You are on page 1of 58

The Concise

Valve Handbook
The Concise
Valve Handbook
Sizing and Construction

Volume I

Michael A. Crabtree

MOMENTUM PRESS, LLC, NEW YORK


The Concise Valve Handbook: Sizing and Construction, Volume I

Copyright © Momentum Press®, LLC, 2018.

All rights reserved. No part of this publication may be reproduced, stored


in a retrieval system, or transmitted in any form or by any means—­
electronic, mechanical, photocopy, recording, or any other—except for
brief quotations, not to exceed 400 words, without the prior permission
of the publisher.

First published by Momentum Press®, LLC


222 East 46th Street, New York, NY 10017
www.momentumpress.net

ISBN-13: 978-1-94708-366-0 (print)


ISBN-13: 978-1-94708-367-7 (e-book)

Momentum Press Automation and Control Collection

Cover and interior design by Exeter Premedia Services Private Ltd.,


Chennai, India

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


To my wife Pam—for her love and patience.
Abstract

Research studies, within the process industry, routinely indicate that the
final control element is responsible for 60% to 70% of poor-functioning
control systems.
Although valves themselves are consistently wrongly selected,
­regularly misapplied, and are often incorrectly installed, a large proportion
of the blame may also be attributed to a number of associated ancillaries:
the valve actuator, I/P converter, and positioner.
Levelled at anyone working at a technical level in the process ­control
industry, Volume I: Sizing and construction provides a total in-depth insight
into valve technology. While studying both liquid and gas valve sizing, the
guide also presents a methodology to ensure the optimum selection of type,
size, body and trim materials, components, and ancillaries—­covering:
control valves, check valves, shut-off valves, and solenoid valves.
Volume II: Actuation, Maintenance, and Safety Relief, takes an
in-depth look at actuators and positioners. This volume also explores
a variety of maintenance and diagnostic issues including: testing for
dead-band/hysteresis, stick-slip and non-linearity; on-line diagnostics;
­
signature analysis; and correct procedures for calculating the spring
“wind-up” or “bench set.”
A complete section is also devoted to the whole field of safety relief
devices.
Lastly, this volume covers a number of topics which are all too
often  ignored: acoustics; water hammer; and even classification of
­stainless steel.

KeyWords

ball valves, butterfly and plug valves, cavitation and flashing, charac-
terization, construction, gas sizing, globe valves, leakage, liquid sizing,
­material selection, noise treatment, trim
Contents

List of Figures xv
List of Tables xxv
Foreword xxvii

Volume I
1   Basic Principles 1
1.1  The Final Control Element as Part of the Control Loop 2
1.2  Basic Theory 3
1.3  Equation of Continuity 3
1.4  Bernoulli’s Equation 5
1.5  Choked Flow 8
1.6  Pressure Recovery 9
1.7  Turndown Ratio and Rangeability 11
1.8  Velocity Profiles 12
1.9  Reynolds Number 13
1.10 Flashing and Cavitation 14
1.11 Flashing 15
1.12 Cavitation 16
1.13 Leakage Classification 18
1.14 Isolation Valve Leakage Classification 21
2   Liquid Valve Sizing 23
2.1  Practical Considerations 23
2.2  Application of Formulae 24
2.3  Sizing Example 1 27
x  •  Contents

2.4  Piping Geometry Factor 29


2.5  Sizing Example 2 31
3   Gas Valve Sizing 33
3.1  Pressure Drop Mechanism 33
3.2  Specific Heat Ratio Factor 38
3.3  Gas Expansion Factor 40
3.4  Valve Sizing 41
3.5  Sizing Example 1 43
4   Valve Construction 47
4.1  Globe Valve 48
4.2  Bonnet Assembly 49
4.3  PTFE (Teflon) 49
4.4  Laminated Graphite 50
4.5  Extended Bonnet 52
4.6  Bellows Seal Bonnet 52
4.7  Valve Trim 54
4.8  Guiding 55
4.9  Post-guiding 55
4.10 Top- and Bottom-guided Double Seat 56
4.11 Single-ported Balanced Globe Valve 57
4.12 Cage-guiding 58
4.13 Split Body Globe 59
4.14 Angle Is 60
4.15 Needle Valve 61
4.16 Bar Stock Body Valve 61
4.17 Gate Valve 61
4.18 Wedge Gate 62
4.19 Slab Valve 64
4.20 Expanding Gate Valve 65
4.21 Knife Edge Gate Valve 67
4.22 Pinch Valve 69
4.23 Diaphragm Valve 72
4.24 Rotary Control Valves 74
Contents   •   xi

4.25 Ball Valve 75
4.26 Trunnion Ball Valve 77
4.27 Characterized Ball Segment Valve 81
4.28 Butterfly Valve 82
4.29 Plug Valve 84
4.30 Eccentric Plug Valve 86
4.31 Check Valves 88
4.32 Valve Sizes and Pipe Schedules 88
4.33 Material Selection 90
4.34 Corrosion 90
4.35 Erosion 94
4.36 End Connections 94
4.37 Screwed End Connections 94
4.38 Flanged End Connections 95
4.39 Hub End Body 96
4.40 Welded End Connections 96
4.41 Lap Joint Flange 98
4.42 Flangeless Connections 99
4.43 Grayloc® Connector 100
5   Valve Trim and Characterization 103
5.1  Inherent Characteristics 103
5.2  Linear Inherent Flow Characteristic 103
5.3  Equal Percentage Inherent Flow Characteristic 104
5.4  Quick Opening Inherent Flow Characteristic 104
5.5  Modified Percentage Inherent Flow Characteristic 105
5.6  Characteristic Profiling 105
5.7  Installed Characteristics 105
5.8  Cavitation Control 108
5.9  Reducing Cavitation 110
5.10 Eliminating Cavitation 112
5.11 Noise Sources 113
5.12 Mechanical Noise 115
5.13 Hydrodynamic Noise 116
xii  •  Contents

5.14 Aerodynamic Noise 117


5.15 Noise Prediction 117
5.16 Noise control 117
5.17 Path Treatment 118
5.18 Insulation 119
5.19 Silencers 120
5.20 Source Treatment 120
5.21 Velocity Control 120
6   Valve Selection 123
Glossary 129
Bibliography 131
About the Author 133
Index 135

Volume II
7    Valve Actuators and Positioners 141
7.1  Pneumatic Control 141
7.2  Flapper–Nozzle Assembly 141
7.3  I/P Converter 142
7.4  Diaphragm Actuators 144
7.5  Springless Diaphragm 145
7.6  Advantages and Disadvantages of Diaphragm Actuators 147
7.7  Cylinder Actuators 147
7.8  Spool Block 149
7.9  Electro-Hydraulic Actuation 149
7.10 Electric Actuation 150
7.11 Torque Limiting 152
7.12 Hammer-Blow Mechanism 153
7.13 Solenoid Valve 153
7.14 Digital Actuators 155
7.15 Transfer Mechanisms 157
7.16 Valve Positioners 161
7.17 Positioner Guidelines 163
Contents   •   xiii

8   Valve Testing and Diagnostics 167


  8.1  Deadband and Hysteresis 167
  8.2  Testing Procedures 169
 8.3  Online Diagnostics 173
 8.4  Electronic Torque Monitoring 176

9   Valve Maintenance and Repair 179


 9.1   In-Line Repairs 180
 9.2   Repairs Under Pressure 180
  9.3   Repairs on Drained Systems 181
 9.4   Packing Replacement 181
  9.5   Replacing or Refinishing Seat Rings 181
 9.6   Other In-Line Repairs 182
 9.7   In-Line Post-Repair Procedures 182
 9.8   Shop Repairs 182
 9.9   Actuator Bench Set 183
 9.10  Spring Calculations 184

10  Safety Relief Valves 187


  10.1  History 187
  10.2  Definitions 190
  10.3  Weight-Loaded Pressure/Vacuum Relief Valves 191
 10.4  Spring-Loaded Relief Valves 192
 10.5  Applications 194
 10.6  Limitations 194
 10.7  Safety Valves 194
  10.8  Basic Operation: Lifting 196
  10.9  Basic Operation: Reseating 198
  10.10 Conventional Safety Relief Valves 200
  10.11 Balanced Safety Relief Valves 204
  10.12 Bellows-Type Balanced Safety Valve 204
  10.13 Piston-Type Balanced Safety Valve 206
  10.14 Non-Reclosing Pressure Relief Devices 211
  10.15 Conventional Rupture Disc 215
  10.16 Scored Tension-Loaded Rupture Disc 215
xiv  •  Contents

