing appears to be wrong with
pumping unit or system
8. Pump operates satisfactorily during start, but performance deteriorates in a relatively
short time
9. Pump is operating with noise, vibrations, or both
347
348 Appendix A

10. Stuffing box leaks excessively


11. Packing has short life
12. Mechanical seal has short life
13. Mechanical seal leaks excessively
14. Bearings have short life
15. Bearings overheat
16. Bearings operate with noise
17. Pump overheats, seizes, or both
18. Impeller or casing, or both, has short life
19. Loud blow is heard each time pump is started or stopped
20. Casing bursts each time pump is started or stopped
21. Gaskets leak during pump operation
22. Flow-rate periodically decreases, or stops completely, then returns to normal
23. Pump develops cavitation when the available NPSH is increased

CHECKLISTS OF CAUSES OF PROBLEMS

1 Pump does not develop any head, nor does it deliver liquid

Possible Causes See Also

1. Pump not primed (not full of liquid) Chapter 1


2. Shaft is broken Chapters 18, 22
3. Broken or disengaged connection between driver and pump
4. Impeller key broken or missing
5. No impeller in pump

2 Pump develops some pressure but delivers no liquid

Possible Causes See Also

1. Air pockets in pump or pipelines Chapters 1, 4, 15


2. Suction line clogged Chapters 5, 18, 23
3. Foot valve stuck to seat or clogged Chapters 5, 18, 23
4. Strainer covered with solid, usually stringy, matter Chapters 5, 18, 23
5. Strainer filled with solid matter such as sand Chapters 5, 18, 23
6. Discharge pressure required by system is higher than maximum
pressure developed by pump Chapters 1, 23
7. Operating speed too low Chapters 1, 6, 25
8. Wrong direction of operation Chapters 19, 23
9. Available NPSH inadequate Chapters 1, 5, 6
10. Excessive amounts of gas or air entrained in pumped liquid Chapters 4, 13
11. Outer diameter of impeller machined to a too small diameter Chapter 25
Checklist of Problems 349

3 Pump delivers less liquid than expected

Possible Causes See Also

1. Air enters pump during operation, or pumping system not


deaerated before starting Chapters 12, 15, Figs. 12-5,
12-6,12-15,12-17,15-1,15-7
2. Insufficient speed Chapter 25
3. Wrong direction of rotation Chapters 19, 23
4. System requires higher pressure than that developed by pump Chapters 1, 23
5. Measuring instruments not properly calibrated or incorrectly
installed Chapter 12
6. Available NPSH too low Chapters 1, 5, 6
7. Excessive amount of air or gas entrained in pumped liquid Chapters 1, 12, 15
B. Excessive leakage through wearing rings or other sealing faces Chapter 7, Figs. 7-7-7-12
9. Viscosity of liquid higher than that for which pump has been
designated ChapterB
10. Impeller or casing partially clogged with solid matter Chapters 12, 15
11. Fins, burrs, or sharp edges in path of liquid Chapter 15, Figs. 12-2, 12-15
12. Impeller damaged Chapter 15, 26
13. Outer diameter of impeller machined to a smaller dimension
than specified Chapter 25
14. Faulty casting of impeller or casing Chapter 15, 19
15. Impeller incorrectly installed Chapters 15, 19, Fig. 15-17
16. Pump operating too far out of the head-capacity curve Chapters 6, 20, Figs. 6-10, 20-2
17. Obstruction to flow in suction or discharge piping Chapters 5, 1B, 23
lB. Foot valve clogged or jammed Chapters 5, 1B, 23
19. Suction strainer filled with solid matter Chapters 5, 1B, 23
20. Suction strainer covered with fibrous matter Chapters 5, 1B, 23
21. Incorrect layout of suction or discharge piping Chapter 14
22. Incorrect layout of suction sump Chapter 14
23. Excessive leakage through stuffing box or seal Chapter 7, Figs. 7-7-7-12
24. Excessive amount of liquid recirculated internally to stuffing box
lantern or seal
25. Excessive leakage through hydraulic balancing device Chapter 19, Figs. 10-4, 10-5,
1M, 10-B
26. liquid level in suction tank or sump lower than originally
specified Chapters 5, 23
27. In a system with more than one pump, operation of one pump
may affect operation of others Chapters 14, 23

4 Pump does not develop enough pressure

See checklist 3.

5 Shape of head-capacity curve differs from rated curve

See checklist 3.
350 Appendix A

6 Pump consumes too much power

Possible Causes See Also

1. Speed too high Chapter 25


2. Pumped liquid of higher specific gravity than originally quoted
3. Pumped liquid of higher viscosity than originally quoted Chapter S
4. Oversized impeller Chapter 25
5. Total head of system either higher or lower than anticipated Chapter 23
6. Misalignment between pump and driver Chapters 15
7. Rotating parts rubbing against stationary parts Chapters 15, 16
S. Worn or damaged bearings Chapter 16
9. Packing improperly installed Chapter 17, Fig. 12-21
10. Incorrect type of packing Chapter 17, Appendix B-3
11. Mechanical seal exerts excessive pressure on seat Chapter 17
12. Gland too tight Chapter 17
13. Improper lubrication of bearings Chapter 16
14. Too much lubricant in bearings Chapter 16
15. Bent shaft Chapter 16, Fig. 15-16
16. Uneven thermal expansion of different parts of pumping unit ChapterS
17. Faulty power-measuring instruments Chapter 12
lS. Power-measuring instruments incorrectly mounted or connected Chapter 12
19. Wrong direction of rotation Chapters 19, 23
20. liquid not preheated to keep viscosity below specified limits ChapterS
21. Impeller or casing partially clogged with solid matter Chapters 5, 15
22. Wetted surfaces of impeller or casing very rough Chapter 7
23. Damaged impeller Chapter 15
24. Faulty casting of impeller or casing Chapter 15
25. Impeller incorrectly located in casing Chapter 19, Fig. 15-17
26. Impeller inversely mounted on shaft Chapter 19, Fig. 10-6
27. Pump operating too far out on head-capacity curve Chapter 6, 20, Figs. 6-10, 20-2
2S. Incorrect layout of suction sump Chapter 14
29. Breakdown of discharge line

7 Pump does not perform satisfactorily, although nothing appears to be wrong with pumping system

Possible Causes See Also

This is usually due to incorrect testing. The reasons for this may
be as follows:

1. Incorrect measuring instruments Chapter 12


2. Measuring instruments damaged during installation Chapter 12
3. Measuring instruments mounted in wrong locations Chapter 12
4. Tubing that leads from pipelines to measuring instruments
clogged Chapter 12
5. Instrument-connecting tubing that should be full of liquid not
deaerated completely Chapter 12
6. Instrument-connecting tubing that should be full of air contains
some liquid Chapter 12
Checklist of Problems 351

7 (continued)

Possible Causes See Also

7. Leakage in instrument-connecting tubing or in its fittings Chapter 12


8. Burrs or fins at mouth of connections between tubing and piping Chapter 12, Fig. 12-7
9. Incorrect connections of wiring to electrical instruments Chapter 12
10. Connections of wires to terminals too loose Chapter 12
11. Dirty electrical terminals or connections Chapter 12
12. Dust or dirt in torque bar Chapter 12
13. Torque bar incorrectly mounted Chapter 12
14. In a dynamometer, misalignment or dirt in bearings produces
false readings Chapter 12
15. In a dynamometer, excessive friction in pivots or pulleys that
guides the levers and cables produces false readings Chapter 12
16. In a dynamometer, weight and stiffness of the electrical
cables affect torque readings Chapter 12
17. Cavitation in measuring instruments Chapter 12
18. Cavitation in pipelines where instruments are hooked up Chapter 12
19. Actual inner diameter of piping different from nominal diameter Chapter 12

8 Pump operates satisfactorily during start, but performance deteriorates in a relatively short time

Possible Causes See Also

1. Air leaks into pump Chapters 4, 15


2. Pumped liquid contains high percentage of entrained air or gas Chapter 4
3. Waterfall-like supply of liquid into suction sump draws air into pump Chapter 4, Fig. 14-22
4. Air pocket in suction line has moved into pump Chapters 4, 14, Figs. 12-5, 12-6,
12-15, 15-1-15-7
5. Air funnels in suction sump Chapter 14

9 Pump is operating with noise or vibrations, or both (see also checklist 16)

Possible Causes See Also

1. Misalignment between pump and driver Chapter 15


2. Rotating parts rubbing against stationary parts Chapter 15
3. Worn-out bearings Chapter 16
4. Wrong direction of rotation Chapters 19,23
5. Available NPSH too low Chapters 1, 5, 6
6. Impeller or casing partially filled with solid matter Chapters 5, 18, 23
7. Fins, burrs, or sharp edges in waterways causing cavitation Chapter 13
8. Damaged impeller Chapter 15
9. Impeller incorrectly mounted Chapters 15, 19, Fig. 15-17
10. System requirements too far out on head-capacity curve Chapters 6, 20, Figs. 6-10, 20-2
11. Suction strainer filled with solid matter Chapters 5, 18, 23
12. Strainer covered with fibrous matter Chapters, 5, 18, 23
352 Appendix A

9 (continued)

Possible Causes See Also

13. Incorrect layout of suction sump Chapter 14


14. Air enters pump during operation Chapters 1, 13
15. Mutual interaction of several pumps within one common
system Chapters 14, 21-23
16. Incorrect layout of suction or discharge piping Chapter 14
17. Piping imposes strain on pump Chapters 15, 23
18. Pump operating at critical speed Chapter 18
19. Rotating elements not balanced Chapter 15
20. Excessive radial forces on rotating parts Chapters 10, 17, 27
21. Too small distance between impeller outer diameter and
volute tongue Chapter 27
22. Faulty shape of volute tongue Chapter 27
23. Undersized suction or discharge piping and fittings causing
cavitation somewhere in system Chapter 13, Fig. 13-11
24. Loose valve disc in system Chapter 13, Figs. 13-12, 13-14
25. Bent shaft Chapter 15
26. Impeller bore not concentric with its outer diameter or not
square with its face Chapter 15, Fig. 15-16
27. Misalignment of pump parts Chapter 15
28. Pump operates at very low flow rates Chapter 9, Figs. 9-8--9-10
29. Improperly designed base plate or foundations Chapter 15
30. Resonance between pump speed and natural frequency of
base plate or foundations Chapter 18
31. Resonance between operating speed and natural frequency
of piping Chapter 18
32. Resonance between operating speed and valve discs Chapter 18
33. Loose bolts Chapter 15
34. Uneven thermal expansion Chapter 8
35. Improper installation of bearings Chapters 15, 16
36. Damaged bearings Chapters 15, 16
37. Improper lubrication of bearings Chapter 16
38. Obstruction to flow in suction or discharge piping Chapters 13, 15, Figs. 12-15-
12-17,15-1
39. Total head of system either higher or lower than expected Chapter 2, 23
40. Excessive amount of air or gas entrained in liquid Chapters 4, 15
41. Waterways of impeller or casing badly eroded or rough Chapter 23
42. Cavitation in pipelines Chapter 13, Figs. 13-11, 13-12,
13-14

10 Stuffing box leaks excessively

Possible Causes See Also

1. Worn out bearings Chapters 14,16


2. Improperly installed packing Chapter 17, Fig. 12-21
3. Incorrect type of packing Chapter 17, Appendix B-3
Checklist of Problems 353

10 (continued)

Possible Causes See Also

4. Rotating element not balanced Chapter 15


5. Excessive radial forces on rotating parts Chapters 10, 27
6. Bent shaft Chapter 15
7. Bore of impeller not concentric with outer diameter, or not square
with face Chapter 15, Fig. 15-16
8. Misalignment of pump parts Chapter 15
9. Rotating parts running off-center Chapter 15
10. Water-seal pipe clogged Chapter 17
11. Seal cage improperly located Chapter 17, Fig. 12-21
12. Shaft sleeve worn or scorched at packing Chapter 17
13. Failure to provide cooling liquid to water-cooled stuffing boxes Chapter 17, Fig. 12-21
14. Excessive clearance at bottom of stuffing box (between shaft
and box bottom) Chapter 17
15. Dirt or grit in sealing liquid Chapter 17

