You are on page 1of 1121

cover next page >

title : Foundations of Vacuum Science and Technology


author : Lafferty, J. M.
publisher : John Wiley & Sons, Inc. (US)
isbn10 | asin : 0471175935
print isbn13 : 9780471175933
ebook isbn13 : 9780585339368
language : English
subject Vacuum, Kinetic theory of gases.
publication date : 1998
lcc : QC166.F68 1998eb
ddc : 621.5/5
subject : Vacuum, Kinetic theory of gases.
cover next page >
< previous page page_iii next page >
Page iii

Foundations of Vacuum Science and Technology

Edited By
James M. Lafferty

< previous page page_iii next page >


< previous page page_iv next page >
Page iv

This book is printed on acid-free paper.

Copyright © 1998 by John Wiley & Sons, Inc. All rights reserved.
Published simultaneously in Canada.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means,
electronic, mechanical, photocopying, recording, scanning or otherwise, except as permitted under Sections 107 or 108
of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization
through payment of the appropriate per-copy fee to the Copyright Clearance Center, 222 Rosewood Drive, Danvers,
MA 01923, (508) 750-8400, fax (508) 750-4744. Requests to the Publisher for permission should be addressed to the
Permissions Department, John Wiley & Sons, Inc., 605 Third Avenue, New York, NY 10158-0012, (212) 850-6011, fax
(212) 850-6008, E-mail: PERMREQ @ WILEY.COM.

Library of Congress Cataloging in Publication Data:


Foundations of vacuum science and technology/edited by J. M. Lafferty.
p. cm.
"A Wiley-Interscience publication."
Includes bibliographical references and index.
ISBN 0-471-17593-5
1. Vacuum. 2. Kinetic theory of gases. I. Lafferty, J. M.
(James Martin), 1916.
QC166.F68 1997
621.5′5dc21 96-29895

Printed in the United States of America

10 9 8 7 6 5 4 3 2 1

< previous page page_iv next page >


< previous page page_v next page >
Page v

Preface

The industrial and scientific importance of vacuum technique has continued to grow during the past 30 years, even with
the demise of television and radio receiving tubes and many gas-filled tubes. The replacement of vacuum and low-
pressure gas discharge tubes by semiconductors and integrated circuits has placed an even greater demand and more
stringent conditions on vacuum technique for the processing and manufacture of these devices. This has led to the
development of a number of "dry" vacuum pumps to produce a "clean vacuum" free of hydrocarbons. The space
programs, high-energy accelerators, analytical instruments, freeze-drying of foods and drugs, and the manufacture of
color television picture tubes, incandescent and metal vapor lamps, high-power vacuum, and x-ray tubes all continue to
require the need for vacuum technique.

While the material in this book is totally new, it follows in the tradition set by Scientific Foundations of Vacuum
Technique by Saul Dushman, published in 1949. That book enjoyed unprecedented success and is now a classic in its
field. By 1960 it was badly in need of revision. This editor had the privilege of participating as editor of the revised
edition, which was published in 1962. This second edition was brought up to date by a number of contributors with
specialized knowledge in the disciplines involved. An attempt was made to introduce the new developments made in
vacuum technique but keep the original plan of the book and retain much of the material that was still of current interest.

The editor was encouraged by Leonard Beavis of the American Vacuum Society Education Committee and the
publishers of the previous editions to undertake the publication of the present volume. The advances made in vacuum
science and technology during the past three decades has required a complete reworking of the material in the previous
volume. However, every effort has been made to follow the unique style of the original bookthat is, to present a survey
of fundamental ideas in physics and chemistry that would be useful to both scientists and engineers dealing with
problems associated with the use, production, and measurement of high vacuums. This volume is a critical survey of
important developments in vacuum technique with many references for those who seek a better understanding and more
detailed information in the field. It is not a vacuum handbook, many of which are listed in the Appendix of this book.

Every effort was made to select on a worldwide basis a number of outstanding vacuum specialists who were willing to
take time to contribute to this volume. With

< previous page page_v next page >


< previous page page_vi next page >
Page vi

the curtailment of vacuum research in the major industrial laboratories, one now only finds vacuum experts as
independent consultants or in companies manufacturing vaccum systems and components, a few educational
institutions, and government laboratories.

While the basic laws of kinetic theory of gases have not changed over the years, a better understanding of gas flow over
a wide range of pressures has necessitated an expanded chapter on the subject. It now encompasses all flow regimes
from free molecular flow to atmospheric pressure. It treats compressible flow through tubes and orifices under choked
and nonchoked conditions as well as turbulent flow in ducts of any cross section.

Many topics that were only mentioned in the second edition of Scientific Foundations of Vacuum Technique now have
full chapters devoted to them. The progress made in vacuum pumps over the past three decades, for example, is
remarkable. Three chapters are now devoted to this subject. Detailed information is given for the first time on liquid
ring pumps, dry pumps, turbo pumps, getter pumps, and cryopumps.

The subject of leak detection, which had only a few paragraphs devoted to it in the old edition, now has a full chapter
describing leak detectors as a rugged industrial tool for everyday use capable of quantitative measurements.

Information on the design of high-vacuum systems has been expanded to help the reader in selecting pump sets for
various system applications and in predicting their performance.

Pressure measurements continue to be important on every vacuum system. This subject is fully covered in the chapter
on vacuum gauges. While the ionization gauge continues to be the principal pressure sensor for measuring total pressure
in high- and ultrahigh-vacuum systems, modern solid-state electronics has simplified its use. The accuracy of this
device in measuring pressure depends on a knowledge of the composition of the gas being measured. The partial
pressure analyzer has become a far more sophisticated way to measure pressure and give the vacuum system operator an
insight of what is occurring within the system. The invention of the quadrupole mass spectrometer with solid-state
electronics has done much to make partial pressure measurements relatively simple and inexpensive. A full chapter is
devoted to this subject.

In discussing pressure measurements, a word about pressure units seems appropriate. While use of the pascal, the ISO
unit of pressure, has been encouraged in this book, many of the European contributors strongly preferred using the
millibar (mbar). The mbar falls in a class of units that are temporarily accepted for use by the ISO. You will find both
units in this book. The advantage of the mbar is that it is nearly equal in magnitude to the Torr or mmHg found in earlier
publications and is familiar to many readers (1 mbar = 0.75 Torr). When several orders of magnitude of pressure are
plotted on a log scale, the mbar and Torr plots are nearly indistinguishable. Some of the figures in this book that have
been copied from earlier publications may still have the pressure plotted in Torr.

Ultrahigh-vacuum technique had its infancy in midcentury. Today it is a matured procedure used in a great variety of
applications and in commercially available equipment. Researchers do not appear to have reached a limit yet in their
quest to produce and measure a perfect vacuum. This work is described in the chapter devoted to ultrahigh and extreme
high vacuum. Considerable progress has been made in this area by pushing vacuum techniques to their limit and gaining
a better

< previous page page_vi next page >


< previous page page_vii next page >
Page vii

understanding of the gassurface interactions and diffusion in solids as described in the chapter on this subject. While
pressures as low as 1011 Pa have been measured in the laboratory, this is still at least three orders of magnitude higher
than that in interstellar space.

The final chapter is devoted to calibration and standards. It describes the physical background and state of the art of
today's primary vacuum standards in the various national laboratories. It should be useful reading not only for those
involved in calibration and quality control but for those interested in the accuracy limitations of various vacuum
instruments.

The editor is indebted to several people for suggestions concerning this volume. Special mention is made of the late
Hermann Adam for helpful discussions and for suggesting a number of German contributors for the book. John Weed, a
member of the American Vacuum Society Education Committee, solicited suggestions for the volume from a number of
A.V.S. members. Nigel Dennis coordinated chapters three and four on vacuum pumps, and Benjamin Dayton and Paul
Redhead have made many helpful suggestions.

J. M. LAFFERTY
SCHENECTADY, NEW YORK

< previous page page_vii next page >


< previous page page_ix next page >
Page ix

Contributors

Helmut Bannwarth, LEDERLE GmbH, Gundelfingen, Germany

Benjamin B. Dayton, Consultant, East Flatrock, North Carolina, USA

Nigel T. M. Dennis, Edwards High Vacuum International, Crawley, West Sussex, England

Johan E. de Rijke, Vacuum Technical Services, Morgan Hill, California, USA

Robert E. Ellefson, Leybold Inficon, Inc., East Syracuse, New York, USA

Bruno Ferrario, SAES Getters S.p.A., Lainate (Milano), Italy

Werner Grosse Bley, Leybold Vakuum GmbH, Cologne, Germany

Hinrich Henning, Leybold Vakuum GmbH, Cologne, Germany

Jörgen Henning, intervac Henning, GmbH, Kreuzwertheim, Germany

John B. Hudson, Materials Science and Engineering Department, Rensselaer Polytechnic Institute, Troy, New York,
USA

Karl Jousten, Physikalisch-Technische Bundesanstalt, Berlin, Germany

R. Gordon Livesey, Edwards High Vacuum International, Crawley, West Sussex, England

R. Norman Peacock, MKS Instruments, HPS Division, Boulder, Colorado, USA

Paul A. Redhead, Institute for Microstructural Sciences, National Research Council, Ottawa, Ontario, Canada

Wolfgang Schwarz, Leybold Systems GmbH, Hanau, Germany

< previous page page_ix next page >


< previous page page_xi next page >
Page xi

Contents

Preface v

Contributors ix

Acronyms xxiii

1. Kinetic Theory of Gases 1


Benjamin B. Dayton

2
1.1. Ideal Gas Law

6
1.2. Avogadro's Number

8
1.3. Molecular Collisions; Mean Free Path; MaxwellBoltzmann Distribution Laws

16
1.3.1. Relation Between Molecular Velocities and Velocity of Sound

17
1.3.2. Determination of Avogadro's Constant from Distribution of Particles in Brownian Motion

18
1.4. Gas Pressure and Rate at Which Molecules Strike a Surface

22
1.5. Rate of Evaporation and Vapor Pressure

26
1.6. Free Paths of Molecules

29
1.7. Relation Between Coefficient of Viscosity, Mean Free Path, and Molecular Diameter

37
1.7.1. Viscosity at Low Pressures

39
1.7.2. Molecular Diameters
39
1.7.3. Application of the van der Waals Equation

40
1.7.4. From the Density of the Solid or Liquid

41
1.7.5. Cross Section for Collision with Electrons

41
1.8. Heat Conductivity of Gases

44
1.9. Thermal Conductivity at Low Pressures

46
1.9.1. Free-Molecule Conductivity (Knudsen)

50
1.9.2. Temperature Discontinuity (Smoluchowski)

< previous page page_xi next page >


< previous page page_xii next page >
Page xii

53
1.10. Thermal Transpiration (Thermomolecular Flow)

57
1.11. Thermal Diffusion

62
1.12. Theory of Diffusion of Gases

65
1.12.1. MaxwellLoschmidt Method for Determination of Diffusion Coefficients

67
1.12.2. Effect of Pressure of Gas on Rates of Evaporation of Metals

69
1.13. Random Motions and Fluctuations

71
1.14. Scattering of Particle Beams at Low Gas Pressures

73
References and Notes

2. Flow of Gases Through Tubes and Orifices 81


R. Gordon Livesey

83
2.1. Flow Conductance, Impedance, and Gas Throughput

85
2.2. Molecular Flow

86
2.2.1. Conductance of an Aperture

87
2.2.2. General Considerations for Long Ducts

87
2.2.3. General Considerations for Short Ducts

88
2.2.4. Uniform Circular Cross Section

90
2.2.5. Duct of Uniform Rectangular Cross Section
92
2.2.6. Tube of Uniform Elliptical Cross Section

93
2.2.7. Cylindrical Annulus (Flow Between Concentric Cylinders)

94
2.2.8. Uniform Triangular Section (Equilateral)

94
2.2.9. Other Shapes

96
2.2.10. Combinations of Components

102
2.2.11. Cases of Unsteady Flow

105
2.3. Continuum Flow

108
2.3.1. Viscous Laminar Flow

112
2.3.2. Turbulent Flow

116
2.3.3. Compressible Flow

119
2.3.3.1. Flow through an Aperture or Short Duct

121
2.3.3.2. Approximation for Flow Through an Aperture

121
2.3.4. Corrections for Flow Obstructions

122
2.3.5. ApproximationsEntrance Correction Model

124
2.3.6. ApproximationsKinetic Energy Model

126
2.3.7. Long Duct Criteria

128
2.4. Transitional Flow

129
2.4.1. Transitional Flow in Long Ducts
134
2.4.2. Long Duct Criterion in Transitional Flow

135
2.4.3. Transitional Flow through Apertures and Short Ducts

137
Symbols

139
References

< previous page page_xii next page >


< previous page page_xiii next page >
Page xiii

3. Positive Displacement Vacuum Pumps 141

Part I. Oil-Sealed Vacuum Pumps 143


Nigel T. M. Dennis

143
3.1. Oil-Sealed Vacuum Pumps

143
3.1.1. Pump Design

144
3.1.2. Gas Ballast

147
3.1.3. Pump Oil

148
3.1.4. Oil Suckback

148
3.1.5. Power Requirements and System Protection

149
3.1.6. Accessories

Part II. Liquid Ring Pumps 151


Helmut Bannwarth

151
3.2. Liquid Ring Pumps

151
3.2.1. Mechanism

152
3.2.2. Single-Stage Liquid Ring Vacuum Pumps

153
3.2.3. Two-Stage Liquid Ring Vacuum Pumps

154
3.2.4. The Operating Liquid

154
3.2.5. Operating Ranges of Liquid Ring Gas Pumps
154
3.2.6. Cavitation and Protection Against Cavitation

156
3.2.7. Types of Operation; Conveyance of Operating Liquid

157
3.2.8. Materials of Construction

157
3.2.9. Sealing

157
3.2.10. Drives

158
3.2.11. Accessories

Part III. Dry Vacuum Pumps 159


Nigel T. M. Dennis

159
3.3. Dry Vaccum Pumps

159
3.3.1. Roots Pump

162
3.3.2. Claw Pump

164
3.3.3. Screw Pump

167
3.3.4. Scroll Pump

169
3.3.5. Piston and Diaphragm Pumps

170
References

171
General References

4. Kinetic Vacuum Pumps 173

Part I. Diffusion and Diffusion-Ejector Pumps 175


Benjamin B. Dayton

176
4.1. Diffusion Pumps
176
4.1.1. History of Development

181
4.1.2. Diffusion Pump Design

< previous page page_xiii next page >


< previous page page_xiv next page >
Page xiv

183
4.2. Diffusion-Ejector Pumps

185
4.3. Performance of Vapor-Jet Pumps

185
4.3.1. Pumping Speed

187
4.3.2. Limiting Forepressure for Maximum Speed

190
4.3.3. Influence of Nozzle and Entrance Chamber Design on Speed

192
4.3.4. Ultimate Pressure

194
4.3.5. Backstreaming and Back Migration of Pump Fluid

198
4.3.6. Throughput

202
4.4. Theory of Pump Performance

202
4.4.1. Speed

204
4.4.2. Limiting Forepressure

205
4.4.3. Vapor-Jet Flow Pattern

221
4.4.4. Ultimate Pressure

Part II. Molecular Drag and Turbomolecular Pumps 233


Jörgen Henning

233
4.5. Molecular Drag Pumps

234
4.5.1. Theoretical Considerations and Performance Data
237
4.5.2. Design Considerations

237
4.5.3. Typical Performance Data of Commerical Pumps

238
4.5.3.1. Compression

238
4.5.3.2. Pumping Speed

238
4.5.3.3. Ultimate Pressure

238
4.6. Turbomolecular Pumps

239
4.6.1. Theoretical Considerations and Performance Data

241
4.6.2. Design Considerations

241
4.6.2.1. Rotor and Stator Geometry

242
4.6.2.2. Rotor Suspension

242
4.6.2.3. Lubrication of Mechanical Bearings

242
4.6.2.4. Magnetic Rotor Suspension

243
4.6.2.5. Balancing and Vibration

243
4.6.2.6. Rotor Materials

243
4.6.2.7. Drive Systems

243
4.6.3. Applicational Considerations

243
4.6.3.1. Venting

244
4.6.3.2. Baking
244
4.6.3.3. Cooling

244
4.6.3.4. Operation in Magnetic Fields

245
4.6.3.5. Pumping Corrosive Gases

245
4.6.3.6. Pumping Toxic or Radioactive Gases

245
4.6.3.7. Turbomolecular Pumps in Combination with Other Pumps

< previous page page_xiv next page >


< previous page page_xv next page >
Page xv

245
4.6.4. Performance Data of Commercial Pumps

246
4.6.4.1. Compression

246
4.6.4.2. Pumping Speed

247
4.6.4.3. Ultimate Pressure

247
4.7. Combined Molecular Drag and Turbomolecular Pumps

248
4.7.1. Design Considerations

248
4.7.2. Typical Performance Data for Commericial Combined Molecular Drag and Turbomolecular
Pumps

248
4.7.2.1. Compression

248
4.7.2.2. Pumping Speed

248
4.7.2.3. Ultimate Pressure

248
4.8. Backing Pumps

Part III. Regenerative Drag Pumps 251


Nigel T. M. Dennis

251
4.9. Regenerative Drag Pumps

251
4.9.1. Mechanism

254
References

5. Capture Vacuum Pumps 259


Part I. Getters and Getter Pumps 261
Bruno Ferrario

261
5.1. Types of Gas Surface Interactions

262
5.2. Basic Concepts of Getter Materials

263
5.3. Adsorption and Desorption

265
5.4. Bulk Phenomena

265
5.4.1. Diffusion

267
5.4.2. Solubility

268
5.5. Equilibrium Pressures

269
5.6. Getter Materials

269
5.6.1. Basic Characteristics of Getter Materials

269
5.6.2. Sorption Speed and Sorption Capacity

271
5.6.3. Principal Types of Getter Materials and Their General Working Conditions

275
5.6.4. Interaction of Getters with Common Residual Gases

275
5.6.5. Evaporable Getters

276
5.6.5.1. Ba Getters

291
5.6.5.2. Titanium Sublimation Getter Pumps

297
5.6.6. Nonevaporable Getters
305
5.6.6.1. Ternary Alloys

310
5.6.6.2. Other Ternary and Multicomponent Alloys

< previous page page_xv next page >


< previous page page_xvi next page >
Page xvi

310
5.7. Getter Configurations

313
5.8. Getter Applications

314
5.8.1. Nonevaporable getters Versus evaporable getters

315
5.8.2. Start-up and Working Conditions of Getters

Part II. Sputter Ion Pumps 317


Hinrich Henning

317
5.9. Gas Discharge Vacuum Pumps

319
5.10. The Penning Discharge

321
5.10.1. Pump Sensitivity

323
5.10.2. Ion Motion

324
5.10.3. Electron Cloud

326
5.10.4. Secondary Electrons

327
5.10.5. Transition from HMF Mode to HP Mode

328
5.10.6. Transition from LMF Mode to HMF Mode

329
5.10.7. Sputtering

329
5.11. SIP Characteristics

329
5.11.1. Gettering
330
5.11.2. Ion Burial

331
5.11.3. Volume Throughput

335
5.11.4. Pumping Mechanism

338
5.11.5. Bakeout

338
5.11.6. Types of SIPs

342
5.11.7. Starting Properties

343
5.11.8. Memory Effect

343
5.11.9. Ultimate Pressure

345
5.11.10. Magnets

Part III. Cryopumps 347


Johan E. de Rijke

348
5.12. AdsorptionDesorption

352
5.13. Cryotrapping

353
5.14. Pumping Speed and Ultimate Pressure

355
5.15. Capacity

357
5.16. Refrigeration Technology

359
5.17. Pump Configuration

363
5.18. Regeneration
364
5.19. Partial Regeneration

364
5.20. Sorption Roughing Pumps

368
References

373
General References

< previous page page_xvi next page >


< previous page page_xvii next page >
Page xvii

6. Vacuum Gauges 375


R. Norman Peacock

377
6.1. Pressure Units Used in Vacuum Measurements

378
6.2. Liquid Manometers

379
6.3. McLeod Gauge

381
6.4. Piston Pressure Balance Gauge

382
6.5. Bourdon Gauge

384
6.6. Capacitance Diaphragm Gauges

385
6.6.1. Sensitivity of the Capacitance Method

386
6.6.2. Deflection of a Thin Tensioned Membrane

387
6.6.3. Accuracy of Commercial Gauges

388
6.6.4. Thermal Transpiration

388
6.6.5. Conclusions

389
6.7. Viscosity Gauges

391
6.7.1. Spinning Rotor Gauge

391
6.7.1.1. Theory

394
6.7.1.2. Commercial Gauges
397
6.7.1.3. Stability

399
6.7.1.4. Secondary or Transfer Standard

401
6.7.1.5. Use Precautions

401
6.7.1.6. Advantages and Disadvantages

402
6.7.2. Oscillating Quartz Crystal Viscosity Gauge

402
6.7.2.1. Advantages and Disadvantages

403
6.8. Thermal Conductivity Gauges

404
6.8.1. Theory

406
6.8.2. Calibration

408
6.8.3. Lowest Useful Pressure

409
6.8.4. Constant Pressure Pirani

410
6.8.5. Calibration Dependence Upon the Gas

410
6.8.6. Upper Pressure Limit

411
6.8.7. Ambient Temperature Compensation

412
6.8.8. Comparison of Pirani and Thermocouple Gauges

412
6.8.9. Stability

412
6.8.10. Thermistor Pirani Gauges and Integrated Transducers

413
6.8.11. Commercial Gauges and Applications
414
6.9. Ionization Gauges

414
6.9.1. Hot-Cathode Gauge Equation

419
6.9.2. Geometric Variations in the BayardAlpert Gauge

421
6.9.3. Modulated BayardAlpert Gauge

422
6.9.4. Extractor Gauge

423
6.9.5. Helmer Gauge

424
6.9.6. Long Electron Path Length Gauges

425
6.9.7. Secondary Standard Hot-Cathode Gauges

426
6.9.8. High-Pressure Ionization Gauges

427
6.9.9. Cold-Cathode Gauges

435
6.9.10. Ionization Gauge Accuracy

< previous page page_xvii next page >


< previous page page_xviii next page >
Page xviii

438
6.9.11. Gauge Constant Ratios for Different Gases

439
6.9.12. Ionization Gauge Controllers

441
References

7. Partial Pressure Analysis 447


Robert E. Ellefson

447
7.1. Ion Sources

448
7.1.1. Electron-Impact Ionization Process

449
7.1.2. Open Ion Source

452
7.1.3. Closed Ion Source

454
7.2. Ion Detection

454
7.2.1. Faraday Cup Ion Detection

454
7.2.2. Secondary Electron Multiplier Detection

456
7.2.3. Microchannel Plate Detector

456
7.3. Mass Analysis

456
7.3.1. Quadrupole Mass Analyzer

460
7.3.2. Magnetic Sector Analyzer

464
7.3.3. Time-of-Flight Mass Analyzer
465
7.3.4. Trochoidal (Cycloid) Mass Analyzer

466
7.3.5. Omegatron

467
7.4. Optical Measurement of Partial Pressures

468
7.4.1. Photoionization Measurement of Partial Pressure

469
7.4.2. Infrared Absorption Measurement of Partial Pressure

470
7.5. Computer Control, Data Acquisition, and Presentation

471
7.6. Residual Gas Analysis

474
7.7. Pressure Reduction Sampling Methods for Vacuum Process Analysis

475
7.8. Calibration of Partial Pressure Analyzers

477
References

8. Leak Detection and Leak Detectors 481


Werner Grosse Bley

482
8.1. Principles of Vacuum Leak Detection

482
8.1.1. Types of Leaks and Leak Rate Units

484
8.2. Total Pressure Measurements

486
8.3. Partial Pressure Measurements

486
8.4. Measurement of Leakage Rates with Helium Leak Detectors

487
8.5. Helium Leak Detection of Vacuum Components
490
8.6. Helium Leak Detection of Vacuum Systems

493
8.7. Special Methods and Other Tracer Gases

493
8.8. Mass Spectrometer Leak Detectors

494
8.8.1. Mass Spectrometer System for Helium Leak Detection

494
8.8.2. Direct-Flow Helium Leak Detectors

496
8.8.3. Simple Counterflow Helium Leak Detectors

< previous page page_xviii next page >


< previous page page_xix next page >
Page xix

498
8.8.4. Advanced Counterflow Helium Leak Detectors

499
8.8.5. Oil-Free and Dry Helium Leak Detectors

500
8.9. Specifications of Mass Spectrometer Leak Detectors

502
8.10. Quantitive Leakage Rate Measurements

504
8.11. Mass Spectrometer Leak Detectors for Other Tracer Gases and Future Developments in Leak
Detection

505
References

9. High-Vacuum System Design 507


Wolfgang Schwarz

507
9.1. Calculations of Vacuum Systems

508
9.1.1. Basic Pumpdown Equations

511
9.1.2. Process Pressure

513
9.2. Gas Loads in High-Vacuum Systems

513
9.2.1. Outgassing

516
9.2.2. Leaks

516
9.2.3. Permeation

518
9.2.4. Process Gas

519
9.3. Design of High-Vacuum Pump Sets
519
9.3.1. Forepump sets

519
9.3.1.1. Fore-Vacuum Pumps

520
9.3.1.2. Roots Combinations

524
9.3.2. High-Vacuum Pump Sets

526
9.3.2.1. Turbomolecular Pump Sets

528
9.3.2.2. Diffusion Pump Sets

531
9.3.2.3. Pump Sets with Cryosurfaces

535
9.3.2.4. Cryopump Sets

537
9.4. Calculation Methods for Vacuum Systems

538
9.4.1. Analytical Approximations

541
9.4.2. Numerical Methods

541
9.4.2.1. Dedicated Software

542
9.4.2.2. Network Approach

546
General References

10. GasSurface Interactions and Diffusion 547


John B. Hudson

548
10.1. Adsorption

548
10.1.1. Basic Equations
551
10.1.2. Adsorption Isotherms

567
10.1.3. Heat of Adsorption

568
10.1.4. Observed Behavior

572
10.1.5. Adsorption Kinetics

575
10.1.6. Chemisorption Kinetics

< previous page page_xix next page >


< previous page page_xx next page >
Page xx

582
10.1.7. Kinetic Measurements

584
10.1.8. Capillarity Effects

589
10.2. Absorption

590
10.2.1. Equilibrium Solubility

590
10.2.2. Diffusion Rates

591
10.2.3. Kinetics of Absorption and Permeation

592
10.2.4. Steady-State Permeation

595
10.2.5. Transient Permeation

600
10.2.6. Effect of Desorption Kinetics on Permeation

606
10.3. Surface Chemical Reactions

614
10.4. Outgassing Behavior

614
10.4.1. Desorption of Adsorbed Gases

616
10.4.2. Dissolved Gases

616
10.4.3. Overall Pumpdown Curves

619
10.4.4. Mitigation of Outgassing

619
10.4.5. Surface Treatments During Construction
620
10.4.6. In Situ Surface Treatments

620
10.4.7. Bakeout Processes

622
References

11. Ultrahigh and Extreme High Vacuum 625


Paul A. Redhead

628
11.1. Limits to the Measurement of UHV/XHV

629
11.1.1. Residual Currents

636
11.1.2. Effects at Hot Cathodes

639
11.1.3. Gauges with Long Electron Paths

641
11.1.4. Comparison of UHV/XHV Gauges

642
11.2. Limits to Pumps at UHV/XHV

642
11.2.1. Kinetic Pumps

643
11.2.2. Capture Pumps

646
11.2.3. Comparisons of Pumps for UHV/XHV

647
11.3. Leak Detection at UHV/XHV

648
11.4. Outgassing

648
11.4.1. Reduction of Outgassing Rates

652
11.5. UHV/XHV Hardware
652
References

12. Calibration and Standards 657


Karl Jousten

658
12.1. Primary Standards

659
12.1.1. Liquid Manometers and Piston Gauges

661
12.1.2. Static Expansion

< previous page page_xx next page >


< previous page page_xxi next page >
Page xxi

665
12.1.3. Continuous Expansion

670
12.1.4. Molecular Beam Expansion

673
12.2. Calibration by the Comparison Method

676
12.3. Calibration of Vacuum Gauges and Mass Spectrometers

676
12.3.1. Capacitance Diaphragm Gauges

680
12.3.2. Spinning Rotor Gauges

683
12.3.3. Ionization Gauges

686
12.3.4. Mass Spectrometers

689
12.4. Calibration of Test Leaks

692
12.5. Measurement of Pumping Speeds

695
References

Appendix 701

701
Graphic Symbols for Vacuum Components

708
Conversion Factors for Pressure Units

709
Vapor Pressure of Common Gases

711
Vapor Pressure of Solid and Liquid Elements

714
General Reference Books on Vacuum Science and Technology
Index 715

< previous page page_xxi next page >


< previous page page_xxiii next page >
Page xxiii

Acronyms

Acronyms used in the text are listed under the chapter where they occur.

Chapter 5
Part I. Getters and Getter Pumps

bcc Body-centered cubic (crystal structure)


CCRT Color cathode ray tube
CRT Cathode ray tube
fcc Face-centered cubic (crystal structure)
FED Field emission display
hcp Hexagonal close-packed (crystal structure)
HPTF High-porosity thick film (getter)
HT High temperature
HV High vacuum
LN2 Liquid nitrogen
PDP Plasma display panel
rf Radio frequency
RT Room temperature
SEM Scanning electron microscope
UHV Ultrahigh vacuum

Chapter 5
Part II. Sputter Ion Pumps

BA BayardAlpert (gauge)
DI Diode sputter ion pump with two cathode materials
FEM Finite element method
HMF High magnetic field (mode)
HP High pressure (mode)
LMF Low magnetic field (mode)
SIP Sputter ion pump
TM Transition (mode)
< previous page page_xxiii next page >
< previous page page_xxiv next page >
Page xxiv

UHV Ultrahigh vacuum


VTP Volume throughput

Chapter 5
Part III. Cryopumps

BET BrunauerEmmettTeller (adsorption model)


GM GiffordMcMahon (thermodynamic cycle)

Chapter 6
Vacuum Gauges

BAG BayardAlpert gauge


CCD Cold-cathode gauge
CDG Capacitance diaphragm gauge
ESD Electron-stimulated desorption
FS Full scale
HCG Hot-cathode gauge
JHP Jauge haut pression (Choumoff gauge)
NIST National Institute of Standards and Technology (USA)
PTB Physikalisch-Technische Bundesanstalt (Germany)
QBG Quartz helix Bourdon gauge
SCR Silicon controlled rectifier
SRG Spinning rotor gauge
UHV Ultrahigh vacuum
XHV Extreme high vacuum

Chapter 7
Partial Pressure Analysis
AP Appearance potential
CRDS Cavity ringdown spectroscope
IP Ionization potential
IR Infrared
MCP Microchannel plate
MS Mass spectrometer
PPA Partial pressure analyzer
QMS Quadrupole mass spectrometer
rf Radio frequency
RGA Residual gas analyzer
SEM Secondary electron multiplier
TOF Time of flight
TOFMS Time-of-flight mass spectrometer
UHV Ultrahigh vacuum
XHV Extreme high vacuum

< previous page page_xxiv next page >


< previous page page_xxv next page >
Page xxv

Chapter 11
Ultrahigh and Extreme High Vacuum

BA BayardAlpert (gauge)
BAG BayardAlpert gauge
ESD Electron stimulated descorption
EXB Cycloidal mass spectrometer
GIP Getter-ion pump
MG Magnetron gauge
IMG Inverted magnetron gauge
MS Mass spectrometer
MBAG Modulated BayardAlpert gauge
NEG Nonevaporable getter
RGA Residual gas analyzer
SIP Sputter-ion pump
TMP Turbomolecular pump
TSP Titanium sublimation pump
UHV Ultrahigh vacuum
XHV Extreme high vacuum

Chapter 12
Calibration and Standards

AVS American Vacuum Society


CDG Capacitance diaphragm gauge
DIN German industry standard
DKG German calibration service
IG Ionization gauge
IMGC Institutodi Metrologia ''G. Colonnetti" (Italy)
ISO International Organization for Standards
NIST National Institute of Standards and Technology (USA)
NPL National Physical Laboratory (England)
NPL National Physical Laboratory (India)
PTB Physikalisch-Technische Bundesanstalt (Germany)
QBS Quartz Bourdon spiral manometer
SI Système Internationale
SRG Spinning rotor gauge
UHV Ultrahigh vacuum
XHV Extreme high vacuum

< previous page page_xxv next page >


< previous page page_1 next page >
Page 1

1
Kinetic Theory of Gases

Benjamin B. Dayton

For a proper understanding of phenomena in gases, more especially at low pressures, it is essential to consider these
phenomena from the point of view of the kinetic theory of gases [1]. This theory rests essentially upon two fundamental
assumptions. The first of these postulates is that matter is made up of extremely small particles, which in the gaseous
state at moderate temperatures are monatomic or polyatomic molecules and at higher temperatures may be entirely
dissociated into atoms or even into positive ions and electrons to form a "plasma." The second postulate is that the
molecules of a gas are in constant motion, and this motion is intimately related to macroscopic properties known as the
temperature and pressure, which characterize the state of the gas in a given small region.

The center of mass of the molecule is assumed to move in a straight line (neglecting the force of gravity) with a constant
velocity between collisions with other molecules, and the forces between molecules are negligible except when the
molecular centers approach within a distance known as the mean molecular diameter. Velocity is a vector which must
be measured with respect to some "fixed" reference frame. The magnitude of the velocity vector is called the speed of
the molecule. In the "laboratory reference frame" the motion of the particles may be divided into bulk or fluid motion
due to pressure or concentration gradients which generate mass flow or diffusive flow and the random velocity
components associated with the concept of temperature as measured in a reference frame moving with the fluid flow
velocity. Kinetic energy is proportional to the square of the molecular velocity and is a scalar quantity.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_1 next page >


< previous page page_2 next page >
Page 2

Temperature is defined as a quantity proportional to the average (kinetic) energy of translation of the particles in a
reference frame moving with any fluid flow at a given small region. In the case of monatomic molecules (such as those
of the rare gases and the vapors of most metals), the effect of increased temperature is evidenced by increased
translational energy of the molecules. In the case of diatomic and polyatomic molecules, an increase in temperature also
increases through intermolecular collisions the rotational energy of the molecule about one or more axes, as well as
vibrational energy of the constituent atoms with respect to mean positions of equilibrium. However, in the following
discussion, only the effect on translational energy will be considered.

1.1
Ideal Gas Law

According to the kinetic theory, a gas exerts pressure on the enclosing walls because of the impact of molecules on
these walls. Since the gas suffers no loss of energy through exerting pressure on the stationary solid wall of its
enclosure, it follows that each molecule is thrown back from the wall with the same speed as that with which it
impinges, but in the reverse direction with respect to the normal; that is, the impacts are perfectly elastic.

Suppose a molecule of mass m to approach a flat wall surface lying in the x, y plane with velocity component vz
perpendicular to the wall. Since the molecule rebounds with the same speed, the change of momentum per impact is
2mvz. If ν molecules strike unit area in unit time with velocity component vz, the total impulse exerted on the unit area
per unit time is 2mvzν. But the pressure, P, on a wall is defined as the rate at which momentum is imparted to a unit area
of surface. Hence,

where the summation is over all values of vzν and it is assumed that all molecules have the same mass m.

It now remains to calculate ν. Of all the molecules within a volume ∆V extending outward from a small area, ∆x∆y, of
the wall by a distance |vzdt|, where dt is a short interval of time, at equilibrium only one half of the molecules will be
moving with velocity components vz toward the wall. Let nz denote the number of molecules per unit volume in the gas
within ∆V that have a velocity component of either vz or vz. Then the flux rate against the wall will be ν = nzυz/2, and
Eq. (1.1) becomes

The total speed of a molecule with velocity components vx, vy, and vz is the square root of the quantity

Defining the averages

< previous page page_2 next page >


< previous page page_3 next page >
Page 3

where n is the total number of molecules per unit volume regardless of velocity components, we have by symmetry
and

Then from Eq. (1.2) the basic equation for the pressure is

where ρ denotes the gas density, Eq. (1.6) can be expressed in the form

which shows that, at constant temperature, the pressure varies directly as the density, or inversely as the volume. This is
known as Boyle's law.

Now it is a fact that no change in temperature occurs if two different gases, originally at the same temperature, are
mixed. This result is valid independently of the relative volumes. Consequently, the average kinetic energy of the
molecules must be the same for all gases at any given temperature, and the rate of increase with temperature must be the
same for all gases. We may therefore define temperature in terms of the average kinetic energy per molecule, and this
suggestion leads to the relation

for each of the three degrees of freedom of translational motion, where T is the absolute temperature (degrees Kelvin),
defined by the relation T = 273.15 + t (t = degrees Centigrade), and k is a universal constant, known as the Boltzmann
constant. The total mean translational energy is then

where υr is known as the root-mean-square velocity. The total kinetic energy of the molecules in a volume V will be

Combining Eq. (1.8) with (1.5) it follows that Boyle's law can be expressed in the form

where the units chosen must be consistent with either the cgs system or the SI system. When n is expressed as
molecules/cm3 and k is in erg/K, then P will be the pressure in
< previous page page_3 next page >
< previous page page_4 next page >
Page 4

dyne/cm2. Also, from Eq. (1.6) and Eq. (1.11), it follows that

which is known as Charles' law or Gay-Lussac's law.