  10.17 Composite Rupture Disc 216


  10.18 Graphite Rupture Disc 217
  10.19 Burst Disc Applications and Installation Practices 218
  10.20 Performance Tolerance 219
  10.21 Maximum Operating Pressure 221
  10.22 Cyclic/Pulsating Duties 221
  10.23 Case A: 276 kPa (g) or Higher 222
  10.24 Case B: Lower than 276 kPa (g) 223
  10.25 Standards 223
Appendix A: J–T Valve 225
Appendix B: Basic Acoustics 227
Appendix C: Block and Bleed 241
Appendix D: Water Hammer 245
Appendix E: Stainless Steel 255
Glossary 263
Bibliography 265
About the Author 267
Index 269
List of Figures

Figure 1.1. The four basic elements of a control system: process,


transducer (sensing element plus transmitter), final
control element, and controller. 2
Figure 1.2. Velocity and pressure distribution of a fluid flowing
through a restriction (e.g., a valve orifice). 4
Figure 1.3. To allow the same amount of liquid to pass the velocity
must increase. 4
Figure 1.4. Actual flow versus √∆P. 9
Figure 1.5. Streamlined valves dissipate less energy, and therefore
have a higher pressure recovery than less streamlined
valves.10
Figure 1.6. Streamlined valves have higher velocities at the vena
contracta with subsequent lower pressures. 10
Figure 1.7. Comparison of rangeability and turndown. 11
Figure 1.8. A flat ‘ideal’ velocity profile. 12
Figure 1.9. A laminar ‘parabolic’ velocity profile. 12
Figure 1.10. A turbulent velocity profile. 13
Figure 1.11. Flashing occurs when the downstream pressure is at or
below the fluid vapor pressure. 14
Figure 1.12. Damage due to flashing appears smooth and polished. 15
Figure 1.13. Cavitation is caused by the pressure dropping to the
vapor pressure of the fluid and rising to a higher
pressure further downstream. 16
Figure 1.14. Cavitation damage occurs when the bubbles collapse
on or near solid surfaces within the valve or piping
and material is chipped away. 16
Figure 1.15. Cavitation damage shows up as very dull, rough,
and pitted—sponge-like in appearance. 17
xvi  •   List of Figures

Figure 1.16. Typical liquid noise characteristic (courtesy


Fluids handling: Principles & Practice). 18
Figure 2.1. Flow is only proportional to √∆P within the subcritical
region.26
Figure 2.2. Data on Fisher Emerson easy-e® ES (flow up)
cage-guided globe valve having a linear characteristic
with a valve opening to 100% of its total travel. 29
Figure 3.1. Initially, there is a straight-line relationship between
the mass flow (w) and the pressure drop ratio (x). 35
Figure 3.2. As the pressure drop ratio (x) increases, there will
come a point where the mass flow will no longer
­continue to increase at the same rate and will start
to deviate from the straight line. 35
Figure 3.3. This deviation will continue until no further increase
in the ‘pressure drop ratio’ will yield any additional flow.
This is termed ‘choked flow’ and the point at which it
occurs is the called terminal pressure drop ratio (XT). 35
Figure 3.4. At low flows, as the velocity increases, the pressure
decreases and reaches a minimum at the vena contracta. 36
Figure 3.5. At some point, before choked flow occurs, the flow rate
is still increasing and the velocity at the vena contracta
becomes sonic. 37
Figure 3.6. As the pressure drop ratio increases, the vena contracta
moves toward the physical restriction and its cross-­
sectional increases. When it backs up to the physical
­restriction, no further increase in flow is possible and
is said to be choked. 37
Figure 3.7. Test performed on a globe valve set at 70% open. 38
Figure 3.8. Performance of two 50 mm valves having almost
identical CV values of 60 when 70% open—one a
­linear globe valve and the other a high performance
­butterfly valve. 39
Figure 3.9. XT not only varies with different styles of valves,
but also with the valve opening. 39
Figure 3.10. Data on Fisher Emerson easy-e ES (flow up)
®

cage-guided globe valve having a linear characteristic


with a valve opening to 100% of its total travel. 45
Figure 4.1. Simple classification of valve types. 48
List of Figures   •   xvii

Figure 4.2. A typical globe valve (courtesy Emerson–Fisher). 48


Figure 4.3. A typical packing box assembly comprising a number of
plastic or composition washers that are retained by a
packing flange. 50
Figure 4.4. Typical standard V-ring packing arrangement. 50
Figure 4.5. Typical double graphite packing arrangement. 51
Figure 4.6. Electrochemical series showing the electrode potentials
referenced to hydrogen. 52
Figure 4.7. Extended bonnet in which the stuffing box is located
further away from the process medium allows the
­process medium temperature to be extended up to
800°C or more (courtesy Emerson–Fisher). 52
Figure 4.8. Bellows stem seals are designed to eliminate valve
leakage and normally supplement conventional packing. 53
Figure 4.9. The bellows are available as either a formed or
welded construction. 53
Figure 4.10. In the seat retainer, the seat ring is clamped into the
body by the bonnet and seat retainer (courtesy Valtek
­Control Products). 55
Figure 4.11. Single seat top-guided contact valve. 56
Figure 4.12. A double seat bottom-guided globe valve. 57
Figure 4.13. In the single-ported balance globe valve, the valve
plug is made as a piston having an internal passage
that permits the fluid pressure to communicate to
both sides of the plug. 58
Figure 4.14. Cage-guided control valve. 59
Figure 4.15. The split-body, stem-guided globe caters for difficult
flows with high viscosity. 59
Figure 4.16. Single-seat angle valve. 60
Figure 4.17. Split-body needle valve (courtesy Fisher Rosemount). 61
Figure 4.18. Bar stock bodies are often specified for corrosive
applications (courtesy Emerson–Fisher). 62
Figure 4.19. A typical gate valve comprising a disk that slides up
and down between seats. 63
Figure 4.20. The gate faces can be in the form of a wedge shape,
sliding up and down between tapered seats. 63
Figure 4.21. (a) One-piece solid wedge. (b) Flex wedge. 64
xviii  •   List of Figures

Figure 4.22. In the slab valve, the gate is pushed against the seal by
the flow pressure. 64
Figure 4.23. The moving section comprises a slab-shaped gate
with a single port drilled through it. 65
Figure 4.24. The expanding gate valve comprises two segmented
assemblies that are attached to each other by either a
spring or lever mechanism. 65
Figure 4.25. When the gate is in its final upper (open) position, with
the front segment against the upper stop, the lower back
angles are in contact with each other (courtesy Daniel
Valves).66
Figure 4.26. Exaggerated view showing how the opposing forces
cause the gate and segment assembly to expand, sealing
against both seats. 66
Figure 4.27. When the gate is in its final lower (closed) position,
with the front segment against the bottom stop, the
upper back angles are in contact with each other
(courtesy Daniel Valves). 67
Figure 4.28. Exaggerated view showing how the opposing forces
cause the gate and segment assembly to expand,
­sealing against both seats. 67
Figure 4.29. The knife edge gate valve is an excellent valve for
service that requires either full or no flow. 68
Figure 4.30. The V-insert gate valve. 69
Figure 4.31. Flow/travel characteristics for different variations of
the sliding gate valve. 69
Figure 4.32. Sliding gate regulator valve makes use of slotted
movable disc and a stationary slotted plate (courtesy
Jordan Valve). 70
Figure 4.33. (a) When throttled open, the orifices of the disc align
with the openings of the plate to allow the required flow
to pass through the slots. (b) When the valve is closed,
the disc and plate form a solid barrier to flow (courtesy
Jordan Valve). 70
Figure 4.34. Sliding gate valve using a stationary plate and a
rotating disk. 71
Figure 4.35. (a) The pinch valve comprises a rubber hose, which
­normally provides full flow. (b) When pinched together,
the flow stops. 71
List of Figures   •   xix