11 Packing has short life

Possible Causes See Also

1. Worn bearings Chapter 16


2. Improperly installed packing Chapter 16, Fig. 12-21
3. Incorrect type of packing Chapter 17, Appendix B-3
4. Gland too tight Chapter 17
5. Rotating element not balanced Chapter 15
6. Excessive radial forces on rotating parts Chapters 10, 17, 27
7. Bent shaft Chapter 15, Fig. 15-16
8. Bore of impeller not concentric with its outer diameter or not
square with its face Chapter 15
9. Misalignment of pump parts Chapter 15
10. Rotating parts running off-center from damaged bearings or
other parts Chapter 15
11. Water-seal pipe clogged Chapter 17
12. Seal cage improperly located in stuffing box, preventing sealing
fluid from entering Chapter 17, Fig. 12-21
13. Shaft scorched where it contacts packing Chapter 17
14. Failure to provide cooling liquid to water-cooled stuffing
box Chapter 17
15. Excessive clearance at bottom of stuffing box, between shaft
and stuffing box's bottom Chapter 17
16. Dirt or grit in sealing liquid Chapter 17
17. Improperlubrication of packing Chapter 17
18. Space in stuffing box where packing is located is excentric
to the shaft Chapter 17, Fig. 17-1
354 Appendix A

12 Mechanical seal has short life

Possible Causes See Also

1. Worn out bearings Chapter 16


2. Rotating elements not balanced Chapter 15
3. Excessive radial forces on rotating parts Chapters 10, 27
4. Bent shaft Chapter 15
5. Misalignment of pump parts Chapter 15
6. Rotating elements running off-center from damage to bearings
or other parts Chapters 15, 16
7. Dirt or grit in seal-flushing liquid Chapter 17
8. Sealing face not perpendicular to pump axis Chapter 17
9. Mechanical seal has been run dry Chapter 17
10. Abrasive particles in liquid coming in contact with seal Chapter 17
11. Mechanical seal improperly installed Chapter 17
12. Incorrect type of mechanical seal Chapter 17
13. Misalignment of internal seal parts preventing proper mating
between seal and seat Chapter 17

13 Mechanical seal leaks excessively

Possible Causes See Also

The same factors as in checklist 12, plus the following

1. Leakage between the seal seat and gland from faulty gasket
or O-ring Chapter 17
2. Leakage between seal and shaft from faulty O-ring or lip seal Chapter 17

14 Bearings have short life

Possible Causes See Also

1. Damaged impeller Chapter 15


2. Impeller partially clogged Chapter 15
3. Rotating elements not balanced Chapter 15
4. Excessive radial loads on rotating parts Chapters 10, 27
5. Excessive axial loads Chapters 10, 27
6. Bent shaft Chapter 15, Fig. 15-16
7. Bore of impeller not concentric with outer diameter or not square
with hub face Chapter 15
8. Misalignment of pump parts Chapter 15
9. Misalignment between pump and driver Chapter 15
10. Pump operates for prolonged time at low flow rate Chapter 18, Figs. 9-8-9-10
11. Improper base plate or foundations Chapter 15
12. Rotating parts running off-center from damaged or misaligned
parts Chapter 15, Fig. 15-16
Checklist of Problems 355

14 (continued)

Possible Causes See Also

13. Improper installation of bearings Chapter 16, Figs. 15-10, 15-15


14. Bores of bearing housing not concentric with bores in water end Chapter 15
15. Cracked or damaged bearing housing Chapter 15
16. Excessive grease in bearings Chapter 16
17. Faulty lubrication system Chapter 16
18. Improper workmanship during installation of bearings Chapter 16, Figs. 16-4, 16-5
19. Bearings improperly lubricated Chapter 16
20. Dirt finds access to bearings Chapter 16
21. Water has entered bearing housing Chapter 16
22. Excessive wear of impeller sealing rings reducing the effects
of balancing means Chapters 7,19
23. Excessive suction pressure Chapter 20
24. Too tight fit between line bearing and seat (may prevent it
from sliding under axial load, transferring this load to
the line bearing) Chapter 16, Figs. 15-15, 16-10
25. Inadequate cooling of bearings Chapter 16
26. Inadequate cooling of lubricant Chapter 16
27. Source of cooling media shut-off from bearing housing Chapter 16

15 Bearings overheat

See checklist 14.

16 Bearings operate with noise

Possible Causes See Also

A. Steady high-pitch tone


1. Excessive radial load Chapters 10, 27
2. Excessive axial load Chapter 10
3. Misalignment Chapter 15
4. Too much clearance between bearing and shaft, and/or housing Chapter 16

B. Continuous or intermittent low-pitch tone


1. Bearing brinelled Chapter 16
2. Pitted raceway, from dirt Chapter 16
3. Resonance with other structural pump parts Chapter 17

C. Intermittent rattles, rumbles, and/or clicks


1. Loose machine pats Chapter 18
2. Dirt in bearings Chapter 16
3. Clearance between balls and races too large for given application Chapter 16
4. Bearings that require preloading not adequately preloaded Chapter 16
356 Appendix A

16 (continued)

Possible Causes See Also

D. Intermittent squeal or high-pitch tone

1. Balls skidding from excessive clearance between balls and races Chapter 16
2. Balls skidding from insufficient preloading (whenever required) Chapter 16
3. Shaft rubbing against housing from improper mounting of housing Chapter 16
4. Shaft rubbing against housing from bent shaft Chapter 23
5. Shaft rubbing against housing from having been machined
excentrically Chapter 23

17 Pump overheats or seizes, or both

Possible Causes See Also

1. Pump allowed to run dry Chapters 1,15


2. Vapor or air pockets inside pump Chapter 14, Figs. 15-6-15-9
3. Pump operates near shut-off Chapter 13, Figs. 9-8-9-10
4. Simultaneous operation of poorly matched pumps Chapter 14
5. Internal misalignment from too much pipe strain, poor
foundations, or faulty repair work Chapter 15
6. Internal rubbing of rotating parts against stationary parts Chapter 15
7. Worn or damaged bearings Chapter 16
8. Poor lubrication Chapter 16
9. Rotating and stationary wearing rings made of identical,
galling-prone materials Chapter 16, Appendix B-2

18 Impeller or casing, or both, has short life

Possible Causes See Also

1. Corrosion from chemical interaction with pumped liquid Chapter 18


2. Electrochemical corrosion from difference of electrochemical
potential of different materials of which wetted pump parts
are made Chapter 18
3. Abrasion from solids contained in pumped liquid Chapter 18
4. Fatigue from thermal shocks Chapters 15, 18
5. Fatigue from vibrations Chapter 16
6. Erosion from cavitation Chapter 18
7. Excessive transient stresses during starting or stopping Chapter 14
8. Pump used at excessively high temperatures Chapter 8
9. Excessive stresses imposed on pump by piping Chapters 15, 18
10. Excessive stresses imposed on casing by foundation bolts Chapter 15
11. Pump mishandled during installation Chapter 15
Checklist of Problems 357

19 Loud blow heard each time pump is started or stopped

Possible Causes See Also

1. Water hammer Chapter 14


2. Air or gas entrapped between pump discharge and nonreturn valve Chapter 14, Figs. 15-10, 15-11
3. Slam pressure Chapter 14

20 Casing bursts each time pump is started or stopped

Possible Causes See Also

1. Water hammer Chapter 14


2. Slam pressure Chapter 14

21 Gaskets leak during operation

Possible Causes See Also

1. Uneven thermal expansion of pump parts Chapter 18


2. Loose bolts Chapter 15
3. Unevenly tightened bolts Chapter 15

22 Flow rate periodically decreases, or stops completely, then returns to normal

Possible Causes See Also

1. Periodic fluctuations of liquid level in suction tank Chapter 18


2. During operation pump removes more liquid from suction tank
than rate at which liquid enters the tank Chapter 18

23 Pump develops cavitation when the available NPSH is increased

Possible Causes See Also

This may happen when the increase in the available NPSH has reduced
the system resistance so far that the pump operates far out on the
QH curve. This happens when
1. Oversized impeller installed in pump Chapter 20
2. Pump operates at excessive speed Chapter 20
3. Breakdown or serious leak in discharge line Chapter 20
4. Open bypass in discharge line Chapter 20
5. Extremely large clearances between impeller and casing Chapter 20
6. Hole in casing allowing liquid from pressure side of casing to
return to its suction inlet Chapter 20, Fig. 15-12
APPENDIXB

Tables
Table B-1 Vapor Pressures of Water at Different Temperatures

Temperature Density Vapor Pressure


at
a Equivalent head of water Absolute
Degrees Degrees given at a given temperature pressure
(OC) (0C) temperature (m) (ft) (kg/cm2)

0 32 1.000 0.0396 0.13 0.0040


4 39 1.000 0.0823 0.27 0.0083
10 50 1.000 0.125 0.41 0.0125
20 68 0.998 0.238 0.78 0.0237
30 86 0.996 0.426 1.4 0.0424
40 104 0.992 0.762 2.5 0.0755
60 140 0.983 2.012 6.6 0.1977
80 176 0.972 5.000 16.4 0.4860
100 212 0.959 10.775 35.6 1.0333
120 248 0.944 21.275 69.8 2.0083
140 284 0.927 39.624 130 3.6731
160 320 0.909 69.799 229 6.3447
180 356 0.889 114.605 376 10.1884
200 392 0.866 184.992 604 16.0203
220 428 0.841 281.635 924 23.6885
240 464 0.814 420.624 1380 34.2388

359
360 Appendix B

Table B-2 Gall Resistance of Material Combinations

Cast Iron S S S S S SSSSSSSSSSSS SSSSSSSSSSSS


3% Ni Cast Iron S S S S S SSSSSSSSSSSS SSSSSSSSSSSS
Ni-Resist ([ype 1, 2) S S S S S SSSSSSSSSSSSSSSSSSSSSSSS
Ductile Iron S S S S S SSSSSSSSSSSSSSSSSSSSSSSS
Ductile Ni-Resist S S S S S SSSSSSSSSSSS SSSSSSSSSSSS
Nickel-Copper Alloy 505 S S S S S FFSSSS~SFSFFSSPSSF~SSSSS
Nickel-Copper Alloy K-500 S S S S S FNFFSSNSFFNNFSFSSFFFSSSS
Nickel-Copper Alloy 400 S S S S S NNFNFFNSNFNNFSFSSFNFSSFS
Nickel-Copper Alloy 506 S S S S S FFFFFFNSNFNNFSFSSFNFSSSS
Nickel-Alluminum Alloy 301 S S S S S FNFFSSNSNFNNFSFSSSFFSSSS
Nickel 2132 S S S S S SFFSSFFSFSFFSSSSSSSSSSSS
Nickel 3052 S S S S S FFFSFSFSFSFFSSSSSSSSSSSS
Nickel-Chromium Alloy 600 S S S S S NNNNFFNFNFNNFSFSSFNFSSSS
Nickel-Chromium Alloy 7053 S S S S S SSSSSSFSFSFFSSSSSSSSSSSS
400 Series
Stainless Steel (Soft) S S S SS FNNNFFNFNFFNFSFSSFNFSFFS
400 Series
Stainless Steel (Hard) S S S S S FFFFSSFSFSFSSSSSSPFSSSSS
300 Series Stainless Steel S S S S S NNNNFFNFFFNNFSFSSFNFSSSS
SAE 1000 to 6000 Steel (Soft) S S S S S NNNNFFNFNSNNSSSSSPNFSSSS
SAE 1000 to 6000 Steel (Hard)S S S S S FFFFSSFSFSFSSSSSSSFSSSSS
Bronze (Leaded)5 S S S S S SSSSSSSSSSSSSSSSSSSS~~S~
Ni-Vee Bronze "A"4 S S S S S FFFFSSFSFSFSSSFSSFFSSPSS
Ni-Vee Bronze "B" S S S S S SSSSSSSSSSSSSSFSSFSSSSSS
Ni-Vee Bronze "D" S S S S S SSSSSSSSSSSS SSSSSSSSSSSS
Ni-Al Bronze6 S S S S S F- F- F- F- S S F- S F- ~ F- ~ S S F P S N F P S F S S
HASTELLOYl Alloys A, B S S S S S FNNFSSNSNFNNFSFSSFNFSSSS
HASTELLOY Alloy C S S S S S FFFFSSFSFSFFSSSSSPFFSSSS
HASTELLOY Alloy D S S S S S SSSSSSSSSSSSSSSSSSSSSSSS
Nitrided S S S S S S S S S S S S S F S S S S S F5S S F S S S S S S
Chrome Plate7 S S S S S S S S S S S S S S S S S S S S S S S S S S S ?7 S
STELUTEl S S S S S SSSSSSSSSSSS SSSSSSSSSSSS