Lastly, let us consider equal volumes of any two different gases at the same values of P and T. Since P and V are
respectively the same for each gas, and is constant at constant value of T, it follows from Eqs. (1.10) and (1.11)
that n must be the same for both gases. That is, equal volumes of all gases at any given values of temperature and
pressure contain an equal number of molecules. This was enunciated as a fundamental principle by Avogadro in 1811,
but it took about 50 years for chemists to understand its full significance.

On the basis of Avogadro's law the molecular mass, M, of any gas or vapor is defined as that mass in grams, calculated
for an ideal gas, which occupies, at 0°C and 1 atmosphere, a volume [2]

V0 = 22,414.10 cm3.

This is therefore designated the molar volume, and the equation of state for an ideal gas can be written in the form

where W is the mass in grams, M is the molecular mass in grams, and R0 is a universal constant in units which depend
on the choice of units for the volume V, the pressure P, and the absolute temperature T. It is then convenient to express
Eq. (1.14) in the form

where nM denotes the number of moles (corresponding to M in grams) in the volume V under the given conditions of
temperature and pressure.

Excellent summaries of the various proposals prior to 1967 regarding the unit of pressure to be used in vacuum science
and technology have been given by Thomas and Leyniers [3]. When the units of length [l], mass [m], and time [t] are
chosen for the three basic units of a coherent system, the pressure unit is expressed by

[p] = [l]1 [m][t]2

or by

[p] = [F][l]2,

where F is the force with dimensions [l][m][t]2. In the cgs system the unit of force is the dyne and the pressure unit is 1
dyne/cm2. In the MKS or SI system the unit of

< previous page page_4 next page >


< previous page page_5 next page >
Page 5

force is the newton (abbreviated N), defined as 1 kg · m · s2 and the pressure unit is 1 N · m2. In 1962 the French
government approved the term "pascal" (abbreviated Pa) for the newton per square meter [4], and this term was
officially adopted by the International Standards Organization in 1978 while deprecating the use of older pressure units
such as mmHg and the torr (or Torr) [5]. Decimal multiples and submultiples of the Pa are accepted, and some have
special nomenclature such as

1 bar = 105 Pa,

1 millibar (mbar) = 100 Pa.

The previously used unit of 1 microbar (µbar) = 106 bar for the dyne/cm2 equals 101 Pa or 7.5 × 104 Torr. The main
reason for now using the mbar in vacuum technology is that its value is close to that of the Torr and the mmHg, which
were used for many years in the vacuum industry and the scientific literature. In 1954 both the British [6] and American
Committees on Vacuum Nomenclature [7] recommended replacing the mmHg by the Torr, defined as exactly
1,013,250/760 dyne/cm2, thus making it independent of the changing values for the measured density of mercury and
the acceleration of gravity. The term torr had previously been used in the German literature as a substitute for mmHg,
and at the First International Congress on Vacuum Techniques in Belgium in 1958 the newly defined Torr was favored
by the Germans and the Americans while the pascal was favored by the Belgians and the French representatives [8].
The Torr then became widely used throughout the world except in France until 1978, when the International Standards
Organization deprecated the use of Torr and defined the standard atmosphere as exactly 101,325 Pa and recommended
the pascal. Thus

In practical use the conventions of 1 Torr = 133.322 Pa and 1 Torr = 1333.22 µbar are sufficient. It may be noted that
the Torr is not identical to the unit mmHg used prior to 1955 which is based on the equation

where ρ = 13.5951 g/cm3 (at 0°C) is the density of mercury, g = 980.665 cm/s2 was the standard acceleration of gravity
accepted at that time, and h is the height (in mm) of a mercury column in a tube above some reference point such as the
surface of the mercury in the reference tube of a U-tube manometer or a McLeod gauge. One bar = 750.06 mmHg.
Another unit no longer used is the micron = 103 mmHg.

The basic unit of volume in the cgs system is 1 cm3, and that in the MKS or SI system is 1 m3. Following the Twelfth
Conférence Général des Poids et Mesures in 1964 the liter (or litre) is defined as 1 dm3 or 103 cm3 or 103 m3. The
preferred abbreviation for the liter is L, but l and the script l have been used. In the United States the volume unit
associated with industrial vacuum chambers and gas flow through mechanical pumps is commonly the cubic foot, which
equals 28.316847 liter. The basic unit of mass in the cgs system is the gram (g), and that in the MKS system is

< previous page page_5 next page >


< previous page page_6 next page >
Page 6

the kilogram (kg). The pound avoirdupois is seldom used for mass, but the pound force (poundf) is used in industrial
engineering in connection with the pressure unit pound per square inch (lb/in2.). 1 lb/in2. = 0.0689476 bar or 6.89476
kPa.

Units of temperature are the Celsius degree (°C) (also known as the degree centigrade) and the degree Kelvin (K) on the
absolute or thermodynamic scale. The zero of the Kelvin scale is 273.15°C, so that temperature T (in K) = temperature t
(in °C) + 273.15. The Fahrenheit temperature scale is seldom used in vacuum physics.

From the equation of state for an ideal gas we obtain the following values for the universal gas constant R0:

R0 = 760 · 22.41383/273.15 = 62.3632 Torr · L · K1 · mol1


= 1013250 · 22414.10/273.15 = 8.314511 · 107 erg · K1 · mol1
= 8.314511 joule · K1 · mol1
= 8.314511/4.1840 = 1.9872 cal (thermochemical) K1 · mol1.

In dealing with gases at low pressures, it is convenient to express the quantity of gas (Q = PV) in Pa · m3 at standard
room temperature 23°C. Note that 1 Torr · L = 0.133322 Pa · m3. The quantity of gas thus defined flowing per unit time
through a given cross section of a pipe has been called the throughput in the British and American Glossaries of terms
used in vacuum technology [9].

The unit of throughput is then

1 Pa · m3 · s1 = 1 N · m · s1 = 1 J · s1 = 1 W.

Objections have been raised against the use of pressurevolume products for quantity of gas on the basis that the
temperature of the walls of a vacuum system is not always clearly specified, and the suggestion is made that flow of gas
should preferably be expressed in mol/s to avoid any ambiguity due to unspecified temperatures [10]. The term
throughput should be reserved for flow in Pa · m3 · s1 at some specified temperature, such as a "standard" room
temperature of 23°C, and a term such as molar flow rate should be used for flow in mol · s1. Then

1 mol · s1 = 8.31451 × 296.15 Pa · m3 · s1 = 2462.342 Pa · m3 · s1 at 23°C.

1.2
Avogadro's Number

Avogadro's law states that the number of molecules per gram-molecular mass is a constant, which is designated NA.
Although a number of different methods have been used for the determination of this constant, the most accurate
method depends upon the determination of both the Faraday (F) and the charge of the electron (e).

For the deposition of one gram-equivalent the most accurate value [2] is

F = 96,485.309 absolute coulombs.

< previous page page_6 next page >


< previous page page_7 next page >
Page 7

Assuming that each univalent ion has a charge equal in magnitude to that of one electron,

NA = F/e.

The value of the charge of the electron is

e = 4.8032068 · 1010 absolute esu


= 1.60217733 · 1019 absolute Coulomb

since 1 absolute Coulomb = 1/10c times the charge in esu absolute coulombs, where c = velocity of light = 2.99792458
· 108 m · s1. Hence

NA = 6.0221367 × 1023 mol1.

From Eqs. (1.11) and (1.15) it follows that for one mole of gas in the volume V, nM = 1 and n = NA/V so that

From Eq. (1.9) it also follows that the average kinetic energy per molecule is given by

The mass per molecule is evidently

From Eq. (1.11) it follows that the number of molecules per cubic centimeter is given by

where Pµb is the pressure in microbars and PPa is the pressure in pascal units, while

where Pτ is the pressure in Torr.

Table 1.1 gives values of n for a series of values of T, Pµb, PPa and Pτ (Torr). The value n = 2.687 × 1019 cm3 is
known as the Loschmidt number.
< previous page page_7 next page >
< previous page page_8 next page >
Page 8

Table 1.1. Number of Molecules per Cubic Centimeter


T(°K) Pµb PPa Pτ n
273.15
1.01325 × 106 1.01325 × 105 760 2.687 × 1019
298.15
1.01325 × 106 1.01325 × 105 760 2.461 × 1019
273.15
1.333 × 103 1.333 × 102 1 3.535 × 1016
298.15
1.333 × 103 1.333 × 102 1 3.239 × 1016
273.15
10 1 7.50 × 103 2.652 × 1014
298.15
10 1 7.50 × 103 2.429 × 1014
273.15
1 0.1 7.50 × 104 2.652 × 1013
298.15
1 0.1 7.50 × 104 2.429 × 1013

1.3
Molecular Collisions; Mean Free Path; MaxwellBoltzmann Distribution Laws

It is evident that there must be a nonuniform distribution of velocities among all the molecules in a given volume because of the constant
occurrence of collisions. In an elementary treatment of the collision process the two molecules are assumed to be solid spheres of mass m1
and a well-defined diameter δ1 for one molecule and m2 and δ2 for the second molecule. The collisions are assumed to be elastic; that is, no
translational kinetic energy is lost by excitation of molecular rotation or atomic vibrations or by excitation of molecular vibrations when one
of the molecules is part of the wall of the enclosure. While the molecular diameters on impact can only be defined in terms of the fields of
attractive and repulsive force around each molecule and the relative momenta, it is assumed for simplicity that δ1 and δ2 have fixed values.
The distance between centers on impact will then be

and the mutual collision cross section is defined as

If we consider a single molecule of diameter δ1 (cm) moving at high speed, v, through a gas composed only of molecules of diameter δ2
(cm) moving relatively slowly and having a concentration of n molecules per cm3, then it is obvious that the average distance (in cm)
between collisions, known as the mean free path, will be given by

and the collision rate will be v/Lc = nσcv.

In an actual gas at equilibrium where all molecules have random velocities with an average speed va, the mean speed of a first molecule
relative to one of the other molecules depends on the angle between their respective directions of motion and the distribution law for
molecular velocities and will be equal to 21/2va in a homogeneous

< previous page page_8 next page >


< previous page page_9 next page >
Page 9

gas as calculated by Maxwell [11]. The collision rate is therefore

and the mean free path in a homogeneous maxwellian gas is

where in this case δ1 = δ2 = δ and σc = πδ2. This quantity is of fundamental importance in vacuum physics since the
Knudsen number defined by the ratio

where Dc is a characteristic length in the vacuum system, such as the diameter of a cylindrical tube, determines the
physics of the gas flow through tubes and ducts. From Eq. (1.11) and Eq. (1.26) in cgs units we obtain

and for nitrogen [12] we have δ = 3.78 × 108 cm, giving

At T = 298 K and Pτ = 103 Torr, this gives a mean free path for nitrogen L = 4.92 cm. At this temperature and Pµb = 1
µbar = 0.1 Pa = 7.5 × 104 Torr, L = 6.56 cm. This pressure is regarded as the upper limit of the high vacuum region [13]
in which the molecules may collide with the walls of the vacuum device more frequently than with other molecules in
the gas phase. We therefore must also consider the case in which a molecule of the gas collides with a molecule which
is part of the wall.

In the laboratory reference frame the velocity vector for the first molecule before the collision may be represented as

and for the second molecule it may be expressed as

where u1 and u2 are the velocity components perpendicular to the line of centers at the time of impact while v1 and v2
are the velocity components along this line of centers before collision as shown in Fig. 1.1.

Similarly, after the collision the velocity vectors for the first and second molecules, respectively, may be represented by
< previous page page_9 next page >
< previous page page_10 next page >
Page 10

Fig. 1.1
Velocity components
in elastic collision
between spheres.

where V1 and V2 are the respective velocity components along this line of centers after the collision, while U1 and U2
are the respective velocity components perpendicular to this line of centers after the moment of collision. For a collision
to take place, v1 and v2 must be antiparallel with vectors directed toward each other; but if they are parallel, one
velocity component must be greater than the other, and the relative positions, must be such that the molecule with the
larger component overtakes the other molecule in this direction.

The conservation of linear momentum in elastic collisions requires

Also, for the momentum components we require

The conservation of translational energy requires


since u1 and v1 are orthogonal components. Similarly,

< previous page page_10 next page >


< previous page page_11 next page >
Page 11

From Eqs. (1.36), (1.37), (1.38), and (1.39) we have

Also, from Eqs. (1.34) and (1.36) we obtain

Dividing Eq. (1.40) by Eq. (1.41) gives

Combining Eq. (1.42) and Eq. (1.43), we obtain

and similar calculations give

Equations (1.44) and (1.45) together with Eq. (1.36) give the vector components after the collision in terms of the
components before the collision. If m1 = m2, then V1 = v2 and V2 = v1 so that the velocity components along the line of
centers are interchanged. The resultant speeds after the collision are then

If the initial velocity component v2 was greater than v1, then the collision will cause the first molecule to increase in
velocity and the second molecule will suffer a decrease in velocity.

To illustrate how some molecules can acquire very high velocities while others are brought to near zero velocity as a
result of collisions, we consider a special class of orthogonal collisions between like molecules in which the first
molecule has an initial velocity u1 orthogonal to the line of centers at impact while the second molecule has an initial
velocity v2 parallel to this line of centers. In this special case we have v1 = 0, u2 = 0, V1 = v2, V2 = 0, U2 = 0 and the
resultant velocity of the second molecule is zero while the speed of the first molecule after the collision has been
increased from u1 to W1 as given by Eq. (1.46). Note that we could have assumed the second molecule to have any
value of a velocity component u2 other than zero orthogonal to
< previous page page_11 next page >
< previous page page_12 next page >
Page 12

the line of centers, but this component would have no effect on the final velocity of the first molecule as given by Eq.
(1.46).

If we consider only orthogonal collisions in which the first molecule is repeatedly struck in successive collisions by one
of the other molecules with a velocity component v2 which is a constant equal to the most probable velocity, vm, and
assume that the initial u1 = vm also, then N successive orthogonal collisions would give a final speed

By a process of reasoning similar to that for calculating the rate of absorbtion of a molecular beam by scattering in a gas
[14], it can be shown that the probability of N + 1 or more successive orthogonal collisions, without collisions of
another type, is

which is the probability that the speed of m1 is equal to or greater than W1(0 < W1 < ∞), where b is a constant to be
determined. When N is large, the velocity W1 can be much greater than vm but the probability is quite small. This
resembles the actual distribution law for molecular speeds as determined by J. C. Maxwell and L. Boltzmann.

The form of the distribution law differs according to the particular type of velocity distribution of interest. If we let
designate the components, along the three coordinate axes, of the randomly directed velocity v, then

and the distribution function, with respect to, say, is given by the relation

where N = number of molecules in the volume under consideration.

The distribution function for all three components of v has the form

With respect to the polar coordinates θ and φ, we have

The most important distribution function is that with respect to v in a random direction, which is given by the relation
Differentiating fv with respect to v, it is observed that the maximum value occurs for

< previous page page_12 next page >


< previous page page_13 next page >
Page 13

Hence α corresponds to the value of the most probable velocity, which is given by the relation

In terms of c = v/α, we can express Eq. (1.53) in the form

where c varies from 0 to ∞, and dy (= fcdc) corresponds to the fraction of the total number of molecules which have
values of c ranging between c and c + dc. Hence

From Eq. (1.57) we derive the value of the arithmetical average velocity, va = αca, where

and, using Eq. (1.55), we obtain

The root-mean-square velocity, vr, corresponds to the square root of the average value of v2 as derived from Eq. (1.9)
and is therefore given by the relation

The second column in Table 1.2 gives values of fc for a series of values of c, and Fig. 1.2 shows a plot of these data, on
which the values of fc are indicated for the values c = 1, 1.1284, and 1.2247, respectively.
< previous page page_13 next page >
< previous page page_14 next page >
Page 14

Fig. 1.2
Plots illustrating Maxwell-Boltzmann distribution laws. Plot fc shows distribution
function for random velocity, c expressed in terms of the most probable velocity α;
plot fx shows distribution function for energy, E, in terms of x = E/(kT); y corresponds
to the fraction of the total number of molecules for which the random velocity
(expressed in terms of α) is less than or equal to a given value c.

gives the fraction of the total number of molecules which have a random velocity equal to or less than that
corresponding to the value c, or to v = αc. The third and fourth columns in Table 1.2 show values of y and of ∆y, where
∆y gives the fraction of the total number which have velocities (in terms of α as a unit) ranging between c and the
immediately preceding value of c. Thus, 8.35% of the molecules have velocities between c = 1 and c = 1.1, and 42.76%
have velocities equal to or less than the most probable value. The values in parentheses are those of (1 y). A plot of y
versus c is shown in Fig. 1.2. It is evident that y corresponds to the area under the curve for fc from the origin to the
given value of c.

From Eq. (1.53) the distribution formula for translational energy (E) can be derived. It has the form

Substituting the variable x = E/(kT), Eq. (1.61) becomes


< previous page page_14 next page >
< previous page page_15 next page >
Page 15

Table 1.2. Values of fc, y, and fx, Illustrating Application of Distribution Laws
c fc y ∆y x fx
0
0 0 0 0
0.1
0.0223 0.0008 0.0008 0.05 0.2401
0.2
0.0867 0.0059 0.0051 0.1 0.3229
0.3
0.1856 0.0193 0.0134 0.2 0.4131
0.4
0.3077 0.0438 0.0245 0.3 0.4578
0.5
0.4393 0.0812 0.0374 0.4 0.4785
0.6
0.5668 0.1316 0.0504 0.5 0.4839
0.7
0.6775 0.1939 0.0623 0.6 0.4797
0.8
0.7613 0.2663 0.0724 0.7 0.4688
0.9
0.8129 0.3453 0.0790 0.8 0.4535
1.0
0.8302 0.4276 0.0823 0.9 0.4352
1.1
0.8142 0.5101 0.0835 1.0 0.4152
1.2
0.7697 0.5896 0.0795 1.2 0.3722
1.3
0.7036 0.6634 0.0738 1.4 0.3294
1.4
0.6232 0.7286 0.0642 1.6 0.2882
1.5
0.5350 0.7878 0.0602 1.8 0.2502
1.6
0.4464 0.8369 0.0491 2.0 0.2160
1.7
0.3624 0.8772 0.0403 2.2 0.1855
1.8
0.2862 0.9096 0.0324 2.5 0.1464
1.9
0.2204 0.9348 0.0252 3.0 0.0973
2.0
0.1652 0.9540 0.0192 3.5 0.0637
2.2
0.0864 0.9784 0.0244 4.0 0.0413
2.5
0.0272 0.9941 0.0157 4.5 0.0266
3.0 (4.2 × 104)
0.0024 0.0055 5.0 0.0170
4.0 4.1 × 106 (5.1 × 107)
6.0 0.0069
5.0 7.8 × 1010 (7.9 × 1011)
7.0 0.0027
6.0 1.9 × 1014 (4.4 × 1016)
8.0 0.0011

The last two columns in Table 1.2 give values of fx as a function of x, and Fig. 1.2 shows a plot of this function. By differentiating fx with
respect to x and equating the result to zero, it is readily shown that fx has a maximum value for x = 0.5; that is, fE has a maximum value for
E = 1/2k T. On the other hand, as stated in Eq. (1.9), Eav = 3/2kT.

Since

it is possible, from the plot for y in Fig. 1.2, to determine the fraction of the total number of molecules which have an energy equal to or less
than that corresponding to a given value of E.

It follows from the equations above that the value of v for which fv is a maximum increases with T½, while that of E for which fE is a
maximum increases with T.

Values of va, at 0°C and 25°C, for a number of gases and vapors are given in Table 1.3.

< previous page page_15 next page >


< previous page page_16 next page >
Page 16

Table 1.3. Masses, Velocities, and Rates of Incidence of Molecules


104 · νa

Gas or Vapor M 1023m 1010 0°C 25°C 1017ν1 1020 105G1 102
H2
2.016 0.3347 0.8878 16.93 17.70 11.23 14.97 0.3759 0.5012
He
4.003 0.6646 1.7631 12.01 12.56 7.969 10.63 0.5297 0.7062
CH4
16.04 2.663 7.063 6.005 6.273 3.981 5.308 1.060 1.414

NH3
17.03 2.827 7.498 5.829 6.089 3.865 5.152 1.092 1.456
H2O
18.02 2.992 7.936 5.665 5.919 3.756 5.007 1.124 1.498
Ne
20.18 3.351 8.886 5.355 5.594 3.550 4.733 1.190 1.586

CO 12.34
28.01 4.651 4.543 4.746 3.012 4.016 1.402 1.868
N2 12.34
28.02 4.652 4.542 4.745 3.011 4.015 1.402 1.868
Air 12.77
28.98a 4.811 4.468 4.668 2.962 3.950 1.425 1.900

O2 14.09
32.00 5.313 4.252 4.442 2.819 3.758 1.497 1.996
Ar 17.59
39.94 6.631 3.805 3.976 2.523 3.363 1.675 2.230
CO2 19.38
44.01 7.308 3.624 3.787 2.403 3.204 1.756 2.342

CH3Cl 22.23
50.49 8.38 3.385 3.356 2.244 2.991 1.881 2.508
SO2 10.64 28.21
64.06 3.004 3.139 1.992 2.656 2.118 2.825
Cl2 11.77 31.23
70.91 2.856 2.984 1.893 2.524 2.229 2.973

Kr 13.90 36.85
83.7 2.629 2.747 1.743 2.324 2.422 3.229
C7H16 16.63 44.12
100.2 2.403 2.510 1.593 2.123 2.650 3.533
Xe 21.80 57.82
131.3 2.099 2.193 1.392 1.856 3.034 4.044
CCl4 25.54 67.72
153.8 1.939 2.026 1.286 1.714 3.283 4.377
Hgb 33.31 (88.33)
200.6 1.698 1.774 (1.126 1.501 3.750 4.998)
Note: ν1 = rate of incidence of molecules per square centimeter per second, at 0°C and 1 µbar.

= rate of incidence of molecules per square centimeter per second, at 0°C and 1 Torr.
G1 = mass of gas corresponding to ν1 (g · cm2 · s1).

= mass of gas corresponding to (g · cm2 · s1).


m = mass of molecule (g); = density of gas at 0°C and 1 µbar (g · cm3).
va = average velocity (cm · s1).
a Calculated from the value ρ = 1.293 × 103 at 0°C and 760 Torr.
b Since the vapor pressure of mercury at 0°C is 1.85 × 104 Torr (= 0.247 µbar), the values given in parentheses have no physical
significance. Actual values at 0°C, corresponding to saturation pressure, are as follows: ρ = 21.79 × 1010; ν = 2.777 × 1016; G = 9.249 ×
106.

1.3.1
Relation between Molecular Velocities and Velocity of Sound

It is of interest to note that the relations for α, va, and vr can also be expressed in terms of the velocity of sound, which we shall designate by u.

Since

< previous page page_16 next page >


< previous page page_17 next page >
Page 17

where ρ1 = density at 1 µbar, and P = Pµb, we can write the relations for molecular velocities in the forms

On the other hand,

where γ = Cp/Cv = ratio of specific heats (per gram-mole) at constant pressure and constant volume. Hence

For mercury and other monatomic gases, γ = 1.667; for diatomic gases (such as H2, N2, and O2), γ = 1.40
(approximately). Hence,

va/u = 1.236 for monatomic gases


= 1.349 for diatomic gases
and
u/α = 0.9124 for monatomic gas.

Thus the velocity of sound in a gas approaches molecular velocities very closely.

1.3.2
Determination of Avogadro's Constant from Distribution of Particles in Brownian Motion

Under high magnification, all suspensions of very fine particles in gases or liquids exhibit ''Brownian" motions. Einstein
(1905) suggested that the motion of these particles is essentially that to be expected, on the basis of the kinetic theory of
gases, of "large molecules" and therefore subject to the same laws as gas molecules. That is, the average energy per
particle at any given temperature T is 3/2kT, and the average velocity of the particles is given by the relation

where m = mass of particle.

< previous page page_17 next page >


< previous page page_18 next page >
Page 18

Application of the BoltzmannMaxwell laws leads to the following relation for the distribution of particles at different levels in
a gravitational field:

where n0 = number of particles per cubic centimeter at h = 0,


n = number of particles per cubic centimeter at height h (centimeters),
g = 981 dynes,
m′ = apparent mass of particles, which is different from the actual mass because of the buoyancy of the
medium.

Let m = actual mass of particle. Then

where ρ′ = density of the medium and ρ = density of the particles.

Actually m′ is determined from the rate of settling of the particles, by application of Stokes' law.

Thus it is possible to determine k (and consequently the value of NA = R0/k) from observations on the value of m′ and the
relation between n/n0 and h. Using a fine suspension of gum arabic in water, Perrin obtained the value NA = 6.8 × 1023.

Equation (1.71) has been applied to the determination of the variation with altitude of the density of the atmosphere.

Assuming an average temperature of T = 230 K at higher altitudes, we obtain

m′/k = M/R0 = (29 × 107)/8.315,

and hence

where Pmm, is the pressure in mmHg (Torr) at the altitude H (in meters) above sea level.

1.4
Gas Pressure and Rate at Which Molecules Strike a Surface

The pressure in a gas is a tensor quantity and has to be defined with respect to an imaginary stationary plane surface passing
through a point in the gas. The pressure is defined as the net rate at which momentum normal to this surface is transmitted
across it per unit area in the positive direction, momentum transmitted in the opposite direction being counted as negative
[15]. Let ∆S represent a small surface element in this imaginary plane surface through which gas molecules included within a
hemispherical surface of radius L centered on the center of ∆S can pass directly without collision, where L is the mean free
path at the prevailing pressure as shown in Fig. 1.3.

< previous page page_18 next page >


< previous page page_19 next page >
Page 19

Fig. 1.3
Diagram illustrating the
calculation of the pressure in
a gas. L=the mean free path.
v=a particular molecular
velocity.

Consider the gas coming from within the solid angle dσ/4πL2, where

is a surface element on the hemisphere. For a gas at rest, momentum normal to the surface will be mv cos θ for such
molecules where v is the speed of the molecule within the range v and v + dv, and the rate at which this momentum is
transmitted across the surface is mv2 cos2θ. Then for a gas at rest with respect to this imaginary surface the total
momentum transmitted will be 2mv2 cos2θ per molecule since the momentum transmitted through ∆S from the opposite
side is counted as negative. The total momentum transmitted per unit solid angle will be 2nL · ∆S · mv2 cos2θ. Dividing
by L · ∆S to obtain force per unit area of surface, the quantity L · ∆S cancels out. If the molecular velocities are
distributed according to Eq. (1.53), the pressure is given by

This reduces to

using Eq. (1.9).

In a gas at rest with molecular velocities distributed according to the MaxwellBoltzmann equilibrium distribution, the
molecular flux across a plane surface element A due to all molecules having velocity vectors with directions within a
small solid angle dω whose axis makes an angle θ with the normal to A is given by the cosine law [16] formula,
< previous page page_19 next page >
< previous page page_20 next page >
Page 20

where is the average molecular velocity. While collisions occurring within the solid angle dω may scatter
molecules out of the region, in an equilibrium gas the collision processes must result in other molecules entering this
region and having the same direction. If we consider only those molecules crossing an imaginary plane surface in the
gas in the direction of the positive normal to the surface, or alternatively only those molecules which strike a plane solid
or liquid surface and are not scattered backward, then the total molecular flux across or against this surface is

Some of the molecules striking a solid surface will be adsorbed and the remainder are scattered back in various
directions depending on the surface roughness and the intermolecular forces during close approach. The adsorbed
molecules eventually reach accommodation with the prevailing temperature of the material at the surface; and because
they must acquire velocity components perpendicular to the surface due to molecular vibrations which have maximum
values ranging from zero to very high values by processes similar to those considered above for orthogonal collisions
within a gas, some of the molecules will acquire sufficient velocity after a certain time known as the adsorption lifetime
to escape the attractive force fields at the surface.

It has been shown by Comsa and coworkers [17] that the velocity distribution in the gas from these molecules which are
desorbed is not necessarily that of a Maxwellian gas. Comsa and several other investigators [18] have shown
experimentally that the angular distribution does not necessarily obey the cosine law. The experiments show that the
evaporated or desorbed flux is peaked in the direction of the normal to the surface varying as cosnθ with n greater than
1 and as high as 9 for strongly peaked emissions. However, as deduced by Clausing and confirmed by Comsa, under
equilibrium conditions the sum of the distribution of the molecules leaving the surface due to various processes
(adsorptiondesorption, reflection, diffraction, and inelastic scattering) has to obey the cosine law.

Epstein [19] has presented a model of the wall boundary condition in terms of the relation between the distribution
functions of the incident and reflected particles, assuming that a certain fraction reflect diffusely (uncorrelated with the
incident conditions) while the remainder reflect specularly depending on the velocity of the incident particles.

The molecular flux of vapor, as measured in the laboratory frame, in the beam of gas issuing into a vacuum from a
Knudsen cell, comprising a source of vapor at a uniform temperature T enclosed in a box with a small thin-edged orifice
in one wall, obeys the cosine law approximately, but the mean translational energy per molecule in the beam at the exit
is 2kT rather than (3/2)kT because the faster molecules have a higher probability of exiting the orifice and the mean
energy per molecule involves averaging the translational energy with a velocity distribution function which

< previous page page_20 next page >


< previous page page_21 next page >
Page 21

contains the factor v3 rather than v2. In fact, this value 2kT is characteristic of the mean translational energy in the flux
of molecules in an equilibrium gas against any surface and must be considered when calculating heat transfer.

At very low pressures where the mean free path is much greater than the linear dimensions of the vacuum vessel, the
velocity of the molecules in the gas phase is entirely determined by collisions with the walls and the temperature of the
walls. Since such collisions are far less frequent than collisions in a dense gas where the m.f.p is only a small fraction of
1 cm, the relaxation time, or time to restore equilibrium in a gas which has been disturbed from equilibrium by
transitory pressure or temperature gradients, is relatively long. Because of outgassing and readsorption at various
surfaces within a vacuum system and the removal of molecules by vacuum pumps, equilibrium conditions seldom
prevail [20], and the pressure measured by a tubulated vacuum gauge will depend on the orientation of the plane of the
entrance of the tubulation with respect to mass flow vectors [21]. Pressures measured by an ionization gauge, which has
the envelope and tubulation removed down to the base mounted inside the vacuum chamber, are less dependent on
orientation except for the shielding effect of the base.

Muntz [22] has described a method of making localized measurements of the molecular velocity distribution function in
rarefied gas flows.

From Eq. (1.78) the incident molecular flux per unit area is

Substituting for n and va, from Eqs. (1.20), (1.21), and (1.58), we obtain the relations

The volume which strikes unit area per unit time is given by

and is therefore a constant at all pressures, but varies with (T/M)1/2.

In the literature, especially that originating in Germany, Eq. (1.87) is expressed in the form

where ρ1 is the density at 1 µbar.


< previous page page_21 next page >
< previous page page_22 next page >
Page 22

The last four columns of Table 1.3 give values of v1 and G1, the values calculated for a pressure of 1 microbar, and of
the values calculated for a pressure of 1 Torrall at 0°C.

The equations given above for v and G are also applicable to the effusion of gases at low pressures through small holes
in very thin plates. The requisite condition for the application of Meyer's relation to effusion is that the diameter of the
opening should be small compared with the mean free path.

A comparison of relative values of v or G for different gases or vapors streaming through such a hole makes it possible
to obtain relative values of M, since, for constant values of P and T, v varies inversely as M1/2, and G varies directly as
M1/2.

A good check on the above equations was obtained by Knudsen [23] in some experiments in which hydrogen, oxygen,
and carbon dioxide, at pressures ranging from 100 to 0.01 Torr, were made to flow into a vacuum through a 0.025-mm
hole in a 0.0025-mm-thick platinum strip.

Equation (1.87) shows that the volume per unit area per unit time, measured at the pressure P, is always the same. It
follows that FPµb corresponds to the volume at 1 µbar. Hence, if Pµb1 and Pµb2 (> Pµb1) denote the pressures on the
two sides of a very thin-walled orifice of area A, the net quantity of gas (Q) flowing through the orifice per unit time is
given by

where Q = volume in cubic centimeters per second, measured at 1 µbar. That is, Q denotes microbars × cubic
centimeters per unit time.

1.5
Rate of Evaporation and Vapor Pressure

An interesting application of Eq. (1.84) was first made by Langmuir [24] to the determination of vapor pressure from
rates of evaporation in high vacua. Quoting from Langmuir's original paper, The Vapor Pressure of Metallic Tungsten:

Let us consider a surface of metal in equilibrium with its saturated vapor. According to the kinetic theory we
look upon the equilibrium as a balance between the rate of evaporation and rate of condensation. That is, we
conceive of these two processes going on simultaneously at equal rates.

At temperatures so low that the vapor pressure of a substance does not exceed a millimeter, we may consider
that the actual rate of evaporation of a substance is independent of the presence of vapor around it. That is, the
rate of evaporation in a high vacuum is the same as the rate of evaporation in presence of saturated vapor.
Similarly we may consider that the rate of condensation is determined only by the pressure of the vapor.

The rate at which molecules will, in general, condense on a surface is given by αv, where α is known as the
condensation coefficient or sticking coefficient. It represents the ratio between the rate at which molecules actually
condense on the surface and the rate at which they strike the surface. If we let µ denote the rate at which molecules
evaporate from the surface, then, at equilibrium, we obtain

< previous page page_22 next page >


< previous page page_23 next page >
Page 23

Langmuir [25] has shown that for metal atoms condensing on the surface of a metal the value of α may be assumed to
be equal to 1. In a later paper on the vapor pressures of high-boiling point organic liquids, Verhoek and Marshall [26]
showed that the same assumption is justified in respect to these liquids. Hence, we may, in practically all cases of
evaporation from the bulk phase, express the relation for rate of evaporation in the form

For the purpose of calculating the vapor pressure of a metal from a determination of loss of weight per unit area per unit
time, it is convenient to express Eqs. (1.85) and (1.86) in the forms

where G = rate of evaporation in grams per square centimeter per second.

As an illustration of the application of these equations, Table 1.4 gives values of G for tungsten [27] and tantalum [28]
at a series of temperatures (degrees K) together with calculated values of Pµb.

For the evaporation from a wire of diameter d′ (in mils) the loss in weight per second per centimeter length is given by

G1 = 2.54 × 103πd′G.

Hence Eqs. (1.92) and (1.93) assume the forms

Table 1.4. Rates of Evaporation and Vapor Pressures of Tungsten and Tantalum
Metal T(K) G Pµb
Tungsten 2600 8.41×109
M = 183.92 7.23×104
2800 1.10×107
9.81×103
3000 9.95×107
9.18×102
3200 6.38×106
6.08×101
3400 3.47×105
3.41

Tantalum 2400 3.04×109


M = 180.88 2.58×104
2600 5.54×108
4.90×103
2800 6.61×107
6.07×102
3000 5.79×106
5.40×101
3200 3.82×105
3.77

< previous page page_23 next page >


< previous page page_24 next page >
Page 24

Equations (1.92) and (1.93) have also been applied by Knudsen and subsequent investigators to the determination of
vapor pressures from rates of effusion through a small orifice.

Thus let us consider the case in which molecules evaporating from a hot surface pass through a small orifice into
another chamber in which they are condensed. If the pressure of residual gas in this "cool" compartment is extremely
low and the radius of the opening is less than L (the mean free path of the evaporating molecules in the "hot"
compartment), then the rate at which molecules pass through the hole is equal to the rate at which they strike this
opening. Consequently, the vapor pressure for any given temperature will be given by Eq. (1.92) or Eq. (1.93), where G
represents the weight passing through the orifice per unit area, per unit time.

These equations are, however, strictly applicable only if the thickness (l) of the wall, in which the orifice of area πa2 is
located, is vanishingly small compared to a. If the orifice consists of a short tube for which l/a is appreciable, then a
correction factor has to be applied, and instead of Eq. (1.82) we have the relation

where K is a function of l/a which is less than 1 for l/a > 0. The manner in which the value of K varies with l/a is
discussed subsequently in Chapter 2. Hence, if G′ denotes the actual loss in weight, at temperature T, of material of
molecular mass M, through an opening of area A, over a period of t seconds, then we obtain

These equations have been applied by a number of investigators for the determination of vapor pressure at low
temperatures, where the values are of the order of a few pascals. The method has been used, for instance, by Egerton for
such metals as zinc, cadmium, mercury [29], and lead [30].

As an illustration let us consider one such determination made for mercury vapor. In this case the area of the opening
was A = 0.0335 cm2. At 33.7°C, the loss of mercury through this orifice was 0.7867 g over a period of 2370 min. To
correct for the fact that l/a was not negligible, the value of K was found (by means of the relations given in Chapter 2) to
be 0.93.

Hence the corrected value of G is given by

Since T = 306.9 and M = 200.6, it follows from Eq. (1.93) that Pτ = 3.77×103 Torr.

Another interesting application of the above relations, and one which is of increasing importance in industrial
chemistry, is provided by the development of

< previous page page_24 next page >


< previous page page_25 next page >
Page 25
high-vacuum distillation for the separation of certain organic compounds in the pure state from naturally occurring oils. The
great advantage of this process arises from the fact that these organic compounds are unstable at higher temperatures and
therefore they can be distilled only at lower temperatures, at which the vapor pressures are in the range of 104 to 106 atm [31].