Figure 4.36. Double vise mechanism in which the clamps move


together and pinch the valve closed at the center line
(a) open and (b) closed. 72
Figure 4.37. In the air-operated pinch valve, the valve body acts as
a built-in actuator, requiring an actuation pressure of
­approximately 2 to 3 bar over line pressure for closure
(courtesy Red Valve). 72
Figure 4.38. In the Saunders patent, flow through the body is over a
transverse weir (a) and the valve is closed by means of
a flexible rubber or synthetic dome-shaped diaphragm
(b) (courtesy Crane Process Flow Technologies Ltd.) 73
Figure 4.39. Straight-through diaphragm valves, exemplified by the
Saunders full bore K/KB type, have a smooth non-­
­turbulent body design and are designed for m
­ inimum
flow resistance while allowing rodding out and
easy ­cleaning (courtesy Crane Process Flow
Technologies Ltd.). 74
Figure 4.40. In the ‘floating’ ball valve, the ball is completely
supported by two seats that provide bearing support
to the ball s­ egment, with fully closed to fully open
­performed by a 90° rotation of the plug segment. 75
Figure 4.41. In the split-body design, the split between the body
and cap is off-center so that the stuffing box is left
intact when the ball or seating is removed. 76
Figure 4.42. End-entry design in which the body comprises a single
piece, with the ball retained by an insert screwed into
the body at one end. 76
Figure 4.43. The trunnion ball valve distributes excess hydraulic
load into the valve body, rather than through the seating
­(courtesy ITT ­Industries, Inc.). 77
Figure 4.44. Leakage integrity is ensured through the use of floating
spring-loaded seats that are kept in contact with the ball
even in the absence of line pressure. 78
Figure 4.45. As the line pressure increases the seat area creates
a piston effect that forces the seat against the ball.
Sealant injection points allow emergency sealant
injection to be carried out in the event of seat-insert
or stem-seal damage that could lead to external or
­internal  leakage. 78
xx  •   List of Figures

Figure 4.46. In the event that leakage past the seals does occur, and
in order to cater for double block-and-bleed a­ pplications,
a body vent allows the body cavity to be vented to
­atmosphere. 79
Figure 4.47. (a) In its fully open position, the ball valve presents a
single circular orifice to the flowing medium. (b) As the
ball is rotated toward its closed position, the shape of
the opening changes to become two identical elliptical
­orifices that offer two equal restrictions in series.
(c) Closed ­position. 80
Figure 4.48. Pathway fitted with parallel perforated attenuator plates
that produce a smooth gradual pressure reduction
across the valve that minimizes velocity, noise
generation, and cavitation (courtesy Neles-Metso). 80
Figure 4.49. Characterized V-notch ball valve in which the opening
between the ball and seal is modified to provide different
flow characteristics (courtesy Emerson–Fisher). 81
Figure 4.50. Different flow characteristics of the characterized ball
valve.81
Figure 4.51. Ball segment valve with centrally mounted shaft
(courtesy Somas Instrument AB). 82
Figure 4.52. Ball segment valve with eccentrically mounted shaft,
­allowing pressure between the segment and seat to
be increased by increasing torque (courtesy Somas
­Instrument AB). 82
Figure 4.53. Butterfly valve comprises a circular disc-shaped damper
mounted on a shaft. 83
Figure 4.54. Conventional center-disc butterfly valve. 83
Figure 4.55. Offset disc, or high performance butterfly valve. 84
Figure 4.56. (a) conventional disc shape (b) the fishtail disc
(courtesy Emerson–Fisher). 84
Figure 4.57. Flow characteristics of different-shaped butterfly discs. 85
Figure 4.58. Basic plug valve. 85
Figure 4.59. Camflex: eccentric rotary plug valve (courtesy ­
Masoneilan).86
Figure 4.60. Once seating occurs, a positive seal between plug and
seat is achieved by the elastic deformation of the plug
arms (courtesy Masoneilan). 87
List of Figures   •   xxi

Figure 4.61. In the hinged ‘swing check valve,’ the sealing disc is
attached to a hinge, which is free to rotate around the
hinge pin. 88
Figure 4.62. The pinch check valve feature: a typical 30-year
­operational lifespan, a low headloss; they do not rust
or ­corrode; they are not affected by UV; and their
flexibility allows them to compress around trapped
­solids (courtesy Tideflex). 89
Figure 4.63. Screwed end valve connections with tapered female
thread.94
Figure 4.64. Maximum pressure rating versus temperature for
­carbon steel flanges (ANSI B16.5). 95
Figure 4.65. Flat-face flanged end connection. 96
Figure 4.66. The raised-face flange has a circular raised face. 96
Figure 4.67. Hub end or separable flange body. 97
Figure 4.68. Socket welding end. 97
Figure 4.69. Butt welding ends. 97
Figure 4.70. The lap joint flange is used in conjunction with a lap
joint stub end that is butt-welded onto the process
pipeline (courtesy Coastal Flange Inc.). 98
Figure 4.71. Flangeless valves are held between flanges by long
through-bolts.99
Figure 4.72. The Grayloc® connector comprises three components:
a metal seal ring, a two-part clamp assembly, and two
hubs (courtesy Oceaneering International). 100
Figure 4.73. (a) The clamp assembly fits over the two hubs and
forces them against the seal ring rib. (b) As the hubs
are drawn together, the seal ring lips deflect against
the inner sealing surfaces of the hubs (courtesy
Oceaneering International). 101
Figure 5.1. Inherent flow characteristics. 104
Figure 5.2. Quick opening flow characteristic uses a simple disc
shaped plug (courtesy Fisher Rosemount). 105
Figure 5.3. Plug outlines used to obtain (from left to right) equal
percentage; linear; and quick-opening flow
­characteristic ­(photographs courtesy of Mitech). 106
xxii  •   List of Figures

Figure 5.4. For cage-guided trim, flow characterization is determined


by the shape of windows in the cylindrical cage (a) equal
percentage (b) linear and (c) quick-opening flow
­characteristics. 106
Figure 5.5. Differential pressure drop is actually distributed across
both the valve and other parts of the system. 106
Figure 5.6. Installed linear flow characteristic. 107
Figure 5.7. Installed equal percentage flow characteristic. 108
Figure 5.8. Trim selection guide—liquid applications (courtesy
Mitech).109
Figure 5.9. Plug hard facing variations: from left to right: seat
surface; full contour; lower guide area; and full
­contour and lower guide (courtesy Valtek Control
­Products). 110
Figure 5.10. High differential pressure across the valve can lead
to cavitation. 110
Figure 5.11. Cavitation across the valve can be avoided by installing
a choke. 111
Figure 5.12. Cavitation control seat retainer (courtesy Mitech). 111
Figure 5.13. Mitech’s ZZ seat retaining energy dissipating device
for small diameter valves and pressures above 40 bar
(courtesy Mitech). 112
Figure 5.14. Disk stack design of globe valve trim (courtesy
Mitech).113
Figure 5.15. Some relative noise levels for common sounds and
­activities. 114
Figure 5.16. Typical liquid noise characteristic (courtesy Fluids
­handling: Principles Practice). 116
Figure 5.17. There are two main elements to noise treatment:
path treatment and source treatment. 118
Figure 5.18. Cut-away section of an in-line silencer making use
of an inlet ­diffuser to break up turbulence. (courtesy
Valtek ­Control Valves, Flowserve Corporation). 120
Figure 5.19. Simple low noise retainer device (courtesy Mitech). 121
Figure 5.20. Simple diffuser plate (courtesy Mitech). 121
List of Figures   •   xxiii

Figure 5.21. Tortuous path in which the fluid undergoes a series of


contractions, expansions, direction changes and splits
in axial, radial and circumferential directions (courtesy
Emerson-Fisher).122
Figure 6.1. Control valve selection guide for gases and liquids
(courtesy Mitech). 126
Figure 6.2. Control valve body selection guide (courtesy
CSD Controls). 127
Figure 6.3. Control valve trim selection guide (courtesy
CSD Controls). 127
Figure 6.4. Actuator and accessory selection guide (courtesy
CSD Controls). 128
List of Tables