Degree of Resistance: S=Satisfactory F=Fair N= Little or None Courtesy of Goulds Pumps, Inc.
1 Trademark of Union Carbide Corporation. 5 Loaded Bronze-85-5-5-5 or 80-10-10_ Hard materials
2 Nickel 213 and Nickel 305 have better gall resistance might "bite" into softer bronzes.
than Nickel 210. Both are comparable in gall resistance 6 Nickel-aluminum-bronze is generally somewhat
but Nickel 305 will stand heavier loads. inferior to Ni-Vee "N' in gall resistance and coefficient
3 Nickel-chromium alloy 705 is superior to Nickel 305 of friction, but will stand heavier loads in slower
and nickel-copper alloy 505 in gall resistance. motion.
4 The Ni-Vee bronzes are 5% nickel, 5% tin, cast and heat- 7 Chromium plate varies greatly in gall resistance. To be
treatable, similar in balance of composition to the 88- its best it must be backed up by hard material and the
10-2 Cu Sn Zn type. A, No load. B, 1% load. D, 10% load. plating must bond well to the backing.
Tables 361

Table B-3 Packing Selections

Standard

Maximum Pressure
Style Description Temperature Remarks

Garlock 5022 AFP or PfFE, Impregnated Copper, 5000 PSI/550°F


equal (C06) Wire Braided Ring (%6 in. Good for most liquids
through 2 in. Plungers) except bromine,
PfFE, Impregnated, Braided chlorine, and
Garlock 8922 or equal
(C06) Ring, (2 112 in. through 47/16 in. oxygen compounds
Plungers)

Optional

Maximum Pressure
Style Description Temperature Remarks

G8048, G432 or equal Nitrile/Fabric, V Ring 2000 PSI/200° F Aqueous solutions


(Neo Duck) except aromatics or
aqueous solutions of
acids or bases

Crane CVH or equal PfFE, V Ring with PfFE 5000 PSI/500°F All liquids except
(8764) Adaptors flourine and its
components

Crane 829 or equal Nitrile/Fabric with Brass 7500 PSI/250°F Mineral oils, petroleum
Adaptors (other adaptor products, water
material available) emulsion solutions

SPecial Common specials noted below. Contact Application Engineer

Maximum Pressure
Style Description Temperature Remarks

Crane CI055 or equal PfFE Yarn, Braided Ring 2000 PSI/500°F Food products

Grafoil 235A or equal Graphite, Comp-Split Ring 4000 PSI/1500°F Strong corrosive,
heat transfer liquids

Courtesy of Milton Roy Company, A Sundstrand Subsidiary.


Notes:
1. Pressures over 2000 psig require hardened plungers and close clearance rings. V-Ring-type packing requires metal
adaptors.
2. When flushing is required, use a V-ring-type packing.
3. Use Neo Duck packing for lima and diatomaceous earth slurries with ceramic plungers and flush connections.
4. Milroyal D% in. & %in. plungers are only available with 25% carbon-filled PTFE packing.
362 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals eF)
and
Formula
z
o
~
~
""z
u
o
u

Acetaldehyde Conc. 120 C C 350 200 C C C C B C C C CC B BA BBA A AA A


CH3CHO
Acetic Acid 60% 73 73 175350 140 C C C CCCCCCCCCCCAAAAA A
CH3COOH
Acetic Acid 85% C 120 73 150 350 140 C C C CCCCCCCCCCCAAAAA A
CH3COOH
Acetic Acid Glacial C 120 73 120350 140 C C C CCCCCCCCCCCCABAA A
CH3COOH
B
Acetic Anhydride C
C 350 C 70 200 to C B C C C C C C C C C C C B B B B A
(CH 3 CO)zO 70
B
Acetone C 73 C C 350130 C to C C AAAAAAAAAAAAAAAAAA
Ch3COCH3 70
Acetylene GAS 70 73 140250 250 200 140 70 70 200 A C C C C C A A A A A A A A A A A A
HC~CH 100%
Acrylonitrile C C 73 350 C C 140 C C B A AAA A A A A AA A A A A A
H2C:CHCN
Allyl Chloride C 212350 C C 70 C C
CH2CHCH2Cl
Aluminum Hydroxide Sat'd 185 140280 250 210 180 100200 CCCCCBBC BBAAAB
Al03·3H20)
Aluminum Nitrate Sat'd 185 180 140280 250 210 180 100 100 100 C C C C C C C C C C C A A AC
Al(N03)3 . 9H20
Aluminum Sulfate (Alum) Sat'd 185 180 140280 250 210 200 160 140 185 C C C C C C C C C CC B A
Al2 (S04h
Ammonia Anhydrous 250 200 100 C CCCCCA A AAAA A
NH3
B
Ammonia liquid 100% 73 C
C 400 210 to 70 70 C A C C C C C A AAAAA A
NH3 70
B B A
Ammonium Carbonate Sat'd 180 140280 400 210 140 140250 to to C C toC BBBBB A
CNH4)HC03· (NHJC02NH2 21270 140
Ammonium Chloride Sat'd 185 180 140280400 210 180200 160250 B C CCCCCCCBCBB B
NH 4 Cl
B
Ammonium Hydroxide 10% 185 180 140225 400 210 to 200 70 70 B C C C C C BAAAB A
N0 4 0H 70
(The information given is indended as a guide only. See page 374 for further information)
Tables 363

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
z
o

~~ z z
u
Z
o
u ~ ~ ~ ~ ~ ~ ~ ~
Anunonium Nitrate Sat'd 185 180 140280 400 250 180200 160 100 C C C C AAA A
NH4N03
Anunonium Sulfate 185 180 140280 400 210 180200 160200 C C C C C C B B C B B B B B A B A
(NH4hS04
Ammonium Sulfide Dilute 125 350 210 140200 160 ACCCCCCCC C B BB
(NH4hS
Ammonium Thiocyanate 50-60% 140275 70 70 70 185 B C C C C C C C C C A AAB A
NH4SCN
B
Amyl Acetate C C 125100 to C C C C A B BBB B BB ABAAAAAAA
CH3COOCsHl1 70

n-Amyl Chloride C 280400 C C C C 200 C AAAAAAAAAAAAAAAA


CH3(CH2hCH2Cl
Aniline C 180 C 120200 140 C 70 C C CCCCCCBBCBBAAAAB A
CsHsNH2
Arsenic Acid 80% 185 140280400 185 160200 180200 C CCCCCCC CBABAA A
H3As04 . %H20
Barium Carbonate Sat'd 140280 400 250 180200 160250 AAAABBBBBAAAAA
BaC03
Barium Chloride Sat'd 140280400 250 180200 160300 AAAAABBCBBBA AA
BaCh·2H 20
Barium Hydroxide Sat'd 140280400 200 180200 140300 CCCCBBC BAAAAA A
Ba(OHh
Barium Sulfate Sat'd 185 140280400 200 100200 160300 BBBBBBA BAAAAA
BaS04
Barium Sulfide Sat'd 140280 400 140 C 200 160300 CCCCCBBC BAAAAA
BaS
Beer C 180 140200300 200 70 200 140200 AAAAAACCC CAAAAA

Beet Sugar Liquors 180 140225 210 100 200 160 185 A A BBB AAAA

Benzene C C C C 170250 C C C C 150 AAAAAAAAAAAAAAA


CsHs
Benzoic Acid All 73 140230350 C C 200160200 CCCCCCC CAAAAA A
CsHsCOOH
Black Liquor Sat'd 185 140175225 180180 70 70 200 CCCCBBB BBABAB
364 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
!!l ~
z z
z o-l B
~
o 0 "'!!l0
uz§!'i ZW;:Jcr.i U