In this operation, evaporation takes place from a very thin film of liquid, which is renewed continuously, and condensation
occurs on an adjacent cooled surface. In a sufficiently high vacuum (pressure of residual gas less than 1 Pa) the rate of transfer
of distilland is in accordance with Eqs. (1.85) and (1.86). As the pressure of residual gas is increased, however, the rate of
distillation is decreased because of collisions between the molecules of the distilland and those of the gas. This is illustrated by
the data shown in Table 1.5, taken from Hickman's paper.

The distilland used was Octoil, which has the chemical formula C6H4 (COOC8H17)2 and molecular weight M = 390.3. From
Eq. (1.86), it follows that the rate of evaporation, W, in grams per second per square meter, is

where T is the absolute temperature corresponding to the vapor pressure of Pτ in Torr or Ppa in pascals.

It is evident from these observations, as well as from observations of a similar nature mentioned in the next section, that
molecules leaving the surface of the distilland are prevented from reaching the surface of condensation because of collisions
with the molecules of the residual gas. As a result of such collisions, many of the molecules leaving the hot film are driven
back, the number of such molecules increasing with the magnitude of residual pressure.

Table 1.5. Variation of Rate of Distillation with Pressure (Hickman)a


Pressure (mTorr) of Residual Gas W (g · s1 · m2)
(Air)
Pτ = 103 Pτ = 3×103 Pτ = 102
T = 368 K T = 383 K T = 393 K
0.3 0.6 1.85 6.4
4.0 0.46 1.59 5.7
7.0 0.38 1.37 5.2
10.0 0.32 1.18 4.6
15.0 0.25 0.95 3.8
25.0 0.21 0.70 2.1
50.0 0.12 0.40 1.67
aThe values given for T were taken from a plot of log Pτ versus 1/T and are therefore only approximate, which
accounts for the fact that values of W calculated by means of Eq. (1.99) for extremely low pressure are slightly less
for 3×103 and 102 Torr than those given in the first row.

< previous page page_25 next page >


< previous page page_26 next page >
Page 26

1.6
Free Paths of Molecules

Although the individual molecules in a gas at rest possess very high velocities, as shown previously, it is a matter of
ordinary observation that gases diffuse into one another very slowly. This is explained on the kinetic point of view by
assuming that the molecules do not travel continuously in straight lines, but undergo frequent collisions. The term
"collision" naturally leads to the notion of free path. This may be defined as the distance traversed by a molecule
between successive collisions. Since, manifestly, the magnitude of this distance is a function of the velocities of the
molecules, we are further led to use the expression "mean free path" (denoted by L), which is defined as the average
distance traversed by all the molecules between successive collisions.

However, this definition assumes that the molecules actually collide like billiard balls; that is, the molecules are
assumed to be rigid elastic spheres possessing definite dimensions and exerting no attractive or repulsive forces on one
another. But this concept can certainly not be in accord with the facts. It is probably impossible to state definitely the
diameter of a hydrogen atom or molecule, much less that of a poly-atomic molecule. Also there is no doubt that the
molecules exert attractive forces on one another for certain distances and repulsive forces when they approach
exceptionally close. Otherwise, how could we explain surface tension, discrepancies from Boyle's law, and a host of
related phenomena? To speak of collisions among molecules, such as these, is impossible. What meaning, therefore,
shall we assign to the free path under these conditions?

Let us consider at t = 0 a group of N0 "tagged" molecules moving in a given direction. As time goes on, these molecules
will suffer random collisions and a number will disappear from the original group. Let N denote the number which, after
a period t, are still identified with the original group, and let ω denote the collision frequency. Then

Integrating this equation, we obtain the result,

If we let l (= vat) designate the path that has been traversed by a molecule without suffering collision during the interval
t, then Eq. (1.101) can be written in the form

Furthermore, we can write

where L is a distance covered between collisions. Then it follows that Eq. (1.102) assumes the form

< previous page page_26 next page >


< previous page page_27 next page >
Page 27

That is, φ(l) is the fraction of the original group of molecules that are still traveling without having suffered a collision
in the distance l.

Furthermore, it follows from Eq. (1.102) that

represents the fraction of all the free paths that have a length between l and l + dl. (Hence the omission of the negative
sign in the differentiation.)

It follows from Eq. (1.105) that the average value of the free path is

For l = L, φ(l) = εl = 0.3679. This result shows that 63.21% of the molecules collide with other molecules in a distance
equal to or less than L. Furthermore, it is seen from Eq. (1.101) that this 63.21% of collisions occur in the interval τ = 1/
ω. Thus 1/ω is a constant of the same nature as the ''decay" constant in radioactive disintegrations, while 1/L may be
regarded as an "absorption" coefficient similar to the coefficient that measures the decrease in intensity of a beam of
light in passing through a medium.

Equation (1.104) indicates an experimental method for the determination of L′ which has been used by Born [32] and
Bielz [33] and which is described by Fraser [34]. A beam of silver atoms is sent into nitrogen or air, and a determination
is made of the amount of silver deposited by the beam in a given time t on a surface distant l from the source. If we let
I0 designate the intensity of the beam at the source, then the intensity at the collector is

where LP is the mean free path of silver atoms in the gas at the pressure in the collecting chamber.

Measurements of the mean free path of potassium in nitrogen have also been reported by Weigle and Plesset [35].

As Fraser [34] points out:

With noncondensable gases, it is not possible to measure I0 directly. We assign therefore to I0 a different
meaning: namely, the intensity which the beam would have if it were not, as is actually the case, weakened
through scattering by the alien molecules present in the collimator chamber. Now, clearly I0 is directly
proportional to the quantity of gas issuing from the source slit per second; but so also is the pressure P in the
collimator chamber, if a constant pump speed is assumed. We can therefore set I0 = c·P. On the other hand,
LP is inversely proportional to P; that is, LP = L/P, where if P is measured say in mTorr, L is the mean free
path at a pressure of 1 mTorr. I can therefore be expressed as a function of the pressure P; thus if l is the
distance between source slit and image slit,

I = c · P eP · l/L,

< previous page page_27 next page >


< previous page page_28 next page >
Page 28

it being assumed that the pressure in the observation chamber is negligibly small. I is a maximum for that
value of P which makes P · l/L = 1. To make a measurement, the intensity is plotted as a function of the
pressure in the collimator chamber, and the value of the latter at the maximum intensity is observed. Then

LP = L/P = l.

At this value of P, I = 0.3679I0.

It is of interest to observe, as Fraser emphasizes, that the requisite condition for obtaining a directional effect of the
molecules passing through the slit is that LP must not be less than d, the width of the slit.

The determination of mean free path for hydrogen has been carried out by Knauer and Stern [36]. The value they
obtained, however, is only about 0.44 times that derived from viscosity relations (see discussion in the following
section). The reason, as Fraser [34] points out, is that

the standard methods require an intimate encounter in order that the molecules may exchange energy and
momentum in amounts capable of affecting the viscosity or heat conductivity of the gas. The molecular ray
method on the other hand counts as a collision an approach of two molecules sufficiently close to deflect them
very slightly out of their paths; with narrow slits angular deflections of less than 104 are detectable.

In this connection the reader will find an interesting description of the many uses of molecular beams in a paper by
Taylor [37]. As he states, "Molecular beams, narrow rays of molecules formed by a slit system and moving in one
direction in an evacuated apparatus, may be used to advantage in many types of research." Among these are
determinations of molecular velocities (involving experimental tests of the validity of the MaxwellBoltzmann
distribution law), mean free paths, vapor pressures, accommodation coefficients, and mechanism of chemical reactions
and of adsorption.

Evidently the mean free path must depend upon the molecular diameter, and simple considerations indicate that the
length of the mean free path must vary inversely as the total cross-sectional area of the molecules per unit volume.
Again, the magnitudes of the coefficients of viscosity, heat conductivity, and diffusivity of gases are intimately bound
up with the length of the free path; whether it be transference of momentum from one layer to another as in viscosity, or
transference of increased kinetic energy of the molecules as in heat conductivity, the rate of this transference must
depend upon the number of collisions that each molecule experiences as it passes from point to point. It is therefore to
be expected that there should exist very similar relations between the values of the mean free path and those of the
coefficients of viscosity, heat conductivity, and diffusion. However, in attempting to deduce such relations, the
theoretical physicist has found himself confronted with the problem regarding the laws governing the variation with
distance of attractive and repulsive forces between molecules. As a result of successive attacks on this problem, by a
number of investigators, the exact forms of these relations have been modified from time to time. The reader is referred
to Chapman and Cowling [1d] for a detailed discussion of the whole problem.

< previous page page_28 next page >


< previous page page_29 next page >
Page 29

1.7
Relation between Coefficient of Viscosity, Mean Free Path, and Molecular Diameter

A gas streaming through a narrow-bore tube experiences a resistance to flow, so that the velocity of this flow decreases
uniformly from the center outwards until it reaches zero at the walls. Each layer of gas parallel to the direction of flow
exerts a tangential force on the adjacent layer, tending to decrease the velocity of the faster-moving and to increase that
of the slower-moving layers. The property of a gas (or liquid) by virtue of which it exhibits this phenomenon is known
as internal viscosity.

As a simple working hypothesis we may assume, as Newton did, that the internal viscosity is directly proportional to the
velocity gradient in the gas. Furthermore, the viscosity must depend upon the nature of the fluid, so that in a more
viscous fluid the tangential force between adjacent layers, for constant velocity gradient, will be greater than in a less
viscous fluid. We thus arrive at the following definition of the coefficient of viscosity:

The coefficient of viscosity is the tangential force per unit area for unit rate of decrease of velocity with distance (i.e.,
per unit velocity gradient).

With this definition we are in a position to deduce the approximate form of the relation between the coefficient of
viscosity and the free path.

Let u denote the velocity of flow of the gas at a distance d from a stationary surface. In uniform flow along a surface,
the velocity will decrease uniformly to zero as the surface is approached. We can therefore represent (see Fig. 1.4) the
velocity at distance OA = d by the ordinate AB = u and represent velocities at intermediate distances by the
corresponding ordinates of the line OB.

We shall imagine the gas divided into layers parallel to the surface, each having a depth equal to the free path, L.

Fig. 1.4
Diagram illustrating the derivation of simple
relation between the coefficient of viscosity (η)
of gas and the molecular mean free path (L).
< previous page page_29 next page >
< previous page page_30 next page >
Page 30

Let us denote the tangential force per unit area between adjacent layers by B. By definition:

where η denotes the coefficient of internal viscosity.

But, according to the kinetic theory, the tangential force per unit area is measured by the rate at which momentum is
transferred per unit area between adjacent layers. Because of the relative motion of the layers, the molecules moving
from a faster- into a slower-moving layer possess more momentum in the direction of flow than those moving in the
opposite direction.

Let us consider any layer, CE or EH, of thickness equal to L. We have chosen this particular value of the thickness so
that we may be justified, as a first approximation, in assuming that the molecules starting at either of the planes CD and
HK reach the plane EF without suffering collisionthat is, without change of momentum.

The momentum, parallel to the surface, of any molecule reaching the plane EF from the plane CD is m(u′ + v), where u′
denotes the velocity of flow at the plane CD and v is the mean velocity of the molecules.

The momentum, parallel to the surface, of a molecule reaching the plane EF from the plane HK is

m(u′ + v + 2uL/d).

The number of molecules that cross unit area per unit time in any direction in a gas at rest is equal to (1/6) nv; and this
must be the same for the molecules traveling in a direction perpendicular to the plane EF, because the velocity of flow
is assumed to be so small that the density remains constant throughout the different layers.

Hence the net rate of transference of momentum across unit area of the plane EF is equal to

From Eqs. (1.108) and (1.109) it follows that

The dimensions of η are evidently ml1t1, and in the cgs system the unit of viscosity is 1 poise = 1 g·cm1·s1 = 1
dyne·s·cm2. This is the unit of coefficient of viscosity used in this volume.

In deducing Eq. (1.110) it has been assumed that all the molecules possess the same velocity v and the same free path L.
Introducing the law of distribution of velocities, Boltzmann (1881) deduced the relation

where va = average velocity, and LB is defined as the average free path.

< previous page page_30 next page >


< previous page page_31 next page >
Page 31

O. E. Meyer, in his Kinetic Theory of Gases, used a different method of calculation and derived a relation of the form

where LM is also defined as the average free path.

From these equations an interesting conclusion may be deduced regarding the dependence of viscosity on pressure. As
has been mentioned, it is evident from very simple considerations that L must vary inversely as the number of molecules
present per unit volume. Consequently the product pL is constant and independent of the pressure. The velocity, v,
depends only upon the temperature and molecular weight. It therefore follows that, for any gas at constant temperature,
the viscosity is independent of the pressure and must increase with the temperature. The confirmation of these two
deductions has been justly regarded as one of the most signal triumphs of the kinetic theory of gases. As is well known,
the viscosity of all ordinary liquids decreases with increase in temperature. That the viscosity of gases must increase
with temperature was therefore regarded as a remarkable conclusion.

At both extremely low pressures and very high pressures, the conclusion that the viscosity is independent of the pressure
is not in accord with the observations, but this is due to the fact that the same derivation as has been presented above is
not valid under those conditions where either attractive forces between the molecules come into play or the pressure is
so low that a molecule can travel over the whole distance between the walls of the enclosure without suffering collision.

Both Eqs. (1.111) and (1.112) have been used by physicists, until comparatively recently, for deriving values of L from
η. However, the work of S. Chapman and D. Enskog, since about 1911, has led to the following relations, which are
discussed by Chapman and Cowling [1d] in their treatise.

To derive a more exact relation between η and L, it is necessary to introduce the relation between L and δ, the molecular
diameter, which has been shown to be of the form

where n = number of molecules per cubic centimeter.

For smooth rigid elastic spherical molecules, it is shown that

Substituting in this equation the relations va = (8kT/pm)½ and ρ = mn, it follows that

A further approximation leads to the conclusion that the right-hand side of the last equation should be multiplied by the
factor 1.016, and consequently

< previous page page_31 next page >


< previous page page_32 next page >
Page 32

Combining this with Eq. (1.113), the result is the relation used in the following discussion:

It follows directly from Eq. (1.114) that, for two gases having approximately equal values of δ, the viscosities should be
in the same ratio as the square roots of the molecular masses. This conclusion has been confirmed by observations on
the relative viscosities of H2 and D2 (deuterium) [38]. At room temperature we have ηD2/ηH2 = 1.39, which is
approximately equal to 21/2.

The difference between Eq. (1.116) and the equations of Boltzmann and Meyer arise from the fact that the two latter
investigators failed to take into account the existence of forces of attraction and repulsion between molecules. The
accurate form of the relation between η and δ has been a topic of considerable discussion by theoretical physicists [39].

One interesting contribution to this subject, to which reference is made in a subsequent section, is the model suggested
by W. Sutherland [40]. Let us assume that the molecules are "smooth rigid elastic spheres surrounded by fields of
attractive force."

The relation between coefficient of viscosity and molecular diameter then has the form

where the constant C is a measure of the strength of the attractive forces between the molecules. In this equation the
notation δm has been introduced to indicate that the value of the molecular diameter derived in this manner is different
from the value δ used in Eq. (1.114). From a comparison with Eq. (1.114) it follows that Eq. (1.117) is equivalent to the
relation

Equation (1.117) leads to the relation, used for calculating C, which has the form

< previous page page_32 next page >


< previous page page_33 next page >
Page 33

where ηT and η0 are the values of the coefficient of viscosity measured at T and T0 respectively. Equation (1.122) is
also often expressed in the form

By combining Eqs. (1.123) and (1.118) it is seen that the value of δm thus deduced is independent of T and actually
corresponds to the value of the molecular diameter at infinitely high temperature, whereas, as pointed out below, the
value of δ decreases with increase in T. For this reason, Eqs. (1.118) and (1.119) were used formerly by many writers on
this subject. However, following the procedure of Chapman and Cowling and of Kennard, Eqs. (1.115) and (1.116) are
used in the following discussion for the calculations of the values of δ and L, respectively. The conversion to values
deduced by application of Sutherland's theory may then be made by means of Eqs. (1.120) and (1.121).

From Eq. (1.116) the following numerical relations are derived:

where Pµb and Pτ denote the pressure in microbars and Torr, respectively, and η is expressed in poises. These relations
have been used to calculate values of L and, from these, values of δ given in Tables 1.6 and 1.7.

From Eq. (1.115) we obtain the following relations for δ2 and Sc = πδ2, the latter being defined as the mean equivalent
cross section for viscosity [41]:

Where δ has been determined by some other method [42], an approximate calculation of L may be made by means of
the relations

Another useful magnitude is the collision frequency per molecule, which is given by the relation
Values of ω are shown in the last row in Table 1.6.

< previous page page_33 next page >


< previous page page_34 next page >
Page 34

Table 1.6. Mean Free Paths, Molecular Diameters, and Related Data for a Number of Gases
Gas: H2 He Ne Air O2 Ar CO2 Kr Xe
107η15
871 1943 3095 1796 2003 2196 1448 2431 2236
x
0.69 0.64 0.67 0.79 0.81 0.86 0.95 0.85 0.92
107η0
839 1878 2986 1722 1918 2097 1377 2372 2129
107η25
892 1986 3166 1845 2059 2261 1496 2502 2308

103
8.39 13.32 9.44 4.54 4.81 4.71 2.95 3.69 2.64

11.04 17.53 12.42 5.98 6.33 6.20 3.88 4.85 3.47

9.31 14.72 10.45 5.09 5.40 5.31 3.34 4.06 2.98

12.26 19.36 13.75 6.69 7.10 6.67 4.40 5.34 3.93


108δ
2.75 2.18 2.60 3.74 3.64 3.67 4.65 4.15 4.91
C
84.4 80 56 112 125 142 254 188 252
1014Ns
15.22 24.16 17.12 8.24 8.71 8.54 5.34 6.69 4.78
109ω
14.45 7.16 1.68 6.98 6.26 5.70 8.61 6.48 5.71

Table 1.7. Mean Free Paths, Molecular Diameters, and Related Data for Water and Mercury Vapors

t(°C) Pt 105η 108δt 1014Ns


H2O 2.90 6.34 × 104 4.68 5.27
0 4.58 8.69
15 12.79 9.26
3.37 1.42 × 104
25 23.76 9.64

Hg 6.28 1.99 × 104 4.27 6.32


219.4 31.57 46.66
4.87 1.74 × 103 4.50 5.70
150.0 2.807 39.04
3.93 1.44 × 102 4.70 5.22
100.0 0.2729 33.56
2.66 1.45 5.11 4.42
25.0 0.0018 25.40
6.26(J)
0.0 16.2(J)
Note: In Tables 1.6 and 1.7 the following notation has been used:
Pτ = vapor pressure at t (°C) in Torr.

= mean free path in centimeters at 0°C and 1 Torr.

= mean free path in centimeters at 25°C and 1 Torr.


= mean free path in centimeters at t°C and P Torr.
ω = collision frequency (per second) at 25°C and 760 Torr.

Although Sutherland's equation (1.123) has been used most frequently to express the temperature variation of viscosity, several other relations
have been derived in the literature. The simplest of these is the exponential relation

where a and x are constants characteristic of each gas. This relation has been recommended especially for relatively small ranges of temperature.

Equations (1.123) and (1.133), as well as some other relations, have been tested by Licht and Stechert [43] for a number of typical gases and
vapors, using for this purpose data published in the LandoltBornstein tables [44].

< previous page page_34 next page >


< previous page page_35 next page >
Page 35
According to these investigators, "For twenty-four representative gases and vapors at atmospheric pressure, Sutherland's equation has been found
to fit extensive experimental data with an average error of less than 1 percent." Values of the constants C and K in Eq. (1.123) and of a and x in
Eq. (1.133), taken from the original discussion, are shown in Table 1.8. These constants were used to derive values of η25 shown in the fifth and
eighth columns of the table.

The values of C given in Table 1.6 are those deduced by Schuil [45], and it is of interest to compare them with those given in Table 1.8.

In deriving the values of L and δ shown in Tables 1.6 and 1.7, the values of η used are those given by Kennard [46] for 15°C. Values for 0°C and
25°C (η0 and η25, respectively) were derived from the values for 15°C by means of Eq. (1.133), using the values of x, taken from Kennard's
table, which are given in the second row of Table 1.6. The values of δ were derived from those of η0 and therefore apply strictly only at 0°C.

Combining Eq. (1.133) with Eq. (1.114), it follows that at room temperature, δ varies as T0.5(x0.5). Since x is greater than 0.5, it also follows
that the calculated value of δ must decrease with increase in temperature.

An interesting method of calculating values of L for air as a function of the temperature has been used by Tsien [47]. From Eqs. (1.116) and (1.67)
it follows that

L = 1.255ηγ1/2/ρu.

Thus L may be expressed in terms of the "kinematic viscosity," η/ρ, and the velocity of sound, u. By means of this relation, Tsien [47] has
calculated values of the mean free path for air for the range 0500°C. A few of the values thus deduced (in centimeters) are as follows:

t (°C): 0 20 40 100 200 300 400 500


106L(P/P0): 5.89 6.48 7.06 8.78 11.73 14.86 17.78 20.83

Table 1.8. Characteristic Constants for ViscosityTemperature Functions


Substance Temperature Sutherland Equation Exponential Equation
Range (°C)
C 106 K 105η25 106a x 105η25
Ammonia
77441 472 15.42 10.30 0.274 1.041 10.30
Argon
183827 133 19.00 22.67 2.782 0.766 21.83
Benzene
0313 403 10.33 7.58 0.299 0.974 7.71
Carbon dioxide
981052 233 15.52 15.03 1.057 0.868 14.86
Helium
258817 97.6 15.13 19.66 4.894 0.653 20.18
Hydrogen
258825 70.6 6.48 9.04 1.860 0.678 8.85
Mercury
218610 996 63.00 25.03 0.573 1.082 27.20
Methane
18499 155 9.82 11.14 1.360 0.770 10.92
Nitrogen
191825 102 13.85 17.80 3.213 0.702 17.60
Oxygen
191829 110 16.49 20.78 3.355 0.721 20.42
Water vapor
0407 659 18.31 9.84 0.170 1.116 9.82

< previous page page_35 next page >


< previous page page_36 next page >
Page 36

In the lower row, P0 denotes the standard pressure, 1 atmosphere, and P denotes any other value of the pressure in
atmospheres. The value 5.89 for 0°C is to be compared with the value 5.98 given for in Table 1.6.

In the case of H2O, for which values of L and δ for a series of temperatures are given in Table 1.7, the Sutherland
relation was used with C = 650 and η15 = 9.26 × 105 cgs unit.

In the case of Hg (see Table 1.7), the values of η used were based on that given by Braune, Basch, and Wentzel [48] for
t = 219.4°C. Values at other temperatures were derived by means of Sutherland's relation, with C = 942.2 [49]. It should
be observed that in their publication these authors used Eq. (1.118) to calculate values of δm.

The value of η for 0°C is quoted by Jeans (from Kaye and Laby, Physical Constants, 1936 edition) in his book (p. 183).

In addition, Tables 1.6 and 1.7 give values of Ns, the number of molecules per square centimeter, to form a
monomolecular layer at 0°C. On the assumption that the spacing is that of a close-packed (face-centered) lattice, we
have

Formulas for the viscosity of mixed gases are given by Kennard [50]. As he points out, the viscosity of a binary mixture
does not necessarily lie between the values for the pure components; it may be below or above both these values.

The relations for collision frequency per unit volume between molecules are of importance in many problems of
interaction between molecules. Let ZAA and ZAB denote the number of collisions between like and unlike molecules,
respectively, per cubic centimeter, per second. Then,

where nA = number of molecules per cubic centimeter of A at pressure PA (microbars), with similar definition for nB
and δAB = ½ (δA + δB).

For instance, for nitrogen (M = 28.02, δ = 3.62 × 108) at T = 298.2 K, we have

Z = 8.157 × 1016 cm3·s1 at 1 microbar


= 8.357 × 1028 cm3·s1 at 1 atmosphere.
< previous page page_36 next page >
< previous page page_37 next page >
Page 37

For the mean free paths of the molecules A and B in a mixture, the following general relations have been derived:

where vA and vB refer to the average velocity of each type of molecule.

For TA = TB, Eq. (1.138) becomes

and similarly for 1/LB.

For and we obtain

For we obtain

For and TA not identical with TB, we obtain

The last equation has been used by Gaede [51] to calculate the mean free path of nitrogen (A) in the blast of a mercury-
vapor pump. In this case we may assume the following: TA = 300 K; and the temperature of the mercury vapor, TB, is
400 K. At this temperature the pressure of mercury vapor is about 1 Torr. Hence, nB = 2.414 × 1016.

From the data in Tables 1.6 and 1.7, we have

108 · δAB = ½(3.78 + 4.70) = 4.24.

Substituting these values and those for MA and MB in Eq. (1.143), the result is LA = 6.73 × 103 cm, whereas from the
data for in Table 1.7 the value derived for the mean free path of mercury molecules in the saturated vapor at 400 K
is L = 4.4 × 103 cm.

1.7.1
Viscosity at Low Pressures
As mentioned previously, the coefficient of viscosity is not independent of pressure when the pressure decreases to a
low value. Under those conditions it was observed by Kundt and Warburg (1875) that the damping of a vibrating
surface by the surrounding gas is decreased, as if the gas were slipping over the surface. If one surface is at rest and
another surface, at a distance d, is moving parallel to the first surface with

< previous page page_37 next page >


< previous page page_38 next page >
Page 38

a uniform velocity, u, the viscous drag upon each surface at normal pressures is given in accordance with the definition
of η by the relation

At low pressures, however, the observed value of the tangential force B is less than that given by Eq. (1.144),
corresponding to an increase in the value of d. That is, the equation assumes the form

where ζ is known as the coefficient of slip.

Considerations based on the kinetic theory of gases lead to the relation

which, combined with Eq. (1.116), leads to the relation

where f is a numerical coefficient which has a maximum value of 1. It was introduced by Maxwell with an interpretation
given by Kennard [52] as follows:

The value of f, the transfer ratio for momentum, will presumably depend upon the character of the interaction
between the gas molecules and the surface; it may vary with the temperature. We can imagine a surface that is
absolutely smooth and reflects the molecules ''specularly" with no change in their tangential velocities; in such
a case f = 0 and ζ = ∞, viscosity being unable to get a grip upon the wall at all. On the other hand, we can
imagine the molecules to be reflected without regard to their directions of incidence and therefore with
complete loss of their initial average tangential velocity.

In this case, f = 1 and ζ = L [53].

More generally, ζ = βL, where β is of the order of unity; and at very low pressures, where Eq. (1.145) becomes

Thus, at very low pressures the rate of transference of momentum from a moving surface to another surface adjacent
and parallel is directly proportional to the pressure and to the velocity of the moving surface. This conclusion was
applied by Langmuir [54] to the design of a molecular gauge for the measurement of low pressures.

The quantity ρva/4, which, as noted previously, corresponds to the mass of gas incident on unit area per unit time, has
been designated by Kennard as the free-molecule viscosity of the gas between the plates.
< previous page page_38 next page >
< previous page page_39 next page >
Page 39

The concept of coefficient of slip has also been used to interpret observations on the flow of gases at very low pressures
through capillaries.

1.7.2
Molecular Diameters

Mention has been made in the previous section of the two relations for deducing values of δ and of δm from viscosity
measurements. These relations are as follows:

As illustrated in Table 1.7 by the values of δ for mercury, these values exhibit a considerable decrease with increase in
temperature, and in this case it is found that δm = 2.50 × 108. The variation with T is obviously greater for those
molecules for which the Sutherland constant C, has a large value.

In spite of the fact that the values of δ thus deduced exhibit a variation with T, Chapman and Cowling [1d] have been
followed in this discussion in choosing Eq. (1.150) rather than Eq. (1.151). While, as pointed out by Chapman and
Cowling, such a variation in the value of δ with T "receives a simple explanation on the hypothesis that the molecules
are centers of repulsive forces, not hard spheres," it is of interest to compare the results obtained by means of Eq.
(1.150) with those obtained by means of Eq. (1.151) and also by other methods.

There are several such methods, and only a few of the more important ones can be mentioned briefly, together with
some of the results deduced.

1.7.3
Application of the van der Waals Equation

Near the critical temperature and pressure the behavior of gases can be described very satisfactorily by a modified form
of Eq. (1.14), deduced by van der Waals, which is as follows:

In this equation, V is the volume per mole, the term a/V2 is a correction term which takes into account the attractive
forces between the molecules, and the constant b is a measure of the actual volume of the total number of molecules in
accordance with the relation

< previous page page_39 next page >


< previous page page_40 next page >
Page 40

The values [55] of the constant b may be determined for any given gas from the values of the critical temperature (Tc) and
critical pressure (Pc) by means of the relation

1.7.4
From the Density of the Solid or Liquid

Assuming that the molecules are closely packed, as in a face-centered cubic lattice, the projected area per molecule is given
by the relation [56]

where m = mass of molecule = M/NA and ρ = density of condensed phase.

But, according to Eq. (1.134),

Equation (1.156) has been used to calculate the number of molecules per unit area required to form a unimolecular layer (or
monolayer). (See last column of Table 1.9.)

Table 1.9. Values of Molecular Diameter (cm·108)


Gas From η From From Electron Collision 1014·Ns
b ρ from ρ
δ0 δm
H2 2.75 2.10 2.76 4.19 2.2
6.58
He 2.18 1.69 2.66 4.21 1.7
6.49
Ne 2.60 2.16 2.38 3.40 2.2
9.98
Ar 3.67 2.42 2.94 4.15 3.6
6.71
O2 3.64 2.50 2.93 3.73 3.4
8.30
Hg 6.26 2.50 2.38 3.26
10.86
CO2 4.65 3.32 3.24 4.05 4.4
7.04
H2O 4.68 2.45 2.89 3.48 3.8
9.53
C6H6 7.65 4.71 4.51 5.89
3.33
CH4 4.19 3.31 3.24 4.49
5.73
C2H6 5.37 3.87 3.70 5.01
4.60
C3H8 6.32 4.45 4.06 5.61
3.67
n-C4H10 7.06 4.84 4.60 6.10
3.18
n-C5H12 7.82 5.05 4.89 6.45
2.78
n-C6H14 8.42 5.22 5.16 6.74
2.54

< previous page page_40 next page >


< previous page page_41 next page >
Page 41

1.7.5
Cross Section for Collision with Electrons [57]

A cathode-ray beam of initial intensity I0 is decreased to intensity I, after passing through a layer of the gas of thickness
x, in accordance with the relation

where, as shown previously in Eq. (1.104), α is a measure of 1/Le, where Le = mean free path for electrons = 4(2)½L.
Hence it follows from Eq. (1.113) that

Thus the collision cross section is given by α/n. However, as has been observed experimentally, the value of α varies in
a rather complex manner with the potential used to accelerate the electrons, with the result that it is actually impossible
to assign a definite value to δ as derived from electron-collision measurements.

The structure of molecules has also been determined from measurements of dipole moments and from electron-
diffraction experiments, all of which are discussed at length by Stuart [57] in his book.

Table 1.9 gives, for comparison, values of δ for a number of gases and vapors as deduced by at least four different
relations. The second and third columns give values of δ0 and δm as deduced by means of Eqs. (1.126) and (1.120),
respectively, from the values of η (extrapolated to 0°C) and the values of C given by Schuil [45]. The fourth column
gives values of δ calculated by means of Eq. (1.154) from the values of the van der Waals constant, b; the fifth column
gives values deduced by means of Eq. (1.158) from values of ρ at extremely low temperatures (in general for the solid
state). The sixth column gives values, derived from observed values of α by means of Eq. (1.160), for 36-volt electrons
[57], and the last column gives values of Ns calculated by means of Eq. (1.157) from the values of δ in the fifth column
[58]. The values thus derived are to be compared with those deduced in Tables 1.6 and 1.7 from kinetic-theory values of
δ.

1.8
Heat Conductivity of Gases

The kinetic theory of gases achieved a great triumph when it led to the conclusion that the viscosity is independent of
the pressure. It led to still further important results when it predicted the existence of simple relations between the
properties of viscosity, heat conductivity, and diffusivity.

From the kinetic point of view it is the same whether the molecules transfer momentum or translational energy from one
layer to another. The equations are quite analogous.

As in the case of viscosity, we consider any two layers CE, EH (Fig. 1.4), each of thickness L, between two plates
whose temperatures are T1 and T2 and distance apart d. Let cv denote the heat capacity per unit mass at constant
volume. The relative temperature drop between the planes CD and HK is equal to

2(T1 T2)L/d.

< previous page page_41 next page >


< previous page page_42 next page >
Page 42

Hence the heat transferred per unit area is

Therefore the coefficient of heat conductivity

If cv is expressed in calories per gram, the unit of λ is 1 cal · cm1 · s1 · deg1 · Comparing the last equation with Eq.
(1.110), it follows that

As in the case of the relation for η, a more careful consideration of the mechanism of energy transfer by means of the
molecules leads to the relation

where, according to Eucken [59, 60], we have

and γ = ratio of specific heat at constant pressure to that at constant volume.

For monatomic gases, γ = 5; and for polyatomic gases, γ tends to approach the value 1, with increase in total number of
atoms per molecule.

Hence

1 ≤ε≤ 2.5.

Table 1.10 gives data published by Kannuluik and Martin [61, 62] on values of λ0 (conductivity at 0°C) and of ε, as
derived from observation by means of Eq. (1.164) and as calculated by means of Eq. (1.165). Similar data for these
gases and a number of others are given in the treatise by Chapman and Cowling [1d].

One important conclusion that follows from Eq. (1.164) is that the thermal conductivity of a gas is independent of
pressure, which is valid as long as the pressure is higher than the range in which molecular flow occurs. (See the
following section).

With regard to the variation in λ with T the following remarks may be made. To a first approximation the variation in
value of λ follows that in the value of η, since ε exhibits only a slight variation with T. However, for larger ranges of T,
account must be taken of the increase with T in the value of cv. Denoting the molecular specific heat, at constant
volume, by Cv, Partington and Shilling [63] give the relations shown in Table 1.11 for different gases:

Since the variation in η with T is given by the Sutherland relation, Eq. (1.123), it follows that, in terms of λ0 in calories
per centimeter per second per degree,
< previous page page_42 next page >
< previous page page_43 next page >
Page 43

Table 1.10. Values of Heat Conductivity Compared with Coefficients of Viscositya


Gas cv (cal·g1) ε (obs) ε (calc)
105λ0 105η0
He
34.3 18.76 0.746 2.45 2.44
Ne
11.12 29.81 0.150 2.50 2.44
Ar
3.82 21.02 0.0745 2.44 2.44

H2
41.3 8.50 2.43 2.00 1.90
Air
5.76 17.22 0.171 1.96 1.91
O2
5.83 19.31 0.157 1.92 1.95
CO
5.37 16.65 0.178 1.81 1.91

CO2
3.43 13.74 0.153 1.64 1.72
N2O
3.61 13.66 0.155 1.71 1.73
a To convert values of λ0 from cal · cm1 · s1 · deg1 to watts·cm1 · deg1 multiply values in Table 1.10 by 4.186.

Table 1.11. Molecular Specific Heat and Specific Heat Ratio (Partington and Shilling)
Gas γ = Cp/Cv Cv
Air
1.4034 4.924 + 1.7 × 104T + 3.1 × 107T2
N2
1.405 4.924 + 1.7 × 104T + 3.1 × 107T2
O2
1.396 4.924 + 1.7 × 104T + 3.1 × 107T2
CO
1.404 4.924 + 1.7 × 104T + 3.1 × 107T2
H2
1.408 4.659 + 7.0 × 104T
CO2
1.302 5.547 + 4.5 × 103T 1.02 × 106T2
H2O
... 6.901 1.19 × 103T + 2.34 × 106T2
Hg, Ar, etc.
1.667 2.990

where α and β are determined from the expression for Cv as a function of T, and lT is expressed in watts per centimeter per degree. For
hydrogen,
< previous page page_43 next page >
< previous page page_44 next page >
Page 44

Similar expressions for gases for which λ has not been determined can be derived, as is evident, from determinations of
η, using the observed values of C and values of ε calculated from those of γ by means of Eq. (1.165).

It will be observed that the heat conductivities of hydrogen and helium are much greater than those of heavier gases,
such as oxygen and carbon dioxide.

For the case of a wire of radius a and length l, suspended along the axis of a cylinder of radius r, the energy loss per unit
time due to thermal conduction by the gas is [64]

where T T0 is the difference in temperature and r/a is not "excessively" large [65] while lm is the average conductivity
over the temperature range T T0.

The units in which E is usually expressed are watts per square centimeter.

The energy loss per unit area per unit time is

and the energy loss per unit length of wire per unit time is

Since the total energy loss from a heated wire is the sum of that lost by radiation (which varies as ) and that
lost by conduction of the gas, the former has to be subtracted from the total energy loss in order to obtain the amount
due to the latter. Furthermore, in the case of short wires especially, a correction has to be made for the loss by
conduction at the ends.

For wires of low emissivity, such as platinum, operating at a temperature below about 500°C, the loss due to radiation is
negligible compared to that due to conduction.

Since the thermal-conduction loss in the case of mixtures varies with both the nature of the gas and the composition, this
fact has been applied to the analysis of gases [66]. An instrument devised for this purpose by Shakespear [67], known as
a katharometer, consists of a platinum spiral filament in a copper block. This instrument has been used extensively by
English investigators for determinations of rates in gaseous reactions and for experiments on thermal transpiration
(which is discussed in Section 1.10) [68].