Table 1.1. Comparison of the flow coefficient, CV, of a number


of different types of valve at different sizes 8
Table 1.2. Typical numerical values of FL for some different
valve styles 11
Table 1.3. Control valve seat leakage classifications according
to ANSI/FCI 70-2-1991 19
Table 1.4. Class VI maximum seat leakage according to
ANSI/FCI 70-2-1991 20
Table 1.5. Leakage rates for different-sized isolation valves
based on API 598 21
Table 2.1. Saturated vapor pressure levels for water at
different temperatures 28
Table 3.1. Results of test show that XT (the choked value of x)
remains constant at 0.72 in each instance 38
Table 4.1. Preferred series of ANSI and DN pipe sizes 89
Table 4.2. Example of some pipe schedules based on ASME
standards B36.10M and B36.19M 91
Table 4.3. Some of the most common cast materials 93
Table 4.4. Preferred series of ANSI class and PN pressures 95
Table 5.1. Valve characteristic selection guide 108
Table 5.2. Permissible personnel noise level exposure and
time levels according to OSHA 114
Table 5.3. Typical pipe wall attenuation (dBA) for a carbon
steel pipe according to different pipe schedules
and pipe sizes 119
xxvi  •   List of Tables

Table 6.1. Comparison table: rating from 1 down to 5


(1 = highest rating 5 = poorest rating) (courtesy
Crane Process Flow Technologies Ltd.) 124
Table 6.2. Comparison of different control valves where:
1 = good; 2 = average; 3 = poor; and x = not suitable
(courtesy Mitech) 125
Foreword

In this book, ‘The Concise Valve Handbook—Part 1. Sizing and con-


struction,’ I have made use of a building-block approach, presenting
material in a form suitable for two distinct classes of reader: the begin-
ner, with no prior knowledge of the subject and the more advanced
specialist.
The complete text is suitable for the advanced reader. However, those
parts of the text that involve a mathematical treatment, which are not
required by the beginner, are indicated by a mark ► at the beginning and
◄ at the end. Consequently, for the beginner, the text may be read, with
full understanding, by ignoring the marked sections.
I offer no apologies for my preference for metric-based measurement:
the SI system. Apart from the United States, only two other countries in the
world still adhere to the fps system (foot-pound-second)—the so-called
imperial system first defined in the British Weights and Measures Act of
1824—Myanmar (Burma) and Liberia.
I have tried to mix it up as far as possible, and I have got a units con-
version table right in the front of the book. But, for the moment, just try
for the following:

1 bar = 100 kPa ≈ 1 atmosphere ≈ 14.7 psi


1 inch = 25.4 mm
20°C = 68 °F
100°C = 212 °F

And lastly, while I have made some compromises (analog instead


of analogue; program instead of programme), I reserve my right to spell
according to the British system:
xxviii  •  Foreword

English United States


Metre Meter
Litre Liter
Fibre Fiber
Colour Color
Vapour Vapor

Unit Conversions

Quantity SI United States customary


Distance 25.5 mm 1 in
1 millimetre 0.03937 in
1m 39.37 in
1m 3.281 ft
0.9144 m 1 yd
Area 1 square metre (m2) 1550 in2
1 square metre (m2) 10.76 ft2
1 square metre 0.00155 in2
­millimetre (mm2)
Volume 1 cubic metre (m3) 61.02 in3
1 cubic metre (m3) 35.31 ft3
0.02832 m3 1 ft3
1 litre 61.02 in3
1 litre 0.03531 ft3
1 litre 0.2642 gal
3.785 litres 1 gal
Mass 1 kg 2.205 lb
454 g 1 lb
Force 1N 0.2248 lbf
4.448 N 1 lbf
Pressure 1 bar 14.504 lbf/in2 (psi)
1 kPa (kN/m2) 0.145 lbf/m2 (psi)
6.895 kPa 1 psi
1 psi 0.0361 inches H2O
(in WC)
Foreword   •   xxix

Quantity SI United States customary


Temperature K 1.800 °R
°C 1.8 °C + 32 = °F
Flow rate 1 m3/h 4.403 gal/min (gpm)
1 kg/h 2.205 lb/h
CHAPTER 1

Basic Principles

In any process control loop, the final control element is the mechanism
that changes the value of the manipulated variable in response to the out-
put signal from the control unit.
The final control element comprises the actuator, with its associated
and linkage and positioner, and the final control element proper—valve,
pump, transformer, motor, variable speed drive, and so on. It is axiomatic
that many of the problems associated with control loop performance can
also be laid at the door of the final control element.
Fluid properties can vary enormously from industry to industry. The
fluid may be toxic, flammable, abrasive, radioactive, explosive, or corro-
sive; it may be single-phase (clean gas, water, or oil) or multi-phase (e.g.,
slurries or dust-laden gases). The pipe carrying the fluid may vary from
less than 1 mm to many in diameter. The fluid temperature may vary from
close to absolute zero to several hundred degrees Celsius, and the pressure
may vary from high vacuum to high pressures.
Many different types of valves have been developed to suit these
variations in fluid properties and flow applications. However, only a few
valves have found widespread application, and no one single valve can be
used for all applications.
The basic purpose of a control valve is to control the flow of a medium
in a pipe, either turning it on or off or varying it continuously. However,
a control valve designed primarily to throttle energy is not necessarily
designed for shut-off purposes, and these two requirements often have to
be balanced or realized in separate systems.

In the fairly certain knowledge that prefaces or forewords are


rarely read, I’m taking this opportunity to invite you to go back
and peruse the foreword since there are a couple of fairly import-
ant points that you really need to understand.
2  •   The Concise Valve Handbook

1.1 The Final Control Element as Part of


the Control Loop

Routinely, research studies within the process industry indicate that the
final control element is responsible for 60% to 70% of poor-functioning
control systems. The problems lie not just with the valve itself, but also
with the valve actuators, I/P converters, and positioners. However, prob-
ably the majority of problems can be attributed to oversized valves and
undersized actuators.
The first step in successful application is to gain an understanding
of the basics of a control system. As shown in Figure 1.1, the four basic
elements of any control system comprise:

• process;
• transducer (sensing element plus transmitter);
• final control element; and
• controller.

Again referring to Figure 1.1, the controlled variable is called the


Process Demand (PD) or Manipulated Variable (MV), or simply the
OP, while the measured variable is called the Process Demand. If the
PD is subject to a step change, by how much will the PV change? This is
determined by what is called the process gain (KP) and is given by divid-
ing the percentage change in the PV by the percentage change made by
the PD:
Equation Chapter (Next) Section 1

PV (%)
Kp =  (1.1)
PD (%)

Controller

Process
Demand (PD)

Process
Variable (PV)

Process

Transducer (sensor + transmitter) Final control element

Figure 1.1.  The four basic elements of a control


system: process, transducer (sensing element plus
transmitter), final control element, and controller.
Basic Principles   •  3

Thus, for example, if we make a step change of 20% to the PD and the
PV also changes by 20%, then the process gain (KP) is 1. However, if the
PV only changes by 10%, then the process gain (KP) is 0.5. Alternatively,
if the PV changes by 60%, then the process gain (KP) is 2.
Generally, the process gain should lie between the 0.5 and 2.0. If it
is less than 0.5, then typically, the transmitter span is too wide for good
control. If the process gain is greater than 2, this is usually an indication
that the control valve is oversized.

1.2 Basic Theory

A control valve can be simply represented as a restriction in a pipeline


that creates a pressure drop, or head loss, that bears a relationship to the
flow rate.
Because the flow entering the restriction is equal to the flow exiting
the restriction, a reduction in cross-sectional flow area at the restriction
must yield an increase in fluid velocity. The velocity increase is accom-
panied by a proportional decrease in pressure that results from an energy
tradeoff where potential energy is converted to kinetic energy. The point
of minimum cross-sectional flow area, maximum velocity, and minimum
pressure is called the vena contracta.
This physical phenomenon is based on two well-known equations:
the equation of continuity and Bernoulli’s equation.
Fluid flow through a valve orifice is illustrated in Figure 1.2.
Assume that a fluid of density ρ flowing in the pipe of area A1 has a
mean velocity v1 at a line pressure P1. It then flows through the restric-
tion of area A2 where the mean velocity increases to v2 and the pressure
falls to P2.

1.3 Equation of Continuity

The equation of continuity states that, for an incompressible fluid, the


volume flow rate, Q, must be constant. Very simply, this indicates that
when a liquid flows through a restriction, then in order to allow the same
amount of liquid to pass (to achieve a constant flow rate), the velocity
must increase (Figure 1.3).
►Mathematically:

Q = v1.A1 = v2.A2 (1.2)


4  •   The Concise Valve Handbook

V1 VVC V3
P1 PVC P2

Vena Contracta

Pressure Unrecoverable pressure loss

∆P
Velocity

Distance

Figure 1.2.  Velocity and pressure distribution of


a fluid flowing through a restriction (e.g., a valve
orifice).