~ Z Z
Z
W
Z ~
~ Iil
~
z~w~z
~o
~~~
o~
z~
Cicr.i
ZCioo
lil~zlil~~ ~o-l§
§
~
tl <:
~ ~ ~~~§~~~§~~~ffi~~~~~~
;21
z
8
00
~
~u 8:: ~ ~
~
Ci
Il.
W
Z
;:J
~
0
W
Z
° :;JU~1Ji:;J~"CiU~Z~C;;~ ;21
Borax Sat'd 180140280 210 140200 140 185 AAAAAABAAAAAAAAA
N&B407 . 10H20
Boric Acid Sat'd 185 180 140280 210 140200 140 185 BBBBCCB CBABAA A
H3B03
Butane 50% 73 140250 350 C 70 200 70 185 A A A A A A A A A A A A A A A A A A
CilllO
Butylene (C) liquid 140280 400 C 70 C C 100 A AAA A AAAAAAA
CH3CH:CHCH3
Butyric Acid 180 73 230 300 140 C C 70 AAAACCCCCBAAAA
CH3CH2CH2COOH
Calcium Bisulfite 185 180 140280350 C 70 200 70 185 CCCCCCC CBA AC A
Ca(HS03h
Calcium Carbonate 185 180 140280 350 210 100 70 70 300 A C CCC B BB BAA A A A A
CaC03
Calcium Chlorate 140280350 140 70 70 70 185 C B BBB B BB BBBA AA
Ca(CI03h . 2H20
Calcium Chloride 100 185 180 140280 350 210 100200 160250 B B BBB A AC CBA B AAB A
CaCh
Calcium Hydroxide 185 180 140280 250 210 140200 70 250 C C C C C C C C C CAA AAA A
Ca(OHh
Calcium Hypochlorite 30% 185 150 140200 200 70 C 140 185 C C C C C C C C C CBB B BC B
Ca(OClh
Calcium Nitrate 180 140280 200 210 180 100 100 200 B B BBB B B B A AA
Ca(N03h
Calcium Sulfate 100 140280 200 210 180200 160200 A A B B B A A B A AA A A A AA A
CaS04
Cane Sugar 73 140275400 250 180100 160200 AAAAAAAAAAAAAAAA
C12H22011
Carbon Dioxide Dry 100 185 150 140280 400 200 180200 160200 A A A A A A A A A A A A A A A A A
C02 100%
Carbon Tetrachloride C 73 C 73 280 350 C C C C 185 B A A AA C C A CAA AAA A
CCL!
Carbonic Acid Sat'd 185 140280350 210 180 70 70 200 C CCC B BB BBAAAAAB
H2C03
Cellosolve 73 280200 140 C 70 C AAAAAAA A A A A
CICH2COOH
B
Chloral Hydrate All 140 75 70 to C
CChCH(OHh 70
Tables 365

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals eF)
and
Formula r.:!
N ~
~ 6
Z
0 Z
0 s~~
UZ~ Zr.:!;:lui u

~
r.:! Z Z
Z
r.:!
Z
~
~ ~
~~~~~
~o
~~~
o~
z~
@~z~~~
Qui
ZQ~
~~§~
><
0

~
@
:3 ~~~§~~~§~~~~~~~~~~
u
Z
~ ~ :2:5:~ <1:
Z 0 0
U 8:: ~ ~ ~ r.:!
~
0 ;:l r.:!
U ~ ~ Z ~ ~u~U'J~~'-'QU~Z~c;j~ :2:
Chlorobenzene Dry 73 C 170200 C C C C 70 A A AA C C B CAAAAA A
CsIIsCl
Chloroform Dry C C C 125200 C C C C 70 A A AA C C C CAAAAAA
CHCll
Chlorosulfonic Acid C 73 C 200 C C C C C CCCCCCBBCCBCCCBA A
CIS020H
BA A
Chromic Acid 10% 210 150 140 175 350 70 C 140 100 C C C CCC C CC CCtoto to B A
H2CrO, 212 70 125
B B
Chromic Acid 50% 210 180 C 125 200 C C 140 C CCCCCCCCCC C to to C B
H2CrO, 70 212
Citric Acid Sat'd 185 180 140275 200 210 70 140 140200 CCCCCCCC CBAAAA A
CsHS07
Coffee 140100 200 A A AAA C C C AAAA A

Copper Acetate Sat'd 73 73 73 250 350 100 180 C 160 140 CCCCCCC CBA AB A
Cu(C2H302)z . H2O
Copper Chloride Sat'd 185 140280350 210 180 200 160200 CCCCCCCCCCCBA AB A
CuClz
Copper Cyanide 185 140 275 350 210 180 160185 CCCCCCCCACBA AB
Cu(CN)z
B
Copper Nitrate 30% 140280 210 to 200 160 200 CCCCCCCCC CBA AC
Cu(N03)z·3H20 70
Copper Sulfate Sat'd 185 120 140280 210 180 200 160 200 CCCCCCCCC CAAAAC A
CuSO,·5H2O
Creosote 73 73 350 C 73 73 C 73 BBBBBAAAAAAAAAAAA

Cresylic Acid 50% 140150200 C C C C 185 AAAAAAABAAAAAAAA

Cyc10hexane 100 C C C 280300 C C C C 185 AAAA B BA BAAAAAAA


CsHl2
Detergents 185 150 140 250 180 200 160 210 AAAAAAAAAAAAAAAA
(Heavy Duty)
Dow ThermA C 212 C C C C C AAAABAA AAAAAAA

Ether C 73 C 125 C C C C AAAA BBBAAAAAAAA


ROR
366 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (oF)
and
Formula w
N ~
Z
0 Z
0 Sf:l:§
uz~
G
...l
u
~
ZW;::lui

Z
W
Z ~ ~
;::l
fIlo
OO~~
;;~;::l
o~
z~
0
Clui
ZCl~
ZO W
§>-
~~NZ~OO.~O~S~oo.~~§~
Z 0
Z 0
uW
Z
<i: ~ ~
~
~
~ ~ 0 ~~zo;:;;:~>; ~~ ~ui~OZ
0
U
Z ~

~
~
U 8:: ~ ~ ~
~
W
Z
;::l
~
0
W
Z
g
~ ~8~~~~~8u~~§~~~~~
Ethyl Acetate C 120 C C 200 70 C C C C AAB AAA AAAAABA
CH3COOC2H5
Ethyl Alcohol (Ethanol) 140 180 140280300 170 180200 70 AAAAAAAAAAAAAAAA
C2H 50H
B B
Ethyl Chloride Dry 73 C 280350 to C 70 to 140 A AB AAAAAAAAAABA
C2H5Cl 70 70
Ethylene Chloride Dry C 73 C 280350 C C 70 A AA
ClCH2CH2C1
Ethylene Glycol 73 185 120 140280 210 180 200 160 250 AAAAAAAA AAAAAAAA
CH20HCH20H
Ethylene Oxide C C C 400 C C C C C AA BAA A A AB A
CH2CH20
Fatty Acids 73 120 140280 400 C 140 C 140 185 C C C CC C C C C A AA A
R-COOH
A
Ferric Chloride (Aqueous) Sat'd 185 180 140280 400 225 180200 160200 C C C C C C C C C C C C C C C to
FeCl:! 175
Ferric Nitrate Sat'd 185 180 140280 400 210 180 140 160200 C C CCC C C C CBA AA C A
Fe(N03h . 9H20
Ferric Sulfate 180 140280 200 210 140 140 140185 C C C CC C C C CBA A AC
Fe2(S04h
Ferrous Chloride Sat'd 185 180 140280400 200 180 200 C C C C C C C C C C C C C C C B
FeCb
Ferrous Sulfate 70 185 180 140280 400 200 180 140 160200 B C C B CCCCCAAAAA A
FeS04
Fluosilicic Acid 50% 73 140280 300 140 100 200 100 210 CCBB CCC CBBBAA A
H2SiF6
Formic Acid 73 73 73 250300 200 C 70 140 C C CB CCCBCAAAAA A
HCOOH
Freon 11 100% C 73 140 200 300 C 70 130 C 70 A A AAA B B B BAAAAA

Furfural C C 75 300 140 C 70 70 C A A A AA A A A AAAAAA A


C4H30CHO
Gasoline,Leaded C C 140275200 C 70 70 70 100 AAAAAAAAAAAAAAAAAA

Gasoline, Unleaded C C 140280200 C 70 70 100 AAAAAAAAAAAAAAAAAA


Tables 367

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
~ ::l
z ~ Z G
0 z
0 ,,~o ....l
Zl'l;:lui
~
Uzta U

Z
I'l
Z ~ ~
~o
~~~
o~
z~
Clui
ZCI~
>-
0
~ Z~I'l~Z ~::lz~~~ ~ ~~
Z Z Z 0
I'l
~ ::E ?i<I: ~ ~I'lNZ~ ~ o~~ .~>-....l I'l

~ g ::E~ZO::E~~G~~ ~~~~~~~
U U ~ 0
Z 0
:;j Uis: 8:: ~ ~ ~ 52 ~8~~~~~2u~~~~~ ~~
~
0 I'l
U I'l ~ Z ~

Glucose 180 185 180 140280 400 250 180200 160300 A A A A A A A A A A A A A A A A A A


CsH120S . H20
Heptane C 140280300 C 70 70 70 185 AAA AAAAAAAAAAAA
C7H16
Hydrazine C 200 250 70 70 70 C CCCCCCC C A AA
H2NNH2
B
Hydrobromic Acid 20% 73 120 140280 250 140 C 100 to 185 CCCCCCCCCCCCCCCCC
HBr 70
Hydrochloric Acid 35% C 210 150 140280 250 70 C 100 100 CCCCCCCCCCCCBCBCC
HCI
Hydrocyanic Acid 10% 73 140280250 200 70 200 185 CCCCCCCCCCCCABAACA
HCN
Hydrofluoric Acid Dilute 73 180 73 250 300 70 C 150 70 150 CCCCCCCCCCCCCCBACA
HF
Hydrogen Peroxide 50% 185 150 140150300 100 C 200 C 185 CCCCCCCCBCCAAAAA
H202
Hydrogen Sulfide Dry 185 150 140280 100 C 140 C 140 B B B B ABAA A
H2S
Inks 300 70 70 70 A AA CCC C A AA

Iodine 10% 73 150 150200 70 70 70 C 70 CCCCCCCC CCCCBA A


h
B
Kerosene to 185 73 140280 250 C 140 C 70 300 70 A A A A A A A A A A AAAAAA
70
Lactic Acid 25% 150 140125300 70 140 140 70 CCCCCBC BAAAAA
CH3CHOHCOOH
Lactic Acid 80% 150 73 125 300 70 C 140 70 CCCCCBC BAAAAA A
CH3CHOHCOOH
Lead Acetate Sat'd 185 180 140280 300 210 70 100 160 C C C CCC C A AA
Pb(C2H302h·3H20
Lime Slurry 100 100 160 100 AA A A AA
B
Linseed Oil 100 185 150 140280 300 to 180200 70 250 A A A A A A A A A A A A A A A A
70
Magnesium Carbonate 140280225 170 140140 140210 B B BBB BAAAAA
MgC0 2
368 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals eF)
and
Formula
N
~
~
Z
0 Z
0 s~i§
UN~
G
zGl::>'
z~
""zP:) o~ Q"l U
z z
Z Z
~
~ ~ ~~~
~~;P:)~
z~
~~z~~~
zQ~
~~~~
~
~ ~-~NZ-~ __ O-~

~
~
U ~x~~~~
U .... 0~ 0 <j0~~<j
0zU ~ ~~ ::>..:f:z ~~i§8~~~G~~~~~~3z~
~
~ ~ ~ P:) 2 ::> _0<O"1<j0
~

~ U
~
~ ~ z UP:)cn P:)'-'QU~z~;;:;~ ~
Magnesium Chloride Sat'd 185 180 140280 400 170 180200 160170 AAABBCCC CCCCBA A
MgCb
Magnesium Hydroxide Sat'd 185 180 140280 300 170 180200 160225 BCCBBAAA AAAAAAAA
Mg(OHh
Magnesium Nitrate 185 180 140280 300 140 70 140 160225 ACe B AAAAB
Mg(N03h . 2H20
Magnesium Sulfate 185 180 140280300 175 180140 160200 AAAAAAAAAAAAAAAAA
MgS0 4 ·7H2O
Malic Acid 185 150 140250 250 C 100 70 70 200 ABB CCC CAAAAA
COOHCH2eH (OH) COOH
Mercuric Chloride 140 180 140250 300 210 140 140 140 185 eeccccccCCCCCCBC A
HgCb
Mercuric Cyanide Sat'd 140250300 70 70 140 70 70 cccceccc C A AC
Hg(CNh
Mercury 185 150 140275 300 210 140 140 140 185 CCCCCAAA AAAAAB A
Hg
B
Methyl Acetate 100300 to C C C C B B BBB BBA AA A
CH3C02CH3 70
Methyl Acetone C 70 C C C AAAAAAAAAAAAAAAA
C3H60
Methyl Bromide e 280300 C 70 C C 185 C e B CCB B BB
CH3Br
Methyl Cello solve C 280 70 C 70 70 C AAB BBB AA AAAA
HOCH2CH2oCH3
Methyl Chloride Dry C e 280250 C C C C 70 C AACCAAAAAAAAAAAA
CH3Cl
Methyl Ethyl Keytone C C C C C 200 70 C C C C A A AAA A AA AAAAAAAA
(MEK)
CH3COC2Hs
Methylene Chloride C C 250 C C C C 70 B BB BBB AAAA A
CH2Cb
Molasses 73 140 150 300 100 150 150 150 185 AAAAAAAA AAAAAA A

Monochloroacetic Acid 50% 73 73 140150200 C 70 e C 70 CCCCCCCC CCCCBB


CH7CICOOH
Morpholine 75 200 70 C e e C B B BBB BBBBBB
C4HsONH
Tables 369

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
f!l ~
z
0 z
0
z
sf!lo
u z §3
~ G
-
Z~;:lui U

~ Z Z
Z
~
Z
~
~ ~
~ Z~~~~
~o
e~~ z~
o~
~~z~~~
~ui
Z~~
~~§~
:><
0

~
~
U
Z
~
~
~ :::E <1:
13
0 0
~~~§~~~6~~~~~~g~~~
""'""' ~ ~
~
~
0
U ;:j U ~ ~ p:)
~
Z ~ ~O~ ~ ;:l
UP:)U1 _8<O""~o
P:)"~u~z ... ;;;;:; :::E