1.9
Thermal Conductivity at Low Pressures

As mentioned in the previous section, the heat conductivity of gases should theoretically be independent of pressure.
That this is at least approximately confirmed
< previous page page_44 next page >
< previous page page_45 next page >
Page 45
by observation is illustrated by the data shown in Table 1.12, obtained by Dickins [69].

The values under W represent the total heat loss (in calories) by conduction, from a platinum wire (a = 3.765 × 103 cm, l = 20.09
cm) suspended along the axis of a Pyrex glass tube (inside radius, r = 0.3346 cm). The tube was maintained at about 0°C by
external cooling, and ∆t is the temperature differential between the wire and the wall. The pressure of the gas in Torr is given in
the third column, and the value indicated by W∞ represents the heat loss extrapolated for P = ∞. The values under λt give the
thermal heat conductivity at the mean temperature indicated in the last column. The values of λt (in calories) were derived by
means of the relation, deduced from Eq. (1.167),

It will be observed that over a range of pressures the values of W did not exhibit any considerable decrease.

That the heat conductivity is practically constant over a large range of pressures is also shown by the plots of the energy loss (in
watts) at constant temperature (about 99°C) from a platinum filament (Fig. 1.5). The filament was a 14-cm length of 3-mil wire
located along the axis of a glass tube 2.54 cm in diameter. The wall temperature was 0°C.

As will be observed, the heat loss in hydrogen was about 10 times that in argon. The loss at a pressure of about 1 m Torr was
0.006 W. A comparison of the relative losses in the three gases at 760 Torr yields values which agree, within a few percent, with
the values of λ given in Table 1.10.

Table 1.12. Variation in Thermal Conduction with Pressure (Dickins)


Gas ∆t Pτ W W∞ 105λt t (°C)
H2
17.866 625.0 0.21323 0.21494 42.77 9.04
442.0 0.21258
302.3 0.21145
229.9 0.21038
168.2 0.20875
129.8 0.20692

Air
23.832 91.7 0.04021 0.04052 6.044 11.94
52.1 0.03997
31.3 0.03961
22.3 0.03926
17.1 0.03891
11.9 0.03825

CO2
23.80 83.3 0.02461 0.02473 3.694 11.91
40.5 0.02447
19.2 0.02417
11.1 0.02379
< previous page page_45 next page >
< previous page page_46 next page >
Page 46

Thus, over the range 100760 Torr, the heat loss in hydrogen increased only about 7%, whereas in the range below 50
Torr the decrease was nearly 100%. Careful measurements show that at very low pressures the thermal conductivity
decreases linearly with the pressure.

The theory of heat conduction at these pressures has been developed from two different points of view. The first of
these, due to Knudsen [70], involves a consideration of the mechanism of energy transfer by individual molecules
incident on the hot surface.

The second point of veiw, due to Smoluchowski [71], is based upon the concept of a temperature discontinuity which is
the thermal analogue of the phenomenon of "slip" discussed in Section 1.7.

1.9.1
Free-Molecule Conductivity (Knudsen)

When molecules originally at a temperature Ti strike a hot surface at temperature Ts (> Ti), complete interchange of
energy does not occur at the first collision. In fact
Fig. 1.5
Plots illustrating the variation in thermal conductivity
with pressure, for nitrogen, argon, and hydrogen. Ordinates
give values of total watts conducted from a platinum
filament located along the axis of a cylindrical glass
tube. Scale of watts for hydrogen should be multiplied
by 10. Abscissas give pressures in centimeters
of mercury (10 Torr).

< previous page page_46 next page >


< previous page page_47 next page >
Page 47

it may often require many collisions for this to occur. Therefore Knudsen introduced a constant, known as the accommodation coefficient,
designated by α, which "can be defined as standing for the fractional extent to which those molecules that fall on the surface and are reflected or
re-emitted from it, have their mean energy adjusted or 'accommodated' towards what it would be if the returning molecules were issuing as a
stream out of a mass of gas at the temperature of the wall [72].

The molecules re-emitted or reflected from the hot surface consequently possess a mean energy which corresponds to a temperature lower than
Ts, which we shall designate as Tr, and the accommodation coefficient is defined by the relation

For α = 1, Tr = Ts; for α < 1, Ts > Tr > Ti. It should be noted that the temperature Tr is not clearly defined, as Blodgett and Langmuir [73]
have pointed out, unless the molecules leaving the surface have a Maxwellian distribution of velocities.

Because the faster molecules emitted from a surface carry more energy than the average, the Maxwell distribution function must be replaced by

= 2(m/2kT)2v3 exp( mv2/2kT)

so that the mean translational energy transferred from a surface at temperature T is given by

(instead of 3/2kT, which is the average energy of the molecules in a volume).

Let us now consider heat transfer in a monatomic gas at low pressure:

E0 = energy transfer from hot to cold surface (at temperature Ti) per square centimeter of hot surface per second

where vi = average velocity at Ti.

Thus the rate of energy transfer at low pressure is proportional to the pressure and the temperature difference.

For diatomic and polyatomic gases, the molecules striking the hot surface acquire not only increased translational energy but also increased
amounts of both rotational and vibrational energy. The amount of the vibrational energy possessed by molecules as compared with the amount
of translational energy is measured by the value of γ. In these cases, a detailed calculation leads to the relation

which for γ = 5/3 (case of monatomic gases) becomes identical with Eq. (1.173).

< previous page page_47 next page >


< previous page page_48 next page >
Page 48

In this equation, α has the value [74]

where α1 and α2 are the values of the accommodation coefficient for the two surfaces.

Substituting for vi from the relation in terms of Ti and M, Eq. (1.174) assumes the form

where Λ0 = free-molecule heat conductivity at 0°C

(In terms of calories per second, Λ0 should be multiplied by 0.2389.)

Table 1.13 gives values of Λ0 in watts per square centimeter per degree per mTorr for a number of gases and vapours
[75], and, for comparison, values of 4.186 λ0 (derived from the values in Table 1.10 and from other sources) which
correspond to the conductivity at 0°C in terms of watts per centimeter per degree.

In terms of calories (15°C) per mole, we have

R0 = 8.3145 × 107/4.1855 × 107 = 1.9865 cal · deg1 · mol1·


Table 1.13. Values of Molecular Heat Conductivity
Gas M γ 106Λ0 4.186 × 104λ0
H2
2.016 1.41 60.72 17.30
He
4.003 1.67 29.35 14.36
H2O
18.016 1.30 26.49
Ne
20.18 1.67 13.07 4.66
N2
28.02 1.40 16.63 2.38
O2
32.00 1.40 15.57 2.44
Ar
39.94 1.67 9.29 1.60
CO2
44.01 1.30 16.96 1.44
Hg
200.6 1.67 4.15

< previous page page_48 next page >


< previous page page_49 next page >
Page 49

Since γ + 1 = (2Cv + R0)/Cv and γ 1 = R0/Cv, where Cv = molar specific heat in calories (at constant volume) and R0
has the above value, in all of the above expressions we can set

Hence Eq. (1.177) can be expressed in the form

For coaxial cylinders of radii r1 and r2 (r2 > r1), the rate of energy transfer per unit area from the inner cylinder or
wire, at temperature Ts and at low pressures, is given by the relation

and when α1 = α2 = α this becomes [76]

Jeans [77] has pointed out that, according to general dynamical considerations, the constant α should be determined by
a relation of the form

where m′ = mass for molecules striking the surface for which the molecular mass is m.

Morrison and Tuzi [78] have used a molecular beam method with mass spectrometer detector to measure the thermal
accommodation coefficient for translational energy of a number of vapors on Pyrex glass.

For completely roughened surfaces, α = 1. Values of α which have been observed for different surfaces are given in
Table 1.14 [79].

Keesom and Schmidt [80] have measured the thermal accommodation coefficients for He, Ne, H2, and N2 on
Thuringian glass at 0°C at pressures from 0.04 to 0.2 Torr. Average values were 0.335 for He, 0.286 for H2, 0.853 for
N2, and 0.670 for Ne.
Klett and Irey [81] have measured the thermal accommodation coefficients for air, nitrogen, and helium on commercial
copper surfaces at pressures from 2 × 104 Torr to 4 × 103 Torr, obtaining 0.799 for N2 at 77.4 K, 0.760 for N2 at 243
K, 0.823 for air at 77.4 K, 0.698 for air at 243 K, 0.564 for He at 77.4 K, and 0.407 for He at 243 K.

< previous page page_49 next page >


< previous page page_50 next page >
Page 50

Table 1.14. Values of the Accommodation Coefficient for Several Gases


Surface H2 O2 CO2
Polished Pt
0.358 0.835 0.868
Pt slightly coated with black
0.556 0.927 0.945
Pt heavily coated with black
0.712 0.956 0.975

H2 N2 Ar Hg Air
Tungsten
0.20 0.57 0.85 0.95
Ordinary Pt
0.36 0.89 0.89 0.90

1.9.2
Temperature Discontinuity (Smoluchowski)

In Section 1.7, mention was made of the phenomenon of ''slip" which occurs in a gas at moderately low pressures. In
Eq. (1.145) we introduced a coefficient of slip ζ = βL to account for the apparent decrease in viscosity at low pressures.

An analogous phenomenon was observed by Smoluchowski in the case of thermal conduction at low pressures. For two
parallel plane surfaces separated by a distance d, the heat loss per unit area per unit time may be represented at these
pressures by a relation of the form

where λm is the mean conductivity over the range T1 T0.

In this equation, g is a coefficient defined by the relation

where ∆T represents the temperature discontinuity at any one surface, and dT/dx is the temperature gradient normal to
the surface.

As shown by Kennard,

where the numerical constant ε has a value, according to Eq. (1.165), which varies between 1 and 2.5, and β′ is a
numerical constant of the order of unity.
Hence Eq. (1.183) may be expressed in the form

where α is assumed to be the same for both surfaces, L = c/P, and the symbol E0P indicates that the conductivity per
unit area is a function of P.

< previous page page_50 next page >


< previous page page_51 next page >
Page 51

A comprehensive treatment of the heat transport between parallel plates in a gas of rigid sphere molecules for the whole
range of rarefaction conditions from a dense Clausius gas to the extreme rarefaction of a gas with a Knudsen number
has been published by Frankowski, Alterman, and Pekeris [82]. Willis [83], using a relaxation-type model for
intermolecular collisions, has derived equations based on the Krook [84] kinetic model for the heat transfer in a rarefied
gas between parallel plates at large temperature ratios. He has also considered heat transfer and shear between coaxial
cylinders for large Knudsen numbers [85].

The problem of heat transfer between concentric spheres in the near free-molecule region has been treated by Havekotte
and Springer [86] and other authors cited in their article.

For a hot wire of radius a situated along the axis of a cylindrical tube of radius r, Eq. (1.186) must be replaced by the
relation

where it is assumed that α is the same for both the wire and inside surface of the cylinder.

For very low values of P, Eq. (1.186) becomes

and Eq. (1.187) becomes

That is, at very low pressures E0P is independent of d and varies linearly with P. This conclusion is identical with that
deduced above by Knudsen for the case of free-molecule flow. That Eq. (1.188) is identical with Eq. (1.176) may be
shown as follows.

Substituting for β′ from Eq. (1.185), substituting for λm from Eq. (1.164), and utilizing the relation η = 0.5ρvaL, we
obtain the result
where α/(2 α) takes the place of α used in Eq. (1.176).

< previous page page_51 next page >


< previous page page_52 next page >
Page 52

Equations (1.186) and (1.187) can be expressed in a form which is very convenient for calculation. We shall consider
especially Eq. (1.187), which is more important in practical measurements of heat conductivity. This equation may
obviously be written in the form

That is, if 1/E0P is plotted against 1/P, a straight line is obtained. From the intercept for 1/P = 0, which we shall
designate by 1/E0 ∞, the value of λm may be obtained, while the slope is given by the relation

where X is a constant determined by the relation

Equation (1.190) thus applies to the transition range of pressures in which E0 changes from being independent of
pressure (at higher pressures) to varying linearly with the pressure (at very low pressures).

In a very interesting paper published in 1912 [87, also see 88], Langmuir formulated a theory of the mechanism of heat
conduction from wires (as well as plane surfaces). Observations on the heat losses in hydrogen from tungsten wires
heated to incandescence showed that whether the wire was in a horizontal or vertical position made a difference of only
a few percent (at constant temperature) in the heat loss. "This," states Langmuir, "was strong indication that the heat
loss was dependent practically only on heat conduction very close to the filament and that the convection currents had
practically no effect except to carry the heat away after it passed out through the film of adhering gas."

In the case of a wire of radius a, Langmuir assumed that the energy loss will occur through a "stationary" film of
adherent gas, the outer edge of which is at a distance b from the axis of the wire. As shown by Langmuir, the value of b
then satisfies the relation

and the energy loss per centimeter length of wire is given by the relation

< previous page page_52 next page >


< previous page page_53 next page >
Page 53

in which λ(T) is given by Eq. (1.166).

Introducing into Eq. (1.195) the relation expressed by Eq. (1.192), it follows that

From measurements of the energy losses from platinum wires in air [89], Langmuir derived the conclusions that, for air
at 20°C and 760 Torr, B = 0.043 cm and is "surprisingly independent" of T. He also concluded that B should be
proportional to the viscosity and inversely proportional to the density, "for it is the viscosity that causes the existence of
the film and it is the difference of density between hot and cold gas (proportional to the density itself) that keeps the
film from becoming indefinitely large."

The problem of heat conduction at low pressures from a wire of radius a, suspended coaxially in a glass tube of radius r,
where r is very much greater than a, has been discussed from the point of view of Langmuir's film theory by Jones [88].
Brody and Körösy [90] have shown that although 90% of the wattage lost can be accounted for by assuming pure
conduction, the film is not "stationary," but convection currents exist with the velocity of convection increasing as the
surface of the filament is approached up to the limit (0.25 mm) of their measurements for incandescent tungsten
filaments in nitrogen.

1.10
Thermal Transpiration (Thermomolecular Flow)

From Eq. (1.82),

it follows that the rate of efflux of a given gas through a small opening varies as ρ(T)½. If we have two chambers A and
B, separated by a small orifice, at two different temperatures, TA and TB, transpiration will occur until a steady state is
established at which

the limiting ratio of pressure is

It was expected that the same result would be obtained when the chambers are separated by a wall containing a porous
plug or by a single long tube of diameter 2a
< previous page page_53 next page >
< previous page page_54 next page >
Page 54

such that the Knudsen number because under this "molecular flow" condition the probability of
passage of a molecule through a long tube is the same in both directions when the molecules enter the ends with a
cosine law distribution. Also, when the effect should be independent of the wall temperature of the
connecting tube. However, Hobson et al. [91] found that, while Eq. (1.199) is obeyed for apertures, this lower limit is
not necessarily reached when the two chambers are separated by a long tube in which the pressure is so low that
. Their data showed that the measured lower limit was aRm with a ranging from 1.1 to 1.3, with lower values
of a as the tube diameter is reduced and the molecular weight M increased. Miller and Buice [92] attempted to explain
the need for the correction factor a based on a model in which the faster moving molecules have a higher probability of
being specularly reflected. They applied a Monte Carlo method to calculate the flow through the tube with a special law
of scattering at the wall, but their results were not in good agreement with the experimental data of Hobson and
coworkers.

Thermal transpiration is of importance in the application of vacuum gauges and vacuum microbalances at low pressures.
For instance, if the chamber A is a part of a system at liquidair temperature (TA = 90) and the pressure is measured by
means of a gauge at room temperature (TB = 300), then the real value of PA is given by the relation

PA = PB(90/300)1/2 = 0.55PB.

Because the envelope of a tubulated ionization gauge is usually at a temperature much higher than room temperature, if
the gauge is not used under the same temperature conditions as those during its calibration, a correction based on Eq.
(1.199) may be required [93]. Poulis et al. [94] have shown that longitudinal Knudsen forces associated with the thermal
transpiration effect can give spurious mass variations in vacuum microbalances when temperature gradients exist along
the balance case.

When the mean free path L is very much smaller than 2a, so that collision between molecules become predominant over
collisions against the walls, then the condition of equilibrium is

There exists therefore a range of pressures in which the pressure ratio PA/PB changes from that given by Eq. (1.200) to
that given by Eq. (1.199) in the case of an aperture, and to a(TA/TB)1/2 in the case of narrow tubes, with a being an
experimentally determined factor between 1.0 and 1.3.

The theory of the phenomenon was first discussed by Maxwell (1879) and subsequently by Knudsen [95], who derived
equations for the pressures gradient along the connecting tube as a function of the tube diameter and pressure in the hot
chamber. For the pressures and temperatures in the two vessels, Knudsen proposed

< previous page page_54 next page >


< previous page page_55 next page >
Page 55

the semiempirical relation

which for reduces to Eq. (1.199), and for reduces to the equilibrium condition PA = PB. For a temperature
gradient along the connecting tube results in a boundary layer flow of gas (thermal creep) along the wall of the tube from the cold to the
hot side balanced by a return flow through the interior from the hot side to the cold side [96], and complicated theoretical equations for
the thermal transpiration effect in this case have been proposed by various authors.

Bennett and Tompkins [97] have recommended that an equation proposed by Liang [98, 99] be used. The equation is supported by a
kinetic argument, and Bennett and Tompkins found it to fit experimental data over a wide range. It should be noted, however, that these
workers used glass systems. Los and Fergusson [100] observed that the nature of the wall affects the magnitude of the correction, as
Liang had predicted [99].

The form of Liang's equation given by Bennett and Tompkins can be expressed as follows:

in which

R = the ratio of the pressures in the two regions at different temperatures, TA and TB,
b = 1 for connecting tubes of internal diameter, d < 1 cm
= 1.22 for connecting tubes of internal diameter, d > 1 cm,
X = PBd,
αHe = 3.70 (1.70 2.6 × 103∆T)2 (∆T = TB TA),
βHe = 7.88[1 (TA/TB)½] for (TA/TB)½≤ 1,

and values of φg are given in Table 1.15.

Bennett and Tompkins also point out that φg values may be calculated from the collision diameters, r0, given by Hirschfelder, Bird, and
Spotz [101], by making use of

Table 1.15. Experimental φg Values for Glass Systems (Bennett and Tompkins)
Gas: He Ne Ar Kr Xe H2 O2 N2 CO CO2 C2H4
φg: 1.00 1.30 2.70 3.90 6.41 1.44 2.87 3.53 3.31 4.52 6.72

< previous page page_55 next page >


< previous page page_56 next page >
Page 56

the equation

Arney and Bailey [102] studied the thermal transpiration effect for air, argon, and helium for Knudsen numbers from
0.01 up to 100 and for temperature ratios from 1.5 to 3.8 by an unusual technique involving two tubes, one with large
diameter and the other with small diameter, joined together by an annular junction at which the temperature could be
raised by an electrical resistance heater while the other ends of the tubes were maintained at a fixed temperatures by
water cooling. The pressure at the cold ends of the two tubes was measured for a series of temperature ratios, and the
pressure at the junction was assumed to be that of the cold end of the larger tube as long as the Knudsen number Kn was
≤ 0.01 for the larger tube. Then when Kn reached higher values in the larger tube, the pressure ratio between the cold
end and the junction could be estimated from the experimentally determined variation of the ratio of pressure at the cold
end of the smaller tube to that at the junction with the Knudsen number for the smaller tube which was determined when
Kn for the larger tube was still less then 0.01. Comparison of their results with the predictions of Knudsen's equation
[103] for dp/dT showed agreement at the upper and lower ends of the range of Kn but somewhat lower values of the
ratio of the pressure at the cold end to that at the hot end for Kn in the middle range from 0.1 to 10. They present a chart
of curves, based on their measurements, of pressure ratios versus temperature ratios for a series of Knudsen numbers
from 0.03 to 10.

Hobson [104] compared the experimental data obtained by Edmonds and Hobson [105] for helium and neon effusing
through an aperture with 2a = 20 mm, TA = 77.4 K, and TB = 295 K with the values predicted by Eq. (1.201) and by Eq.
(1.202). For the mean free path in Eq. (1.201), Hobson used the relation of Weber and Schmidt [106]:

where (PL)0 = 0.1339 Torr · mm and n = 0.147 for helium while (PL)0 = 0.08841 Torr · mm and n = 0.20 for neon, and
T was chosen as 295 K. An appropriate choice of the constants αHe, βHe, and φg in Liang's equation gave a curve of
pressure ratio R versus PB in very good agreement with the experimental data, while the curve based on the simple Eq.
(1.201) agreed with the experimental data at the lower limit 0.512 and upper limit 1.0 but was slightly higher at values
of PB in the middle range.

Podgurski and Davis [107] obtained results that agreed with Bennett and Tompkins' equation [Eq. (1.202)] for argon
and xenon but deviated appreciably from this equation for neon and hydrogen at low pressures.

Sharipov [108] has made extensivse calculations based on a modified Boltzmann equation of the exponent γ in the
equation

as a function of the Knudsen number (which he defines as Kn = L/a) and other factors, for long capillary tubes of radius
a under conditions of a balance between the

< previous page page_56 next page >


< previous page page_57 next page >
Page 57

flow due to the pressure gradient and the flow due to the temperature gradient. He tried to compare his calculations with
the experimental results of Edmonds and Hobson as recalculated using a rarefaction factor proportional to the inverse
Knudsen number at PB and TB but found a disagreement for different capillaries.

Dietz [109] studied the thermal transpiration effect in the mass spectrometer inlet system for argon in the leak tube and
found that sensitivity response was not affected if gas mixtures and calibrating gases are run at the same pressure.

Variable capacitance diaphragm transducers used as transfer standards in an intercomparison of pressure standards in
the low vacuum region of 1 to 10 Torr are operated at a stabilized elevated temperature to reduce inherent sensitivity to
ambient temperature changes, and thermal transpiration causes a pressure differential between the transducer and the
vacuum system which must be determined in order to achieve maximum accuracy of calibration comparisons. Jitschin
and Röhl [110] have studied this problem quantitatively and applied correction procedures based on Eq. (1.202) above
using both the Liang expression for f(X) and the more recent formula of Takaishi and Sensui [111]

where the parameters A*, B*, and C* depend on the gas species and are determined experimentally. The correction
according to the Liang formula was in poor agreement with their experimental data, but the correction using the formula
of Takaishi and Sensui reproduced the experimental data within about 4%. The degree of agreement seemed to depend
on the material used for the connecting tube, and by adopting an effective temperature (TA)eff and an effective tube
diameter they were able to obtain agreement to better than 0.1%.

Williams [112] derived equations for thermal transpiration in the near-continuum regime starting with the NavierStokes
equations and using the slip boundary condition at the wall including thermal creep. Comparison of his equations with
the experimental data of various authors showed good agreement for low Knudsen numbers but gave too small values as
the Knudsen number increased.

1.11
Thermal Diffusion

The thermal-diffusion effect has been described by Ibbs [113] as follows:

If a temperature gradient is applied to a mixture of two gases of uniform concentration, there is a tendency for the
heavier and large molecules (mass m1, diameter δ1) to move to the cold side, and for the lighter and smaller molecules
(mass m2 and diameter δ2) to move to the hot side. The separating effect of thermal diffusion (coefficient DT) is
ultimately balanced by the mixing effect of ordinary diffusion (coefficient D12) so that finally a steady state is reached
and a concentration gradient is associated with the temperature gradient. The amount of thermal separation thus
produced by a given difference in temperature depends upon the ratios m1/m2 and δ1/δ2 and upon the proportion by
volume of the heavier gas f1 and of the lighter gas

< previous page page_57 next page >


< previous page page_58 next page >
Page 58

f2 (where f1 + f2 = 1) and also upon the nature of the field of force operating between the unlike molecules.

The effect was predicted by Enskog (1917) and, independently, by Chapman and Dootson [114]. Following the earlier
experimental investigations of Ibbs and others, Clusius and Dickel [115] devised a continuous method for separating
mixtures of gases and isotopes which has been applied extensively by subsequent investigators.

In this section the discussion will be limited largely to the experimental data, since the mathematical theory is quite
complex and beyond the scope of the objectives of this chapter.

The coefficient of thermal separation is defined by the relation

where DT and D12 have been defined above. Now

If kT is a constant, then we obtain from Eq. (1.207) the relation

for the amount of separation, where T1 and T2 denote the hot and cold temperatures, respectively.

Figures 1.6 and 1.7 illustrate results obtained by Ibbs [116] with mixtures of H2 and N2. The concentrations of H2 in
the different mixtures varied from 4.7% for No. 2
Fig. 1.6
Relation between degree of separation by thermal diffusion
and log(T1/T2) for mixtures of H2 and N2.

< previous page page_58 next page >


< previous page page_59 next page >
Page 59

Fig. 1.7
Relation between composition of H2N2 mixtures
and degree of separation for log(T1/T2) = 0.2.

to 50.5% for No. 7. It will be observed that these observations are in accord with Eq. (1.208), since ∆f is found to be a
linear function of log(T1/T2). Figure 1.7 shows the variation in value of ∆f with composition for a constant value of the
ratio T1/T2.

Chapman has shown that kT can be expressed in the form [117]

where the coefficients A to D are functions of the molecular weights and of the ratio δ1/δ2. These can be calculated
from the force constants s12 in the repulsive force relation F = c/rv, where v = s12 for unlike molecules and r = 1/2(δ1 +
δ2).

The separation in a mixture of H2 and CO2 was investigated by Ibbs and Wakeman [118] over a range of 600°C. For
any given mixture it was observed that the amount of separation was in accordance with Eq. (1.208). For a constant
difference in temperature, ∆f increased from 0 for pure H2 to a maximum value of 0.0365 (for ∆T = 600°C), for a
mixture containing 61% H2.

According to Chapman, kT in this case should be given by a relation of the form

where f1 refers to CO2 and f2 refers to H2. Since this relation is deduced on the assumption that the molecules behave
as rigid elastic spheres with a force relation in which s12 has a definite value for the interaction between the two types
of molecules, it was expected that the observed values of kT would differ from those calculated. Table 1.16 gives the
results of a series of measurements. The last column gives values of a constant defined by the relation
< previous page page_59 next page >
< previous page page_60 next page >
Page 60

Table 1.16. Separation Measurements in a Mixture of H2 and CO2


f1 Temperature Range (°C) kT (calc) kT (obs) RT
0.469 0.1575 0.0665 0.422
Below 144
0.1575 0.0939 0.596
Above 144
0.1575 0.1045 0.663
At 470

0.34 0.1685 0.0695 0.415


Below 145
0.1685 0.0929 0.552
Above 145

For CO2N2 mixture, Chapman's expression for kT is of the form

where f1 again refers to the heavier gas. (In all the following equations, f1 has the same significance.)

For f1 = 0.494, the value of kT calculated is 0.0564, whereas the value of RT observed varied from 0.247 for temperatures below 144°C to 0.441
for temperatures above 144°C.

For the mixtures ArHe, NeH2, and N2He, relations similar to Eq. (1.210) and Eq. (1.212) for kT have been published by Ibbs and Grew [119].
All the observed values of RT showed a decrease with decrease in temperature down to 190°C.

According to Chapman and Cowling,

Hence observations on the values of RT should yield values of s12. In fact the observations on thermal diffusion of gases should reveal
interesting information on the forces between molecules. Table 1.17 gives values of the repulsive force constant s12, as determined from
observed values of RT, and, for comparison, values of s1 and s2 for the individual gases, as determined from other phenomena.

In 1938, as mentioned previously, Clusius and Dickel [115] devised an arrangement of hot and cold surfaces that could be utilized practically for
the separation of mixtures of different gases and of isotopes. They used a long vertical tube with a hot wire along the axis. Because of thermal
diffusion, the relative concentration of the heavier molecules is greater at the cold wall. Convection causes the gas at the hot surface to rise to
the top, where it is deflected to the cold wall. There the gas sinks to the bottom and the cycle is then repeated. As a result, the heavier
components concentrate at the bottom, whereas the lighter ones concentrate at the top.

For instance, in a chamber 1 meter in height, for a temperature difference of 600°C, it was possible to separate a mixture containing 40%
(by volume) CO2 and 60% H2, with a yield of practically 100% CO2 at the bottom and 100% H2 at the top. In

< previous page page_60 next page >


< previous page page_61 next page >
Page 61

Table 1.17. Values of ''Force" Constants for Interaction of Molecules


Gases RT s12 s1 s2
O2H2 0.48
7.3 7.6 11.3
N2Ar 0.47
7.2 8.8 7.35
H2N2 0.58
8.2 11.3 8.8
HeAr 0.65
9.0 14.6 7.35
H2Ne 0.74
11.4 11.3 14.5
H2CO2 0.47
7.2 11.3 5.6
NeAr 0.54
7.9 14.5 7.35
HeNe 0.80
11.4 14.6 14.5

a chamber 2.9 meters in height, for a temperature difference of 600°C, air (21% O2, 78% N2) could be separated, yielding 85% O2 at the
bottom. Finally, from a mixture of 23% HCl37 and 77% HCl35, a mixture containing 40% of the heavier isotope was obtained at the bottom of
the diffusion chamber.

Brewer and Bramley [120] used a heated inner tube 1 cm in diameter and an outer cooled concentric tube 2 cm in diameter, each 1 meter long.
With a 350°C difference in temperature, He and Br2 could be separated, so that after 15 min no Br2 could be detected at the top. Under
similar conditions a 5050 mixture of CH4 and NH3 was enriched 25% in NH3 at the bottom. Brewer and Bramley deduced the following
relation: "If l denotes the cylinder length, r the radius of the outer tube, and d the difference in radii, then, to a first approximation, the rate of
separation varies as rd and the purity as l/d."

In a thermal-diffusion column developed by Nier [121] for the concentration of C13H4, a Nichrome heating wire threaded through porcelain
insulators was inserted along the axis of a steel tube 3/4 in. in outside diameter, which was concentric with a steel tube 1¾ in. in outside
diameter having a wall thickness of 0.035 in. This tube constituted the inner wall of the annular space containing the gas, and the outer wall
consisted of a 2-in.-outside-diameter brass tube, which was water-cooled. The column, which was 24 ft in length, is used as a return lead for the
current flowing through the Nichrome wire. At the lower end, the two steel tubes were brazed concentrically to a circular steel plate. The
average temperature of the outer wall was 27°C, and that of the inner wall was approximately 375°C. The samples of methane taken at the
bottom and top were analyzed by means of a mass spectrometer for the ratio of C13H4 to C12H4. At a pressure of 656 Torr and after operating
for 23.5 hours, the values of this ratio at the bottom and top were 0.0215 and 0.0054 respectively, so that the ratio of C13H4 at the bottom to that
at the top was 3.99.

The results obtained were found to be in agreement with conclusions deduced from theoretical considerations by Furry, Jones, and Onsager
[122]. They derived the relation

< previous page page_61 next page >


< previous page page_62 next page >
Page 62

where

represent equilibrium concentrations of C13H4 in the lower and upper ends of the column, respectively,
a and b are constants for the gas used and the dimensions of the column, and p denotes the pressure. Thus εx
corresponds to the separation factor. In the investigation under consideration, this factor was observed to pass through a
maximum at p = 0.6 atm, approximately.

The separation of C13H4 and C12H4 has also been investigated, in a column 40 ft in length, by Taylor and Glockler
[123], who found that the separation factor could be represented, in agreement with the theory, by the above equations.

The rate at which equilibrium is approached in such a column has been discussed by Bardeen [124] and the conclusions
reached have been found to be in good agreement with experimental results obtained by Nier.

Although the above discussion has dealt largely with the separation of gases which differ in value of M, it follows from
the theory of thermal diffusion that it should be possible by this method to effect a certain degree of separation in a
mixture of gases which have the same value of M but differ with respect to the magnitude of the molecular diameter.
Results in agreement with this prediction have been obtained by Wall and Holley [125].

1.12
Theory of Diffusion of Gases

As in the case of viscosity and heat conduction, approximate kinetic-theory considerations lead to the conclusion that
the coefficient of self-diffusion, D, is given by a relation of the form

which indicates that, at constant temperature, D varies directly as the mean free path or inversely as the pressure.

Combining Eq. (1.216) with the relation

Introducing the correction for persistence of velocities and that for Maxwellian distribution of velocities, it is found that
the more accurate relation is

According to Chapman and Enskog [126], the coefficient in Eq. (1.217) should have a value lying between 1.200 for
hard spheres and 1.543 for inverse-fifth-power repulsion.

< previous page page_62 next page >


< previous page page_63 next page >
Page 63

Actually observed values of D yield the following values of the coefficients [127]:

CO, 1.30; H2, 1.37; O2, 1.40; CO2, 1.58.

For the derivation of a relation for the coefficient of interdiffusion, D12 (corresponding to the fact that molecules 1 and
2 are involved), we shall follow the StefanMaxwell derivation [128], which, as will be evident, is based on the same
considerations as those used in deriving Eq. (1.216) above.

Letting n1 and n2 represent the number of molecules of each kind per unit volume at a location (x, y, z) and assuming
the concentration gradients to occur only with respect to the x-axis, then for a gas at rest we obtain

dn1/dx + dn2/dx = 0.

The rate of flow, J1, of molecules of the first type countercurrent to the flow of molecules of the second type at the rate
J2 is defined by

where D12 is the mutual diffusion coefficient or coefficient of interdiffusion. The Meyer formula [129] for D12 is

where v1 and v2 are average velocities.

The StefanMaxwell formula for the mean free path, L1, in a mixture of two gases, designated 1 and 2, according to Eq.
(1.140), is

However, as Stefan and Maxwell have pointed out, the collisions between like molecules do not influence diffusion, and
hence for this case the mean free paths are given by the relations

Since for thermal equilibrium


< previous page page_63 next page >
< previous page page_64 next page >
Page 64
and

On this basis it is found that the relation for the coefficient of interdiffusion is

where n = n1 + n2.

It is evident that for like molecules this relation becomes

Combining this with the relation

η = 0.499ρva/(21/2πnδ2),

the result is

which is in substantial agreement with Eq. (1.217).

Table 1.18 [130] gives the values of D12 (cm2 · s1) observed for several pairs of gases, at 0°C and 1 atm, values of δ12 calculated
from the values of D12 by means of Eq. (1.224), and, for comparison, values of δ12 calculated from viscosity measurements (see
Table 1.6) for each of the two gases.

For the first six pairs, which are comprised of "hard" molecules (with a value for the repulsive force constant greater than 7.4), the
agreement is very satisfactory. Even in the case of "softer" molecules the agreement is fair, as shown by the values for the second
set of five pairs.

Table 1.18. Coefficients of Interdiffusion and Average Molecular Diameters


Gases D12 (obs) 108δ12 (calc from D12) 108δ12 (calc from η)
H2air
0.661 3.23 3.23
H2O2
0.679 3.18 3.17
O2air
0.1775 3.69 3.68
O2N2
0.174 3.74 3.70
COH2
0.642 3.28 3.25
COO2
0.183 3.65 3.70

CO2H2
0.538 3.56 3.69
CO2air
0.138 4.03 4.20
CO2CO
0.136 4.09 4.22
N2OH2
0.535 3.57 3.69
N2OCO2
0.0983 4.53 4.66

< previous page page_64 next page >


< previous page page_65 next page >
Page 65

Summerhays [131] has used the katharometer (see Section 1.8) to measure the diffusion coefficient of water vapor in
air. The value observed was D = 0.282 cm2 · s1 at 16.1°C.

As will be noted, Eq. (1.224) indicates that the interdiffusion coefficient should be independent of composition.
However, a series of experiments carried out at Halle [132] to test this conclusion show that there is a variation of not
more than a few percent with variation in the ratio n1/n2.

Mention should be made in this connection of a relation for the coefficient of interdiffusion, given by Loeb [133], which
is of the form

in which β is a numerical factor, the value of which is between 1.00 and 1.50, and n = n1 + n2.

From the equations for D it follows that to a first approximation D should vary as T3/2 and as P1. Since, however, η
varies with T in accordance with Sutherland's relation, Eq. (1.122), it is expected that the exponent of T would exceed
3/2. Lonius [132] gives the following relation for D as a function of T:

where DT = value at T (in degrees Kelvin) and Pt (in Torr) and D = value at 15°C and 760 Torr. The values of x used by
Lonius for several binary mixtures are all in the neighborhood of 1.75.

Ivakin and Suetin [134] have measured the temperature dependence of the diffusion coefficients of 18 pairs of gases in
the temperature range 290470 K. Using their experimental data they calculated the interaction potential parameters for
potential functions of the Lennard-Jones type and of the modified Buckingham type and also for point center repulsion,
φ(r) = d/rδ. For the latter case they used the formula

D12 = D0TN,

where D0 is given by a complicated formula taken from the treatise by Hirschfelder, Curtiss, and Bird [135], while N =
3/2 + 2/δ. The calculated δ values ranged from 4.33 for H2Ar to 14.29 for Heair. For comparison they also present
potential parameters obtained from viscosity measurements and the second virial coefficient. The potential parameters
obtained from the diffusion equations gave the best fit to the temperature behaviour.

1.12.1
MaxwellLoschmidt Method for Determination of Diffusion Coefficients

In view of the fact that in vacuum technique a problem arises occasionally regarding the time required for gases to mix,
it is of interest to review briefly one method which has been used for the determination of D for gases.