QIN = v1. A1 QOUT = v2. A2


Velocity

Distance
Figure 1.3.  To allow the same amount of
­liquid to pass the velocity must increase.

where:
v1 and v2 and A1 and A2 are the velocities and cross-sectional areas of the
pipe at points 1 and 2, respectively.
Basic Principles   •  5

1.4 Bernoulli’s Equation

In its simplest form, Bernoulli’s equation states that, under steady flow
conditions, the total energy (pressure + kinetic + gravitational) per unit
mass of an ideal fluid (i.e., one having a constant density and zero viscos-
ity) remains constant along a flow line.

v12 P1 v 22 P2  (1.3)
+ = +
2 ρ 2 ρ
where:
v = velocity at a point in the streamline;
P = pressure at that point;
ρ = fluid density;
g = acceleration due to gravity;
z = level of the point above some arbitrary horizontal reference plane
with the positive z-direction in the direction opposite to the gravitational
acceleration; and
k = constant
In the restricted section of the flow stream, the kinetic energy (dynamic
pressure) increases due to the increase in velocity, and the potential energy
(static pressure) decreases.
Relating this to the conservation of energy at two points in the fluid
flow, then:

v12 P1 v 22 P2  (1.4)
+ = +
2 ρ 2 ρ
Multiplying through by ρ gives:

1 1
ρ .v12 + P1 = .ρ.v 22 + P2  (1.5)
2 2
or:

1 1
P1 − P2 = .ρ.v 22 − .ρ.v12  (1.6)
2 2
or:

1 1
∆P = ρ .v 22 − .ρ.v12  (1.7)
2 2
6  •   The Concise Valve Handbook

where:

∆P = P1 − P2  (1.8)

Now from the continuity equation (1.2), we can derive:

Q
v1 =  (1.9)
A1
and:

Q
v2 =  (1.10)
A2
substituting in (1.7):

2 2
1  Q 1  Q
∆P = .ρ.   − .ρ.    (1.11)
2  A2  2  A1 
Solving for Q:

2∆P
ρ
Q = A2 . 2
 (1.12)
A 
1−  2 
 A1 

Since it is more convenient to work in terms of the diameters of


the restriction (d) and the ID (inside diameter (D)) of the pipe, we can
­substitute for:

À.D 2
A1 =  (1.13)
4
and:

π.d 2
A2 =  (1.14)
4
to give:

2.∆P
π.d2
ρ
Q= . 4
 (1.15)
4  d
1−  
 D
Basic Principles   •  7

The term:
1
4
 (1.16)
 d
1−  
 D

is called the ‘Velocity of Approach Factor’ (EV), and by substituting in


(1.15), we have:

2.∆P
Q = E v .d 2 .  (1.17)
ρ

Unfortunately, equation (1.17) only applies to perfectly laminar,


inviscid flows. In order to take into account the effects of viscosity and
turbulence, a term called the discharge coefficient Cd is introduced that
marginally reduces the flow rate (Q).
The full equation for an incomprehensible fluid thus becomes:

2∆P
Q = Cd .E v .d 2 .  (1.18)
ρ

In valve technology, this equation has been modified as follows:

∆P  (1.19)
Q = Cv
SG f

Q = flow rate;
CV = valve flow coefficient;
∆P = differential pressure (P1–P2); and
SG = specific gravity of fluid (water at 60°F = 1.0).
The valve flow coefficient, CV, is an index used to measure the capac-
ity of a control valve. CV is determined experimentally, using water as the
test fluid, for each style and size of valve with the valve either fully open
or at a given valve opening, usually stated as a percentage of maximum
travel.
Numerically, CV is defined as “the number of US gallons per minutes
of water at 60°F that will pass through a given flow restriction, with a
pressure drop of 1 psi across the valve.”
CV is thus an index that allows the liquid capacities of different
valves to be compared under a standard set of conditions. Table 1.1
compares the flow coefficient of a number of different types of valve at
different sizes.
8  •   The Concise Valve Handbook

Table 1.1.  Comparison of the flow coefficient, CV, of a number of


different types of valve at different sizes
Cv (typical at 100% open)
Valve Size
Valve type 50 mm 100 mm 200 mm
Globe: cage-guided 70 240 850
Globe: stem-guided 45 180 N/A
Full ball N/A 500 2,200
V-notch ball 175 600 1,820
Butterfly: standard 90 520 2,820
Butterfly: high performance 90 490 2,170

The metric equivalent of CV is KV, defined as “the number of cubic


metres per hour of water at 15°C that would pass through a valve with a 1
bar pressure drop across it.”
Although many valve manufacturers now list the KV values in their
data sheets, CV still remains the almost universal standard, despite its
imperial background. Nonetheless, many users prefer to work with metric
standards and make use of CV:

∆P
Q = 0.87 C v  (1.20)
SG f

where the volumetric flow rate (Q) is expressed in m3/hr and the differen-
tial pressure drop (∆P) is expressed in bars.

1.5 Choked Flow

Equation (1.19) implies that, for a given valve, simply increasing the pres-
sure differential across the valve can continually increase the flow. In real-
ity, this relationship only holds true for a limited range. This is illustrated
in Figure 1.4 that shows a typical plot of actual flow versus √∆P through a
flow restriction. This indicates that flow is only proportional to √∆P within
the sub-critical flow region.
If the differential pressure is further increased, a point is reached
where no further flow increase occurs, despite increasing the differen-
tial pressure. This is termed as choked flow (also known as the crit-
ical flow) and is the maximum flow rate possible through that valve.
Decreasing the downstream pressure will not result in an increased flow
Basic Principles   •  9

Sub-critical Critical

Choked flow
Incipient cavitation

Flow
Cv

∆P

Figure 1.4.  Actual flow versus √∆P.

rate, although the valve can handle the higher pressure drop with no
detrimental effects.
In gases, choked flow occurs when the velocity reaches the speed of
sound (Mach 1). For liquids, the speed of sound is extremely high, and
practically speaking, incompressible fluids do not choke. In practice, how-
ever, as the differential pressure is increased and the velocity increases,
the pressure at the vena contracta decreases. If the vena contracta pressure
falls to below the vapor pressure of the liquid, partial vaporization occurs
and the sonic velocity of the resultant liquid/vapor mixture falls dramati-
cally. At this point, choked flow occurs.
The practical consideration of choked flow is that, when calculat-
ing the CV required for a particular application, only the choked pressure
drop can be used in the formulae and not the actual pressure drop. This
results in a larger CV requirement than would otherwise be the case. If
choked flow is not taken into account, it is possible to select a valve that
is too small.

1.6 Pressure Recovery

The ability of the control valve to reconvert kinetic energy downstream


of the restriction back into pressure is a characteristic known as pressure
recovery.
The degree of pressure recovery varies from valve to valve.
Considering two valves with equal flow, a streamlined valve will dissi-
pate less energy through the restriction and will, therefore, have more
energy downstream for recovery to a higher pressure. Conversely, in a
less streamlined valve, larger amounts of energy are dissipated through
10  •   The Concise Valve Handbook

the restriction, and therefore, less energy will be available downstream for
recovery to a higher pressure (Figure 1.5).
Alternatively, streamlined valves produce relatively higher veloci-
ties through their restriction than do less streamlined, restrictive valves.
Velocity, being inversely proportional to pressure, suggests lower pressure
at the vena contracta with high-recovery streamlined valves (Figure 1.6).
The amount of pressure recovery varies with valve style and stroke
and is a function of the upstream, vena contracta, and downstream pres-
sures. It is quantified by what is called the pressure recovery coefficient
FL—a dimensionless expression of the pressure recovery ratio in a control
valve that is mathematically represented as:

P1 − P2
FL =  (1.21)
P1 − Pvc

P1
High recovery
Pressure

P2

Low recovery
PVC

Distance

Figure 1.5.  Streamlined valves dissipate less


energy, and therefore have a higher pressure
recovery than less streamlined valves.