B
Naphtha to 73 73 140280200 C 140 C C 150 A AAB AAAAAAAAAAAA
70
B
Naphthalene to C 200250 C C C C 170 A A AB AAA A AAAAAAA
ClOHs 70
Nickel Chloride Sat'd 185 180 140280 406 210 180200 160210 CCB CCC A AC A
NiCh
Nickel Nitrate Sat'd 140280 400 210 180 250 C C C CCC AAAAC
Ni(N03h·6H2O
Nickel Sulfate Sat'd 185 180 140280 400 210 200160300 ACCB CCC A
NiS04
Nicotinic Acid 140250 70 140 ABB CCC BBBBB
CSH4NCOOH
Nitric Acid 30% 180 120 140125250 70 C 100 C 185 CCCCCCCCC BA A
HN03
Nitric Acid 70% 73 C 73 C 250 C C C C 100 BCCCCCCCCC CA A
HN03
Nitric Acid 100% C C C 70 C C C C 70 BCCCCCCCCCCCA AC
HN03
Nitrobenzene 73 C 73 400 C C C 70 BB AAA A AA A
C6HsN02
B B
Oleic Acid 185 150 140250250 to 100 70 to 185 ABBA BBC BAAAA A
Cfu(CHv7CH:CH(CH2hCOOH 70 70
Oxalic Acid 50% 185 180 140 125 300 150 C 100100 C CCC CCCCCBAAAA A
HOOCCOOH . 2H20
Palmitic Acid 70% 73 180 73 250300 100 C C 185 ABBBABBB BBAAAA
CH3(CH2h4COOH
Perchloric Acid 10% 140 73 73 200250 70 C 70 70 70 C A AA
HCI04
Perchloroethylene 275200 C C C C 200 BB BBB BAAAAA A
ChC:CCh
Phenol C 73 73 73 125 70 C C C 200 AAC CCC CAAAAA A
C6HsOH
Phosphoric Acid 10% 210 180 140275 300 140 70 200 140200 C C C CCC C CC CCBAAAC
H3P04
370 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
z z
o 0
~ z z
~
~
~ z z ~ @
~
u 0 ;;E <i:
u P-.
z (fJ
5: u
~
Ii 0 z 0
0
o
u ~ u
P-.
P-. 5: ~ ~
P-.
~
::>
~ ::r::
~
z ~
Phosphoric Acid 50% 210 180 140275 300 70 C 200 70 200 C C C C C C C C C C C B A A A C
H3P0 4
Phosphorus Trichloride C 200300 C C C C A AA
PCh
Potassium Bicarbonate Sat'd 73 170 140200400 170 70 200 160200 A A A A
KHC03
Potassium Bromate 180 140275 400 180140 140250 CAA A A A
KBr03
Potassium Bromide 180 140280 400 170 180200 160200 BBB CCC A AA
KBr
Potassium Carbonate 70 180 140280400 170 180200 160200 B B B BB A A A A AA A A A AA
K2C03
B
Potassium Chlorate 180 140200 400 140 to 140 100 140 B B B AAAA AAAAAA
KCl03 (Aqueous) 70
Potassium Chloride 185 180 140280 400 210 180200 160200 B A A BB B B C B BBA AAA A
KCI
Potassium Chromate 140280 400 170 140 70 70 200 AAB BBB B AAAA
KzCr04
Potassium Cyanide 185 140280 400 140 180200 160 185 C C C C C C B B B B AA AAA A
KCN
Potassium Fluoride 140275 400 140 180 250 A AA
KF
B
Potassium Hydroxide 25% 185 150 140 150 400 210 to 140 160 140 C C C C B B B B AA AAA
KOH 70
Potassium Nitrate 140280 400 210 180 140 140250 A A A BB B B B B A A A A AA A
KN03
Potassium Permanganate 10% 150 140250 400 210 C 100 100 140 B B AAA A A A A AA
KMn04
Potassium Sulfate 180 140280200 210 140140 140250 A A A BB A A A A BA A A A AA A
K2S04
Potassium Sulfide 275300 100 70 100 C C C CC C C C B B B B B C A
KzS
Potassium Sulfite 300 140 70 70 200 BBB CCC A AB
KzS03·2H20
Tables 371

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
z z
0
0
~ ~ ~
zf-< z z
Z
0
Z
[l
u
~ ~
"" 9
u ::;s
~
u ~ p..,
z
0
(fJ

~
s: p..,
~ ~ ""~
Q
p..,
Z
:=>
co
0
~
0
:=>
.....l

""
U U p.., ~ Z

B
Propane 73 140 280 300 C 70 to 70 70 A A A A A A A A A A A A A A AA A
C3Hs 70
Propyl Alcohol 150 350 140 140 140 140250 A A A A A A A A A A A A A A AA A
CH 3CH2CH20H
Rosin 200 70 70 70 100 C C CCC CAAAAA

Silver Nitrate 70 185 180 140280 350 210 140 200 160250 C C C CC C C C CBA A C
AgN03
Soaps 70 185 73 140 400 210 180 140 140250 BBA BBB BAAAAB A

Sodium Acetate Sat'd 185 180 140280400 170 C 70 C AAB BBC BBA AA A
NaC2H302
Sodium Aluminate Sat'd 300 200 180 140 140200 CCB BBA B A AA A
NazAi203
Sodium Bicarbonate 70 185 180 140280 400 250 180200 160300 A A A BB A A C AA A A A AA A
NaHC03
Sodium Bisulfate 70 180 140280 200 180 100 140250 C C C CC C C C CBA AA
NaHS0 4
Sodium Bisulfite 185 180 140280 400 200 180200 140250 B B CCC C A AC A
NaHS0 3
Sodium Borate (Borax) Sat'd 73 300 140 70 100 100140 AA B B BAAAAA
Na2B407·lOH20
Sodium Carbonate 70 185 180 140280 400 140 140 140 140300 C A A BB A AA A A A A A ABA
Na2C0 3
B B B
Sodium Chlorate Sat'd 180 73 250 350 to to to 100 AAC BBB BBAAAC
NaCl03 140 70 140
Sodium Chloride 210 180 140280 350 140 140 100 160200 B A A A B B B B CAB B B B A A A A
NaCI
Sodium Chlorite 25% 73 C 250 200 C C 140 C
NaCl02
Sodium Chromate 200 70 70 70 70 AA BBB BAAAAA
Na2Cr04 . lOH20
Sodium Cyanide 185 180 140280350 140140140 140200 C C C CC A A A A AA A A A
NaCN
Sodium Fluoride 140 185 140280 350 140 70 140 70 140 AAB CCC A AA
NaF
372 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals eF)
and
Formula
~ ~
z
0 z
0
~ z
::s~o . . .:l B
uztiS Z~;:lui U

~ Z Z
Z
~

~
~
~
~
;:l
~o
~~~
~~;:l
o~
z~
Oui
~!il~Z~(FJ~~i5~~P2cri:Z::~.....:l
zo~
0 0 §§>-
~
~
U ../:
Z
0
u
(FJ
~ u~ 8: ~ ~
~

~
~
6:
~
Z
;:l
~
0
~
Z
0
g
~
~~i5§~~~§~~~~~~~~
~utiSoo~~~o ~z§~~ ~~ ~
B
Sodium Hydroxide 15% 70 185 180 140170400 210 140200 160 to C A A AA BAAAAAAA
NaOH 100
B
Sodium Hydroxide 30% 70 210 180 140 73 350 210 100 140 160 to C A B BB BAAAAA
NaOH 100
(Caustic Soda)
Sodium Hydroxide 50% 70 210 180 100 C 350 180 C 140 160 C C B B C C C B B B B B A A A A A
NaOH
Sodium Hydroxide 70% 70 180 100 C 350 70 C 100 100 C C B C C C C B B B B B A A A A A
NaOH
Sodium Hypochlorite C 185120 73 200350 70 C 150 C 140 C C C C C C C C C C C C A AA A
NaOCI· 5H20
Sodium Nitrate Sat'd 185 180 140280 400 210 140 140 140225 B A A BB A AA AAAA A AABA
NaN03
Sodium Perborate 73 140 350 70 70 70 70 70 C C BBB AAAAA A
NaB02·3H20
B
Sodium Peroxide 140 200 250 140 to 200 70 185 B C C CC C C C AAAAA B
N~02 70
Sodium Silicate 180 280 200 140200 140200 CCB AAA AAAAAAAA
2Na20· Si02
Sodium Sulfate Sat'd 70 185 150 140280 400 140 140 140 140200 A A A BB A A A A AA A A A AA A
Na2S04
Sodium Sulfide 70 185 150 140 280 350 140 180 200 140 200 C C C CC B B C B BAA A A A
Na2S
Sodium Sulfite 70 185 180 140280 350 140 140 140 140200 AAC BBB BBAAAC A
Na2S03
Sodium Thiosulphate 150 140280 350 200 140200 160 200 A B BC C CC C A AA
Na2S203·5H20
Starch 140200300170180200 160200 B B BB B B B BAA A AA

Stearic Acid 185 73 140275 350 C 140 70 70 100 A A A CB C C C B CAA A A A A


CH3(CH2h6COOH
Sugar 275 350 140 100 140 140200 CC BC BAAAAA
C6H1206
Sulfur C 140250350 C 70 70 250 C C C C C C B B C B B B A AA A
S
Tables 373

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals ("F)
and
Formula
~
z
0 z ~ Z
::s~o ....l
G
~
0 u z §3 Z~ ~ui U

~ Z
~
Z ~ ~
~
il!lo
~~~
~~~ 0
o~
z~
Clui
ZClOO
ZO ~ 0
§>-
Z Z Z
~ ~ ~~~z~ ~~o~3~ .~~~§~
<
~
~

~~~~~~~~~~~;~~~~~
U

~ '~"' @ 0...
~ 0
Z
0
U
00
~
~
U
0...
0...
Cl
~
Z
~
~
0
~
Z ~
Sulfur Chloride C 73 350 C C 70 C 70 CCCCCCCCCCCCCCBCC
S2Ch
Sulfur Dioxide Dry C 73 140175350 70 C 200 C 100 A A A BAA A A A AAAAAAAA
S02
Sulfur Dioxide Wet C 73 73 150 140 C 200 140 BCCBBC CACAA A
S02
B
Sulfuric Acid Up to to 210 180 140250 250 140 C 100 100200 CCCCCCCCCCCABAACA
H2SO4 30% 100
Sulfuric Acid 50% C 210 150 140250 250 140 140 150 C 200 CCCCCCCCCCCACAACA
H2SO4
Sulfuric Acid 70% C 210 120 140200 200 140 C 150 C 250 CCCCCCCCCCCCBCACCA
H2SO4
B
Sulfuric Acid 100% C C C C C
to C C C C C CCCCCCCCCCCCBCBCC
H2SO4 200
Sulfurous Acid 150 C 230350 C 150 C C CCCCCCCCCCBAAAACA
H2S03
Tannic Acid 10% C 185 180 140225 250 70 100 100 100 100 AA BBCBBBAAAA A
C7ili52046
Tartaric Acid 150 140250250 C 70200 70 70 BAACC CCC CCAAAAA A
HOOC(CHOm2COOH
Toluene (foluol) C C C C 175 200 C C C 70 AAAAAAAAAA AAAAAAA
Ch3Cili5
Trichloroethylene C C C C 280200 C C C C 185 AAAAAABBB AAAAAAA
CHCI:CCh
Turpentine 73 C 140280 C 70 C C 150 AAAAAAAAAA AAAAAAA

Urea 185 180 140250 210 140 140 140 185 C BB C C C AA BC


CO (NH2h
Varnish 250350 C 70 C 70 AAABBCCC BAAAA A

Vegetable Oil 185 73 140275300 C 70 70 70 200 A AA AA AA AAAA

Vinegar 73 150 140 140225 300 140 C 200 70 C CBCCCCCCC AAAAA A

Vinyl Acetate C C 250 350 70 70 C C C BB BBB A AB


CH3COOCH:CH2
374 Appendix B

Table B-4 Chemical Resistance Guide for Valves and Fittings (continued)

Plastics and Elastomers at


Maximum Temperature Metals
Chemicals (OF)
and
Formula
z z
0
0
~ z z
i.Ll
~
~
i.Ll z z 0 gz ~
u :3 Ci .J:Z
:g
'" ~
Z u Po. 0
6: u
'"~ Po. p:)~
0
z B
rfJ
a
~
Po.
6: ~ i.Ll
U U Po. ~
'"
Water, Deionized 70 210 180 140280 400 200 70 160 AABBCCCCC CBAAAA A
H20
Water, Salt 70 210 180 140280 400 250 180200 160 CBBBCCCCCBCBAAAABA
H20
Water, Sea 70 210 73 140280 400 250 180200 160 CBBBCCCCCBCBBAAACA
H20
Whiskey 185 150 140200350 200 140140 140140 A C C B C C C CBA AA A

Wine 185 150 140200350 170140140 140140 C C CCC CBA AA

Xylene (Xylol) C C C C 200 350 C C C C 150 A A A A A A A A A A A A A A A A A A


C6H4(CH3h
Zinc Chloride 185 180 140280 400 180 70 200 160200 CCC CCC CCBBAA
ZnCh
Zinc Sulfate 185 180 140280400 180 140200 140200 A C CB C C C B CAA AAA A
ZnS04·7H20

NIBCO INC., its marketing companies and distributors, and the authors of and contributors to this table specifically deny any
warranty, expressed or implied, for the accuracy and/or reliability of the fitness for any particular use of information contained
herein.