< previous page page_65 next page >


< previous page page_66 next page >
Page 66

A tube of length l is divided by means of a wide-bore stopcock into two parts of equal length. This is mounted in a vertical position; the
heavier gas (A) is put in the lower part, and the lighter gas (B) is put in the upper part. At time t = 0 the stopcock is opened and diffusion is
allowed to occur for a definite period, t. The stopcock is then closed, and the proportions of A and B are measured in each compartment.

Let U = amount of A expressed as the fraction of the total number of moles in the lower compartment,
Q = amount of A in the upper compartment, expressed similarly.

Then, on the basis of the equations for diffusion,

For instance, in the case of H2O2 mixture, for l = 99.93 cm and t = 1800 s we obtain U = 0.5982 and Q = 0.4018 at 14.8°C and 749.3 Torr.
Hence, at 15°C and 760 Torr, D = 0.788.

Table 1.19 gives values of UQ for two different values of l as a function of Dt, along with values of t, for l = 100 cm, assuming D = 0.2 cm2
· s1.

For l = 100, only the first term in the series on the right-hand side of Eq. (1.229) is of importance for values of Dt > 500. For l = 25, only the
first term is of importance for Dt > 100. Equation (1.230) shows that for values of l and Dt for which the first term only is sufficiently
accurate, t varies as l2 for constant value of (U Q).

Ziering and Sheinblatt [136] have considered diffusion theory for both thermal gradients and density gradients in rarefied gases valid in the
entire pressure range from the continuum domain to the free-molecular-flow domain as applied to

Table 1.19. Illustrating the Application of Eq. (1.229)


Dt l = 100 cm l = 25 cm t (min) for D = 0.2 cm2 · s1
UQ UQ
100 0.9550 0.2063 8.33
200 0.8401 0.0425 16.67
300 0.7514 0.0088 25.00
400 0.6771 33.34
500 0.6118 41.67
700 0.5014 58.67
1000 0.3728 83.83
1500 0.2276 125.00
2000 0.1390 166.67
3000 0.0518 250.00
5000 0.0071 416.67

< previous page page_66 next page >


< previous page page_67 next page >
Page 67

molecules which obey a center of repulsive force F = K12/rv, where v = 5 (Maxwell molecules) and in the extreme of
rigid-sphere molecules, v = ∞.

Cunningham and Geankoplis [137] have presented equations for diffusion in three-component mixtures in the transition
region between Knudsen and molecular diffusion in an open system. Solution of the equations is, by trial and error,
aided by computer.

1.12.2
Effect of Pressure of Gas on Rates of Evaporation of Metals

In Section 1.5, equations were derived for rates of evaporation in a vacuum. These equations have been applied, as
illustrated in the above connection, to the calculation of vapor pressures of high-melting-point metals from observations
on rate of loss in weight as a function of the temperature. However, in the presence of a gas which does not react
chemically with the metal, the observed rate of evaporation is lower; as is well known, this fact was utilized by
Langmuir in the invention and development of the gas-filled tungsten-filament lamp. This phenomenon has been
explained by Langmuir by assuming the existence of a film adjacent to the evaporating surface through which the atoms
evaporated from the surface diffuse.

Fonda [138] has shown that this theory accounts very satisfactorily for his observations on the rate of evaporation of
tungsten in the presence of argon at various pressures, and the following remarks on the mechanism of evaporation
under these conditions are quoted from his paper:

The filament is considered to be surrounded by tungsten vapor at the same pressure as would be present in a
vacuum. The atoms of this vapor, however, instead of being projected directly from the filament, as in a
vacuum, are pictured as diffusing through the stationary film of gas. Once an atom reaches the outer boundary
of the film, it would be carried away in the convection current of gas and would be lost to the filament; but the
path within the film is so irregular that an atom may in fact return to the filament and be deposited on it, thus
leading to a reduced evaporation as compared with that in a vacuum.

Let dc/dr denote the concentration gradient at the surface of the wire, where r is the distance from the axis, and let D
denote the diffusion constant. Assuming uniform distribution over the surface of the wire, the rate of diffusion per unit
length of the wire is

where D is given by Eq. (1.224).

Since the value of n for tungsten atoms is negligible compared with that of n for argon gas, the expression for D
becomes identical with Eq. (1.216), where L and va refer to the values of the mean free path and average velocity,
respectively, of the tungsten atoms at the temperature of the filament. Thus D varies inversely as the pressure, P of
argon, and Eq. (1.231) becomes

where A is a proportionality constant.

< previous page page_67 next page >


< previous page page_68 next page >
Page 68

For a wire of diameter 2a and gas film of diameter 2b, it follows, from the same considerations as those that lead to Eq. (1.192) for the heat loss, that

Let m denote the rate of evaporation in grams per square centimeter per second. Then it follows from Eqs. (1.232) and (1.233) that

where ca and cb are the concentrations of the tungsten atoms at r = a and r = b, respectively.

Table 1.20 shows results obtained by Fonda for the rate of evaporation from a filament in a mixture of argon and nitrogen, such as is used in
gas-filled lamps. The value m = 230 × 109 g · cm2 · s1 observed for P = 0 is in agreement with the value 250 × 109 observed by Langmuir,
Jones, and Mackay [139].

As Fonda states: "The expression maPcm log (b/a) is sufficiently constant at pressures above 10 cm (100 Torr) to allow of credence being given
the hypothesis developed above."

The constancy of the expression has a further significance, because, as is evident from its derivation, it denotes a constant difference between
the vapor pressure of tungsten at the surface of the filament and at the surface of the film of gas. For constant filament temperature this implies
that the concentration of tungsten vapor at the surface of the stationary film of gas should be constant for all pressures above 10 cm. It is this
vapor which constitutes that which effectively evaporates from the filament, because the rising convection currents of gas carry it off to be
deposited on the bulb. The rate of evaporation at different pressures should be determined therefore by the area of the assumed film of gas if
the concentration of vapor at its border is in

Table 1.20. Rate of Evaporation (m) of Tungsten in 86% Argon, 14% Nitrogen at 2870 Ka
Pcm 109m b (cm) 109maPcm log(b/a) 109m/b
0 230
1 57.5 9.68 3.9
5 23.5 2.42 6.3
10 20.5 1.31 9.8 15.6
25 10.3 0.63 10.4 16.2
50 5.4 0.36 9.6 14.8
70 4.2 0.28 9.6 14.8
165 2.0 0.15 8.8 13.5
aDiameter of filament (2a) = 0.00978 cm; Pcm = pressure in cm of Hg.

< previous page page_68 next page >


< previous page page_69 next page >
Page 69

fact a constant. The constancy of the ratio m/b is therefore a further confirmation of the validity of the theory. Other
evidence, based on the appearance of the surface of the filament, has also been shown by Fonda to be in agreement with
the views expressed above.

From Eq. (1.234) it also follows that at constant pressure and for filaments of different diameters the expression ma log
(b/a) should be constant. The validity of this conclusion was confirmed by Fonda in a series of observations made with
both pure nitrogen and a mixture of nitrogen and argon [140]. In these cases the value actually used for B (the thickness
of the film for a plane surface in gas at atmospheric pressure) was the same as that obtained from data on heat
conduction in nitrogen.

1.13
Random Motions and Fluctuations [141]

Let us consider the case in which a group of N molecules start from the plane z = 0, at the instant t = 0, and diffuse
through the gas. The differential equation for the diffusion is

At any instant, t, the distribution of these molecules along the z axis is given by the relation

which satisfies the condition

Hence, the average value of z is

And similarly it can be shown that the root-mean-square value is given by the relation

That is, the total net displacement in any given direction varies as t1/2.

These equations are of special importance in the determination of diffusion coefficients from observations on the
Brownian motion of small particles.

< previous page page_69 next page >


< previous page page_70 next page >
Page 70

We can also express Eqs. (1.237) and (1.238) in terms of the mean free path L, on the basis of the relation

Substituting in the above equations we obtain the relations

The collision frequency per molecule is

ω = va/L,

while the total length of path actually traversed by a molecule in time t is

That is, the total length of path varies as the square of the net displacement from the point of origin at t = 0.

As an illustration of the above equations let us consider the case of molecules in air at 25°C and atmospheric pressure.
Since L = 6.69 × 106 and va = 4.67 × 104, it follows from Eq. (1.239) that D = 0.104 cm2 · s1, and for t = 60 s we
obtain zr = 3.53 cm, while l = 6.98 × 109 t = 4.19 × 1011 cm.

That l = 1.19 × 1011zr is obviously due to the fact that after each collision the probability of a displacement toward
lower (or more negative) values of z is just as great as that of a displacement toward more positive values of z.

The equations for random motion have been applied to determine the value of Avogadro's constant from observations
on rates of diffusion of Brownian particles. It may be demonstrated that the rate of diffusion of spherical particles is
given by the relation
where a = radius of particle and η = viscosity of medium.

< previous page page_70 next page >


< previous page page_71 next page >
Page 71

Also, the mean square of the displacement per second is expressed as

Values of NA obtained by application of these relations, though not nearly as accurate as those obtained from
determinations of the unit electric charge (the method used by Millikan), are in good agreement with them.

1.14
Scattering of Particle Beams at Low Gas Pressures

As calculated on the basis of classical collision theory of smooth elastic spheres, a neutral projectile particle of mass m1
traveling with uniform velocity w1 before scattering by a target molecule of mass m2 moving with average speed, w2,
much smaller than the magnitude of the projectile velocity so that we can set w2 = 0, will suffer a fractional loss of
energy (see Section 1.3.).

so that, using Eqs. (1.39) and (1.44),

where θ is the angle of incidence between w1 and the line of centers at collision in the laboratory system.

When the target particle is at rest in the laboratory system, the center of mass moves in this system with the constant
velocity

In the center-of-mass coordinate system, CM, the velocity of the projectile particle is w1 w12 and the velocity of the
target particle is w12. By combining the various vectors in both systems in one diagram it can be shown [142] that

where β is the angle of scattering in the center-of-mass system. Then the fractional energy loss can be written as

where βL is the angle of scattering in the laboratory system.

< previous page page_71 next page >


< previous page page_72 next page >
Page 72

For a parallel monoenergetic beam of projectile particles uniformly distributed in a plane normal to the direction of
motion, the probability of a particle being scattered in the angle between βL and dβL is

so that the integral of this probability over all angles from 0 to π equals 1. There is no backwards scattering from an
initially stationary target particle in the laboratory reference frame.

When as in the case of elastic scattering of an electron, the scattering angle βL is approximately equal to β
and the laboratory system is almost the same as the center-of-mass system.

For a molecular beam emerging into a ''perfect vacuum" there will still be some scattering as the faster molecules
overtake the molecules moving more slowly. By dividing the beam molecules into two groups, each with a narrow
range of differing velocities but following a Maxwellian law of distribution, Troitskii [143] has calculated that the mean
free path for collision of molecules within a unidirectional beam (far from the source opening) is about three times the
mean free path in a stationary gas of the same particle density as in the beam, neglecting second collisions with particles
diffusely scattered by a first collision.

When the projectile particles are charged and moving with very high speeds, the target particle becomes the nucleus of
an atom in the gas, and the scattering is designated as Coulomb scattering. At pressures of 106 Torr in the vacuum
chamber of a synchrotron, or the beam tube of particle accelerators in general, the mean free path for a hydrogen atom
in nitrogen might typically be of the order of 104 cm, but the "mean free path" for collision of a hydrogen nucleus
(proton) with the nucleus of a nitrogen atom might be more like 107 cm, depending on the velocity. When the projectile
particle is charged, the effect of scattering is compensated to a large degree by focusing magnetic fields which redirect
the particle away from the walls while being accelerated by electromagnetic fields over a total path length of nearly cta,
where ta is the acceleration time. For ta = 1 s, this path length is 3 × 1010 cm, so that many collisions with the nuclei in
gas molecules at 106 Torr will occur. Without magnetic focusing, even deflection through a small scattering angle β
would be sufficient to scatter the projectile particle into the wall of the vacuum chamber.

In the presence of the magnetic fields, elastic scattering gives rise to "betatron" oscillations with amplitudes which
increase slowly with time until the particle is captured by the wall of the vacuum chamber. Blachman and Courant [144]
in 1948 published formulas which make possible an estimate of the fraction of the original particles in a circular
synchrotron which is scattered to the wall. They begin with the Rutherford formula for the scattering cross section, dσ,
of a particle of charge e by the nucleus of a target particle with charge Ze, into any solid angle, dω, at the angle β,

where m is the relativistic mass of the projectile particle and v0 is its velocity. Because of the finite size of the nucleus,
there is an upper limit on β given by λ/b, where

< previous page page_72 next page >


< previous page page_73 next page >
Page 73

λ = h/mv0 is the de Broglie wavelength and b is the "radius" of the nucleus. These authors give an equation for the
allowable pressure in Torr inside the vacuum chamber of a synchrotron which is directly proportional to the parameter

when the effects of damping are included, where is the mean square maximum amplitude of the vertical betatron
oscillations and A is the vertical semiaperture of the chamber cross section. If no more than a 10% loss through vertical
scattering is to be permitted, then η must be no more than 0.089 and the calculated upper limit of pressure is of the order
of 1.3 × 104 Pa (106 Torr) in typical synchrotrons. In 1949 Blachman and Courant [145] derived new equations to apply
to the "racetrack" modification of the synchrotron by the insertion of straight sections.

In 1951 Greenberg and Berlin [146] considered the gas scattering problem for protons and for electrons in great detail
and estimated the limiting pressures at which 50% of the protons are lost and at which 10% of the protons are lost in the
Brookhaven and Birmingham Synchrotrons. The values depend on the method of particle injection, but in general
indicate that pressures less than 1.3 × 104 Pa (106 Torr) are required to avoid loss of more than 10% in these machines.
Similar calculations were made in 1955 by Moravcsik and Sellen [147] for the strong focusing electron synchrotron at
Cornell.

In 1953 Courant [148] revised the previous gas scattering theory to include effects of scattering through angles large
enough for immediate particle loss. Kheifets [149] has used scattering formulas to estimate the lifetime in a storage
device of a beam of electrons. Various methods for computing particle loss by gas scattering in cyclical accelerators
were compared by Didenko and Serdyutskii [150] in 1963. Orlov and Kheifets [151] have used revised formulas with
appropriate boundary conditions for estimating the particle loss due to multiple Coulomb scattering in a cyclical
accelerator.

References and Notes

1. In connection with this chapter, use has been made of the following treatises:

a. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics. McGraw-Hill, New York,
1938.

b. L. B. Loeb, Kinetic Theory of Gases. McGraw-Hill, New York, 1934.

c. J. H. Jeans, An Introduction to the Kinetic Theory of Gases. Macmillan, New York, 1940; Cambridge University
Press, Cambridge, UK, 1960.

d. S. Chapman and T. G. Cowling, The Mathematical Theory of Non-Uniform Gases. University Press, Cambridge,
UK, 1939; 3rd ed., 1970.

Some other treatises which are recommended for further details are:

G. A. Bird, Molecular Gas Dynamics. Oxford University Press (Clarendon), Oxford, 1978.

C. Cercignani, Mathematical Methods in Kinetic Theory. Plenum, New York, 1969.

E. A. Guggenheim, Elements of the Kinetic Theory of Gases. Pergamon, New York, 1960.

M. N. Kogan, Rarefied Gas Dynamics. Plenum, New York, 1969.

L. C. Woods, An Introduction to the Kinetic Theory of Gases and Magnetoplasmas. Oxford University Press,
Oxford, 1993.
T. Wu, Kinetic Equations of Gases and Plasmas. Addison-Wesley, Reading, MA, 1966.

< previous page page_73 next page >


< previous page page_74 next page >
Page 74

2. The values given for V0 and other constants in this and subsequent sections are those recommended by CODATA: E.
R. Cohen and B. N. Taylor, Rev. Mod. Phys. 59, 11211148 (1987); Phys. Today, August, BG9 (1996).

3. E. Thomas and R. Leyniers, in: Adam (ed.) Trans. 3rd Int. Congr., 1965, Vol. 2, Part 1, pp. 279285 Pergamon,
London (1966); Trans. Czech. Conf. Electron. Vac. Phys., 3rd, Prague, 1965, pp. 457470 (1967).

4. E. Thomas, Vacuum 13, 376 (1963).

5. Document ISO 31/111-1978(E). International Standards Organization, London.

6. Glossary of Terms Used in Vacuum Technology, British Standard 2951. Part 1. British Standards Institution, London,
1969 (first published in 1958).

7. B. B. Dayton, Trans. Natl. Vac. Symp., 1st, Asbury Park, NJ, 1954, pp. 132147 (1955).

8. E. Thomas, ed., Advances in Vacuum Science and Technology (Proc. 1st Int. Congr., Namur, Belg., 1958), Spec.
Meet. Pressure Units, pp. 155158. Pergamon, London, 1960.

9. M. S. Kaminsky and J. M. Lafferty, eds., Dictionary of Terms for Vacuum Science and Technology, Surface Science,
Thin Film Technology and Vacuum Metallurgy. AIP Press, Woodbury, NY, 1980.

10. C. D. Ehrlich, J. Vac. Sci. Technol. A 4, 2384 (1986); C. D. Ehrlich and J. A. Basford, ibid. 10, 1 (1992); G. M.
Solomon, ibid. 4, 327 (1986); G. Lewin, J. Vac. Sci. Technol. 5, 75 (1968).

11. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 110. McGraw-Hill, New
York, 1938.

12. A. Roth, Vacuum Technology, p. 45. North-Holland Publ., Amsterdam, 1976; 3rd ed., 1990.

13. M. S. Kaminsky and J. M. Lafferty, eds., Dictionary of Terms for Vacuum Science and Technology, Surface
Science, Thin Film Technology and Vacuum Metallurgy, p. 71. AIP Press, Woodbury, NY, 1980.

14. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 101. McGraw-Hill, New
York, 1938.

15. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics. p. 6. McGraw-Hill, New
York, 1938.

16. M. Knudsen, Ann. Phys. (Leipzig) [4] 48, 1113 (1915).

17. G. Comsa, R. David and K. D. Rendulic, Phys. Rev. Lett. 38, 775 (1977); G. Comsa, Surf. Sci. 299/300, 77 (1994).

18. W. Van Willigen, Phys. Lett. 28A, 80 (1968); R. L. Palmer, J. N. Smith, Jr., H. Saltsburg and D. R. O'Keefe, J.
Chem. Phys. 53, 1666 (1970); T. L. Bradley, A. E. Dabiri and R. E. Stickney, Surf. Sci. 29, 590 (1972); 38, 313 (1973);
26, 522 (1971); F. C. Hurlbut and F. S. Sherman, Phys. Fluids 11, 486 (1968); S. Nocilla, in Rarefied Gas Dynamics (J.
H. de Leeuw, ed.), Vol. 1, p. 315. Academic Press, New York, 1966.

19. M. Epstein, AIAA J. 5 (October issue), 1797 (1967).

20. D. Da and X. Da, J. Vac. Sci. Technol. A 5, 2484 (1987); R. A. Haefer, Vak.-Tech. 16, 149, 185 (1967); M. Lunc, in
Rarefied Gas Dynamics (J. A. Laurmann, ed.), Vol. 1, pp. 94101. Academic Press, New York, 1953.
21. B. B. Dayton, Ind. Eng. Chem. 40, 795 (1948); Vacuum 15, 53 (1965); W. W. Stickney and B. B. Dayton, Trans.
Natl. Vac. Symp. 10, 105116 (1963); R. A. Haefer, Vak. Tech. 16, 149, 185, 210 (1967); W. Steckelmacher and D.
Turner, J. Sci. Instrum. 43, 893 (1966); S. Komiya and T. Ikeda, in Adv. Vac. Sci. Technol., Proc. Int. Congr., 1st,
Namur, Belg., 1958, E. Thomas (ed.), pp. 323328 Pergamon, London (1960); D. H. Holkeboer, 10th Natl. Vac. Symp.,
292, Macmillan Co., New York (1963).

22. E. P. Muntz, Phys. Fluids 11, 64 (1968).

23. M. Knudsen, Ann. Phys. (Leipzig) [4] 28, 75 [4] (1909).

< previous page page_74 next page >


< previous page page_75 next page >
Page 75

24. I. Langmuir, Phys. Rev. 2, 329 (1913).

25. I. Langmuir, Phys. Rev. 2, 329 (1913), and subsequent papers.

26. F. H. Verhoek and A. L. Marshall, J. Am. Chem. Soc. 61, 2737 (1939).

27. H. A. Jones and I. Langmuir, Gen. Electr. Rev. 30, 310, 354, 408 (1927).

28. D. B. Langmuir and L. Malter, Phys. Rev. 55, 748 (1939).

29. A. C. Egerton, Philos. Mag. [6] 33, 33 (1917).

30. A. C. Egerton, Proc. R. Soc. London, Ser. A 103, 469 (1923).

31. A comprehensive review of this topic has been published by K. C. D. Hickman, one of the pioneers in this field, in
Chem. Rev. 34, 51 (1944).

32. M. Born, Phys. Z. 21, 578 (1920).

33. F. Bielz, Z. Phys. 32, 81 (1925).

34. R. G. J. Fraser, Molecular Rays. Cambridge University Press, Cambridge, UK, 1931; also Molecular Beams.
Methuen, London, 1937.

35. J. J. Weigle and M. S. Plesset, Phys. Rev. 36, 373 (1930).

36. F. Knauer and O. Stern, Z. Phys. 53, 766 (1929).

37. J. B. Taylor, Ind. Eng. Chem. 23, 1228 (1931); also see W. H. Bessey and O. C. Simpson, Chem. Rev. 30, 239
(1942).

38. A. C. Torrey, Phys. Rev. 47, 644 (1935); A. B. Van Cleave and O. Maass, Can. J. Res., Sect. B 12, 57 (1935); 13,
384 (1935).

39. This topic is discussed at length by Chapman and Cowling [1d] in their treatise on this subject. It will be observed
that LM = 1.611L = 1.131LB and LB = 1.425L.

40. S. Chapman and T. G. Cowling, The Mathematical Theory of Non-Uniform Gases, Sect. 10.41. University Press,
Cambridge, UK, 3rd ed., 1970.

41. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 147. McGraw-Hill, New
York, 1938.

42. See sub-section 1.7.2 on molecular diameters in Section 1.7.

43. W. Licht, Jr. and D. G. Stechert, J. Phys. Chem. 48, 23 (1944).

44. LandoltBornstein Physikalisch-Chemische Tabellen, 5th ed., Springer, Berlin, 19231935.

45. A. E. Schuil, Philos. Mag. [7] 28, 679 (1939). This paper gives data on values of η for gases and vapors over the
range 0250°C.

46. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 149. McGraw-Hill, New
York, 1938.
47. H.-S. Tsien, J. Aeronaut. Sci. 13, 653 (1946).

48. H. Braune, R. Basch, and W. Wentzel, Z. Phys. Chem. Abt. A 137, 447 (1928).

49. Value given in Ref. [48], which differs somewhat from that given in Table 1.8.

50. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, pp. 160162. McGraw-Hill,
New York, 1938.

51. W. Gaede, Ann. Phys. (Leipzig) [4] 46, 357 (1915).

52. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 297. McGraw-Hill, New
York, 1938.

53. A. Timiriazeff [Ann. Phys. (Leipzig) [4] 40, 971 (1913)] assumed that f, which may be regarded as an
accommodation coefficient for transfer of momentum, has the same value for any gas-surface combination as α the
accommodation coefficient for heat transfer (discussed in Section 1.9.1); but B. Baule [Ann. Phys. (Leipzig) [4] 44, 145
(1914)], disagreed with this assumption and concluded that the value of the ratio between ζ and L [denoted by β in Eqs.
(1.148) and (1.149)] must be a complicated function of the diameters of the molecules in the gas and those constituting
the surface. The theoretical

< previous page page_75 next page >


< previous page page_76 next page >
Page 76

investigations on this topic are discussed by Kennard and Loeb. Loeb quotes results obtained by R. A. Millikan
[Phys. Rev. 21, 217 (1923)], which lead to values of f ranging from 0.87 to 1.00.

54. I. Langmuir, Phys. Rev. 1, 337 (1913).

55. Values of b for a large number of gases and vapors are given in the Handbook of Chemistry and Physics, published
by CRC Press, Boca Raton, FL. Since these values are expressed in terms of the volume at 0°C and 1 atm as unity, they
should be multiplied by 22,414 to give values for use in Eq. (1.154).

56. H. K. Livingston, J. Am. Chem. Soc. 66, 569 (1944); also S. Brunauer, The Adsorption of Gases and Vapors, p. 287.
Princeton University Press, Princeton, NJ, 1942.

57. H. A. Stuart, Molekülstruktur, pp. 49 et seq. Springer, Berlin, 1934.

58. See also P. H. Emmett and S. Brunauer, J. Am. Chem. Soc. 59, 1553 (1937).

59. A. Eucken, Phys. Z. 14, 324 (1913).

60. L. B. Loeb, Kinetic Theory of Gases, pp. 234252. McGraw-Hill, New York, 1934.

61. W. C. Kannuluik and L. K. Martin, Proc. R. Soc. London, Ser. A 144, 496, (1934).

62. See also the following references on this topic: H. Gregory and C. T. Archer, Proc. R. Soc. London, Ser. A 110, 91
(1926); 121, 284 (1928); H. Gregory and S. Marshall, ibid. 114, 354 (1927); 118, 594 (1928); B. G. Dickins, ibid. 143,
517 (1934); H. A. Daynes, Gas Analysis by Measurement of Thermal Conductivity. Cambridge University Press,
Cambridge, UK, 1933, gives a very comprehensive table of relative thermal conductivities (air = 1) for a large number
of gases, including hydrocarbons.

63. J. R. Partington and W. G. Shilling, The Specific Heat of Gases. Benn, London, UK 1924; see also G. N. Lewis and
M. Randall, Thermodynamics, p. 80. McGraw-Hill, New York, 1923; Bulletin 30, Cornell University, Engineering
Experiment Station, October, 1942, gives specific heats of a number of gases over a wide range of pressures and
temperatures.

64. ln = loge = 2.303 log10.

65. The exact meaning of this requirement will appear in the subsequent discussion.

66. See A. Farkas and H. W. Melville, Experimental Methods in Gas Reactions, p. 190. Cambridge University Press,
Cambridge, UK, 1939; also A. Farkas, Orthohydrogen, Parahydrogen, and Heavy Hydrogen. Cambridge University
Press, Cambridge, UK, 1935, for illustration of the application of this method.

67. G. A. Shakespear, Proc. R. Soc. London, Ser. A 97, 273 (1920); see also the comprehensive discussion of this
instrument by H. A. Daynes, Gas Analysis by Measurement of Thermal Conductivity. Cambridge University Press,
Cambridge, UK, 1933.

68. See especially a description of a modified construction by T. L. Ibbs, Proc. R. Soc. London, Ser. A 99, 385 (1921);
also 107, 470 (1925); W. E. Summerhays [Proc. R. Soc. London 42, 218 (1930)] has used the katharometer to measure
the coefficient of diffusion of water vapor.

69. B. G. Dickins, Proc. R. Soc. London, Ser. A 143, 517 (1934).

70. M. Knudsen, Ann. Phys. (Leipzig) [4] 31, 205 (1910); 32, 809 (1910); 33, 1435 (1910); 34, 593 (1911); [5] 6, 129
(1930); see also Loeb [1b, pp. 310325] and Kennard [1a, pp. 311320].
71. M. von Smoluchowski, a number of papers published before 1911 and Ann. Phys. (Leipzig) [4] 35, 983 (1911); also
discussed by Loeb [1b] and Kennard [1a].

72. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, pp. 311312. McGraw-Hill,
New York, 1938.

73. K. B. Blodgett and I. Langmuir, Phys. Rev. 40, 78 (1932).

74. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 316. McGraw-Hill, New
York, 1938.

< previous page page_76 next page >


< previous page page_77 next page >
Page 77

75. A table of values of γ is given by Loeb [1b, p. 445]; also see Partington and Shilling [63].

76. M. von Smoluchowski, Ann. Phys. (Leipzig) [4] 35, 983 (1911).

77. J. H. Jeans, An Introduction to the Kinetic Theory of Gases, p. 193. MacMillan, New York, 1940.

78. J. A. Morrison and Y. Tuzi, J. Vac. Sci. Technol. 2, 109 (1965).

79. The data for platinum are from Loeb [1b, p. 321]. The data for tungsten are from H. A. Jones and I. Langmuir [Gen.
Electr. Rev. 30, 354 (1927)], while those for "ordinary" platinum are from B. G. Dickins [proc. R. Soc. London, Ser. A
143, 517 (1934)]. That the value of the accommodation coefficient is greatly affected by the nature of the surface film
has been shown by K. B. Blodgett and I. Langmuir [Phys. Rev. 40, 78 (1932)]. In the case of a tungsten wire in
hydrogen at 0.20 Torr, the value of α changes from 0.534 for bare tungsten to 0.094 for tungsten with an adsorbed film
of oxygen. The general problem of the interaction between molecules and solid surfaces, which is obviously involved in
the interpretation of α, has received considerable attention from a number of investigators. For detailed discussion,
especially with regard to the role of α in adsorption phenomena, see Loeb [1b, p. 311] and also J. K. Roberts (Some
Problems in Adsorption. Cambridge University Press, Cambridge, UK, 1939). Values of α for air and a number of
different metals having etched, polished, and machined surfaces have been determined by M. L. Wiedemann [Trans.
ASME 68, 57 (1946)]. These values range from 0.87 to 0.97.

80. W. H. Keesom and G. Schmidt, Physica (Amsterdam) 3, 590, 1085 (1936).

81. D. E. Klett and R. K. Irey, Adv. Cryog. Eng. 14, 217 (1969).

82. K. Frankowski, Z. Alterman, and C. L. Pekeris, Phys. Fluids 8, 245 (1965).

83. D. R. Willis, in Rarefied Gas Dynamics (J. A. Lauermann, ed.), Vol. 1, pp. 209225. Academic Press, New York,
1963.

84. P. L. Bhatnager, E. P. Gross, and M. Krook, Phys. Rev. 94, 511 (1954).

85. D. R. Willis, Phys. Fluids 8, 1908 (1965).

86. J. C. Havekotte and G. S. Springer, AIAA J. 7, 782 (1969).

87. I. Langmuir, Phys. Rev. 34, 401 (1912).

88. H. A. Jones, Gen. Elec. Rev. 28, 650 (1925).

89. Of course the loss due to radiation was subtracted from the total observed energy loss.

90. I. Brody and F. Körösy, J. Appl. Phys. 10, 584 (1939).

91. J. P. Hobson, T. Edmonds, and R. Verreault, Can. J. Phys. 41, 983 (1963); T. Edmonds and J. P. Hobson, J. Vac.
Sci. Technol. 2, 182 (1965).

92. G. A. Miller and R. E. Buice, J. Phys. Chem. 70, 3874 (1966).

93. Tentative Recommended Practice for Ionization Gauge Application to Space Simulators, ASTM E296-66T,
Appendix A2. Am. Soc. Test. Mater., Philadelphia.

94. J. A. Poulis, B. Pelupessy, C. H. Massen, and J. M. Thomas, J. Sci. Instrum. 41, 295 (1964).

95. M. Knudsen, Ann. Phys. (Leipzig) [4] 31, 205, 633 (1910); 83, 797 (1927).
96. L. B. Loeb, Kinetic Theory of Gases, p. 358. McGraw-Hill, New York, 1934.

97. M. J. Bennett and F. C. Tompkins, Trans. Faraday Soc. 53, 185 (1957).

98. S. C. Liang, Can. J. Chem. 33, 279 (1955); J. Phys. Chem. 57, 910 (1953).

99. S. C. Liang, J. Appl. Phys. 22, 148 (1951). See also J. P. Hobson, J. Vac. Sci. Technol. 7, 351 (1970) for measured
effect of surface roughness.

100. J. M. Los and R. R. Ferguson, Trans. Faraday Soc. 48, 730 (1952).

101. J. O. Hirschfelder, R. B. Bird, and E. L. Spotz, J. Chem. Phys. 16, 968 (1948); Chem. Rev. 44, 205 (1949).

< previous page page_77 next page >


< previous page page_78 next page >
Page 78

102. G. D. Arney and A. B. Bailey, AIAA J. 1, 3863 (1963).

103. M. Knudsen, Kinetic Theory of Gases, p. 37. Methuen, London.

104. J. P. Hobson, Vacuum 15, 543 (1965).

105. T. Edmonds and J. P. Hobson, J. Vac. Sci. Technol. 2, 182 (1965).

106. S. Weber and G. Schmidt, Kamerlingh Onnes Laboratory, Commun. No. 246C. University of Leiden, 1936.

107. H. H. Podgursky and F. N. Davis, J. Phys. Chem. 65, 1343 (1961).

108. F. Sharipov, J. Vac. Sci. Technol. A 14, 2627 (1996).

109. L. A. Dietz, Rev. Sci. Instrum. 27, 817 (1956).

110. W. Jitschin and P. Röhl, J. Vac. Sci. Technol. A 5, 372 (1987).

111. T. Takaishi and Y. Sensui, Trans. Faraday Soc. 59, 2503 (1963).

112. J. C. Williams, III, J. Vac. Sci. Technol. 8, 446 (1971).

113. T. L. Ibbs, Physica (Amsterdam) 4, 1135 (1937).

114. S. Chapman and F. W. Dootson, Philos. Mag. [6] 33, 248 (1917); see also Jeans [1c, p. 251]. A mathematical
treatment of the subject is given by Chapman and Cowling [1d, pp. 140147, 351], as well as by W. H. Furry, R. Clark
Jones, and L. Onsager [Phys. Rev. 55, 1083 (1939)] and R. Clark Jones and W. H. Furry [Rev. Mod. Phys. 18, 151
(1946)]. A "simple theory" has been published by L. J. Gillespie [J. Chem. Phys. 7, 530 (1939)], and a much more
elaborate mathematical discussion by S. Chapman [Proc. R. Soc. London, Ser. A 177, 38 (19401941)].

115. K. Clusius and G. Dickel, Naturwissenschaften 26, 546 (1938); 27, 148 (1939).

116. T. L. Ibbs, Proc. R. Soc. London, Ser. A 107, 470 (1925).

117. See L. J. Gillespie, J. Chem. Phys. 7, 530 (1939), for a discussion of the derivation of this equation.

118. T. L. Ibbs and A. C. R. Wakeman, Proc. R. Soc. London, Ser. A 134, 613 (1932).

119. T. L. Ibbs and K. E. Grew, Proc. R. Soc. London, Ser. A 43, 142 (1931).

120. A. K. Brewer and A. Bramley, Phys. Rev. 55, 590 (1939).

121. A. O. Nier, Phys. Rev. 57, 30 (1940).

122. W. Furry, R. Clark Jones, and L. Onsager, Phys. Rev. 55, 1083 (1939).

123. T. I. Taylor and G. Glockler, J. Chem. Phys. 8, 843 (1940).

124. J. Bardeen, Phys. Rev. 57, 35 (1940).

125. F. T. Wall and C. E. Holley, J. Chem. Phys. 8, 949 (1940).

126. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, pp. 195196. McGraw-Hill,
New York, 1938.
127. J. H. Jeans, An Introduction to the Kinetic Theory of Gases, p. 216. MacMillan, New York, 1940.

128. J. H. Jeans, An Introduction to the Kinetic Theory of Gases, pp. 207210. MacMillan, New York, 1940.

129. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, p. 189. McGraw-Hill, New
York, 1938.

130. J. H. Jeans, An Introduction to the Kinetic Theory of Gases, pp. 217218. MacMillan, New York, 1940.

131. W. E. Summerhays, Proc. Phys. Soc., London, 42, 218 (1930).

132. The results are summarized by A. Lonius, Ann. Phys. (Leipzig) [4] 29, 664 (1909).

133. L. B. Loeb, Kinetic Theory of Gases. McGraw-Hill, New York, 272 (1934).

134. B. A. Ivakin and P. E. Suetin, Sov. Phys.Tech. Phys. (Engl. Transl.) 9, 866 (1964).

< previous page page_78 next page >


< previous page page_79 next page >
Page 79

135. J. Hirschfelder, C. Curtiss, and R. Bird, Molecular Theory of Gases and Liquids. Wiley, New York, 1954.

136. S. Ziering and M. Sheinblatt, Phys. Fluids 9, 1674 (1966).

137. R. S. Cunningham and C. J. Geankoplis, Ind. Eng. Chem., Fundam. 7 (August issue), 429 (1968).

138. G. R. Fonda, Phys. Rev. 21, 343 (1923).

139. I. Langmuir, H. A. Jones, and G. M. J. Mackay, Phys. Rev. 30, 211 (1927).

140. G. R. Fonda, Phys. Rev. 21, 343 (1923).

141. E. H. Kennard, Kinetic Theory of Gases with an Introduction to Statistical Mechanics, Chapter 7. McGraw-Hill,
New York, 1938.