P1
Pressure

Low recovery P2

PVC

High recovery

Distance

Figure 1.6.  Streamlined valves have higher


velocities at the vena contracta with subsequent
lower pressures.
Basic Principles   •  11

Table 1.2.  Typical numerical values of FL for some different valve styles
Cage-guided Standard
globe globe Disk 60° Disk 90° Ball 90°
FL 0.9 0.85 0.75 0.5 0.6
FL2 0.81 0.72 0.56 0.25 0.36

where:
P1 = upstream pressure;
PVC = vena contracta pressure; and
P2 = downstream pressure.
Evaluation of the preceding expression suggests that high-recovery
valves will result in less pressure drop, P1–P2. Therefore, high-recovery
valves have a low value of FL, and low recovery valves have high values
of FL—where FL is always less than 1.0.
FL is determined by laboratory tests and cataloged by most valve man-
ufacturers for use in more precise determination of valve capacity during
critical flow. It also is useful in predicting damaging phenomena such as
cavitation. Some typical numerical values of FL are given in Table 1.2.
Note: FL is the ISA nomenclature and that sometimes the symbol Cf
is used. Fisher Rosemount formerly made use of the symbol Km where:
FL = K m  (1.22)

1.7 Turndown Ratio and Rangeability

In our discussions on CV, it should be noted that two definitions are


­frequently used and confused.
Turndown relates to the application and is the ratio of the calculated
CV at maximum conditions to the calculated CV at minimum
Rangeability applies to the valve and is the ratio of the CV of the valve
fully open to the minimum CV at which it can control. On this basis, the
rangeability of the selected valve must exceed the turndown requirements
of the application. This is illustrated in Figure 1.7.

Calculated CV at min Calculated CV at max


Turndown of application

Min. controllable C CV fully open


Rangeability of selected valve

Figure 1.7.  Comparison of rangeability and turndown.


12  •   The Concise Valve Handbook

1.8 Velocity Profiles

One of the most important fluid characteristics affecting valve performance is


the shape of the velocity profile in the direction of flow since the predictions
used in sizing calculations are based on a fully developed turbulent flow.
In a frictionless pipe in which there is no retardation at the pipe walls,
a flat ‘ideal’ velocity profile would result (Figure 1.8) in which all the fluid
particles move at the same velocity.
Real fluids, however, do not ‘slip’ at a solid boundary, but are held to
the surface by the adhesive force between the fluid molecules and those
of the pipe. Consequently, at the fluid/pipe boundary, there is no relative
motion between the fluid and the solid.
At low flow rates, the fluid particles will move in straight lines in a
laminar manner, with each fluid layer flowing smoothly past adjacent layers
with no mixing between the fluid particles in the various layers. As a result,
the flow velocity increases from zero, at the pipe walls, to a maximum value
at the center of the pipe, and a velocity gradient exists across the pipe.
The shape of a fully developed velocity profile for such a laminar
flow is parabolic, as shown in Figure 1.9, with the velocity at the center
being equal to twice the mean flow velocity. Clearly, this concentration of
velocity at the center of the pipe can compromise the flow computation if
not corrected for.

Figure 1.8.  A flat ‘ideal’ velocity profile.

Figure 1.9.  A laminar ‘parabolic’ velocity


profile.
Basic Principles   •  13

Figure 1.10.  A turbulent velocity profile.

For a given pipe and liquid, as the flow rate increases, the flow of a
fluid will cease to be laminar and becomes turbulent. In turbulent flow, the
paths of the individual particles of fluid are no longer straight, but inter-
twine and cross each other in a disorderly manner so that thorough mixing
of the fluid takes place.
As shown in Figure 1.10, the velocity profile for turbulent flow is
flatter than for laminar flow, and thus closer approximates to the ‘ideal’ or
‘one dimensional’ flow.

1.9 Reynolds Number

The onset of turbulence is often abrupt, and in order to be able to predict


the type of flow present in a pipe, for any application, use is made of the
Reynolds number, Re—a dimensionless number given by:

ρ .V.D
Re =  (1.23)
µ
where:
ρ = density of fluid (kg/m3);
µ = viscosity of fluid (Pa.s);
v = mean flow velocity (m/s); and
D = diameter of pipe (m).
In practice, the Reynolds number is a ratio of the viscous and inertial
forces. If the viscous forces dominate (Re < 2,000), the flow is laminar,
and if the inertial forces dominate (Re > 3,000), the flow is turbulent.
At Re 2,000 to 3,000, the flow is said to be transitional.
The major significance of changes in the flow regime between tur-
bulent and laminar flow is that, for turbulent flow, the pressure loss is
proportional to the square of velocity and in laminar flow the losses are
linearly proportional to the velocity. This means that, for equivalent flow
14  •   The Concise Valve Handbook

rates, the differential pressure across the valve will be different for each
flow regime. As stated earlier, the predictions used in sizing calculations
are based on a fully developed turbulent flow in which inertial forces dom-
inate. If, due to a change in viscosity, the flow regime changes, then the
predicted calculations are no longer valid. As a result, a correction factor
can be applied where:

CVR = FV .CV  (1.24)

where:
CVR = required CV;
FV = Reynolds number factor; and
CV = valve flow coefficient.

1.10 Flashing and Cavitation

Flashing and cavitation are two related physical phenomena and are the
most common causes of control valve failure. Flashing and cavitation
cause structural damage to the valve and adjacent piping, and in order
to reduce or compensate for these undesirable effects, it is important to
understand the changes that occur to the medium as it passes through
the valve.
Flashing and cavitation only occur within liquids and takes place
whenever the internal pressure of the liquid falls below the vapor pres-
sure. This is illustrated in Figure 1.11 that shows a pressure gradient curve
through a valve where PV represents the vapor pressure of the flowing

P1
Pressure

Vapour pressure PV

P2

PVC

Distance

Figure 1.11.  Flashing occurs when the


downstream pressure is at or below the fluid
vapor pressure.
Basic Principles   •  15

fluid. As the liquid passes through the restriction, there is a decrease in


pressure. When the pressure at the vena contracta reaches the vapor pres-
sure of the fluid, vapor bubbles begin to form. In effect, the liquid is said to
be boiling. In water, at 100°C, this occurs at normal atmospheric pressure
of 1.01 bar. However, at 20°C, the pressure would need to be reduced to
23.8 mbar before vaporization occurs.

1.11 Flashing

If, as shown in Figure 1.11, the outlet pressure (P2) of the valve is at or
below the vapor (PV), the vapor bubbles remain intact and proceed fur-
ther downstream. This is known as flashing. The vaporization of liquid
causes a large increase in volume, and therefore higher overall fluid
velocity.
Liquid droplets suspended in a high-velocity vapor flow stream
impinging on metallic surfaces can cause physical damage to carbon
steel and cast iron. The damage is smooth and polished, similar to erosion
(Figure 1.12).
The majority of flashing damage occurs at the point of highest veloc-
ity—at or near the seat line of the valve plug or seat ring. The process is
two-stepped: a corrosion film forms at the surface that is subsequently
‘swept’ away by the high-velocity liquid flow. This cycle is then repeated.
The noise associated with flashing is a high-pitched hissing sound,
similar to that of sand passing through the valve.

Figure 1.12.  Damage due to flashing appears smooth and polished.


16  •   The Concise Valve Handbook

1.12 Cavitation

If, further downstream, the outlet pressure (P3) of the valve recovers to a
point above the vapor pressure of the fluid, the vapor bubbles will collapse
(Figure 1.13). This two-stage phenomenon, vapor bubble formation and
their subsequent collapse, is known as cavitation.
Implosion of the vapor bubbles produces large pressure shocks due to
microjet or spherical shock waves. If the vapor bubbles are close to or in
contact with a solid wall, pressure shocks of the order of 100,000 bar (1.5
million psi) are generated, producing both noise and physical damage. If
the bubbles collapse on or near solid surfaces, material is chipped away
(Figure 1.14). The amount of damage in a short period of time can be

P1

P2
Pressure

Vapour pressure PV

PVC

Distance
Figure 1.13.  Cavitation is caused by the pressure
dropping to the vapor pressure of the fluid and
rising to a higher pressure further downstream.