MATERIAL RATING FOR THERMOPlASTICS & ElASTOMERS


Temp. in OF UN' rating, maximum temperature which material is recommended, resistant under normal conditions
B to Temp. in OF Conditional resistance, consult factory
C Not recommended
Blank No data available

MATERIAL RATINGS FOR METALS

A Recommended, resistant under normal conditions


B Conditional, consult factory
C Not recommended
Blank No data available
Bibliography

[1) Yedidiah. S.: Effect of Scale and Speed on Cavitation in Centrifugal Pumps. ASME Symposium on
Fluid Mechanics in the Petroleum IndustIy. pp. 61-69. Houston. Tx. December 1973.
[2] Acosta. A. J. and Bowerman. R. D.: An Experimental Study of Centrifugal Pump Impellers. Trans.
ASME. 81: pp 1821-1839. 1957.
[3) Yedidiah. S.: A Study of Suction Specific Speed. ASME Cavitation Forum. pp. 32-34. Chicago Ill.
1967.
[4] Yedidiah. S.: Some Observations Relating to Suction Performance of Inducers and Pumps. Trans.
ASME. Basic Engng .. pp. 567-574. September 1972.
[5) Chanaud. R. C.: Measurements of Mean Flow Velocity Beyond a Rotating Disc. ASME Paper No. 70-
FE-C. 1970.
[6) Bennet, T. P. and Worster. R. C.: The Friction on Rotating Discs and the Effect on Net Radial Flow
and Externally Applied Whirl. BHRA Publication No. RR-691. 1961.
[7] Yedidiah. S.: Effect of Energy Losses on the Head Developed by a Rotodynamic Pump. ASME
Pumping Machinery Symposium. FED Vol. 81. pp. 181-186. La Jolla. Ca. 1989.
[8] Yedidiah. S.: ReCirculation in Centrifugal Pumps. AIAA Publication No. 86-1124. Atlanta. Ga. 1986.
[9) Yedidiah. S.: Cause and Effect of Recirculation in Centrifugal Pumps. Parts I and II. World Pumps.
pp. 267-295. September 1985.
[l0] Yedidiah. S.: A Study of Recirculation at the Inlet of Centrifugal Pumps. Joint ASME--JSME
Conference on Thermal Engineering. Hawaii. 1987.
[11] Worster. R. C. and Thorne. E. W.: Roughness and Friction Effects on Performance in Centrifugal
Pumps. BHRA Publication No. SP-564. 1957.
[12) Varley. F. A.: Effect of Impeller Design and Surface Roughness on the Performance of Centrifugal
Pumps. Proc. lnst. Mech. Engng.. 175(21): pp. 955-989.1961.
[13) Ishida. M. and Senoo. Y. : On the Pressure Losses due to the Tip-Clearance of Centrifugal Pumps.
ASME Paper No. 80-GT-139. 1980.
[14] Yedidiah. S.: Some Causes of Unstable Performance Characteristics of Centrifugal Pumps. 17th
International Gas Turbine Conference (an ASME Paper). pp. 5-14. San FranCisco. Ca. 1972.
[15] Yedidiah. S.: Certain Unexplained Phenomena. Observed in Centrifugal Pumps. ASME Paper No. 85-
FE-3. Albuquerque. NM. 1985.
[16] Yedidiah. S.: The Recirculation Theory of Regenerative Pumps. ASME FED. Vol. 154. Pumping
Machinery. pp. 355-358. Washington. DC. 1993.
[17) Yedidiah. S.: Cause of a Dip in the QH Curve of a Centrifugal Pump. ASME Symposium on Unsteady
Flow. Anaheim. Ca. 1986.
[18) Yedidiah. S.: A Possible Cause of Surge in NPSH-Requirements of Centrifugal Pumps. ASME
Cavitation and Multiphase Forum. FED. Vol. 36. pp. 39-41. Atlanta. Ga. 1986.
[19) Yedidiah. S.: Certain Effects of Recirculation on Cavitation in Centrifugal Pumps. Proc. Inst. Mech.
Engng. Vol. 200(A4). pp. 283-292. 1986.
[20) Yedidiah. S.: Effect of Impeller-Inlet GeometIy on the Intensity of Recirculation and on Cavitation in
a Centrifugal Pump. ASME Cavitation and Multiphase Forum. FED. Vol. 153. pp. 197-203.
Washington. DC. 1993.

375
376 Bibliography

[21] Peck. J. F.: Investigations Concerning Flow Conditions on a 6-inch Experimental Centrifugal Pump.
British Hydromechanics Research Association (BHRA) Publications. National Engineering Laboratory.
East Kilbride. Glasgow. TN-12. 1949.
[22] Schweiger. F.: Stability of the Centrifugal Pump Characteristics at Part Capacity. International
Conference on Pump and Turbine Design. NEL. Paper No. 3-3. September 1976.
[23] Paulon. J .. Fradin C. and Poulain. J.: Improvement of Pump Performance at Off Design Conditions.
ASME Paper No. 85-GT-200. Houston. Tx. 1985.
[24] Guiton. P.: Actual Behaviour of Pumps Outside their High Efficiency Range. Von Karman Institute
for Fluid Dynamics. Lecture Series 1978-3.
[25] Zanker. K. J.: Experiments with Back Vanes used for Balancing Axial Thrust in Centrifugal Pump
Impellers. BHRA Publication No. RR-729. April 1962.
[26] Zanker. K. J.: Axial Thrust in Centrifugal Pumps. BHRA Publication No. RR-746. November 1962.
[27] Stepanoff. A.: Centrifugal and Axial Flow Pumps. John Wiley & Sons. New York. 1957.
[28] Worster. R. C.: Flow in the Volute of a Centrifugal Pump and Radial Forces in the Impeller. BHRA
Publication No. RR-543. 1956.
[29] Agostinelli. A.. Nobles. D .. Mockridge. C. R. et al.: An Experimental Investigation of Radial Thrust in
Centrifugal Pumps. ASME-Paper No. 59-HYD-2. 1959.
[30] Osterlei. R. E.: Motor Efficiency Test Methods-Apples and Oranges. Power Transmission Design.
May 1980.
[31] Pfleiderer. C.: Die Kreiselpumpen. 3rd edn .. Springer-Verlag. p. 314. 1949 (in German).
[32] Taylor. I.: Two Pump Applications Need Extra NPSH Available. ASME Polyphase Forum. pp. 38-41.
San Francisco. Ca. 1972.
[33] Yedidiah. S.: Radial Thrust in Centrifugal Volute Pumps. The author's private notes. 1967.
[34] Yedidiah. S.: Factors Affecting the Suction Performance of Centrifugal Pumps. ASME Symposium on
Fluid Mechanics in the Petroleum Industry. pp. 53-60. Houston. Tx. December 1975.
[35] Yedidiah. S.: Alternate Vane Cavitation in an Impeller. ASME Cavitation Forum. pp. 12-13. Atlanta.
Ga. 1973.
[36] Yedidiah. S.: Some Observations Relating to Suction Performance of Inducers and Pumps. Basic
Engng .. pp. 567-574. September 1973.
[37] Minami. Sungo. Kyai Kawaguchi and Tetsou Homma: Experimental Study of Cavitation in
Centrifugal Pump Impeller. J. Soc. Mech. Engng. Vol. 3(9): pp. 19-28. 1960.
[38] Yedidiah. S.: Oscillation at Low NPSH. Caused by Flow Conditions in the Suction Pipe. ASME
Cavitation Forum. pp. 27-28. Montreal. Canada. 1974.
[39] Knapp. R. T. et al.: Cavitation. McGraw-Hill. New York. 1970.
[40] Rees. R. P. and Trevence. D. H.: The Effects of Temperature and Viscosity on the Critical Tension of
Liquids. ASME Cavitation Forum. p. 1. Chicago. Ill .. 1967.
[41] Yedidiah. S.: Effect of a Sharp Edge on the Appearance of Vapor Bubbles in a Flowing Liquid. ASME
Cavitation and Multiphase Forum. FED Vol. 194. pp. 101-103. Lake Tahoe. Nv .. 1994.
[42] Denny. D. F.: Vortex Formation in Pump Sumps. BHRA Publication No. SP-436. 19??
[43] Campbell. J. M.: Development of a Pipe Bend having Good Outlet Velocity Distribution. and the
Effect of Subsequent Contractions. BHRA Publication No. RR-658. 1960.
[44] Levi. E.: A Universal Strouhal Law. Joint ASME-ASCE Mechanics Conference. Boulder. CO. 1981.
[45] Yedidiah. S.: Effect of a Sharp Edge on the Appearance of Vapor Bubbles in a Flowing LiqUid. ASME
Cavitation and Multiphase Flow Forum. 1994. FED Vol. 194. pp. 101-103. Lake Tahoe. Nv .. 1994.
[46] Yedidiah. S.: A Study of Suction Specific Speed. ASME Cavitation Forum. 1967. pp. 32-35. Chicago.
Ill.. 1967.
[47] Hammitt. F. G.: Observation of Cavitation Scale and Thermodynamic Effects in Stationary and
Rotating Components. Journal of Basic Engineering. Trans ASME Series D. Vol. 85. pp. 1-16. 1963.
[48] Jekat. W. K.: Reynolds Number and Incidence Angle Effects on Inducer Cavitation. ASME Paper No.
66-WA/FE-31. 1966.
[49] Yedidiah. S.: Effect of Impeller Width on the Suction Capability of Centrifugal Pumps. ASME
Cavitation and Multiphase Forum. 1988.
Bibliography 377

[50] Worster, R. C.: The Flow in Volutes and its Effect on Centrifugal Pump Performance, Proc. Inst.
Mech. Engng. Vol. 77(31), pp. 843-876, 1963.
[51] Anderson, H. H.: Centrifugal Pumps, 3rd ed., Trade & Technical Press, 1980.
[52] Copley, D. M. and Worster, R. C.: Pressure Measurements at the Blade Tips of a Centrifugal Pump
Impeller and the Effects of Tip Profile on Pump Performance, BHRA Publication No. RR-71O, 1961.
[53] Yedidiah, S.: Effect of Blade-Geometry on the Head Developed by a Rotodynamic Impeller, Fluid
Machinery Forum, ASME Summer Meeting, FED Vol. 222, pp. 25-34, Hilton Head, S.C., 1995.
[54] Yedidiah, S.: The Theoretical Head-Capacity Curve of a Centrifugal Impeller, presented at the Fluid
Machinery Forum, ASME Summer meeting, Hilton Head, S.C., 1995.
[55] Yedidiah, S.: Calculation of Head Developed by a Centrifugal Impeller, ASME Paper No. 89-FE-9, San
Diego, Ca., 1989.
[56] Yedidiah, S.: About the Validity of a Slip-Factor for Predicting the Head of a Centrifugal Pump, FED
Vol. 119, pp. 7-9, Portland, Or., 1991.
[57] Yedidiah, S.: An Alternate Method for Calculating the Head Developed by a Centrifugal Impeller, FED
Vol. 107, pp. 131-138, Portland, Or., 1991.
[58[ Saalfield, K.: Einige neuere Gedanken zur Laufradberechnung von radialen und halbaxialen
Kreiselpumpen, KSB Teechnische Berichte 11, August 1966 (in German).
[59] Yedidiah, S.: A Correlation between Aerofoil Theory and Euler's Equation for Calculating the Head of
a Constant Pitch Axial-Flow Inducer, Proc. Inst. Mech. Engng., Vol. 205(C5), pp. 357-363, 1987.
[60] Yedidiah, S.: A Study of Application of the Aerofoil Theory for Calculating the Head Developed by an
Axial-Flow Impeller, Proc. CSME Engineering Forum, Vol. 1, pp. 25-29, Toronto, Canada, 1990.
[61] Varghese, G., Mohana Kumas, T. C., Rao, Y. V. N. et al.: Influence of Surface Roughness on the
Performance of Centrifugal Pumps, ASME Joint Applied Mechanics, Fluid Engineering and
Bioengineering Conference, Paper 77-FE-8, New Haven, Connecticut, 1977.
[62] Worster, R. C.: The Effects of Skin Friction and Roughness on the Losses in Centrifugal Pump
Volutes, BHRA Publication No. RR-557, 1957.
[63] Myles, D. J.: An Analysis of Impeller and Volute Losses in Centrifugal Fans, Proc. Inst. Mech.
Engng., Vol. 184, Pt. 1, No. 14, pp. 253-279, 1970.
[64] Yedidiah, S.: Beware of Pitfalls in Testing of Centrifugal Pumps, POWER. pp. 85-87, September
1986.
[65] Yedidiah, S.: Centrifugal Pumps, Problems and Cures, Pennwell Books, Tulsa, Ok., 1980.
[66] Janigro, A. and Ferrini, F.: Inducer Pumps, Von Karman Institute For Fluid Dynamics, Lecture
Series 61 (3 lectures), 1973.
[67] Yedidiah, S.: Approximate Method for Calculating the Head Developed by an Impeller with a Finite
Number of Blades, ASME Paper No. 69-FE-8, 1969.
Index