142. E. W. McDaniel, Collision Phenomena in Ionized Gases. Wiley, New York, 1964.

143. V. S. Troitskii, Sov. Phys.JETP (Engl. Transl.) 14, 281 (1962).

144. N. M. Blachman and E. D. Courant, Phys. Rev. 74, 140 (1948).

145. N. M. Blachman and E. D. Courant, Rev. Sci. Instrum. 20, 596 (1949).

146. J. M. Greenberg and T. H. Berlin, Rev. Sci. Instrum. 22, 293 (1951).

147. M. J. Moravcsik and J. M. Sellen, Rev. Sci. Instrum. 26, 1158 (1955).

148. E. D. Courant, Rev. Sci. Instrum. 24, 836 (1953).

149. S. A. Kheifets, Instrum. Exp. Tech. (Engl. Transl.) 6 (June), 873 (1961).

150. A. N. Didenko and V. A. Serdyutskii, Sov. Phys.Tech. Phys. (Engl. Transl.) 7, 679 (1963).

151. Y. F. Orlov and S. A. Kheifets, Sov. Phys.Tech. Phys. (Engl. Transl.) 7, 671 (1963).

< previous page page_79 next page >


< previous page page_81 next page >
Page 81

2
Flow of Gases through Tubes and Orifices

R. Gordon Livesey

The nature of gas flow in pipes and ducts changes with the gas pressure and its description is generally divided into
three parts or regimes. The flow dynamics are characterized by λ, the molecular mean free path, in relation to some
characteristic dimension such as the diameter of a pipe. The flow regime cannot be determined from the mean free path
alone but only from the relation of this parameter to the characteristic dimension. The relation is known as the Knudsen
number, defined as*

Three regimes are generally identified:

1. Free Molecular Flow. The mean free path is of the same order as, or greater than, the characteristic dimension (the
range of relatively large Knudsen numbers), and gas dynamics are dominated by molecular collisions with the walls of
the retaining vessel or pipe.

2. Continuum Flow. The mean free path is small compared with the characteristic dimension (the range of small
Knudsen numbers), and intermolecular collisions are much more frequent than wall collisions. In this regime the
properties of the gas

* In the literature the Knudsen number may be variously defined, for a cylindrical tube, as λ/d, d/λ, λ/R, or R/
λ.
Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.
ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_81 next page >


< previous page page_82 next page >
Page 82
(temperature, density, flow velocity) do not vary significantly over several mean free paths and the gas can be considered a continuous medium.
The gas dynamics are therefore described and analysed hydrodynamically. Flow in this regime is often referred to as viscous flow, although there
are circumstances (such as flow through short ducts) in which. viscosity plays no part.

3. Transitional Flow. The transition between continuum and free molecular flow occurs at intermediate values of the Knudsen number where
both wall collisions and intermolecular collisions are influential in determining the flow characteristics.

Expressed in terms of pressure and characteristic dimension the Knudsen number is

For air at 20°C, with d in mm and P in mbar we obtain

Table 2.1 shows the generally accepted range of Knudsen numbers for the three regimes. There is no sharp transition between the regimes, and
somewhat different values may be quoted by different authors. The gas factor Fg in the table, used to correct for different gases, is the ratio of
the mean free path for air to that of the gas under consideration (at the same pressure and temperature) and can be calculated from

Values of the gas factor for a number of common gases are shown in Table 2.2.

Applications of vacuum technology range from the lowest pressures attainable (< 1014 mbar) through to around atmospheric pressure, so that all
of the regimes described are of interest to workers in this field. One of the main aims of this chapter is to enable calculation of flow under as wide
a range of conditions as possible, so that a large number of flow equations is presented. The widespread availability of scientific calculators,
personal computers, and mathematical software means that some of the more cumbersome formulae are considerably less daunting than in times
past. However, the "back of an envelope" is still a much favored tool of scientists and engineers and rough calculations are often sufficient, so
that approximations will be given wherever possible. Equations for the molecular, continuum and transitional

Table 2.1. Flow Regimes versus Knudsen Number and Pressure


Regime Kn (λ/d) P (mbar), d (mm)
Molecular
Kn > 0.5 PdFg < 0.133
Transitional
0.5 > Kn > 0.01 0.133 < PdFg < 6.6
Continuum
Kn < 0.01 PdFg > 6.6

< previous page page_82 next page >


< previous page page_83 next page >
Page 83

Table 2.2. Properties of Some Common Gases at 20°C


Gas Relative Molecular Mass Viscosity Viscosity Ratio (Air/Gas) Fg
Pa·s × 106
H2
2 8.8 2.07 0.543
He
4 19.6 0.929 0.345
H2O (vapor)
18 9.7 1.88 1.48
N2
28 17.6 1.03 1.02
Air
29 18.2 1 1
O2
32 20.4 0.892 0.937
Ar
40 22.3 0.818 0.959
CO2
44 14.7 1.24 1.53

regimes are discussed in Sections 2.2, 2.3, and 2.4, respectively. For the most part, derivations are not given since there are textbooks and
papers where the basic theories are discussed extensively; several references are listed in each section for the reader interested in studying the
subject in more detail.

All equations are written at least once in the text in SI units; where numerical coefficients are given, the units used are stated.

Except for the discussion of adiabatic compressible flow, it is generally assumed throughout this chapter that isothermal conditions apply.

2.1
Flow Conductance, Impedance, and Gas Throughput

In the field of vacuum science and technology it is common practice to express gas flow rate as throughput in pressurevolume units. The symbol
is normally used and the throughput of gas at a particular pressure is then

If the volumetric flow rate is due to a pump

where S is the speed (or volumetric rate) of the pump at the pressure P.

The pumping speed available at a chamber will be affected by restriction due to connecting pipework. One of the most common problems in
vacuum technology is to estimate the loss in speed due to such restrictions (system design is covered in Chapter 9).

Knudsen [1] first introduced the notion of a pipe as an impedance or resistance in the electrical sense and Dushman [2] introduced the concept
of conductance, which is defined by the relation

< previous page page_83 next page >


< previous page page_84 next page >
Page 84

where Pu is the upstream pressure and Pd is the downstream pressure. These pressures normally refer to values in
(perhaps notional) plenums at the entrance and exit of a duct or a system fitting such as a valve. Gas flow conductance
is thus analogous to electrical conductance, with pressure difference being the analogue of voltage difference and the
analogue of current. The reciprocal of conductance (resistance or impedance, Z = 1/C) could equally well be used;
however, conductance has come into common usage in vacuum technology mainly because of its intuitive relation to
volume flow rate and pumping speed.

Applying this concept to a set of pipes or components in series, the net conductance is found from

The net speed of a pump in series with a component or pipe is found in a similar way:

In practice it is seldom quite so easy as these equations imply. Some care is needed with combinations of components,
and this is discussed in Section 2.2.10 in relation to molecular flow. In continuum flow, conductance depends in a
complicated way on the flow conditions, and a better approach is to calculate the pressure ratio (Kp) across a component
or series of components.

It is usually assumed that continuity applies through a system; that is, the throughput is the same through all sections.
This will be the case as long as sufficient time has elapsed (from opening a valve or starting a pump for example) and
there are no temperature differences between the points of interest. In many common situations, steady conditions are a
reasonable assumption. (Some cases of unsteady molecular flow are covered in Section 2.2.11.) If Pd is the inlet
pressure to a pump of speed S which is connected via a pipeline or component to a chamber, then, assuming steady
conditions, the speed (Sn) and pressure (Pu) at the chamber are simply related by

In this way, the net pumping speed can be found if the pressure ratio can be calculated.

From the definition of conductance

Dividing through by the downstream pressure Pd and rearranging gives


< previous page page_84 next page >
< previous page page_85 next page >
Page 85

This calculation is often easier than taking the reciprocal of a sum of reciprocals, and Kp is the factor by which the
pumping speed is reduced.

2.2
Molecular Flow

One of the fundamental assumptions in the derivations of molecular flow conductance is that molecules scatter from a
surface according to a cosine distribution. This is also referred to as diffuse or random scattering and means that there
are no favored directions. A scattered molecule has the same probability of emerging in any direction, and this is
unrelated to its direction of incidence. There are special circumstances in which nondiffuse scattering may occur, but for
microscopically rough surfaces the diffuse scattering law is well established theoretically and experimentally [35].

Figure 2.1 shows the distribution of molecules emerging from an aperture and from tubes of various lengths. The
lengths of the vectors are proportional to the number of molecules emerging in that direction. It is noticeable that the
longer the tube, the more heavily weighted is the emerging flux to the tube axis. This is an indication of what occurs
inside a tube. An observer close to the entrance (looking upstream) will see the entrance plane as a diffuse source. Deep
inside the tube an observer will see a perturbed flux which is peaked toward the axis. This is often referred to as the
beaming effect of a tube.

The aperture in Fig. 2.1 acts as a plane cosine emitter, and the molecular flux shows a spherical distribution. At small
angles to the plane the flux of molecules is reduced,

Fig. 2.1
Angular distribution of molecules
exiting tubes of various length-to-diameter
ratios. Reproduced with permission from
L. Valyi, Atom and Ion Sources, p. 86.
Copyright 1977, Akadémial Kiadó, Budapest.
< previous page page_85 next page >
< previous page page_86 next page >
Page 86

compared with larger angles. This is not because there is a preferred direction but is a consequence of the angle of view:
The emitting plane does not appear ''dimmer," simply smaller.

In the molecular regime, solution of gas flow problems can be reduced to finding the conductance of the elements
involved since conductance is independent of pressure or flow conditions. The derivation of conductance, by theoretical
or analyticostatistical methods, assumes that molecules arrive at the entrance plane of a duct from a chaotic gas, so that
the entrance plane effectively behaves as a diffuse (cosine) emitter. When vacuum components are connected in series,
this may not be the case and some correction is needed; this will be covered in Section 2.2.10.

Clausing [6] first introduced the concept of transmission probability, denoted by α. If N2 molecules arrive at the
entrance plane of a duct then the number of these which reach the exit plane is N2α, and N2(1 α) return to the entrance.
Similarly, of N1 molecules striking the exit plane (from a downstream chamber), N1α reach the entrance. The net flux
of molecules from entrance to exit is then (N2 N1)α. Although proportional to the pressure difference across the duct,
the net flux is not driven by a pressure difference; it actually consists of two independent fluxes, and there is not a flow
in the usual sense of the word. Some of the molecules which enter the duct will return to the entry plane after one or
more wall collisions. The flow dynamics are thus very different to the continuum flow case in which all molecules
crossing the entrance plane will leave the exit (apart from the possibility of back diffusion which can occur in some
circumstances).

Expressions for conductance are usefully formulated in terms of transmission probability, so that the conductance of a
duct (or other component) is given by the entrance aperture conductance multiplied by the transmission probability

2.2.1
Conductance of an Aperture

The molecular flow conductance of a thin aperture is directly related to the rate of impingement of molecules over the
aperture area A (discussed in Chapter 1):

where R0 is the universal gas constant, T is the thermodynamic (or absolute) temperature, and Mm is the molar mass (e.
g., 0.028 kg/mole for nitrogen).

This gives the familiar result that the molecular flow conductance of an aperture for air at 20°C is 11.6 liters per second
per square centimeter.

For air at 20°C, Eq. (2.15) can be written in the convenient form

and Table 2.3 gives values of the constant ka for several combinations of commonly used units.

< previous page page_86 next page >


< previous page page_87 next page >
Page 87

Table 2.3. Aperture Conductance for Air at 20°C in Various Units


Ca = kaA
Ca ka A
m3·s1
115.6 m2
liter·s1
11.56 cm2
liter·s1
0.1156 mm2
m3·h1
0.4163 mm2
cfm
0.245 mm2
cfm
158.1 in.2
liter·s1
74.62 in.2

2.2.2
General Considerations for Long Ducts

Molecular flow in long ducts was first studied experimentally and theoretically by Knudsen [7]. He deduced a general
relationship for a long duct of length l, varying cross-sectional area A and perimeter B, which can be written as

where va is the mean thermal velocity of molecules.

However, as discussed by Steckelmacher [8, 9], Eq. (2.17) gives the correct result only in the case of a long cylindrical
tube and leads to erroneous results for all other cross sections.

A correct expression was derived by Smoluchowski [10] which may be written

where ρ is a chord making an angle θ with the normal to the perimeter s. Expressions for a number of different cross
sections have been derived from Eq. (2.18).

2.2.3
General Considerations for Short Ducts
For small values of l it is clear that the long duct relation will give values for conductance which are too high. As the
length tends to zero the conductance apparently tends to infinity. Reasoning from the point of view that the entrance of a
duct can be considered as a vacuum circuit element with resistance Za = 1/Ca in series with the duct proper (regarded as
a "long" duct), of resistance Zml = 1/Cml, the net conductance is [applying Eq. (2.8)]

< previous page page_87 next page >


< previous page page_88 next page >
Page 88

Using this principle, the net transmission probability of a short duct then becomes

where al is the long duct transmission probability.

This principle was originally applied by Dushman [2] to short circular cross-sectional tubes as an approximate method
of correcting for the end effect. A similar logic is often applied to short ducts of other cross sections. The maximum
error for a cylindrical tube is about 12% (too high). Errors of this order are expected for blocky cross sections, but for
other cross sections the errors may be more serious. In the case of narrow rectangular ducts the errors can be greater
than 50%. The most accurate results are obtained using transmission probabilities which have been derived for a
number of different shapes either theoretically or via Monte-Carlo methods.

Transmission probability data for cylindrical tubes, from the results of Cole [11], are shown in Table 2.5. It is apparent
that the transmission probability for a unit length increases as the length of the duct increases. Consider a unit length l/d
= 1 for which α = 0.514. For two unit lengths (l/d = 2) the transmission probability, expected from the shorter length,
would be 0.257, whereas the actual value is 0.357. At 10 unit lengths the transmission probability is almost twice the
expected value. This reflects the effect of the random molecular distribution near the entrance compared with the
beamed distribution which evolves further down the tube.

2.2.4
Tube of Uniform Circular Cross Section

Long Tubes. The familiar expression for the conductance of a long cylindrical tube of diameter d was first presented by
Knudsen [7] in 1909:

It can be seen from Eq. (2.21) that the transmission probability for a long cylindrical tube is

Berman [12] derived the solution as an asymptotic expansion, the first four terms of which are

For this reduces to Eq. (2.22) (which needs significant correction for l < 50 d).

Table 2.4 lists values of long tube and aperture conductance for a number of gases.

< previous page page_88 next page >


< previous page page_89 next page >
Page 89

Table 2.4. Conductance of Long Cylindrical Tubes and Apertures for Air at 20°C
Gas Relative Molecular Cml (l/d3) liter · s1 Ca/A liter · s1 Ca/d2 liter · s1
Mass (mm) (mm) (mm)
H2
2 0.461 0.440 0.346
He
4 0.326 0.311 0.245
Air
29 0.121 0.116 0.0908
Ar
40 0.103 0.0985 0.0773

Short Tubes. Transmission probabilities for short tubes derived by Cole [11] are shown in Table 2.5. Berman [12] presented equations for the
direct calculation of α for any length.

Table 2.5. Transmission Probabilities for Cylindrical Tubes


α (Cole [11]) α [Eq. (2.20)] % Error
l/d
0.952399 0.963855
0.05 1.20
0.869928 0.898876
0.15 3.33
0.801271 0.842105
0.25 5.10
0.743410 0.792079
0.35 6.55
0.694044 0.747664
0.45 7.73
0.671984 0.727273
0.5 8.23
0.632228 0.689655
0.6 9.08
0.597364 0.655738
0.7 9.77
0.566507 0.625000
0.8 10.33
0.538975 0.597015
0.9 10.77
0.514231 0.571429
1 11.12
0.420055 0.470588
1.5 12.03
0.356572 0.400000
2 12.18
0.310525 0.347826
2.5 12.01
0.275438 0.307692
3 11.71
0.247735 0.275862
3.5 11.35
0.225263 0.250000
4 10.98
0.206641 0.228571
4.5 10.61
0.190941 0.210526
5 10.26
0.109304 0.117647
10 7.63
0.076912 0.081633
15 6.14
0.059422 0.062500
20 5.18
0.048448 0.050633
25 4.51
0.040913 0.042553
30 4.01
0.035415 0.036697
35 3.62
0.031225 0.032258
40 3.31
0.027925 0.028777
45 3.05
0.025258 0.025974
50 2.83
0.002646 0.002660
500 0.51

Defining

< previous page page_89 next page >


< previous page page_90 next page >
Page 90

we obtain

This gives results which agree with the Cole [11] data to within 0.13%.

Santeler [13] devised a simpler and more convenient formulation and calculates the transmission probability as

where le is an "equivalent length" and

This gives transmission probabilities with a maximum error of less than 0.7% relative to the Cole data.

2.2.5
Duct of Uniform Rectangular Cross Section

The convention used to denote dimensions of rectangular (and elliptical) ducts is as follows: a and b are the cross-
sectional dimensions, with b ≥ a, and l is the length in the direction of gas flow. Thus the cross-sectional area is A = ab.
To avoid any confusion, equations quoted from various authors have been recast to conform with this convention.

Long Ducts. The transmission probability, due to Smoluchowski [10] is

where δ = a/b and

A useful approximation is

This is accurate to better than 1% for aspect ratios (b/a) up to almost 100. The error increases slowly with aspect ratio
but is still only 1.9% and 2.4% for aspect ratios of 1000 and 10,000, respectively.

< previous page page_90 next page >


< previous page page_91 next page >
Page 91

Short Ducts. There appear to be no general expressions which cover the whole range of lengths and aspect ratios. Data are available from the
Monte Carlo calculations of Levenson et al. [14]. The results of Santeler and Boeckmann [15], which cover a greater range of lengths and
aspect ratios, are shown in Table 2.6 (for short ducts the original data are listed to six significant figures). Also shown for comparison in
Table 2.7 are some of the results of Cole [16], derived using a complementary variational method which gave upper and lower bounds for α;
the values listed are the means and are accurate to 1.2% or better.

Table 2.6. Transmission Probabilities for Rectangular Ductsa


l/a b/a
1 1.5 2 3 4 6 8 12 16 24
0.9902 0.9918 0.9926 0.9934 0.9938 0.9942 0.9944 0.9946 0.9947 0.9948
0.01
0.9807 0.9839 0.9854 0.9870 0.9878 0.9885 0.9889 0.9893 0.9895 0.9897
0.02
0.9626 0.9685 0.9715 0.9744 0.9759 0.9774 0.9782 0.9789 0.9793 0.9797
0.04
0.9370 0.9467 0.9515 0.9564 0.9589 0.9613 0.9625 0.9638 0.9635 0.9650
0.07
0.9131 0.9260 0.9326 0.9392 0.9425 0.9458 0.9475 0.9491 0.9500 0.9508
0.1
0.8428 0.8645 0.8757 0.8869 0.8926 0.8982 0.9011 0.9039 0.9053 0.9067
0.2
0.7334 0.7659 0.7829 0.8004 0.8093 0.8182 0.8227 0.8272 0.8295 0.8317
0.4
0.6178 0.6575 0.6793 0.7022 0.7140 0.7260 0.7321 0.7381 0.7411 0.7442
0.7
0.5363 0.5786 0.6026 0.6285 0.6421 0.6560 0.6631 0.6702 0.6737 0.6773
1
0.3780 0.4192 0.4444 0.4733 0.4893 0.5063 0.5150 0.5240 0.5285 0.5330
2
0.2424 0.2759 0.2977 0.3245 0.3404 0.3583 0.3679 0.3781 0.3833 0.3885
4
0.1596 0.1848 0.2020 0.2242 0.2380 0.2545 0.2639 0.2742 0.2796 0.2852
7
0.1195 0.1397 0.1537 0.1723 0.1843 0.1991 0.2078 0.2177 0.2230 0.2287
10
0.0655 0.0776 0.0864 0.0984 0.1066 0.1171 0.1238 0.1319 0.1366 0.1419
20
0.0346 0.041 0.0464 0.053 0.058 0.0652 0.0695 0.075 0.078 0.083
40
0.020 0.024 0.0275 0.032 0.035 0.039 0.042 0.046 0.048 0.052
70
0.014 0.017 0.019 0.023 0.025 0.028 0.030 0.033 0.035 0.038
100
a From Santeler and Boeckmann [15].
Table 2.7. Transmission Probabilities for Rectangular Ductsa
l/a b/a
1 5 10 20 50 100 1000 10000
0.53619 0.66722 0.68266
1
0.24233
4
0.11930 0.21280 0.2372 0.2400
10
0.11207
20
0.07234
40
0.04464
80
0.01438 0.0320 0.0438 0.0464 0.0468
100
0.0224
200
0.01295
400
a From Cole [16].

< previous page page_91 next page >


< previous page page_92 next page >
Page 92

For short slits (or large flat plates): Values of transmission probability were first calculated by Clausing [6]. Berman
[12] devised equations for the direct calculation of transmission probability to a greater accuracy than the tabulated
Clausing values.

For and and putting x = l/a we obtain

If (long, closely spaced slot, again with ), this equation simplifies to

2.2.6
Uniform Elliptical Cross Section

a and b are the minor and major axes,

Long Ducts. The expression derived from Eq. (2.18) by Steckelmacher [8] is

is the complete elliptic integral, and

Steckelmacher [8] has shown that, for the same aspect ratio and cross-sectional area, the expressions for a rectangular
duct (Eq. (2.26)] and an elliptical duct are in close agreement. The rectangular duct approximation was derived on this
basis, so that a similar approximation is available for the elliptical duct:
The constant is chosen to give the correct result for b = a (circular) and minimises the errors for practical aspect ratios
up to 10 (although the errors are less than 1% even for unrealistically large aspect ratios up to 1000).

< previous page page_92 next page >


< previous page page_93 next page >
Page 93

Generally for ducts which have cross sections intermediate between rectangular and elliptical we have

where αr is the rectangular duct transmission probability and

Short Ducts. No data are available for short ducts of elliptical or similar cross section. However, it is expected that the
similarity between long elliptical and rectangular ducts (of the same cross-sectional area and aspect ratio) will also
apply to short ducts. It is suggested that approximate transmission probabilities can be found from

For very short ducts this reduces to αs = αr since the transmission probability of an aperture is independent of shape.

2.2.7
Cylindrical Annulus (Flow between Concentric Cylinders)

Long Ducts.

where K(e2) and E(e2) are the complete elliptic integrals of the first and second kinds. X(e) is listed for a range of
values of e in Table 2.8.

Short Ducts. Table 2.9 presents the results of Berman [17], who calculated transmission probabilities over a large range
using the variational method. Berman also obtained an empirical expression which is more convenient for computer
calculation.

< previous page page_93 next page >


< previous page page_94 next page >
Page 94

Table 2.8. The Function X(e) for a Long Cylindrical Annulus [Eq. (2.31)]
e = d1/d2:
0 0.1 0.2 0.3 0.4 0.5
X(e):
1.3333 1.231 1.1238 1.0116 0.8942 0.7711
e = d1/d2:
0.6 0.7 0.8 0.9 0.95 1.0
X(e):
0.6416 0.5044 0.3576 0.1966 0.1071 0

Defining x = l/(d2 d1), we obtain

The expression is valid in the range 0 ≤ x ≤ 50 and 0 ≤ e ≤ 0.9.

2.2.8
Uniform Triangular Section (Equilateral)

Long Ducts.

Short Ducts. Approximate transmission probabilities can be obtained using the entrance correction principle [Eq. (2.20)].

2.2.9
Other Shapes
Transmission probabilities for a number of geometries are shown in graphical form in Figs. 2.2 [18], 2.3, and 2.4. It is worth
noting that the transmission probability for an

< previous page page_94 next page >


< previous page page_95 next page >
Document
Page 95

Table 2.9. Transmission Probabilities (× 104) for Cylindrical Annulusa


y = l/(R2 R1) R1/R2
0.25 0.5 0.75 0.9
0.1 0.2 0.4 0.6 0.8 0.95
0.5 8017 8022 8030 8037 8043 8046
1.0 6737 6754 6783 6808 6829 6842
1.5 5842 5867 5915 5958 5997 6020
5295 5365
2.0 5175 5206 5266 5324 5378 5413
2.5 4655 4690 4758 4826 4894 4926 4940
3.0 4237 4274 4348 4423 4501 4558
3.5 3893 3931 4007 4087 4174 4241
3661 3761 3872
4.0 3604 3642 3720 3804 3896 3972
5.0 3123 3181 3260 3347 3448 3507 3538
2948 3071
6.0 2791 2828 2906 2994 3100 3201
7.0 2513 2548 2625 2712 2820 2929
2339 2436 2559
8.0 2286 2321 2395 2481 2589 2704
9.0 2099 2132 2204 2288 2496 2515
2081 2200
10.0 1914 1973 2042 2124 2230 2304 2352
1819 1933
12.0
1617
14.0
15.0 1414 1440 1499 1569 1666 1740 1792
1381 1456 1559
16.0
1325
18.0
1216 1310
20.0 1404
25.0 921.7 941.1 984.5 1038 1116 1180 1230
30.0 1019
35 897
40 802
50 496.0 507.6 533.9 567.4 618.2 700.4
100 258.9 265.4 280.1 299.2 328.9 380.5
200 132.7 136.1 144.0 154.3 170.6 200.1
500 53.97 55.4 58.69 63.04 70 82.99
1000 27.15 27.88 29.56 31.77 35.34 42.09
104 2.733 2.807 2.978 3.204 3.57 4.273
105 0.2735 0.2809 0.298 0.3207 0.3575 0.4282
a From Berman [17].

< previous page page_95 next page >


< previous page page_96 next page >
Page 96

Fig. 2.2
Molecular transmission probabilities of an elbow,
from the results of Davies [18].

elbow is almost the same as two short tubes (with length measured at the inside of the elbow) connected by a large
volume. The effect of the elbow is to randomize, at least partly, the molecular distribution.

2.2.10
Combinations of Components

If two components, with transmission probabilities α1 and α2, are connected in series, then the usual method of
determining the net transmission probability is
< previous page page_96 next page >
< previous page page_97 next page >
Page 97

Fig. 2.3
Molecular transmission probabilities of a cylindrical
tube with restricted openings, from the results of Davies [18].

Consider two identical, short tubes, of transmission probability α, connected via a large volume V as shown in Fig. 2.5a.
The inlet of tube 2 and the outlet of tube 1 are also connected to large volumes and, for convenience, the downstream
pressure is taken to be zero. It is supposed that there is no beaming between the tubes, and the effect of the large volume
is to randomize the molecular distribution between the tubes.
The net flux of molecules through tube 2 is (N2 N1)α, and the net flux through tube 1 is N1α. Under steady conditions,
these must be the same, so that N1 = N2/2 and the number of molecules transmitted is N2α/2. The overall transmission
probability of the system is then α/2. The conclusion is the same if the transmission probabilities are combined as in Eq.
(2.38).

< previous page page_97 next page >


< previous page page_98 next page >
Page 98

Fig. 2.4
Molecular transmission probabilities of a chevron baffle.
Reproduced with permission from Levenson et al. [14]
Copyright 1963, Société Française d'Engenieurs
et Techniciens du Vide.

Now consider Fig. 2.5b, in which the two tubes (each with l/d = 1) have been brought together. The transmission
probability of each tube separately is (from Table 2.5) 0.514. However, the transmission probability for the joined tubes
(l/d = 2) is 0.357 and not 0.514/2 = 0.257.

Clearly the method of combining transmission probabilities for the joined tubes is incorrect. In this case every molecule
which crosses plane AA also crosses plane BB and vice versa, but this is not so when the tubes are separated by a large
volume.

Oatley [19] discussed the correct method of combining transmission probabilities, α1 and α2, for two joined tubes of
the same cross section and showed that the net transmission probability is given by
In the example above, this gives the overall transmission probability as 0.346, which is much closer to the correct result.
The Oatley method gives results with a maximum error of 5% or 6% for l/d ~ 2. It is, perhaps, surprising that the
method gives such good results, since the derivation assumes random gas entry into the second tube and thus ignores the
beaming effect. However, in short ducts the molecular distribution is

< previous page page_98 next page >


< previous page page_99 next page >
Page 99

Fig. 2.5
Combination of two short tubes.

not too seriously perturbed from a chaotic distribution, and in long ducts the entrance effect is relatively small.

The transmission probabilities for each of the two tubes, in effect, includes an entrance correction. Thus, when Eq.
(2.38) is applied, the overall transmission probability includes two entrance effects. This is correct when the two tubes
are separated by a large volume but not when they are joined and the Oatley method is equivalent to removing one of
the two corrections.

A typical case is illustrated in Fig. 2.6, in which a pump is connected to a chamber via a tube of the same size as the
pump inlet. The pump speed (S) has been measured, so that any entrance effects are already accounted for (at least in
principle). A pump can be regarded as a conductance with a transmission probability aH equal to its Ho coefficient (the
ratio of the pump speed to the conductance of the pump inlet aperture). If Ca is the pump and tube aperture conductance
and α is the tube transmission probability, then S = aHCa and the tube conductance C equals aCa. The net speed at the
chamber is then Sn = anCa, where

Alternatively,

The effect of this procedure is to remove an entrance correction.


< previous page page_99 next page >
< previous page page_100 next page >
Page 100

Fig. 2.6
Pump connected via a tube
of the same diameter.

If the pump speed is 300 liter · s1 and the connecting tube is 200 mm long and 100 mm in diameter, then Ca = 908 liter
· s1 (for air), α (for l/d = 2) = 0.357, and

which gives Sn = 188 liter · s1 [instead of 155 liter · s1 using Eq. (2.34)].

An addition theorem developed by Haefer [20] enables the calculation of multiple components of differing diameters.
The overall transmission probability of n elements α1n is related to the individual transmission probabilities ai and inlet
areas Ai by

The overall transmission probability α1n is expressed in terms of the inlet tube aperture.

Some cases of the application of this theorem will be discussed with reference to Fig. 2.7.

(a)
Series Arrangement of Tubes of Different Diameters (Fig. 2.7a). The overall transmission probability is
< previous page page_100 next page >
< previous page page_101 next page >
Page 101

Fig. 2.7
Combinations of components.

The diameter increases from tube 1 to tube 2 so that δ2, 1 = 0, but δ3, 2 = 1 since the diameter decreases from tubes 2 to
3. If l/d = 2 for each tube and the diameters are 15 mm, 25 mm, and 20 mm for tubes, 1, 2, and 3 respectively, then α1n
evaluates to 0.214. The aperture conductance of tube 1 (15-mm diameter) is 20.4 liter·1, so the conductance of the
arrangement is 0.214Ca1 = 4.38 liter·1. If the order of the tubes is reversed (20 mm, 25 mm, 15 mm), α1n now
evaluates to 0.121. The inlet aperture conductance (for 20-mm diameter) is now 36.3 liter·s1, so the overall conductance
is 0.121 × 36.3 = 4.38 liter·s1 as expected. The conductance of an arrangement cannot be changed by reversing the
order of components; but note that if the tubes were rearranged in ascending order of size, then α1n = 0.224. It is
generally the case that the highest conductance for a series of components is achieved when they are physically
arranged in order of size. In a calculation, the temptation to reorder the components for mathematical convenience
should be resisted since this can lead to incorrect results.

Equation (2.42) relating transmission probabilities can be expressed in terms of conductances:

which is more convenient if conductance values are given for components. δ has the same meaning as in Eq. (2.42).

The first summation in Eq. (2.44) contains the ''tube only" conductance Cmt discussed by Holland et al. [21]:

that is, a tube with its entrance correction subtracted. If the conductance quoted for a component is the "tube only" value
then the entrance correction should not be subtracted. A typical example would be a quarter swing valve which is not
designed for direct connection to a large chamber but is normally connected via a manifold. Similar remarks apply to
Eq. (2.42) if the tube only transmission probability is given.

< previous page page_101 next page >


< previous page page_102 next page >
Page 102

Either of Eqs. (2.42) or (2.44) can be used to find the net speed of a pump in series with a set of components.

(b)
Pump Connected to a Chamber via Two Tubes or Components (Fig. 2.7b).

Since Cm3 = αHCa3 = S, this reduces to

which can be written

Written in terms of "tube only" conductances, this becomes

The second term on the right-hand side of this equation can be seen as the total correction required to account for the
differing sections (remembering that a correction for A3 is inherent in the pump speed).

If, in Fig. 2.7b, tubes 1 and 2 and the pump all have the same aperture size of conductance Ca, then Eq. (2.48) becomes

Each tube has its entrance correction subtracted and only one correction, inherent in the pump speed, is applied.

(c)
Pump Connected to a Chamber Via a Second Chamber (Fig. 2.7c). This case is equivalent to making Ca2 and Cm2 very
large in Eq. (2.47), giving

In this case the entrance correction for tube 1 is retained.


2.2.11
Cases of Unsteady Flow

Consider a system in which a pump is connected to a vessel via a valve and a pipe of some significant length. It is clear
that, at the instant that the valve is opened, the throughputs at the pump and vessel must be

< previous page page_102 next page >


< previous page page_103 next page >
Page 103

different. Some time is required for continuity to be established. Usually this is very short compared with the timescale
for exhaust of the vessel, especially at continuum flow pressures where the pipe conductance is relatively high.
However, if the "vessel" is a long pipe, then steady conditions are never achieved.

Unsteady flow is of most interest under molecular flow conditions and is particularly relevant to filling, exhaust, or leak
testing of long pipelines. Mathematically, unsteady molecular flow in a constant section duct is analogous to heat
conduction in an infinite slab. Both can be treated as one-dimensional since there is no transverse flow. The cases
described here are intended only as a small sample to illustrate the kind of unsteady flow problems that can be solved.
Solutions for many heat conduction cases, which have practical parallels in molecular flow, are available in the
literature; Carslaw and Jaeger [22] is a good source of reference.

In the equations which follow, V is the volume and C the conductance of the pipe, assumed to be long. Given a relation
for pressure distribution, the throughput (from or into the pipe) can be obtained from the pressure gradient at the end of
a pipe:

Case 1. Pipe with uniform initial pressure P0, closed at x = 0. x = l opened to an environment maintained at constant
pressure Pe at t = 0.

This covers both exhaust and filling of a pipe. Lawson [23] discussed the application of this and similar cases to
pumping of trapped volumes and leakage.

The time constant for the first (slowest) term is

Contrast this with the time constant of chamber of volume V pumped at speed S or through a restrictive conductance C
(covered in Chapter 9).

For air at 20°C (l and d in meters) we have

and for helium we have


< previous page page_103 next page >
< previous page page_104 next page >
Page 104

A 1-meter-long pipe, 10 mm in diameter would have a time constant for helium of ~100 msec. For a pipe length of 10
meters, this increases to ~10 sec. Leak testing can be difficult with a detector at the downstream end of such long lines
since any leak will be located at the position of the helium probe several seconds prior to the indication on the leak
detector gauge.

The series converges quite rapidly and taking only the first term is a good approximation for t > 0.3τ0. Thus, taking
only the first term we obtain, at x = 0

and the throughput

Case 2. Pipe, with constant initial pressure P0, closed at x = 0 and pumped at the other end by a pump of constant speed
S, for t > 0.

where r = S/C and φn are the roots of

A good approximation for the first root is

accurate to within 3% in the worst case.

Taking only the first term of the series for throughput, we obtain

For the time constant becomes V/S and the case reduces to the simple pumpout of a large chamber.
Case 3. Pipe with zero initial pressure. Constant leak into the pipe at x = 0 and constant pumping speed S at x = l
for t > 0.

< previous page page_104 next page >


< previous page page_105 next page >
Page 105

This is again relevant to leak test of a long pipeline since the initial partial pressure of He will be zero.

where r = S/C and φn has the same meaning as in Case 2.

Taking only the first term, the throughput from the pipe is

2.3
Continuum Flow

In the continuum regime, calculation of gas flow through ducts is complicated by the different types of flow which can
occur. Flow may be broadly distinguished into two major types, referred to as viscous laminar flow and turbulent flow.
Since flow through a duct is driven by a pressure difference, all gas flow is compressible. There are circumstances in
which gas can be treated as incompressible, and this leads to considerable simplification of the equations describing
flow. However, there are also many circumstances in which compressibility cannot be ignored, so compressible flow
will also be discussed.

Continuum (or viscous) flow is often thought of as occurring at relatively high pressures. But consider air flowing
through a 100 mm (4 inch) diameter pipe. From Eq. (2.2) the Knudsen number Kn is < 0.01 (and hence the flow
continuum) down to pressures of about 0.1 mbar. Thus many vacuum processes will operate at pressures where
continuum flow conditions prevail.

At relatively low velocities, gas flows smoothly in stream lines, generally parallel to the duct walls, and the flow is said
to be laminar. In long ducts, viscosity of the gas is a controlling factor in the flow rate; this is not the case in short ducts,
although the flow may still be laminar. As the flow velocity is increased, there comes a critical point at which the flow
breaks up into turbulent eddies. These two types of flow, viscous laminar and turbulent, are described by different
equations. It is important to distinguish these flow types in calculation of flow rates; failure to do so can lead to wildly
inaccurate results. In short ducts, or in longer ducts at high flow velocities, compressibility becomes important and use
of incompressible flow formulae can also lead to serious errors.

The primary controlling parameter in the viscous behavior of Newtonian* fluids is the dimensionless Reynolds number

* Newtonian fluids are those in which the shear stress is proportional to the transverse velocity gradient. Most
common fluids (water, oils, and gases) are newtonian.

< previous page page_105 next page >


< previous page page_106 next page >
Page 106

where ρ is the density of the fluid, η is its viscosity, and u is the flow velocity. Dh is the hydraulic diameter* of the duct
(applicable to any cross section) defined by

where A is the cross-sectional area and B is the perimeter.

For a circular cross-sectional tube we obtain

so that the hydraulic diameter is simply the actual tube diameter.

For a concentric annulus we have

Note that, in calculating B, the perimeter of all "wetted" surfaces must be included.

It is useful to write Re in terms of the throughput:

In the case of a circular cross-sectional tube, B = πd. For air at 20°C, Eq. (2.67) can conveniently be written as

and Table 2.10 lists values for the units conversion kR for a variety of commonly used units.