Figure 1.14.  Cavitation damage occurs when the bubbles


collapse on or near solid surfaces within the valve or
piping and material is chipped away.
Basic Principles   •  17

extensive and eventually prevents the control valve from performing its
intended function.
Because cavitation damage occurs where the bubbles collapse, the
effects will be evident downstream of the restriction and will show up
as very dull, rough, and pitted—sponge-like in appearance (Figure 1.15).
Corrosion, often mistaken as cavitation, takes on a similar appearance
with its affected area being more generalized or widespread.
The noise associated with cavitation is a high-pitched hissing
sound—usually accompanied by vibration—similar to that of gravel pass-
ing through the valve. The noise levels produced are rarely a problem to
the surrounding environment, but can be used as a good indication of the
severity of the cavitation (Figure 1.16).
Should cavitation be allowed to continue, the next manifestation
will be a loss of seat tightness as the damage starts to affect the seating
surfaces. Further use will cause the normal controlling position to pro-
gressively reduce, as the valve has to move toward the closed position to
compensate for the wear taking place.
As devastating as cavitation damage is, it is fortunate that its occur-
rence is associated with few process fluids. Hydrocarbon mixtures, such
as gasoline, do not have a fixed vapor pressure and will ‘boil’ over a rel-
atively wide temperature range. Gasoline boils from 40°C to 200°C. It is
felt that this results in an apparent buffeting effect, protecting the body
wall and other vital valve components. On the other hand, high surface
tension associated with water enhances the damage potential due to the
high related implosion stresses. In addition to collapse in the vicinity of
surfaces, the potential for damage is high when flowing fluid:

Figure 1.15.  Cavitation damage shows up as very dull, rough,


and pitted—sponge-like in appearance.
18  •   The Concise Valve Handbook

110

100

90
SPL dBA

80

70 Full cavitation

Incipient cavitation Flashing


60

50
0.02 0.04 0.06 0.1 0.2 0.3 0.4 0.6 0.8 1.0
ΔP
P1 − P 2

Figure 1.16.  Typical liquid noise characteristic (courtesy Fluids


handling: Principles & Practice).

• has well-defined vapor pressure;


• has high surface tension; and
• is not a mixture.

1.13 Leakage Classification

At the beginning of this chapter, we stated that a control valve designed


primarily to throttle energy is not necessarily designed for shut-off pur-
poses. In order to quantify the ability of a control valve to provide adequate
shut-off, when closed, the American National Standards Institute (ANSI)
has produced a set of standardized testing procedures for control valves.
The ANSI/FCI 70-2-1991 standard uses six different classifications
to describe the valves seat leakage capabilities. These are outlined in
Tables 1.3 and 1.4.
What do these figures mean in practical terms?
From equation (1.20), we had:

∆P
Q = 0.87 C v  (1.25)
SG f
Basic Principles   •  19

Table 1.3.  Control valve seat leakage classifications according to ANSI/


FCI 70-2-1991
Maximum
permitted Test Test
Class leakage medium pressures Testing procedure
I No test required provided that both supplier and user agree
II 0.5% of Air or 3 to 4 bar or Pressure applied
rated water maximum to valve inlet,
capacity at 10 to operating with outlet open
III 0.1% of 52°C differential, to atmosphere or
rated whichever connected to a low
capacity is lower head loss measuring
device, with full
IV 0.01%
normal closing
of rated
thrust provided by
capacity
actuator.
V 5 × 10 l/s
-9
Water at Maximum Pressure applied to
of water 10 to rated valve inlet after
per mm 52°C differential filling entire body
orifice pressure cavity and connected
per bar across piping with water
differential valve plug and stroking valve
plug closed. Use net
specified maximum
actuator thrust, but
no more, even if
available during
test. Allow time
for leakage flow to
stabilize.
VI Not to Air or 3.5 bar or Pressure applied
exceed nitrogen maximum to valve inlet.
amounts at 10 to operating Actuator should be
given, 52°C differential, adjusted to operating
according whichever conditions specified
to port is lower with full normal
diameter closing thrust
applied to valve plug
seat. Allow time
for leakage flow to
stabilize and use
suitable measuring
device.
20  •   The Concise Valve Handbook

Table 1.4.  Class VI maximum seat leakage according to ANSI/FCI


70-2-1991
Nominal port
­diameter Bubbles per minute(1)
ins mm ml per minute Bubbles per minute
1 25 0.15 1
1½ 38 0.30 2
2 51 0.45 3
2½ 64 0.60 4
3 76 0.90 6
4 102 1.70 11
6 152 4.00 27
8 203 6.75 45
(1)
Bubbles per minute as tabulated are a suggested alternative based on
a suitably calibrated measuring device; in this case, a 6.3 mm OD × 0.8
mm wall tube submerged in water to a depth of from 3 to 6 mm. The
tube end shall be cut square and smooth with no chamfer or burrs, and
the tube axis shall be perpendicular to the surface of the water. Other
apparatus may be constructed and the number of bubbles per minute may
differ from those shown as long as they correctly indicate the flow in ml
per minute.

Assume a 50 mm globe valve with a CV of 50.


With a differential pressure of 3.5 bar, the maximum water flow rate is:

3.5
Q = 0.87.50 = 81m3 /hr  (1.26)
1

The question now is, how long would it take for a valve meeting
ANSI requirements to fill a 10- litre (about 2.6 gallons) bucket?
Let us look at ANSI II, that is, 0.5% of rated capacity:
0.5% of 81 m3/hr = 0.405 m3/hr
= 405 l/hr
= 6.75 l/min
So, time taken to fill a 10 l bucket would be 1.5 minutes. Not very
good, is it?
What about ANSI III, that is, 0.1% of rated capacity?
0.1% of 81 m3/hr = 0.081 m3/hr
= 81 l/hr
= 1.35 l/min
Basic Principles   •  21

So, the time taken to fill a 10 l bucket would be nearly 7.5 minutes.
Still not very good.
And, ANSI IV, that is, 0.01% of rated capacity?
0.01% of 81 m3/hr = 0.0081 m3/hr
= 8.1 l/hr
= 0.135 l/min
Here, the time taken to fill a 10 l bucket would be just over 74 minutes.
And finally, ANSI V, that is, 5 × 10-9 l/s of water per mm orifice per
bar differential
5 × 10-9 × 50 × 3.5 × 60 × 1000 ml/min = 0.0525 ml/min = 132 days
4.5 hours

1.14 Isolation Valve Leakage


Classification

It should be noted that the foregoing ANSI seat leakage classes apply only
to control valves and not to isolation valves. An isolation valve is designed
solely for that purpose: to stop the flow of a gas or liquid. Consequently,
it is intended for use only in the fully open or closed positions, either to
divert the process media or to isolate it.
While control valves would generally be actuator-operated, isolation
valves can be actuator- or manually operated.
Obviously, the standards for seat leakage rates for isolation valves are
more rigorous than for control valves and are generally governed by API 598.
This states that, for all resilient-seated valves, there shall be no leak-
age at all during the specified test duration, with the period ranging from
15 to 120 s, dependent on the size of the valve. For liquids, 0 drops means
no visible leakage, while for gas, 0 bubbles means less than one bubble
during the test (Table 1.5).

Table 1.5.  Leakage rates for different-sized isolation valves based on


API 598
Metal-seated valves
Valve size Resilient- Liquid test (drops Gas test
(NPS) seated valves per min) (bubbles per min)
≤2 0 0 0
2½–6 0 12 24
8–12 0 20 40
≥ 14 0 2/min/inch NPS 4/min/inch NPS
Index