A B
Abrasion 73 Back vanes 10 1. 328-330
by metallic contact 74. 238. 274 Balancing device 99-104
by packing 221. 238 balancing holes 100. 139
by pumped liquid 237-238 Bearings. function and problems 209
resistance of materials to 73 sliding
Air advantages and disadvantages 209-210.
appearance 31 273
effects of 9. 31. 32. 132. 134-136. 158. lubrication 210
193. 196-197. 278 problems with 210
funnel 163-165. 167 rolling
handling capability of impellers 32. 204 advantages and disadvantages 211
in column 250 failures and their causes 216-219
sources of 31 handling 214-216
Air leakage lubrication 211-214
detection 139. 202 some special problems 219
due to bent shaft 207. 274. 290 Belt drive 120. 208
due to parallel operation 185-186 Bernoulli's equation 164
due to prerotation 168 Bernouli's equation for a rotating system 10-11
due to other sources 202-204 Blow
due to vortices 167 due to cavitation 231. 155-156
through mechanical seal 139-140. 202 due to compressed air 201-202. 206. 250
through stuffing box 139. 202-203 effects of 206-207
Air pockets 127. 128. 194-200 due to hard object 206
effects on check valve 201-202 due to prerotation 154-156
in discharge line 200-202 due to water hammer 178
in donut pumps 200 Boiling 33
in pumps proper 198-200 Borehole pumps (see Deep well pumps)
in suction lines 127-128. 136. 194-195 Bypass. uses of 183. 241
prevention of 195-196
in suction nozzle 199
reduction of 194 c
Air. solubility in liquids 141. 193
Air valve 126. 181 Careless starting 206
Alignment. checking for 290-292 Casing
Analysis of preliminary information 273-280 diffuser type 24. 260. 318
Axial thrust 97-99 effects of geometry 318-322
balancing split 23
multi-stage pumps 101-104 volute 20-21.317
single-stage pumps 99-101 Cavitation. definition of 8-9. 33-34
on closed impellers 97-98 alternate vane 150-151
on semi-open impellers 98-99 and specific speed 38. 51

379
380 Index

Cavitation. caused by: loud blow is heard each time the pump is
cutdown of impeller 37-38 started or stopped 357
heating of parts 205-208 mechanical seal leaks excessively 354
inertia 179-181 mechanical seal has short life 354
prerotation 143. 154-156 packing has short life 353
vibrations 141. 151-153 pump does not develop any head. nor does it
vortices 157-159 deliver liquid 348
water hammer 155. 179-181 pump develops some pressure. but does not
Cavitation. effects of deliver liquid 348
balancing holes 265-269 pump delivers less liquid than expected
discharge line 153 349
impeller cutdown 37-38 pump does not develop enough presure
impeller width 315 349
impurities 160 pump consumes too much power 350
sharp edge 161 pump does not perform satisfactorily.
suction line 157 although nothing appears to be wrong
temperature 159-160 with the pump or with the system 350
valve 152-153 pump operates satisfactorily during start.
Cavitation. external signs of 34. 279 but performance deteriorates shortly
Cavitation. cushioning 158 afterwards 351
Cavitation. effects on pump is operating with noise and vibrations.
performance 147-151 or both 351
wetted parts 237. 285. 287 pump overheats and/or seizes 356
Cavitation. graphical presentations pump develops cavitation under increased
at constant flow rate 41 NPSH 357
at different flows 41 shape of head-capacity curve differs from
at design-flow 47-54 rated curve 349
Cavitation Number stuffing box leaks excessively 352-353
dimensionless 40 Check valve. effect of air pockets 201. 250
Thoma's 37 Chemical compatibility of materials 362-374
Cavitation. occurence Centrifugal pumps. classifications:
at impeller inlet 148-150 by application 19
at impeller outlet 180 deep well 24. 25
at volute (diffusor)-throat 56-58 high temperatures 20. 24. 76
in istruments and their connections 129 irrigation 14
in pumps operating in parallel 186 sewage-disposal 25
in pumps operating in series 185 by design
in special cases 151-153.263-266 diffusor-pumps 24
Cavitation. prevention 181 multi stage 19. 23-26
Cavitation. remedial means 181. 300-302 single-stage 19. 20. 23
Cavitation. resistance of materials 303 volute pumps 20-22
Cavitation. at increased available NPSH by structure
263-265 centerline-discharge 20
Cavitation. at elevated temperatures 159-160 closed-impeller 25
Chain reactions 75. 117-118. 190 double-suction 23
Checklist. of most common problems 347-348 end-suction 20-21
Checklists of most common causes horizontal 21. 23-25
bearings overheat 354-355 open-impeller 25
bearings operate with noise 355-356 side-suction 21
bearings have short life 354-355 single-suction 20-21
casing bursts when the pump is started or split-casing 23
stopped 357 vertical 22. 25
flow rate periodically decreases. or by specific speed 26-28
even stops. then returns to normal Centrifugal pumps. principles of operation 6-8
357 Choice of pumps. effects of 14-17.55
gaskets leak during pump operation 357 Clearances. checking of 243-246
impeller and/or casing has short life 356 Clearances. closed impeller. effects of 67-70
Index 381

Clearances, open impeller, effects of 71-72 Dip in performance curve 87, 113-114
Closed test loop 133, 138, 153 Disk-friction 61-65
Column Discharge line
effects of air 250 air pockets in 201
effects of faulty assembly 250 cavitation in 179-180
effects of faulty machining 250 water hammer 177-179
oil lubricated 251 Discharge nozzle, effect on performance 91
water lubricated 251-253 Donut pumps 258-260
Compression, fracture due to 234 air in 200
Corrosion due to balancing axial thrust 101-104, 259-260
cavitation 237, 285 Drives
chemical affinity 234, 235, 285, 362-374 belt 120, 208
crevice 235 vertical lineshaft 248-249
elevated temperatures 236-237 Drooping curve, appearance and causes 92-95,
electrolytic 235 111-113
intermittend operation 237 effect of discharge nozzle 91
reduced NPSH 237 interaction with system 111-113
Corrosion, effects of pumps operating in parallel 185-188
discontinuities in surface 285 Dynamometer 121
velocity of flow 236
stabilisers 236-237
Critical speed 274, 300
Curve, drooping E
causes of 90-95, 111-113
definition of 14-15 Effects of
effects of 186-188 air pockets 127-128, 194-198
Curves, performance blow with a hard object 206-207
dip in 87, 113-114 blow against a ductile surface 234
as a diagnostic tool 273-278 careless starting 206
steepness of 15 choice of pumping unit 14-18
Curves, system 16,44 improper handling and workmanship 239
Cutdown of impeller-diameter misalignment 207-208, 290-292
with parallel edge 307-310 position of impeller 208, 258-259
with inclined edge 310-312 time 73-74, 134-136, 159, 240-241
Efficiency, definition of 13
shape of curve 17
effect of scale and specific speed 26-28
D effects of output 27
importance of 17
Damage Elbow, effect on flow 255
classification of 233-241 Electrical connections
effects of human factors 239, 283-285 checking 121
Deep well pumps, description 25, 247 safety precautions 290
Deep well pumps, problems caused by Energy, frictional losses
extension of lineshaft 247-248 in casing 65-66
in floating vessels 262 in ducts 61
installation 250 in impeller 61-65
sand 249-250 in rotating disk 61-65
well-geometry 250 Energy transfer 3-6
Deep well pumps, problems related to EqUilibrium, stable 187
all types 247-250 Erosion 234-238 (see also Corrosion)
oil lubrication 251 Errors in measuring of
semi-open impellers 245-246, 253-254 flow rate 121-122
variations in water level 248-249 head 122-130
water lubrication 251-253 power consumption 119-121
Diffusor 260, 318 NPSH 131-132
installation of 258-260 Extrusion 206, 234
382 Index

F flat 14
steep 14
Failed parts. visual inspection of 281-287 Head developed by an impeller 6. 164.334-341
bearings 281 Head increases with flow 277-278
casing 285-287 Head. measurements 122-130
impeller 285. 287 Head-NPSH curves 41-43
seals and packing 221-229. 281-282 Head. reduced 158
shaft 282. 300 Head. shutoff 91-94. 188
wearing rings 282-285 Head. Static 15-16
Failure of parts Head. Total 10. 122. 123. 164
time effects on 134-136. 159. 240-241 Heating of parts
human factors 239 causes 75. 205-208
Faults in effects 75. 190. 293.355-356
assembly 205-208.247.250 Hot liquids 75-76
layout 163-176 Human factors 239. 283-285
Field procedures 289-295 Hydraulic sources of noise 229-231
Flow meter Hysteresis due to recirculation 85-88
cavitation in 122
Flow rate
fluctuationsin 178.187.237 I
increases with head 14-15.92-95.
111-113 Impeller
measurements of 121-122 air-handling capability 32. 204
Flow ratio 42 closed 25
Flow for handling sewage 25
obstruction to 123-124. 136. semi-open 25
resistance to fluctuations 187 variations with specific speed 27
straightener 176-177 Impeller cutdowns. effect on
Foundation. setting of 240-241 head 38. 306-312
Fracture due to NPSH-requirements 38.57-58
bending 234 Impeller. effects of
cerelessnes 206 roughness 65
compression 234 width 313-315
extrussions and dents 207. 234 Impeller
fatigue 234 adjusting clearances 243-245
galling 74. 238-239 checking clearances 244-247
shear 234 inspection. on site 285. 290
tension 233 safety precautions. prior to on-site inspection
Frequency. natural 293 290
Impellers semi-open. problems with
in deep-well pumps 245-246.253-254
G in multi-stage pumps 244-247
in single-stage pumps 243-244
Galling. definition of 74. 238-239
Inducer
Gland. tightening of 238
effect of radial clearance 327-328
Gas (see Air)
effect of inlet-tip shape 328
Gas. solubility in liquids 141. 193
effect on NPSH-requirements 84
Gasket. obstruction to flow 123-124. 195-196
effect on performance 83-86
Gradual reduction of flow 178-182. 249-250
effect on recirculation 84
Graphs. performance
Inspection. visual of
as a diagnostic tool 273-278
bearings 209-219. 281
Grease. for different applications 212. 260
casing 285-287
impeler 285-287
H seals and packings 221-224.281-282
shaft 282
Head-capacity curve wearing rings 282-285
drooping 14-15.92-95.111-113 Inspection. on-site
Index 383