As shown by the work of Senecal and Rothfus [24], the transition from viscous laminar to turbulent flow begins at
Recrit ~2000, although there is no sharp boundary, and flow is normally fully turbulent at a Reynolds number of 3500.
In very smooth pipes with well-rounded entrances, the transition to turbulence may be delayed to higher values of Recrit.

Using the value Recrit = 2000, the value of throughput for the onset of turbulent flow is

* Re defined in terms of diameter is the most commonly used convention. Some texts define Re in terms of
radius and make use of hydraulic radius rather than diameter. When making comparisons, note that hydraulic
diameter = 4 × hydraulic radius (yes, 4!).
< previous page page_106 next page >
< previous page page_107 next page >
Page 107

So for air at 20°C the flow will become turbulent if

Values of kT for a variety of units are given in Table 2.11.

For a cylindrical tube for example, with QT in mbar·liter·s1 and tube diameter in mm, we obtain

For example, in a 25-mm-diameter tube the conditions will be turbulent if the throughput exceeds 600 mbar · liter · s1.
If this tube size were used with a 20 liter · s1 pump, then turbulent flow would persist down to about 30 mbar.

In the continuum flow regime, conductance may be a function of pressure, pressure ratio, or rate of flow depending on
the type of flow and particular circumstances. The term conductance is useful as a conceptual toolto say, for example,
that the conductance of some tubulation is too small does provide a succinct means of characterizing a flow
conditionbut has limited practical value. Calculation of conductance is generally a means to an endcalculation of
throughput (when the end pressures are known) or calculation of pressure difference (when throughput is known), for
example. Thus, relations will be presented for throughput and pressure ratio where these can be expressed explicitly.

Table 2.10. Units Conversions for Reynolds Number and Throughput [Eq. (2.68)] for Air at 20°C

B kR
Pa · m3 · s1
m 2.615
mbar · m3 · h1
mm 72.64
mbar · liter · s1
mm 261.5
torr · liter · s1
mm 348.6
torr · liter · s1
in 13.73
torr · cfm
in 6.478
Table 2.11. Units Conversions for Turbulent Throughput [Eq. (2.70)] for Air at 20°C

B kT
Pa · m3 · s1
m 764.8
mbar · m3 · h1
mm 27.53
mbar · liter · s1
mm 7.648
torr · liter · s1
mm 5.736
torr · liter · s1
in 145.7
torr · cfm
in 308.7

< previous page page_107 next page >


< previous page page_108 next page >
Page 108

Before proceeding to discuss the equations of continuum flow, it is convenient to define

This collection of terms occurs frequently in equations of continuum flow. As is evident, the term is closely related to
Ca, the molecular flow conductance of an aperture [Eq. (2.15)], differing only by a numerical factor. The term therefore
has units of volumetric flow rate and provides a succinct means of expressing many continuum flow relations.

The long tube equations which will be presented for laminar and turbulent flow are of limited practical application.
There are few circumstances in which straight tubulation is employed of sufficient length to meet the long tube criteria
(to be discussed later). In vacuum technology, "short and wide" is beautiful! The long tube equations will be discussed
first since they form the basis for expressions which apply to ducts of any length. This will be followed by discussion of
the equations for compressible flow and then of approximations which are sufficiently accurate for many practical
purposes.

2.3.1
Viscous Laminar Flow

The simplest solutions for viscous flow, in ducts of constant cross section, are based on four assumptions: (1) The fluid
is incompressible; (2) the flow is fully developedthat is, the flow velocity profile is constant throughout the length; (3)
the flow is laminarthat is, in one direction only parallel to the duct axis and there are no turbulent motions; (4) the flow
velocity at the walls is zero.

The assumptions may appear restrictive but are true for newtonian fluids flowing in long ducts at relatively low
velocities. Except for the assumption of incompressibility, this also includes gases, but even for gases it can be shown
that compressibility can be ignored if

where Ma is the Mach number of the flow, defined as the ratio of the flow velocity to the local velocity of sound. The
generally accepted criterion is Ma < 0.3.

When fluid flows into the entrance of a duct, the flow velocity is approximately uniform over the entrance area. As the
fluid moves down the duct, shear stress due to viscous friction retards the flow near the walls and the fluid develops a
velocity profile which, after some entry length, becomes constant and the flow is said to be fully developed. In the case
of a circular cross section, for example, the profile is a parabola of revolution with the greatest fluid velocity at the tube
axis. The pressure drop over the entrance length is greater than that for fully developed flow because of the increased
shear stress and the kinetic energy needed to accelerate the flow. Shah and London [25] have given the following
correlation for entry length:

< previous page page_108 next page >


< previous page page_109 next page >
Page 109

The first term is significant only at very low Reynolds numbers, and the condition is usually expressed as

At the transition to turbulent flow, with Re = 2000, this gives an entry length of 112 diameters for a circular pipe. Since
the viscous laminar flow equations are valid only for fully developed flow, some correction is required unless the duct
length is much greater than the entry length. Corrections for entrance effects and long tube criteria are discussed in
Sections 2.3.4 and 2.3.7.

Viscous laminar flow has proved irresistibly attractive to theorists since the fundamental equations of flow can be
solved analytically; a wealth of solutions is available for an amazing variety of cross-sectional shapes. A few of the
more practical shapes will be discussed.

For incompressible fluids, equations are usually presented for volumetric flow rate which is proportional to the pressure
difference across the pipe. For gases, density varies along the duct, so equations are normally given for throughput in
terms of the average pressure which reflects the average density.

If Cv is written for the viscous flow conductance, then

where the conductance is proportional to the average pressure

and kV is a constant for the particular duct, containing the numerical factors, viscosity and geometric terms.

The throughput is then

Given a relationship for kV, it is no particular problem to determine the throughput if both end pressures are known. If
the throughput is given (because, for example, the downstream pumping speed at a given pressure is known), it is not so
easy to calculate the pressure drop because both pressures are required by Eq. (2.76). However, the equation can be
expressed as, for example,

where KP = Pu/Pd is the pressure ratio across the duct.

The term kVPd has the same units as conductance (i.e., volumetric flow rate); it can be thought of as the conductance of
the duct calculated at the downstream pressure

< previous page page_109 next page >


< previous page page_110 next page >
Page 110

(instead of the average pressure). In order to express flow equations in a form convenient for calculation, it is useful to
define two artificial conductances:

Cvu = viscous flow conductance calculated at the upstream pressure,


Cvd = viscous flow conductance calculated at the downstream pressure.

This artifice enables the equations of flow to be expressed in terms of quantities which are known.

When both end pressures are known, the throughput can be written as

If the downstream pressure and pumping speed (i.e., the volumetric flow rate) are known, then the pressure ratio across
the duct can be found from

If the upstream pressure and downstream pumping speed are known, we obtain

Cv, Cvu, and Cvd can be calculated, for different geometries, from the equations that follow, and Eqs. (2.78) to (2.81)
can be used to calculate throughput or pressure ratio.

Circular Cross-Sectional Tube. The HagenPoiseuille equation for a long circular pipe is probably the best-known
equation in viscous flow. The throughput is

hence the conductance is

For any gas, this can be written as


< previous page page_110 next page >
< previous page page_111 next page >
Page 111

Values of the units conversion constant kv are given in Table 2.12 for a variety of commonly used combinations of units.

It should be apparent from the previous discussion that, for a circular pipe for example, the artificial conductance Cvd is

Rectangular Cross Section.

Although only the first three terms are needed for good accuracy, this expression is rather cumbersome. A good
approximation is [21]

This expression, valid for all aspect ratios, is accurate for a = b and and it shows a maximum error of <3% at
intermediate aspect ratios.

Concentric Cylindrical Annulus.

Eccentric Cylindrical Annulus. It is worth noting that flow rate (for the same pressure difference) increases significantly if
the inner cylinder is offset from the axis. In the extreme case of a narrow annulus with maximum offset,
the flow rate is increased by a factor of 2.5. This explains the observation that flow rate can be very difficult to adjust with
an almost closed needle valve.

Table 2.12. Units Conversions for Viscous Flow Referenced to Air at 20°C [Eq. (2.84)]
Cv d l P kv
m3·s1
m m Pa 54,963
liter·s1
mm m mbar 5.496 × 103
liter·s1
mm mm mbar 5.496
liter·s1
mm m torr 7.326 × 103
liter·s1
mm mm torr 7.326
m3·h1
mm m mbar 0.01978
m3·h1
mm mm mbar 19.78
cfm
in ft torr 21,195
cfm
in in torr 254,363

< previous page page_111 next page >


< previous page page_112 next page >
Page 112

2.3.2
Turbulent Flow

Although particular solutions of the fundamental equations of fluid flow are known (viscous laminar flow, for example),
no general analysis of fluid motion has been successfully developed. The reason is the dramatic change in fluid behavior
which occurs at a critical Reynolds number when the flow becomes turbulent. The chaotic, fluctuating nature of
turbulent flow has defied theoretical analysis since it was first observed by Hagen in 1839. The flow equations which
have been developed are semiempirical, describing the gross mean properties of the fluid and ignoring small-scale
fluctuations. The DarcyWeisbach (1850) equation for head loss in a pipe can be expressed as

where ∆P is the pressure difference, u is the fluid velocity, and fD is the Darcy* friction factor. The friction factor fD is
not a constant (would that life were so easy!) but varies with the Reynolds number and is also a function of the cross-
sectional shape.

The Blasius relation holds for smooth pipes:

However, most real pipes are not smooth. Haaland [26] devised a useful general relation which takes surface roughness
into account. Jones [27] showed that the optimum friction factor correlation with Re for rectangular ducts is based on an
effective hydraulic diameter of Deff = (Dh/SF) which leads to an effective Reynolds number of Re/SF to be used in
calculating the friction factor. This approach is recommended by White [28] for all noncircular ducts. Haaland's
expression, modified to incorporate this proposal, is

where ε is the surface roughness, typically 0.0015 mm for the drawn tubing commonly used for system pipework.

SF is the shape factor, given by

and Ge is the numerical constant and cross-sectional geometric terms in the viscous flow conductance equations of
Section 2.3.1.

* The Fanning friction factor is also in common use in the literature on fluid flow. Darcy friction factor = 4 ×
(Fanning friction factor).

< previous page page_112 next page >


< previous page page_113 next page >
Page 113

Circular cross section:

Rectangular cross section:

Concentric annulus:

The shape factor for some cross sections is shown in Fig. 2.8.

Figure 2.9 shows (for circular cross section) the friction factor plotted against Re for three pipe diameters with ε =
0.0015 mm. Also shown for comparison is the Blasius smooth pipe relation. Larger-diameter pipes (d > 25 mm)
approach the smooth pipe curve for values of Re up to ~ 105, but there is significant deviation for smaller pipe sizes.

For viscous laminar flow the friction factor is a simple function of the Reynolds number:

For a circular tube we have SF = 1, so that fD = 64/Re.


Fig. 2.8
Shape factor for some cross sections.

< previous page page_113 next page >


< previous page page_114 next page >
Page 114

Fig. 2.9
Friction factor for a cylindrical tube.

< previous page page_114 next page >


< previous page page_115 next page >
Page 115

For a very narrow rectangular duct we have SF = 3/2, so that fD = 96/Re.

If the relation for fD is substituted into the turbulent flow Eq. (2.88), then the viscous laminar flow equations appear. It
would be convenient if fD were independent of cross-sectional geometry, because then laminar flow solutions for any
shape could be derived from Eq. (2.88). Unfortunately, the laminar flow solutions must be known in order to find the
shape factor.

For turbulent flow in a long duct, Eq. (2.88) can be rearranged to give the throughput as

Written in terms of Cz [defined by Eq. (2.72)], we obtain

Expressed in this latter form, it is immediately apparent that throughput equates to the product of a volumetric flow rate
term (Cz) and a pressure. In calculating throughput from this relation, it is only necessary to calculate Cz and express
the pressure in whatever units the user finds convenient. The terms under the square root sign are either dimensionless
or dimensionless ratios.

If throughput and downstream pressure are known (and hence the pumping speed or volumetric flow rate), then the
pressure ratio across the duct can be found from

Calculation is straightforward in this case; since the throughput is known, Re and hence fD are readily found. However,
the nature of the variation in friction factor with Re complicates calculation of throughput. A value of fD must be
chosen and calculated, Re can then be calculated, and a new value of fD obtained from Eq. (2.90). The process is
then repeated to obtain to the desired degree of accuracy. This procedure is painful for hand calculation, although it
is a simple matter to program on a personal computer. Often a single iteration will be sufficient and, if only a rough
estimate is needed, a constant value fD ~ 0.03 can be used. In vacuum pumping lines, Re seldom exceeds 105 and will
usually be less. Vacuum system pumping lines are often sized to avoid significant performance loss at the lowest
operating pressure. This usually means that losses are small at pressures where turbulent flow occurs, and accurate
calculation may be unnecessary. If accurate calculation is needed, then one simply has to go through the pain (or speak
to a friendly programmer). This is not the

< previous page page_115 next page >


< previous page page_116 next page >
Page 116

least of calculation problems that may be faced: It should be emphasized that the equations described thus far apply only
to long pipes without obstructions to the flow (such as bends and pipeline components) and to flow velocities of Ma <
0.3.

The entry length in turbulent flow is

At Re = 3000 for example, this gives an entry length of 17 diameters for a circular pipe. The criterion for a long pipe
will be discussed in Section 2.3.7.

2.3.3
Compressible Flow

Unlike other common fluids, such as water and oils, the flow of gases can involve significant changes in density. As
noted in the discussion on flow in long pipes, the effects can be ignored for small flow velocities (Ma < 0.3). At higher
flow velocities and in short pipes the effects of density changes become significant. Density is related to both pressure
and temperature, so that in analysis of gas flow the laws of thermodynamics must be considered in addition to the laws
of motion and continuity.

The flow relations will be discussed with reference to Fig. 2.10, which relates to an adiabatic process. When gas enters a
duct (from a large volume) the flow accelerates across the entrance, and this is associated with a fall in both temperature
and pressure. As the gas proceeds down the duct, pressure and temperature continue to fall and the gas continues to
accelerate, reaching its maximum velocity at the duct exit. If the pressure upstream of a duct is kept fixed and the
downstream pressure is reduced, then the flow rate will increase. As the pressure is reduced further, the flow rate
becomes constant when the pressure ratio across the duct reaches a critical value. Further reduction in the downstream
pressure will produce no further increase in rate of flow. This is due to the limiting velocity which can be achieved by a
gas flowing in a duct of constant cross sectionMach 1, the speed of sound.* In this condition the flow is said to be
choked (or blocked). The flow rate can be changed by changing the upstream pressure; but once the critical pressure
ratio is reached, the flow rate becomes completely independent of the pressure on the downstream side of the duct. The
term choked flow is used because the mass flow rate has reached the maximum value that it is physically possible to
achieve for a given upstream pressure and the duct behaves as though something is preventing, or choking off, any
further increase in flow. The phenomenon is particularly apparent with a small orifice restricting the inlet of a pump.
Once the pumping speed exceeds a certain value, the throughput will become constant and cannot be increased no
matter how large a pumping speed is used. Orifices, nozzles, and short or long ducts can become choked; the necessary
condition is that the flow velocity of the gas reaches the speed of sound. For any particular duct there is a specific value
of pressure ratio at which choking will occur; this critical pressure ratio is generally referred to as the choked pressure
ratio. In the case of an orifice or very short duct, the choked pressure ratio is independent of pressure or flow

* Supersonic velocity can be achieved downstream of the throat of a convergingdiverging duct.

< previous page page_116 next page >


< previous page page_117 next page >
Page 117

Fig. 2.10
Compressible flow through a duct.

conditions. In longer ducts the choked pressure ratio, because of its dependence on the friction factor, will change with
the pressure and type of flow (laminar or turbulent). Note that if the gas flow velocity in a duct does reach March 1, it
will do so at the duct exit. It is not physically possible for an initially subsonic flow (in a constant section duct) to reach
March 1 at any other point.

In the flow of gases through apertures or short ducts, or through longer ducts at high flow velocities, there is little or no
heat exchange between the gas and its surroundings and the process is approximately adiabatic. In longer ducts there
may be significant heat exchange and the flow may be approximately isothermal.

The thermodynamic theory of high-speed flow shows that isothermal conditions cannot be maintained at flow velocities
close to the limiting value because the required rate of energy input to the gas tends to infinity. Approximately
isothermal conditions may prevail for part of the flow through a duct but cannot be sustained if the flow becomes
choked. In long ducts the results of isothermal analysis and adiabatic analysis converge; in short ducts, adiabatic flow is
the most reasonable assumption. Some of the results of compressible flow analysis will be quoted without proof; the
basic theory is covered in a number of textbooks on fluid dynamics [2830].

Using results of thermodynamic analysis, it can be shown that throughput, velocity, and pressure at any point a along
the duct are related by

where Cz is defined by Eq. (2.72). The throughput is referenced to the stationary (or stagnation) conditions in the
upstream chamber (because of temperature changes, throughput is not constant through the duct).

For subsonic flow, the duct exit pressure (Px) must be the same as the downstream chamber pressure (Pd); so,
rearranging Eq. (2.100)

Since Max≤ 1, it is apparent that the flow will be choked if


< previous page page_117 next page >
< previous page page_118 next page >
Page 118

This is a general relation, true for a duct of any length. If the pumping speed S at the downstream end of a duct is
known, it can be immediately determined if the flow is choked. β is simply used as a convenient way of writing the
terms in γ.

It can be shown that the relationship between throughput and entry velocity is

and it can be shown further that the relationship between entry and exit velocity is given by

Choking is a property of this equation in that the variation of flow rate with Mach number exhibits a maximum at Max =
1.

For a finite length duct, the equations of compressible flow cannot be rearranged to give an explicit relation between
throughput and pressure. They can be solved for choked and nonchoked, laminar and turbulent flow, but complex
iterative procedures are needed. Sadly, envelopes (no matter how large) must be discarded and resort made to computer
programming. However, an approximate treatment, giving explicit equations accurate to a few percent, is covered in
Section 2.3.6.

It is instructive to examine how the rate of flow varies with duct length and an example is shown in Fig. 2.11. The
throughput (normalized to the throughput of an aperture) is plotted for a 10% pressure difference across a 12.5-mm-
diameter tube; flow conditions were arranged to be viscous laminar and turbulent. The normalized molecular flow
throughput is also shown for comparison. The striking difference

Fig. 2.11
Variation of normalized throughput with tube length for a
12.5-mm-diameter tube and a pressure difference of 10%.
< previous page page_118 next page >
< previous page page_119 next page >
Page 119

between the molecular and continuum flow cases illustrates the relative insensitivity to length of continuum flow in
short ducts. For larger pressure differences the reduction in continuum flow throughput with length is even smaller.

2.3.3.1
Flow through an Aperture or Short Duct

Since there is no frictional work and the process is adiabatic, flow through an aperture is often referred to as isentropic
flow.

For a duct of zero length, the right-hand side of Eq. (2.104) is zero and an obvious solution is Man = Max.

The critical pressure ratio for choked flow can be found by combining Eqs. (2.101) and (2.103) along with Man = Max
= 1:

The maximum throughput for choked flow, from Eq. (2.103), is

Typical values of Kpca, γ, and the function G(γ) are shown in Table 2.13 for several gas types. The values of γ are
typical of monatomic, diatomic, and so on, gases; but more accurate values can be found in Kaye and Laby [31], for
example.

For sharp-edged apertures the throughput is reduced by a factor of ~ 0.85 because the flow narrows to a vena contracta
which has a cross section smaller than the duct inlet area. This entry loss can be reduced to negligible proportions by a
radius on the entrance edge of ~ 0.2 diameters (or 0.2 × smallest dimension for noncircular ducts).

It is apparent from Eq. (2.106) that a choked aperture has a constant speed, given by

A choked aperture (or short nozzle) is often used a convenient means of providing a constant volumetric flow rate, or
speed, which is independent of pressure. Such a nozzle might be used to provide a 'soft start' pumpdown, to avoid
disturbances due to turbulence or pressure fluctuations in a process chamber during the initial stages of roughing. Any
disturbances occurring in the inlet ducting to the pump (downstream of the choke) will travel at the speed of sound; they
cannot travel upstream through the sonic choke and hence cannot be communicated to the chamber.
Table 2.13. Some Thermodynamic Properties of Gases
Gas Type γ Kpca G(γ)
Monatomic 0.7252
1.66 2.049
Diatomic 0.6847
1.4 1.893
Triatomic 0.6673
1.3 1.832
Polyatomic 0.6284
1.1 1.71

< previous page page_119 next page >


< previous page page_120 next page >
Page 120

For air at 20°C, Sa = 20A liter·s1, with A in cm2. Compare this with 11.6A liter·s1 for the molecular flow conductance
of an aperture.

The downstream pumping speed needed to choke an aperture is [cf. Eq. (2.102)]

For air at 20°C, Sca = 37.5A liter·s1, with A in cm2. Once the pump speed exceeds this value, the pumping speed at the
chamber will remain constant at Sa no matter how large the pump.

At pressure ratios smaller than the critical value, the throughput is given by

or, if the downstream pressure and pumping speed are known, the pressure ratio can be found from

The net speed at the inlet of the aperture will then be

There is no particular value in expressing the conductance of an aperture, since this varies with the pressure ratio. For
example, an aperture which is just choked (pressure ratio 1.89 for air) has a conductance of 42A liter·s1 (with A in cm2)
reducing to 20A liter·s1 for a very large pressure ratio The conductance C increases as the pressure ratio
is reduced and C →∞ as Kp→ 1.

For a choked aperture, it is interesting to note that for a given pressure difference the throughput is reduced by only 19%
with two apertures (of the same area) in series compared with a single aperture. In molecular flow the throughput would
be reduced by 50%

Temperature Changes. In adiabatic flow, the temperature at any point along a duct is

which indicates that there can be significant temperature changes in high-speed flow. For example, in the case of a
diatomic gas such as air, the temperature will fall from 20°C to 29°C if the gas reaches sonic velocity. This can cause
freezing up of choked nozzles due to condensation of water vapor initially present in the source air.

Time to Vent a Chamber Through an Aperture or Short Duct. This useful result can be derived from the choked and
nonchoked throughput relations given above. The
< previous page page_120 next page >
< previous page page_121 next page >
Page 121

source pressure (e.g., atmospheric) is assumed constant and much greater than the initial chamber pressure. The time for
the chamber pressure to equalize with the source pressure is then

where Cz is defined by Eq. (2.72) and Ca is the molecular flow conductance of an aperture [Eq. (2.15)].

The derivation assumes that the temperature of the gas, after entering the chamber, is the same as the source gas
temperature. In fact the gas temperature in the chamber will be greater, and this will reduce the equalization time.
Although the entering gas cools due to adiabatic expansion at the throat of the inlet aperture or nozzle, reheating occurs
as the gas comes to rest in the chamber. If the process were entirely adiabatic (i.e., no exchange of heat between the gas
and walls of the chamber), then the temperature rise (above the external ambient air temperature of T0 Kelvin) would be
T0(γ 1). For an ambient temperature of 293 K, this implies a temperature rise of about 100 K.

Practically, such large temperature rises are not observed due to heat exchange, especially since the thermal capacity of
the gas is much lower than that of the chamber. As an example, rapid venting (< 1 s) of a 22-liter chamber led to an
observed temperature rise of about 40°C. Greater temperature increases may be observed in larger chambers which are
vented rapidly.

2.3.3.2
Approximation for Flow through an Aperture

A simple approximation for nonchoked flow through an aperture can be obtained from Bernoulli's equation:

Taking downstream values for ρ and u leads to

or, if the downstream pumping speed is known,

The maximum errors occur at the choked pressure ratio and are 2.5% (γ = 1.66), 3.1% (γ = 1.4), 5.5% (γ = 1.3), and
10.9% (γ = 1.1). At lower pressure ratios the errors are less.

2.3.4
Corrections for Flow Obstructions

In most practical circumstances, gas and vacuum lines include obstructions such as bends, valves, or other components.
The basis for corrections is Bernoulli's equation

< previous page page_121 next page >


< previous page page_122 next page >
Page 122

and the observation that pressure head losses are proportional to the square of the mean flow velocity over the cross
section.

where nc is the loss coefficient, or number of corrections. This relation is commonly used to estimate head losses in
incompressible flow. The same principle is applied here for gases, with the velocity taken to be the mean value over the
length of the flow path. nc can apply to more than one component, and the corrections are simply summed for all the
components in the pipeline. The implication of taking the mean velocity is that the separate obstructions are regarded as
evenly distributed through the pipeline. It appears that the positioning of flow obstructions makes little difference to the
overall pressure loss. For example, with nc = 4 and maximum flow rate the variation in pressure loss is only about 2%
wherever the obstructions are located. This is likely to be smaller than the uncertainties in applying corrections.

For ducts longer than the entry length, the loss coefficient [32] is nc ~ 0.7 (flat plates) and 1.25 (cylindrical tube), for
viscous laminar flow. About half of this is due to excess shear over the entry length and the remainder due to the
additional pressure difference needed to accelerate a uniform flow into the developed velocity profile. This loss is in
addition to the kinetic energy needed to accelerate the flow from zero velocity. In ducts shorter than the entrance length,
correction is more complex; the subject is discussed in some depth by Shah and London [25]. The additional entrance
loss appears to be less important in turbulent flow. It is suggested to take nc ~ 1 to account for entrance losses. Note that
at high velocities in short ducts the kinetic energy needed to accelerate the flow is much more important than viscous
drag.

Table 2.14 lists suggested corrections to be used in the equations that follow. These are in addition to the kinetic energy
allowance built into the equations which thus assume that gas enters a duct from a large volume.

2.3.5
ApproximationsEntrance Correction Model

In a similar fashion to Dushman's original treatment of short ducts in molecular flow, entrance correction models
imagine a real duct to consist of an aperture in series with an ideal duct.

The total pressure drop is then taken to be the sum of the pressure drop across the entrance plus viscous losses in the
duct:

Pu Pd = (Pu Pi) + (Pi Pd) = entrance loss + viscous loss.

Table 2.14. Loss Coefficients Due to Flow Obstructions


Obstruction nc
Sharp-edged entrance
0.5
Mitred (90°) elbow
1
Standard (90°) elbow
0.8
Tee (used as elbow)
1
Tee (in-line flow)
0.25
Right-angle valve (fully open)
3
< previous page page_122 next page >
< previous page page_123 next page >
Page 123

Santeler [33] discussed a similar model but with the loss considered as an exit loss. For nonchoked flow, as shown
previously, it makes little difference where the correction is placed. Applying the isentropic equation for an aperture to
this model, the variation in calculated pressure difference between aperture-at-entrance and aperture-at-exit is
(for any length of duct). For a duct with a choked exit the Santeler model makes more intuitive sense since an entrance
cannot be choked.

In continuum flow this model is equivalent to allowing for the kinetic energy required to accelerate the gas to the mean
flow velocity:

As discussed previously, this can be generalized for any number of corrections:

Using this expression the following equations can be derived.

Viscous Laminar Flow. When the two end pressures are known the throughput can be found from

When the throughput and downstream pressure are known we obtain

Turbulent Flow. When the two end pressures are known we obtain

When the throughput and downstream pumping speed are known we obtain

These equations will be found useful in many practical circumstances. Strictly speaking, they should not be used for
choked flow or for very short pipes. In this case errors in calculated throughput can be up to around 50% (for an
aperture), although even this may be acceptable for a rough estimate. These equations do not show
< previous page page_123 next page >
< previous page page_124 next page >
Page 124

a maximum with pressure ratio; that is, the equations do not exhibit a choke property unlike the thermodynamic
equations and those which follow.

2.3.6
ApproximationsKinetic Energy Model

The main deficiency of the previous approximations is the failure to take proper account of the kinetic energy acquired
by the gas. The entrance correction approach is valid for an incompressible fluid which must reach its final velocity at
the duct entrance. If the fluid density is constant and the mass flow rate through the length of the duct is constant, there
can be no change in velocity. In contrast, a gas flowing in a duct reaches its final velocity at the duct exit. Thus, the
kinetic energy accounting must relate to the exit velocity.

In this case, the viscous loss term relates to mean density and velocity, but the kinetic energy correction applies to
density and velocity at the duct exit.

The flow equations derived from this expression show a maximum in the throughput with pressure ratio. In other words,
the equations exhibit the property of choking in the same way as the compressible flow relation. Unfortunately, an
explicit expression cannot be found for the critical pressure ratio, and the value of approximate equations is nullified if
iterative methods are required for solution. However, with a little analytical trickery, a solution is found by borrowing a
result from the thermodynamic theory, namely

obtained from Eqs. (2.101) and (2.103) with Ma = 1.

A set of equations can then be derived covering both nonchoked and choked flow including allowance for flow
obstructions. Iteration is needed to find the friction factor for turbulent flow when the two end pressures are known and
the throughput is to be calculated. Otherwise, all the relations are explicit and iteration procedures are not required.
These equations give calculated flow rates and pressure differences to within 7% of values predicted by the
thermodynamic equations.

Viscous Laminar Flow. (a) With both end pressures known, the choked pressure ratio is given by

If Kp > Kpc, so the flow is choked, then the throughput can be found from Eq. (2.123).

< previous page page_124 next page >


< previous page page_125 next page >
Page 125

If the flow is not choked, then

(b) If the throughput and downstream pressure are known, then the test for choked flow is [Eq. (2.102)]

S ≥βCz.

If the flow is choked, then the choked pressure ratio in terms of the known downstream conditions is

and the upstream pressure can then be found from

If the flow is not choked, then the pressure ratio is

Turbulent Flow. The choked pressure ratio is

(a) Both end pressures known: If the flow is choked, then the throughput is given by [Eq. (2.123)]

If the flow is not choked, then the throughput is


< previous page page_125 next page >
< previous page page_126 next page >
Page 126

(b) Throughput and downstream pressure known: The flow will be choked if [Eq. (2.102)]

If the flow is choked, then the upstream pressure can be found by rearranging Eq. (2.123):

If the flow is not choked, then the pressure ratio is

It is interesting to consider the case of flow from atmospheric pressure through a hole, of diameter 0.1 mm and length
10 mm, in a vacuum chamber. This might be, for example, a pinhole leak in a weld. With a pressure of 103 mbar in the
chamber, it is generally assumed that the flow changes from viscous through transitional to molecular at the vacuum end
of the hole. The flow is most likely to be viscous laminar, so Eq. (2.124) can be used to calculate the choked pressure
ratio. In this case, Cz = 2.276 × 103 liter·s1 [Eq. (2.72)], Cvu = 1.349 × 103 liter·s1 [Cv of Eq. (2.83) but calculated at
the upstream instead of the average pressure], and (for γ = 1.4) β = 1.296 [Eq. (2.102)]. Putting these values into Eq.
(2.124) gives Kc = 5.55, so the pressure at the hole exit is 160 mbar and the flow is continuum through the whole
length. The throughput [from Eq. (2.123)] is calculated to be 0.532 mbar·liter·s1; turbulent flow [from Eq. (2.71)]
would require mbar·liter·s1, so the flow conditions are viscous laminar as assumed. An almost identical
result is obtained from solution of the thermodynamic equations.

2.3.7
Long Duct Criteria

In setting out the simple equations for turbulent and viscous laminar flow, it was said that these were applicable only to
long ducts. This raises the obvious question as to what constitutes ''long." This is most easily answered by considering
the approximate equations derived from the kinetic energy model. Equation (2.130) is quite general, covering both
turbulent and viscous laminar flow (with appropriate choice of the friction factor). A duct can be considered long if
viscous drag is the dominant effectin other words, if the viscous drag term greatly exceeds the kinetic energy term so
that kinetic energy losses can be ignored. The long duct condition is then

The maximum pressure ratio is achieved when the flow is choked. Substituting the expression for the choked pressure
ratio [Eq. (2.129)] in this inequality leads to

< previous page page_126 next page >


< previous page page_127 next page >
Page 127

Interestingly, the right-hand term of this inequality is the (approximate) choked pressure ratio for an aperture.
Depending on the value of γ, this then gives

Low-speed flows imply a small pressure difference, Kp ~ 1, so that

Substituting for the friction factor fD in Eq. (2.134) gives

For example, with Re = 250 the criterion gives l 7.8d, meaning that the tube should be at least 78 diameters long
(taking "much greater than" to mean a factor of 10). The case is illustrated in Fig. 2.12, for which a set of pressures was
chosen (using the thermodynamic equations) to maintain constant throughput in a 10-mm-diameter tube under choked
viscous laminar flow conditions. Throughputs were then calculated using the entrance correction [Eq. (2.118)] and
kinetic energy [Eq. (2.123)] models and the uncorrected Poisseuille equation [Eq. (2.82)]. Even at l = 78d the
Fig. 2.12
Comparison of throughput calculated using various equations
and illustrating the long duct criterion.

< previous page page_127 next page >


< previous page page_128 next page >
Page 128

error in the Poisseuille values is still 19%, so the criterion is, if anything, rather conservative. The entrance correction
model gives improved accuracy and does not give the wildly inaccurate values for smaller lengths. The kinetic energy
model gives values within a few percent of the correct value over the whole range.

Thus, for viscous laminar, high-speed flows the long duct criterion is

For air at 20°C, flowing in a cylindrical tube the criterion can be expressed as

For high-speed flows, the likely error is still about 20% and the length would need to be increased by a factor of 2 to
achieve errors of less than 10%. For low-speed flows which also meet the incompressible criterion (Ma < 0.3), the error
in calculated values should be within 10%.

For high-speed, turbulent flow the long tube criterion is

The friction factor varies from 0.04 at Re = 4000 to 0.02 at Re = 50,000, giving l > 500Dh to 1000Dh. For low-speed
flows the values are halved; that is, l > 250Dh to 500Dh. With these criteria the error in calculated values should be
within 10%.

The conclusions to be drawn from this analysis are that a "long" duct is much longer than is commonly thought and that
is generally unsafe to use the simple uncorrected viscous and turbulent flow equations.

2.4
Transitional Flow

In the transition regime, gas flow dynamics are intermediate between free molecular flow and continuum flow. In very
short ducts transition occurs between molecular and isentropic flow and in long ducts between molecular and viscous
laminar flow. The flow cannot be turbulent at or near transitional flow pressures as is easily shown.

The maximum flow velocity possible is equal to the speed of sound, and the Reynolds number at this velocity is

Combining this with Eq. (2.2) for the Knudsen number Kn gives

< previous page page_128 next page >


< previous page page_129 next page >
Page 129

Kn = 0.01 marks the high-pressure boundary of transitional flow, so the maximum possible value of Re is ~ 160, much
less than the value (~ 2000) at which turbulent flow occurs.

Molecular flow analysis is concerned with the effect of bounding walls on the free flight of individual molecules, while
continuum flow analysis is based on hydrodynamic and thermodynamic considerations. It is clear that an analytical
unification between the regimes presents considerable difficulty, but a number of attempts have been made. A number
of analyses based on numerical solution of the Boltzmann equation have been described [3436]. Since, at the most
fundamental level, gas flow dynamics is determined by molecular interactions, attempts [37, 38] have been made to
extend Monte Carlo methods into the transition regime by taking account of moleculemolecule collisions. However,
these methods require considerable computing power even for the simplest geometries. Scherer-Abreu and Abreu [39]
developed a probabilistic three-dimensional model requiring more modest computing power and obtained good
agreement with published results.

2.4.1
Transitional Flow in Long Ducts

In spite of the significant amount of work, there have been no general derivations of flow equations which are entirely
based on first principles. The state of theory was reviewed by Thompson and Owens [40] who discuss, in particular, slip
theory and empirical methods of obtaining an equation for the total flow regime. In viscous laminar flow, a stationary
layer of fluid is assumed to exist adjacent to the duct walls. Slip theory supposes that the velocity of this layer is
nonzero, essentially due to a degree of specular reflection of molecules at the surface. One of the equations for a long
tube, discussed by Thompson and Owens, derived from slip theory can be written as

where Cm is the long-tube molecular flow conductance [Eq. (2.14) or (2.21)] and δ is the fraction of molecules diffusely
scattered at a surface. For large values of Kn the conductance tends to the molecular flow value, and for small values of
Kn the first term inside the brackets becomes dominant and the conductance becomes equal to the Poiseuille, viscous
flow conductance [Eq. (2.83)]. The best correlation with experimental data (for glass and copper tubes) is achieved with
δ = 0.84.

There is considerable evidence that δ ~ 1 under molecular flow conditions, and there is no evidence of slip in the
viscous regime (the Poiseiulle equation has been verified to a high degree of accuracy). Equation (2.141) has the
character of a linear combination of viscous and molecular flows with a suitable weighting function to achieve
agreement with observation. The theory of slip flow is not entirely consistent, and there is some question whether slip
actually occurs; the concept was introduced primarily as a means of extending solution of the NavierStokes equation
into transitional flow. Slip theory does not appear to give any better results than empirical methods.

< previous page page_129 next page >


< previous page page_130 next page >
Page 130

Knudsen [7] was the first to develop an empirical expression for transitional flow in a long circular cross-sectional tube,
which can be written as

where Cv is the Poiseuille long-tube conductance [Eq. (2.83)], Cm is the molecular flow conductance [Eq. (2.21)], and
XK is given by

Expressed in terms of the Knudsen number, we have

With pressure in millibars and diameter in millimeters, Eq. (2.142) becomes

where the first term inside the brackets is Cv/Cm and Fg is the gas factor defined by Eq. (2.4). At low and high
pressures the conductance tends to the molecular and viscous conductances respectively, as expected.