A C
Acoustic insulation, 119 Cage-guided control valve, 55,
Actual pressure drop, 27 58–59
Aerodynamic noise, 117 Camflex eccentric rotary plug
Air-operated pinch valve, 71, 72 valve, 86
Angle valve, 60 Cavitation
damage, 16–17
B elimination, 112–113
Ball segment valve, 81–82 flashing damage, 109
Ball valve liquid noise characteristics,
end-entry design, 76 17, 18
floating ball valve, 75 pressure dropping, 16
‘run’ torque, 76–77 reducing, 110–112
split-body design, 75, 76 Stellite region, 109
top-entry design, 76 trim selection guide, 109
trunnion (see Trunnion ball Center-disc butterfly valve, 82, 83
valve) Check valve, 88, 89
Bar stock body valve, 61, 62 Choked flow, 8–9
Bellows seal bonnet Choked pressure drop, 27
accordion-like tube, 52 Compressibility factor, 43
formed-type bellows, 53 Control valve
leakage detection, 54 basic theory, 3
lifespan, 54 Bernoulli’s equation, 5–8
welded bellows, 53 cavitation, 16–18
Bernoulli’s equation, 5–8 choked flow, 8–9
Bonnet assembly, 49 equation of continuity, 3–4
Butterfly valve final control element, 1–3
center-disc butterfly valve, 82, 83 flashing and cavitation, 14–15
circular disc-shaped damper, 82, fluid properties, 1
83 isolation valve leakage
conventional disc shape, 84, 85 classification, 21
fishtail disc, 84, 85 leakage classification, 18–21
flow characteristics, 84, 85 pressure recovery, 9–11
offset disc, 84 purpose of, 1
136  •   Index

rangeability, 11 Flangeless connections, 99–100


Reynolds number, 13–14 Flat-face flanged end connection,
turndown, 11 96
velocity and pressure Flat ‘ideal’ velocity profile, 12–13
distribution, 3, 4 Flexible wedge gate, 63–64
velocity profiles, 12–13 Flow characteristics
inherent (see Inherent flow
D characteristics)
Diaphragm bellows, 53 installed (see Installed flow
Diaphragm valve characteristics)
advantage, 72 Formed-type bellows, 53
maintenance, 73
operating temperature, 72–73 G
pressure rating, 73 Gas expansion factor, 40–41
Saunders patent valve, 73 Gas valve sizing
straight-through diaphragm compressibility factor, 43
valves, 74 Fisher Emerson easy-e® ES cage-
throttle flow, 72 guided globe valve, 43–46
Discharge coefficient, 7 gas expansion factor, 40–41
Double seat bottom-guided globe mass flow rates, 42
valve, 57 pressure drop mechanism, 33–38
specific heat ratio factor, 38–40
E volumetric flow rates, 41
Eccentric plug valve Gate valve
Camflex eccentric rotary plug expanding gate valve, 65–67
valve, 86 full and no flow conditions, 62
integral extension bonnets, 87 isolation applications, 61
seat portion, 86, 87 knife edge, 67–71
spherically shaped, 87 wedge shape, 62, 63
End connections Globe valve
flanged end connections, 95–96 advantages, 48–49
hub end body, 96, 97 disadvantage, 49
screwed end valve connections, globular-shaped cavity, 48
94–95 size, 49
welded end connections, 96–98 Grayloc® connector, 100–101
Equal percentage inherent flow
characteristics, 104 H
Equation of continuity, 3–4 Hub end body, 96, 97
Expanding gate valve Hydrodynamic noise, 116
centralizing mechanism, 66
segmented assemblies, 65–67 I
Inherent flow characteristics
F cage-guided trim, 105, 106
Final control element, 1–3 curves, 103–104
Flanged end connections, 95–96 equal percentage, 104
Index   •   137

linear, 103 M
modified percentage, 105 Manipulated variable (MV), 2–3
plug outlines, 105, 106 Mechanical noise, 115–116
quick opening, 104 Modified percentage inherent flow
Installed flow characteristics characteristics, 105
differential pressure drop, 105, Molecular weight of gas, 41, 42
106 Mufflers, 120
equal percentage, 108
linear, 107 N
pressure drop ratio, 107 Needle valve, 61
valve selection guide, 108 Noise
Isolation valve leakage acoustic insulation, 119
classification, 21 aerodynamic noise, 117
control, 117–118
K hydrodynamic noise, 116
Knife edge gate valve mechanical noise, 115–116
advantages, 67 path treatment, 118–119
disc alignment, 68, 70 permissible personnel noise level
flow/travel characteristics, 68, 69 exposure, 114
sliding gate regulator valve, 68, prediction, 117
70, 71 silencers, 120
V-insert gate valve, 68, 69 sounds and activities, 113, 114
source treatment, 120–122
L Nominal bore (NB), 88–90
Laminar ‘parabolic’ velocity Nominal pipe size (NPS), 88–90
profile, 12
Laminated graphite packing O
benefits, 51 Offset disc butterfly valve, 84
double graphite packing
arrangement, 51 P
electrochemical reaction, 51, 52 Packing box, 49, 50
sacrificial zinc washers, 51 Pinch check valve, 88, 89
temperature, 50 Pinch valve
Lap joint flange, 98–99 air-operated pinch valve, 71, 72
Linear inherent flow applications, 71
characteristics, 103 double vise mechanism, 71, 72
Liquid valve sizing rubber hose/sleeve, 69, 71
Emerson Fisher easy-e®, 27–29 turbulence, 70
identical inlet and outlet fittings, Piping geometry factor, 29–31
31–32 Plug valve
oversized valves, 23 conventional, 84–85
PC-based software package, lubricated valve, 85
23–24 PTFE-lined valve, 86
piping geometry factor, 29–31 sealing mechanism, 85
valve sizing programs, 24 Pressure drop mechanism
138  •   Index

choked flow, 37 Streamlined valves, 10


different valve styles, 38, 39 Stuffing box, 49
flow rate, 36 Swing check valve, 88
gas mass flow, 33–34
globe valve set, 37, 38 T
terminal pressure drop ratio, 34, Terminal pressure drop ratio, 34
35 Top and bottom-guided double
Pressure drop ratio (PDR), 107 seat, 56–57
Pressure recovery coefficient, Trunnion ball valve
10–11 double block and bleed system,
Process demand (PD), 2–3 78, 79
full port design, 78–79
Q fully open position, 79, 80
Quick opening inherent flow hydraulic load, 77
characteristics, 104 leakage integrity, 77, 78
line pressure, 77, 78
R pathway, 79–80
Raised-face flanged end Turbulent velocity profile, 13
connection, 96 Turndown, 11
Rangeability, 11
Reynolds number, 13–14 V
Rotary control valves, 74 Valve authority, 107
Valve construction
S angle valve, 60
Sacrificial anode, 51 ball segment valve, 81–82
Saunders patent valve, 73 ball valve, 75–77
Screwed end valve connections, bar stock body valve, 61
94–95 bellows seal bonnet, 52–54
Silencers, 120 bonnet assembly, 49
Single-ported balance globe valve, butterfly valve, 82–85
57–58 cage-guided control valve, 58–59
Single-seat angle valve, 60 cavitation control, 108–110
Single seat top-guided contact check valve, 88, 89
valve, 56 control assembly, 47
Slab valve, 64–65 corrosion, 90, 94
Sliding gate regulator valve, 68, diaphragm valve, 72–74
70, 71 eccentric plug valve, 86–87
Solid wedge gate, 62–63 end connections (see End
Specific gravity, 41 connections)
Specific heat ratio factor, 38–40 erosion, 94
air, 38–40 extended bonnet, 52
natural gas, 40 flangeless connections, 99–100
Split body globe valve, 59–60 gate valve (see Gate valve)
Straight-through diaphragm globe valve, 48–49
valves, 74 Grayloc® connector, 100–101
Index   •   139

guiding system, 55 cavitation, 109–113


laminated graphite, 50–52 flow characteristics
lap joint flange, 98–99 inherent (see Inherent flow
material selection, 90, 93 characteristics)
needle valve, 61 installed (see Installed flow
pinch valve, 69–72 characteristics)
plug valve, 84–86 noise (see Noise)
post-guiding, 55–56 noise sources, 113–118
PTFE packing, 49–50 plug, 54
rotary control valves, 74 seat retainer, 54–55
single-ported balance globe seat ring, 54
valve, 57–58 Velocity control
slab valve, 64–65 exit jet independence, 122
sliding stem valves, 47 frequency spectrum shift, 122
split body globe valve, 59–60 low noise retainer device, 120,
top and bottom-guided double 121
seat, 56–57 multistage pressure reduction,
trunnion ball valve, 77–80 121
valve sizes and pipe schedules, passage shape, 121
88–90 velocity management, 122
valve trim, 54–55 Velocity of Approach Factor (EV),
valve types, 48 7
wedge gate, 62–64 Velocity profiles, 12–13
Valve flow coefficient, 7, 8 flat ‘ideal’ velocity profile, 12
Valve selection laminar ‘parabolic’ velocity
actuator and accessory selection profile, 12
guide, 128 turbulent, 13
body selection guide, 127 Venturi seat rings, 60
control valve selection guide, V-notched ball segment, 81
126
decision criteria, 123 W
technical requirements, 123–126 Wedge gate, 62–64
valve trim selection guide, 127 Welded bellows, 53
Valve trim Welded end connections,
cage-guided control valve, 55 96–98

You might also like