before the pump is started 289-292 Misalignment 190, 207-208


during operation 292-294 Model laws
precautionary steps 290 for testing 130
when dismantling the unit 294-295 for pump-sumps 175-176
Instabilities Multistage pumps 19, 23-24, 246-247
in suction line 155-156
at low, partial flow-rates 86-88, 155-156,
241 N
in double-suction pumps 232, 255-256
Intermittend operation 237 Natural frequency 293
Noise
description 229-230
L diagnosing the sources of 293
dry, cracking 279
Leakage of air 139, 167-168, 185-186, from combined sources 231-233
202-205, 224 hydraulic sources of 229-231
Leakage of liquid 202 mechanical sources of 205-208, 231
through open impeller 71-72 sources of, checklist 351-352
through wearing rings 66-70 special cases 233
Leakage, locations 202 NPSH, available 35
Leakage, effect on NPSH, definition 34
pump output 69-70 NPSH, effects of
suction performance 159 air (gas) 134-136, 141, 150, 158
Lineshaft, extension 247-248 blade-tip shape 46, 328
Liquids choice of pump 55
effects on suction performance 141, design flow 53
159-160, 193 design speed 53
thermodynamic properties 159-160, 359 different liquids 141, 160
Losses of energy 61-73 discharge line 137-138
Loop, closed 133, 138 flow-rate 55-56, 146-150
Loop for testing pumps 126, 131-133 impurities in liquid 160
Lubrication of ball bearings 211 leakage 70-71, 159
Lubrication, grease 212-213 prerotation 143
for special duties 212, 260 scale 54
Lubrication: Grease vs, oil 212-214 suction line 136-137, 157
Lubrication of journal bearing 210 temperature 141, 159-160
Lubrication: oil 213-214 time 134-136, 159
vibrations 141, 151-153
NPSH-Flow curves 42
M for a given application 145-150
NPSH-Head curves 35,41, 146-152
Materials, resistance to NPSH, Required 35
abrasion and wear 73 at other than tested flow rates 55-56
cavitation 73 at other than tested speeds 55, 306-307
chemicals 74, 362-374 sudden jump in 268-269
Measurements of at very high flow rates 56-60
flow 121-122 within normal range of flow-rates 55
head 122-130 NPSH-tests 131
errors in 123, 125, 127-129 NPSH, test-loops 131-133
NPSH 123-124, 131-132, 142-143 NPSH, variable 132
power 119-121
suction head 123-124
Mechanical seals o
features of 224
Mechanical seals, problems with 222-223 Oil, lubricating
causes of problems 223 handling 214
diagnostic procedures 223-228 suitability 213-214
384 Index

On-site inspection 289-295 increasing clearance between inducer and


Operation at low flow-rates 205, 241 casing 327-328
Operating conditions, changes in 241 underfiling inducer inlet-tips 328
choice of pump for given conditions 14-17 PH, effects on corrosion 236-237
Operating speed, effects of 180, 305-307 Pipe diameter, actual vs.nominal 129-130
Operation in parallel 185-188 Pipelines, stresses imposed by 188-189, 241,
Operation in series 185 274, 294
Output Power
effects of air or gas 278, 357 consumed at different flows 13-14
decreases gradually, after start 278, 351 consumed by thrust bearing 120
low, combined with low power consumption consumed by belt drive 120
275 consumed by column shaft 120
low, at start only 194-212, 357 consumed by geared head 120
means of modifying 323-330 definition of 4-5, 118-119
periodical fluctuations of 278-279, 357 measurements 119-121
reduced 349 transfer of 3-6.334-341
stopped completely 348 unchanged at reduced output 277
Overheating of parts 75, 205-208 Prerotation 80-83
effects on head-measurements 123-124
effects of flow on appearance 82
p effect of impeller-geometry 154-155
effects on NPSH 88
Packing (see Stuffing boxes) effects on shape of QH-curve 83-85
Packing, for different applications 361 problems caused by 85-88, 154-156
Packing, handling and troubleshooting 221-222 Pressure 9
Performance at blade-tips 326
effects of careless starting 206 difference between both faces of a blade 10,
effects of casing-geometry 317-322 34, 47-48
effects of handling 207-208 increases with flow rate 277
effect of impeller-cutdown 307-312 local drop in 33. 48
effect of position of impeller 208 pulsations, frequency 229-230
effect of reduced impeller-width 313-317 slam 183-184
effects of speed 180,305-307 prevention 184-185
effects of time 73-74, 159,240-241 vapor 33, 359
effect of vanes in suctioin bell 269 Priming 9
effect of volute-geometry 322 Pump, classification
effect of volute-tongue 320-321 deep-well 25
effects of workmanship 205, 207 double-suction 23
effects of a drooping curve 186-188 end-suction 20
effects of suction line 157, 190-191 for high temperatures 20
Performance, in series 185 for hot liquids 76
Performance, in parallel 185-188 for low NPSH 84,204,300-301
Performance curves mUltistage 23, 24
applications 14-17 with open impeller 25
as a diagnostic tool 273-278 sewage handling 25
effect ofleakage 67-72 Single-stage 20-22
for different NPSH 43, 44 split-casing 23
breakoff 278 wet pit 25
factors affecting 273-278 Pumps. description 3, 7
unstable, and their causes 85-88, principles of operation 3-8, 13, 64, 193.
155-156, 232, 255-256 334-341
Performance, means of altering principal parts 6-8
adding back-vanes to shrouds 328-330 Pumps, problems with
blade-tips, overfiling 325-327 deep-well (see Deep-well pumps)
blade-tips, underfiling 323-325 dry pit 261-262, 283-285
changing end-clearance of semi-open due to human factors 239, 256-258,
impellers 71-72, 327 283-285
Index 385

effect of elbow in suction line 254-256


floating vessels 262
s
multistage pumps 10 1-104, 258-259 Sand, in columns 249-250
operating in parallel 185-188 Seal, mechanical (see Mechanical seal)
operating in series 185 Semi-open impellers (see Impellers, semi-open)
propeller pumps 183 Shaft, assembly 250
split-casing 232, 254-258 Shaft, extension 247
special problems 263-268 Shear-fracture signs of 234
Pumps, care of Shock
after shutoff 285 due to cavitation 154-156
during starting 206 due to entrapped air 201
in harsh climatic conditions 260 thermal 76
Pump-system interaction 43-44, 185-188, due to water hammer 177-178
293-294 Shutoff head
Pumping action, principles of 6-8, 13, 193 testing at 123-124
Pumping hot liquids 75-76 Similarity laws for
Pumping viscous liquids 76-78 pump-sumps 175-176
Pump-sumps testing 130
design 166-175 Slam pressure 183-184
model laws 175-176 prevention 184-185
troubleshooting 174-175 Special cases 263-268
Pumping systems Specific speed 26
layout 183-191 effect on design 27
problems with 172, 176-178 Stabilisers 236-237
remedial means 174-176,178,182-185 Stuffing box
Pump selection handling and troubleshooting 139-140,
for given operating conditions 14 221, 238, 282, 352-353
for low NPSH 55 packing materials 221, 361
for low power consumption 17 special problems 222, 274
Stuffing box, air leakage 139-140
Stress
imposed by piping 188-189,241,274,294,
R 352
consequences of 190
Radial thrust 104-110, 321-322 due to inproper workmanship 188, 239,
Rating curves 282
as a diagnostic tool 273-278 due to setting of structures 189
differnt shapes of 14-18 due to changes in temperature 188
drooping HQ-curve 14-16 Stress, means of prevention 188-189
factors affecting their shapes 92-95, Submergence of suction-inlet 166
111-113 Suction head 8-9, 35
Recirculation in presence of prerotation 123-124
at pump-inlet 80-82, 122-123 Suction lift 8
at impeller outlet 88-95 Suction line
definition 80 air pockets in 127-128, 194-198
effect on QH-curve 79, 82-85 air leakage into 202-204
effect on shutoff head 88-95 effects on head-measurements 122, 158,
effect of flow-rate 82 193-194
effect on NPSH 84, 88 effects on performance 176-177, 190-191
from discharge-nozzle 89-95 Suction, process of 8-9
to eye of impeller 83 Suction inlet
to an intermediate radius 88 location in sump 166-167
Reducer, excentric 196 submergence 166
Resonance 233,274,352,393 Suction nozzle 8
in variable-speed drive 233 Suction performance 34-44, 145-153
Rotation, direction of 206 graphical representation 41
Roughness, effects of 61-66 factors affecting 147-150, 158
386 Index

measuring 126-127 multi-stage 10 1-103


testing 131-136 Thrust. radial 104-110.321-322
Suction specific speed Time. effects of 73-74. 134-136. 159. 240-241
definitions of 39-40 Torque-meter 121
effects of design-flow and design speed Total head 10. 122. 123. 164
52-53 Transient conditions 154-156. 177-182.206
effects of operating speed 52 Trouble. detection of causes:
for given flow-ratio 42 analysing information 273-280
scale effects 54 collecting information 273. 275
Suction lift 8 external inspection of pumping unit
Suction sump 163 289-292
design principles. of 166-173 external inspection of site 289-290
effects of inlet duct 168-169 inspection of failed parts 281-285
for more than one. single pump 172-173 observations during operation 292-294
model laws 175-176 preliminary steps 273-280. 289-290
recommended dimensions 167 verifying information 289
troubleshooting 174-175 Troubleshooting. recommended sequence of
Sump. suction see Suction sump activities 273-295
Surface roughness 61-66 certain shortcuts 299-303
System-curve 16 Turbine pumps 247
determination of 16. 44. 293-294 oil-lubricated 251
System-pump interaction 44. 293-294 water lubricated 251-253
at reduced NPSH 44 with semi-open impellers 253-254
System. pumping
layout 183-191
problems 167-173.176-178 u
remedial means 174-175.178.182-185
Unstable performance
due to fluctuation in liquid-level 248-250
T due to interaction of several pumps
172-173
Temperature. effects on due to prerotation 85-88. 154-156
alignement of parts 75 due to sump-layout 169-174
cavitation 159-160 Unsuitable choice of pump 14. 17. 55
NPSH required 159 Unusual case-histories 263-269
pump-parts 293. 355-356
Temperature. elevated 76. 141. 159-160
Temperature. fatigue due to changes in 75-76. v
188
Tension. fracture due to 233-234 Vacuum in column 250
Testing near shutoff 123-124. 141 Valve
Test loops effects of closing 178
for NPSH 131-133 effect of air 201-202. 250
for performance 126. 133 Valve-disk 201
Tests and testing 118-119 Valve. air 126. 181
field 292-295 Vapor bubbles
similarity laws 130 collapse of 33
Test. misleading Vapor pressure 33. 359
causes of 119-125. 157 Vibrations see also Noise
implications 119 Vibrations. sources of
near shutoff 122-123. 141 checklist of 351-352
Tests. problems with 119-125. 157 combined 231-233
Thermal shock 76 diagnosing its sources 293
Thoma's Cavitation Number 37 hydraulic 229-231
Thrust. axial 97-99 mechanical 205-208. 231
Thrust. balancing self-excited 233. 274. 293. 352
single stage 99-101 valve-disk 152-153.278
Index 387

Viscosity due to prerotation 154-157


definition 76 during start 181
effect on NPSH 78 due to shutoff 179-181
effect on performance 77, 206 discharge line 177
Volute frequent starts and stops 178
effect of air in 198-200 pipelines 177
effect of throat-area 318-320 prerotation 154-156
effect of tongue geometry 320-321 suction line 177,254-256
Vortex, cavitation due to 169 damage due to 177
Vortex, effect on performance 157, 169 prevention 178-179, 182
Vortex, formation of 163 Water level, variations in 248-250
Vortex, origin 163-165 Wear resistance 73
periodic shedding of 268 Wearing ring leakage
Vortex, in sumps 163, 167-173 effects on cavitation 70-71
effects on performance 66-70
Well
w deviations from vertical pOSition 250
oil in 252
Water hammer principles 178 straightness of 250
Water hammer caused by Workmanship, effects of 201-203,239
cavitation 155-156

You might also like