Table 2.15 shows values of the conductance (as a ratio to the molecular flow conductance) and the term Z1 for a range
of values of Kn together with the average

Table 2.15. Ratio of Conductance of Cylindrical Tube (Ct) for That for Molecular Flow (Cm) as a
Function of Knudsen Number
Pmean (mbar) Z1 Ct/Cm
Kn for d = 25 mm
0.809
104 26.44 737.1
0.810
103 2.644 74.44
0.811
0.01 0.264 8.174
0.821
0.1 0.0264 1.557
0.831
0.2 0.0132 1.199
0.856
0.5 5.289 × 103 1.003
0.884
1 2.644 × 103 0.958
0.905
1.55 1.706 × 103 0.952
0.917
2 1.322 × 103 0.954
0.955
5 5.289 × 104 0.970
0.974
10 2.644 × 104 0.982
0.986
20 1.322 × 104 0.990
0.994
50 5.289 × 105 0.996
0.997
100 2.644 × 105 0.998

< previous page page_130 next page >


< previous page page_131 next page >
Page 131

pressure (for air at 20°C) in a 25-mm-diameter tube. For Kn < 0.01 the flow is almost entirely viscous, whereas for Kn
= 0.5 the conductance has fallen to the molecular flow value. This is the justification for the values of Kn, delimiting the
flow regimes, discussed in the introduction.

The pressure versus conductance results are also plotted in Fig. 2.13 for a 1-meter-long, 25-mm-diameter tube. At low
pressures the conductance passes through a shallow minimum; the slope then rises through the transition region,
increasing to 45° in the viscous laminar flow region when conductance becomes proportional to pressure. At pressures
above 10 mbar the conductance has increased to more than 1000 liter·s1 and it might be thought that this size of pipe
could be used with a 50- or 100-liter·s1 pump with little loss in speed. However, Fig. 2.13 also shows the effect of using
this pipe with a 50-liter·s1 pump. At pressures ~ 1 mbar, the conductance begins to deviate from the long tube
conductance and flattens sharply at around 10 mbar as the flow becomes turbulent. At 100 mbar the conductance is ~
380 liter·s1 rather than ~ 5000 liter·s1. If a 100 liter·s1 pump were used, the conductance would be reduced to ~ 200
liter·s1. Even at a length of 40 diameters the pipe is not "long," as discussed in Section 2.3.7, and it cannot be assumed
that conductance always continues to increase with pressure.

As an illustration of the application of the Knudsen equation, it is of interest to calculate the rate of leakage of air from
atmosphere through a small hole into a vacuum chamber at 20°C. The pressure in the chamber is assumed to be much
lower than atmosphere (say, 1000 mbar), so the mean pressure is 500 mbar. Consider two sizes of holes:

1. l = 1 mm, d = 2 × 103 mm. The Knudsen number is 0.066, so flow is in the transition regime. The aperture
conductance (from Table 2.3) is 3.63 × 107 liter·s1 and α = 0.00265 [Eq.(2.25)], so Cm = 9.64 × 1010 liter·s1.
Applying Eq. (2.145) gives the conductance as Cm × 1.93 = 1.86 × 109 liter·s1, so the throughput is 1.86 × 106
mbar·liter·s1.

2. l = 1 mm, d = 0.02 mm. The Knudsen number is now 0.0066, so the flow is still transitional. Cm = 9.24 × 107
liter·s1. Applying Eq. (2.145) gives the conductance as Cm × 11.95 = 1.1 × 105 liter·s1, so the throughput is 0.011
mbar·liter·s1. To maintain a pressure of 104 mbar with this leak rate would require a pumping speed of 110 liter·s1.

As will be observed, the ratio of transitional to molecular flow conductance is greater for the larger-bore capillary.
These examples serve to illustrate the fact that, even for a very fine hole, the rate of leakage can be substantial.

The conductance minimum observed by Knudsen at intermediate pressures (Kn ~ 1.6) may not always occur at this
value and may be entirely absent. Pollard and Present [41] have offered a qualitative explanation for the minimum and
have shown that it should depend on the length of the tube. When the pressure is sufficiently low that the mean free path
is much greater than both the diameter and the length of the tube, subpopulations of molecules occur which have near
axial velocity components on entry to the duct or after scattering from a wall. These molecules, which can travel great
distances before a further wall collision, make a disproportionately large contribution to the transmitted flux. As the
pressure is increased, the mean free path

< previous page page_131 next page >


< previous page page_132 next page >
Page 132

Fig. 2.13
Variation of long-tube conductance with pressure and effect of
50 liter·s1 pump for a 25-mm-diameter, 1-meter-long tube.

< previous page page_132 next page >


< previous page page_133 next page >
Page 133

becomes smaller relative to the tube length, and the paths of these molecules are disrupted by intermolecular collisions.
At the same time, the effect of the increased number of intermolecular collisions is to initiate an overall drift velocity.
Pollard and Present reasoned that, since the development of continuum properties (and hence a drift velocity) depends
on λ/d, the decrease in flow due to the shortened mean free path will outweigh the increase due to a drift velocity until
the mean free path is more closely comparable with the tube diameter than with its length. The implication of this model
is that a conductance minimum is likely to occur for long tubes but may not occur for short tubes or, at least, will be less
pronounced.

The conductance minimum is evident in the computed results of Sharipov and Seleznev [42] who presented tabulated
values of normalized flow rate for a range of Knudsen numbers, based on solution of the Boltzmann equation. The
tabulated values can be used in calculation of conductance or throughput over the transition region and give results
which are a maximum of about 6% lower (at round Kn ~ 1) than those given by the Knudsen equation.

The term Z1 in the Knudsen equation varies between 1 and 0.81, depending on the pressure. A simple approximation to
the Knudsen equation takes Z1 = 1, so the equation becomes a straightforward summation of viscous and molecular
flows:

Calculation of throughput is then straightforward, using

If the downstream pumping speed and pressure are known, then the pressure ratio can be found from

There appears to be little information on transitional flow in ducts of noncircular cross section. Dong and Bromley [43]
discussed transition and slip flow and developed empirical equations for rectangular and annular cross sections,
although it is difficult to know how to interpret their results or to say whether the empirical relations are applicable to
ducts with lengths and aspect ratios different to those used in their experiments.

It is therefore suggested that the Knudsen Eq. (2.142) [or the simple approximation of Eqs. (2.146) to (2.148)] be used
as a rough approximation for transitional flow in other cross sections, taking Cv as the viscous conductance for the
particular duct shape.

An alternative to the Knudsen equation is

Like the Knudsen equation, this also exhibits a minimum in the transition region, although significantly deeper. This
may be more appropriate for narrow rectangular

< previous page page_133 next page >


< previous page page_134 next page >
Page 134

ducts which are known to show a more pronounced minimum than circular cross sections [43]. The basis of this
formulation will be covered in Section 2.4.3.

2.4.2
Long Duct Criterion in Transitional Flow

For continuum, high-speed flow, it was shown previously [Eq. (2.132), Section 2.3] that the worst-case condition for a
long tube is

so that, for a circular cross-sectional tube in laminar flow we have

Combining this with the result of Eq. (2.140) for the maximum Reynolds number leads to (for high speed flows)

At the high-pressure limit of transitional flow (Kn ~ 0.01, at the duct exit), this gives

depending on the value of γ. As discussed under continuum flow, this suggests llong ~ 40d even at sufficiently low
pressures that conditions at the duct exit are close to transitional. Since the Knudsen equation is a combination of
viscous and molecular flows, it is expected that a similar criterion is applicable through the regime. This implies that
llong becomes smaller as the pressure is reduced and flow conditions become transitional (Kn > 0.01). At molecular
flow pressures the kinetic energy term must vanish, so the criterion becomes irrelevant.

Sharipov and Seleznev [42] showed that the conditions of applicability of their solution of the Boltzmann equation are

where Kn is the Knudsen number at the mean pressure and ∆P is the pressure difference across the tube. The results of
this solution agree well with those calculated using the Knudsen equation. Thus, these conditions should also apply to
the Knudsen

< previous page page_134 next page >


< previous page page_135 next page >
Page 135

equation and to other empirical transitional flow formulations which are based on a combination of viscous and
molecular flows. These conditions imply that solutions are valid for relatively short tubes if the pressure difference is
small. However, one of the main assumptions underlying solutions of the Boltzmann equation is that the tube is
sufficiently long that end effects can be neglected and the flow considered as one-dimensional. This suggests that l
should not be less than about 20d even well into the transitional flow regime. It is difficult to be more specific because
of the dearth of experimental and theoretical information on transitional flow in short ducts.

2.4.3
Transitional Flow through Apertures and Short Ducts

Transitional flow through a thin slit was studied by Kieser and Grundner [44], who gave an empirical fit to data for air
at 20°C, which can be written as

where Ca is the molecular flow aperture conductance, a is the short dimension of the slit, λ is the mean free path at the
mean pressure, k1 = 0.5, and k2 = 0.3412. In their experiments, ; thus the equation only describes the transition
from molecular to continuum, choked flow. The authors noted fluctuations of the measuring points in the high-pressure
range, which they attributed to small alterations of flow pattern as a function of pressure. The equation gives the
maximum flow rate as 85% of the theoretical isentropic flow rate through an aperture [Eq. (2.106)], which is typical of
thin, sharp edged apertures.

As noted earlier (Section 2.3.3.1), the theoretical flow rate can be achieved with a small radius on the inlet edge of finite-
length ducts. It is suggested that the equation can be generalized for any gas and for apertures with radiused edges, with
k1 = 0.5 and

so that, at high pressures, Eq. (2.155) will give the theoretical value for an orifice.

Kieser and Grundner [44] also studied flow, at large pressure ratios, through rectangular ducts. They regard the duct as a
series combination of entrance aperture and the duct itself. In the modified form suggested by O'Hanlon [45] the flow
rate can be found from

< previous page page_135 next page >


< previous page page_136 next page >
Page 136

where Ct is essentially the Knudsen transitional flow conductance, given by Eq. (2.142) with Cv as the viscous flow
conductance for a rectangular duct. As for the aperture, this equation is valid only for the case when .

Santeler [33, 46] suggested the following formulation for transitional flow through an aperture (again valid only for
large pressure ratios):

where the subscripts ma and 0c refer to molecular flow [Eq. (2.15)] and isentropic flow [Eq. (2.106)] respectively, and θ
is a weighting function which can be written, in terms of Kn at the mean pressure, as

The best fit to the results of Eq. (2.155) is obtained with ks ~ 12 (Santeler used ks = 28). In a fashion similar to that of
Eq. (2.19), short ducts are considered to consist of an ideal duct in series with an exit aperture. The total pressure drop is
then due to purely viscous losses in the duct plus the pressure drop across the transitional/isentropic aperture.

If the formulation of Eq. (2.158) is applied to a finite length duct, then

where are the molecular and viscous flow throughputs. For a long circular section tube, it is interesting to
note that

If ks = 128/3π is taken as the basis for the weighting factor θ, the result is Eq. (2.149).

An expression, due to DeMuth and Watson [47], for nonchoked transitional flow through orifices is

where [Eq. (2.109)] is the isentropic and ma the molecular flow throughput of an aperture. The Knudsen number
is calculated at the mean pressure. The equation is based on studies of air flowing through apertures at pressure ratios
ranging from 1.1 to 1.4. Since the authors appear to define the Knudsen number as R/λ, their equation has been
modified to conform with the definition in this chapter.

For large pressure ratios, the duct + isentropic exit (or entrance) models appear to give good results for short ducts. This
is principally because high-speed flow through short ducts is very insensitive to length (as illustrated in Fig. 2.11) and
the isentropic aperture conductance is independent of pressure. But this type of model breaks down at lower pressure
ratios. As the pressure ratio is reduced, the isentropic aperture conductance increases and tends to infinity at vanishingly
small pressure ratios. The
< previous page page_136 next page >
< previous page page_137 next page >
Page 137
isentropic term, in Eq. (2.158) for example, then makes a disproportionately high contribution to the total flow. For example,
consider a pressure ratio Kp across an aperture. Using the approximation of Eq. (2.114) for continuum flow through an aperture and
dividing Eq. (2.158) by the pressure difference gives the conductance as

For a small pressure difference (say Kp = 1.01) and with Kn = 1 (molecular flow) so that θ = 0.93 [Eq. (2.159) with ks = 13], this
gives

The enhancement factor from Eq. (2.162) is 1.83 for the same conditions. It seems unlikely that the aperture conductance would be
enhanced by such a large a factor under essentially free molecular conditions.

Clearly, simple combinations of isentropic and molecular flow and aperture + entrance models are not a good basis for general
empirical equations applying to short ducts. As discussed under continuum flow, in short ducts a term must be included to account
for the kinetic energy required to accelerate the gas. It is suggested that approaches to empirical equations in the transition regime
could be based on formulations which force the kinetic energy term to vanish under molecular flow conditions.

Symbols

Symbol Meaning Section

a Duct cross-section dimension, minor axis of ellipse 2.2.5


A Cross-sectional area 2.2.2
b Duct cross-sectional dimension, major axis of ellipse, b ≥ a 2.2.5
B Cross-sectional perimeter (includes all wetted surfaces) 2.2.2
C Conductance 2.1
Ca Molecular flow conductance of an aperture 2.2
Cm Molecular flow conductance of a duct 2.2
Cn Net value conductances in series 2.1
Ct Transitional flow conductance of a duct 2.4.1
Cv Viscous flow conductance 2.3.1
Cvd Viscous flow conductance, calculated at downstream (lower) pressure 2.3.1
Cvu Viscous flow conductance, calculated at upstream (higher) pressure 2.3.1

Cz The term 2.3


d Diameter of a cylindrical tube Introduction
Dh Hydraulic diameter 2.3

(table continued on next page)

< previous page page_137 next page >


< previous page page_138 next page >
Page 138

(table continued from previous page)

Symbol Meaning Section

fD Darcy friction factor 2.3.2


Fg Gas factor Introduction
Ge Numerical constant and cross-sectional geometry terms in viscous conductance equations 2.3.2
G(γ) Function of γ defined by Eq. (2.106) 2.3.3.1
k Generally used for constants and units conversions
Kn Knudsen number Introduction
Kp Pressure ratio: (higher pressure)/(lower pressure) 2.1
Kpc Choked pressure ratio 2.3.6
Kpca Choked pressure ratio for an aperture 2.3.3.1
l Length of a duct (in direction of gas flow) 2.2.2
le Equivalent length used in Eq. (2.25) 2.2.4
lentry Entry length 2.3.1
Ma Mach number (ratio of flow velocity to local velocity of sound) 2.3.1
Man Mach number at duct entrance 2.3.3
Max Mach number at duct exit 2.3.3
Mm Molar mass (e.g., 0.028 kg·mole1 for nitrogen) Introduction
n Number density of molecules Fig. 2.5
nc Loss coefficient or number of corrections for flow obstructions 2.3.4
N Number of molecules striking an area 2.2

Mean pressure 2.3.1


P Pressure Introduction
Pd Downstream (lower) pressure 2.1
Pu Upstream (higher) pressure 2.1

Throughput (in pressure × volume units) 2.1

Choked throughput for an aperture 2.3.6

Value of throughput at onset of turbulent flow 2.3


r Generally used for ratios
R Radius of a cylindrical tube 2.2.4
Re Reynolds number 2.3
R0 Universal gas constant (8.314 J·mole1·K1) Introduction
S Pumping speed 2.1
Sa Speed of a choked aperture 2.3.3.1
Sca Pumping speed required to choke an aperture 2.3.3.1
SF Shape factor 2.3.2
Sn Net speed of a pump in series with conductances 2.1
t Time 2.2.11
T Thermodynamic temperature Introduction
u Flow velocity [and dimensionless terms in Eq. (2.36)] 2.3
va Mean thermal velocity of molecules 2.2.2
V Volume 2.1
Z Gas flow impedance 2.1
Z1 Term in Knudsen transitional flow Eq. (2.142) 2.4.1
α Transmission probability 2.2
αn Net transmission probability of components in series 2.2.10

(table continued on next page)

< previous page page_138 next page >


< previous page page_139 next page >
Page 139

(table continued from previous page)

Symbol Meaning Section

αH Ho coefficient of a pump 2.2.10

β The term 2.3.3


γ Ratio of the principal specific heats for a gas 2.3.3
δ Dimension ratio used in Eq. (2.26), diffuse scattering fraction in Eq. (2.141), and factor in Eq.
(2.42)
ε Wall roughness 2.3.2
λ Mean free path Introduction
η Viscosity of a gas Introduction
ρ Gas density 2.3
θ Angle in Eq. (2.18) and weighting factor in Eq. (2.158)
τ Time constant 2.2.11

References

1. M. Knudsen, Ann. Phys. (Leipzig) [4] 28, 9991016 (1909).

2. S. Dushman, Scientific Foundations of Vacuum Technique, Chapter 2. Wiley, New York, 1949.

3. W. Steckelmacher, Vacuum 16, 561584 (1966).

4. W. Steckelmacher, Proc. Int. Vac. Congr., 6th, Kyoto Japan, 1974; J. Appl. Phys., Suppl. 2 (Part 1), 117125 (1974).

5. W. Steckelmacher and M. W. Lucas, J. Phys. D 16, 1453 (1983).

6. P. Clausing, Ann. Phys. (Leipzig) [5] 12, 961 (1932); J. Vac. Sci. Technol. 8, 636646 (1971).

7. M. Knudsen, Ann. Phys. (Leipzig) [4] 28, 75 (1909).

8. W. Steckelmacher, Vacuum 28, 269275 (1978).

9. W. Steckelmacher, Rep. Prog. Phys. 49, 1083 (1986).

10. M. Smoluchowski, Ann. Phys. (Leipzig) [4] 33, 1559 (1910).

11. R. J. Cole, Rarified Gas Dyn. 10 (Part 1), 261272 (1976).

12. A. S. Berman, J. Appl. Phys. 10, 3356 (1965); erratum, ibid. 37, 2930 (1966).

13. D. J. Santeler, J. Vac. Sci. Technol. A 4(3), 338 (1986).

14. L. L. Levenson, N. Milleron and D. H. Davies, Vide 103, 42 (1963).

15. D. J. Santeler and M. D. Boeckmann, J. Vac. Sci. Technol. A 9(4), 2378 (1991).

16. R. J. Cole, Proc. R. Soc. Edinburgh 82A, 211223 (1979).


17. A. S. Berman, J. Appl. Phys. 40, 4991 (1969).

18. D. H. Davies, J. Appl. Phys. 31, 1169 (1960).

19. C. W. Oatley, Br. J. Appl. Phys. 8, 15 (1957).

20. R. A. Haefer, Vacuum 30, 217 (1979).

21. L. Holland, W. Steckelmacher and J. Yarwood, Vacuum Manual. Spon, London, 1974.

22. H. S. Carslaw and J. C. Jaeger, Conduction of Heat in Solids. Oxford University Press, London, 1959.

23. J. D. Lawson, J. Sci. Instrum. 43, 565 (1966).

24. V. E. Senecal and R. R. Rothfus, Chem. Eng. Prog. 49, 533 (1953).

< previous page page_139 next page >


< previous page page_140 next page >
Page 140

25. R. K. Shah and A. L. London, Laminar Flow Forced Convection in Ducts. Academic Press, New York, 1978.

26. S. E. Haaland, J. Fluids Eng. 105, 89 (1983).

27. O. C. Jones, J. Fluids Eng. 98, 173 (1976).

28. F. M. White, Fluid Mechanics, 2nd ed. McGraw-Hill, New York, 1986.

29. B. S. Massey, Mechanics of Fluids, 5th ed. Van Nostrand-Reinhold, Wokingham, Berkshire, England, 1983.

30. A. H. Shapiro, The Dynamics and Thermodynamics of Compressible Fluid Flow, Vols. 1 and 2. Ronald Press, New
York, 1953.

31. G. W. C. Kaye and T. H. Laby, Tables of Physical and Chemical Constants, 15th ed. Longman, New York, 1986.

32. F. M. White, Viscous Fluid Flow. McGraw-Hill, New York, 1991.

33. D. J. Santeler, J. Vac. Sci. Technol. A 4(3), 348 (1986).

34. C. Cercignani and F. Sernagiotto, Phys. Fluids 9, 40 (1966).

35. C. Cercignani, Theory and Application of the Boltzman Equation. Scottish Academic Press, Edinburgh, 1975.

36. S. K. Loyalka, T. S. Storvick and H. S. Park, J. Vac. Sci. Technol. 13(6), 1188 (1976).

37. Masahiro Ota and Hiroyoshi Taniguchi, Vacuum 44, 685 (1993).

38. L. Fustoss, Vacuum 31, 243 (1981).

39. G. Scherer-Abreu and R. A. Abreu, Vacuum 46, 863 (1995).

40. S. L. Thompson and W. R. Owens, Vacuum 25, 151 (1975).

41. W. G. Pollard and R. D. Present, Phys. Rev. 73, 762 (1948).

42. F. M. Sharipov and V. D. Seleznev, J. Vac. Sci. Technol. A 12, 2993 (1994).

43. W. Dong and L. A. Bromley, Trans. Natl. Vac. Symp. 8, 1116 (1961).

44. J. Kieser and M. Grundner, Vide 201, 376 (1980).

45. J. F. O'Hanlon, J. Vac. Sci. Technol. A 5, 98 (1987).

46. D. J. Santeler, J. Vac. Sci. Technol. A 12, 1744 (1994).

47. S. F. DeMuth and J. S. Watson, J. Vac. Sci. Technol. A 4, 344 (1986).

< previous page page_140 next page >


< previous page page_141 next page >
Page 141

3
Positive Displacement Vacuum Pumps

A positive displacement pump is a vacuum pump in which a volume filled with gas is cyclically isolated from the inlet,
the gas being then transferred to an outlet. The first pumps of this type were piston pumps, but in 1905 Gaede [1]
introduced the modern rotary vane (blade) vacuum pump. It has developed in many ways over the last 90 years: First
and very early on, it was oil-sealed and then a number of improvements were incorporated, including reduction of noise
and vibration, handling of condensibles and corrosive materials, elimination of oil being sucked back when stopped,
direct drive (running at 1450 or 1800 rev/min), and so on.

Another type of pump which has been widely used, primarily in the chemical industry, is the liquid ring pump wherein
the pumping action is developed by a rotating liquid. Pumps of this type produce nearly isothermal compression and can
handle dry gases or vapor gas mixtures. By reason of the isothermal compression, it is possible to handle explosive
gases or gases subject to polymerization. With the pumping of vapor-laden gases, condensation occurs in the pump,
resulting in an enhanced capacity.

In the early 1980s it became clear that the pumping of corrosive and inflammable materials, combined sometimes with
abrasive dust, was complicating the operation of the oil-sealed rotary pump and much reduced its normal long life and
performance. It was also becoming necessary to take precautions against backstreaming of the oil into some vacuum
systems to reduce system contamination. These two requirements led to the search for pumps which did not use liquid
sealing in their pumping mechanisms. The resultant pumps were referred to as ''dry vacuum pumps" because of the
absence of liquid sealing. The first pumps were based on a number of stages working either on the Roots Principle
(circa 1861) or on the more recent Claw Principle (circa 1930), but there are now a number of other mechanisms in use
and these are described in a later

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_141 next page >


< previous page page_142 next page >
Page 142

section. An attempt is made to indicate the relative merits of the various mechanisms, but due to the rapid development
in this field they are likely to be modified with time by design changes.

When a positive displacement pump is used to back a secondary pump which has high pumping speed, but in a lower
pressure region, the forepressure requirement of the secondary pumps must be satisfied. In recent years the introduction
of turbo pumps with integral molecular drag stages has resulted in pumps requiring a forepressure of around 10 mbar, so
a two-stage diaphragm pump is sometimes used. When choosing a forepump it is always necessary to ensure that the
effect on the performance of the secondary pump is taken into account (see Chapter 9).

< previous page page_142 next page >


< previous page page_143 next page >
Page 143

Part I
Oil-Sealed Vacuum Pumps

Nigel T. M. Dennis

3.1
Oil-Sealed Vacuum Pumps

Capacities available: 1 to 1500 m3/hr

Operating pressure range:

Single-stage: 1000- to 5 × 102-mbar total and partial pressure


Two-stage: 1000- to 103-mbar total pressure
1000- to 104-mbar partial pressure

Total pressure is influenced mainly by the type and quality of the oil charge, while the partial pressure is due to internal
transfer of gas within the pump.

3.1.1
Pump Design

There are two basic designs of oil-sealed rotary pumps, and these are illustrated in Fig. 3.1. Figure 3.1a shows the two-
vane pump where gas is trapped between the vanes and the stator before it is swept out through the outlet valve. Figure
3.1b illustrates the rotary piston pump where a single vane is slotted into the stator by the use of a hinge pin and is part
of a sleeve which fits around the rotor. The vane is hollow and acts as an inlet valve, closing off the pumping chamber
from the inlet when the rotor is at top center. Smaller pumps tend to be of the two-vane design, while larger pumps are
more frequently of the rotary piston design.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_143 next page >


< previous page page_144 next page >
Page 144

Fig. 3.1
Two basic designs of rotary pump mechanism. (a) Two-vane pump.
(b) Rotary piston pump. A, rotor; B, stator casing; C, rotor sleeve and
single vane; D, vane; E, hinge pin; F, gas ballast port; G, outlet valve;
H, gas entering the swept volume; J, the gas being compressed.

In the two-vane design the main path of leakage between the inlet and the outlet is the rotor to stator gap at the point
between these two ports. To reduce this leakage to a minimum, it is usual to use a long path seal which is provided by a
groove in the stator which is of the same diameter and on the same center as the rotor. This gives a long path clearance
which is sealed with oil between the inlet and the outlet and reduces carryover of gas between the two ports to a
minimum.

In a two-stage pump it is normal to feed outgassed oil from the outlet stage to seal and lubricate the inlet stage, thereby
ensuring that the best ultimate can be achieved.

3.1.2
Gas Ballast [2]

This is a feature useful in reducing the extent of vapor contamination of the oil. Atmospheric air or, if required, a dry or
inert gas is admitted to the pump during the compression stage (see Fig. 3.2) to increase the proportion of
noncondensable gas in the pump by the time compression has progressed to a point (about 1200 mbar) when the outlet
valve lifts. By this means the partial pressure of the vapor being pumped, at the time when the outlet valve lifts, does not
exceed its saturated vapor pressure at pump temperature, so that the vapor is discharged without liquefaction.
Furthermore, due to the extra work done in compressing the gas introduced as gas ballast, the pump temperature rises
and further assists in preventing vapor condensing within the pump. Gas ballasting is a useful technique for purging the
pump oil of condensed or dissolved vapor; it should proceed for a minimum of 20 minutes to allow the pump to warm
up completely. Gas ballasting is very useful for pumping vapors that do not dissolve in the pump oil and is still of some
value for vapors that do dissolve.

< previous page page_144 next page >


< previous page page_145 next page >
Page 145

Fig. 3.2
Method of introducing gas ballast to a bladed rotary pump.

The gas ballast flow is generally about 10% of the free air displacement of the pump. Because the effectiveness of gas
ballast is very dependent on pump temperature, it is desirable to run the pump as hot as possible. Typically, running
temperatures are 60°C to 90°C.

Condensation will occur when the temperature in the pump outlet line reaches the saturated vapor pressure of the vapor
that is being pumped. It is, therefore, important that a catch pot is fitted immediately adjacent to the outlet port so that
fluid condensed in the outlet line does not run back into the pump oil box.

The ultimate pressures attainable when gas ballast is used are as follows:
Single-stage pump: about 0.5-mbar total pressure
Two-stage pump: about 102-mbar total pressure

< previous page page_145 next page >


< previous page page_146 next page >
Page 146

To calculate the minimum (maximum) safe pumping rate it is assumed that the ballast gas just becomes fully saturated
with vapor at the end of the compression as it is forced through the outlet valves and that the vapors are insoluble in the
pump oil [3]. It is therefore necessary to know the volume flow rate, the partial pressure, and the absolute temperature
of the vapor under these conditions (this may be assumed to be the temperature of the pump itself). The partial vapor is
then the saturation vapor pressure at the pump temperature (T, K), and its value is available from vapor pressure data
(Ps, mbar). If the total pressure reached under the outlet valve is PT (mbar), the partial ballast gas pressure is PT Ps.

The volume rate of flow for the vapor is the same as for the ballast gas and is therefore obtained by calculating the
volume flow rate for the ballast gas, in terms of its measured flow rate (V, m3/h) into the pump at atmospheric pressure
and temperature (PA and t, K).

Thus we have the volume/hour of emerging gas:

If the maximum pumping rate is W (kg/h) and if the vapor density of the saturated vapor at T (K) is ρs (kg/m3), then

The ordinary perfect gas equation is used to obtain ρs with adequate accuracy in this pressure range; thus for water
vapor we obtain

Taking PA as 1013 mbar and assuming that the pressure PT under the outlet valve when the mixture is expelled is 1200
mbar and t = 293 K, then for water vapor we obtain

Knowing Ps (the vapor pressure for water) at the pump's temperature T and the volume rate of flow of the ballast gas, it
is now possible to calculate the maximum safe water vapor pumping rate in kg/h. Knowing the speed of the pump (S,
m3/h), it is then possible to calculate the maximum safe inlet pressure of water vapor:

< previous page page_146 next page >


< previous page page_147 next page >
Page 147

If the ballast air is humid, then Ph (mbar) is the partial water vapor pressure, the true mass flow rate of air is reduced in
the ratio (PA Ph)/PA, and water vapor is also entering the pumps with the ballast air at a rate of

Equation (3.1) must be changed to allow for the reduced true ballast air flow and the additional water vapor entering
with the ballast air; thus

This is only applicable when pumping water vapor. If other vapors are pumped, the moisture contributes useful ballast.

The effect of ballast humidity is small at the normal gas-ballasted pump temperature of 70°C or higher and is generally
disregarded.

As the vapor pressure of water is rapidly increasing at these higher temperatures, the pump temperature is very
important in determining its water vapor capacity. In the standard on determining this part of the pump performance, it
is indicated that an initial test is carried out allowing air to be substituted for water vapor at the likely safe inlet pressure,
so that the pump temperature T is adjusted.

High or low ambient temperature has an effect which can be appreciable if it is more than a few degrees away from 20°
C.

An alternative to gas ballast is the use of extremely small air bubbles passing through the oil to carry the contamination
away. This has the advantage of minimal loss of pump oil and a normal ultimate vacuum. The amount of contamination
that can be removed by this method is much less than by normal gas ballasting and does require a small air pump to
produce the bubbles.

3.1.3
Pump Oil [4]

The type most widely used is normally highly refined hydrocarbon oil. For very-heavyduty applications, synthetic
organic fluids are used, and in highly corrosive or flammable situations a perfluoropolyether fluid is used. The oil used
should be of the viscosity recommended by the manufacturer; lower-viscosity oil normally results in the pump being
noisier, while higher-viscosity oil can give difficulty in starting at low ambient temperature and occasionally leads to
seizure.

1. Hydrocarbon Oils: The general fluid used in most applications is a highly refined mineral based oil; it normally has a
vapor pressure around 106 mbar at room temperature. Some oils have additives to help reduce the corrosive effects of
any vapors being pumped, but their addition results in a higher ultimate total pressure. They are also gradually
neutralized, so their effective life is limited. Hydrocarbon oils are slightly hygroscopic, so the pump should be filled
from

< previous page page_147 next page >


< previous page page_148 next page >
Page 148
a bottle or can which has only recently been opened. If the pump is not used for a lengthy period, it is advisable to gas ballast for a while to clear
the oil of any dissolved water vapor.

2. Synthetic Organic-Type Fluids: These are used on heavy-duty applications because certain types have improved resistance to oxidation at the
high pump temperature associated with these applications.

3. Fluorinated Fluids [5]: These are rotary pump grades which are used in applications where their corrosion-resistant properties and
chemical inertness result in greatly increased operational times between pump maintenance. The basic properties of perfluoropolyether fluids
and their advantages for use in vacuum systems are as follows:

Chemical inertness. Ideal for pumping aggressive materials, particularly in semiconductor processing.
Noninflammable. No fire risk.
High thermal No residual "tars" formed by overheating.
resistance. Eventually reduced to gaseous products.
Oxygen-compatible. Allows safe pumping of oxygen. Refer to manufacturers literature for maximum recommended
service temperature and pressure with oxygen.
Immiscible with most common solvents. Allows pumping of solvent in some cases without gas ballast.

The pump must have no traces of hydrocarbon oil when charged with perfluoro-polyether fluid; otherwise, some or all of the above advantages
are adversely affected. With this type of fluid the ultimate vacuum of the pump is not normally as good as with the normal oils. This is due to oil
box gases being dissolved in the fluid and evolving when the fluid reaches the high-vacuum stage of the pump. Note: Perfluropolyether fluids
decompose at 250300°C, and the resultant products are toxic.

3.1.4
Oil Suckback [6]

This occurs if the pump is stopped for a period of time under vacuum. Oil suckback is prevented by incorporating into the pump inlet line a
nonreturn valve. When this is an integral part of the pump, it is operated in a number of different ways. When this arrangement is used, it is
normal for the pump mechanism to be allowed to reach atmospheric pressure so that there is no chance of oil being sucked back due to faulty
sealing of the valve. An alternative way is to keep the pump under vacuum and stop any oil entering the pump by an oil valve and a sealing
outlet valve. The oil and outlet valves are designed so that only a small amount of oil can be sucked into the pump mechanism if the valves fail
to seal.

3.1.5
Power Requirements and System Protection

The power to the pump rises at a maximum of 100300 mbar and then falls to a low level. It is mainly caused by friction within the pump and
motor losses, when the pump reaches 10 mbar and below.

< previous page page_148 next page >


< previous page page_149 next page >
Page 149

Contamination of the system occurs at low gas flows toward the pump, normally when the inlet pressure is below 1.0
mbar. This is due to condensation of both the oil vapor from the pump and its breakdown products in the system which
is at a lower temperature than the pump. Therefore, it is inadvisable to leave a rotary pump on a system at or near its
ultimate vacuum for a prolonged period without a trap or another type of pump between it and the system. The traps
used contain either an activated alumina or a molecular sieve trapping medium. Both these materials absorb water
vapor, and it is therefore best to bypass them during pumpdown because otherwise it becomes difficult to achieve a
good ultimate vacuum.

3.1.6
Accessories [7]

Besides the inlet trap mentioned above, there is a wide range of accessories mainly aimed at preventing dust from
getting into the pump mechanism, decontaminating the pump oil, and preventing oil from being lost from the pump.
Inlet filters are often used on dusty systems, but these reduce the pumping speed considerably especially at the lower
pressures. The filter element is normally removable and of a pleated construction to give a high conductance. In
applications where dust that is too fine to be readily stopped by inlet filters or is being continuously formed at the pump
inlet, an oil circulation system is fitted to the pump oil box. This system includes an oil filter and a circulating oil pump
(sometimes an integral part of the vacuum pump). An oil pressure gauge or switch gives an indication of when the filter
element has to be changed.

On a system where large quantities of vapor are to be pumped, it is normal to use an inlet condenser. The cooling
medium of this condenser is normally water, but it can be cooled down to as low as 80°C depending on the application.
For details see "Sorption Roughing Pumps," Section 5.20. Outlet mist filters are used to capture the extremely fine mist
that is evolved from a pump. Some pumps include a preliminary oil filter directly above the outlet valve, but to
completely eliminate mist under all operating conditions it is still necessary to fit an external filter. In smaller systems
this filter can also include a charcoal deodorizer. In cases where the pumps are running at reasonably high pressure
(about 10 mbar), it is possible to suck the collected oil back from the outlet filter into the inlet of the pump or through
the gas ballast connection; otherwise, it is necessary to periodically empty the oil from the filter and return it to the
pump. If vapors are being pumped, neither of these systems is possible and the oil should not be returned to the pump.

< previous page page_149 next page >


< previous page page_151 next page >
Page 151

Part II
Liquid Ring Pumps

Helmut Bannwarth

3.2
Liquid Ring Pumps

Capacities available: 1 m3/h to 27,000 m3/h vacuum pumps [8]


Operating pressure range (minimum): 1013 to approximately 33 mbar total pressure extended to 5 mbar with gas ejector

3.2.1
Mechanism

With regard to the arrangement of the impeller in the housing, it is necessary to distinguish between the concentrically mounted and the
eccentrically mounted vane-impeller. Machines with cylindrical housings are designated as single-acting or single-chamber liquid ring
pumps. According to their method of operation, machines with oval-shaped or elliptical housing are designated as double-acting or two-
chamber liquid ring pumps. A further classification is made with regard to the direction of the gas inlet and outlet stream in relation to the
impeller. Usually, pump assemblies have axial or radial flow. Liquid ring vacuum pumps are, as a rule, of single-stage or two-stage
design. The functioning of a liquid ring machinefor example, a vacuum pump with eccentrically arranged impelleris described as follows.
The liquid ring vacuum pump with eccentric impeller, arranged in a circular housing, belongs to the group of single-acting or single-
chamber liquid ring pumps. The housing must not be exactly cylindrical, but should have an elliptical profile. Through varying the profile
of the housing, it is possible to attain the optimal adjustment of the machine for the operating requirement. In these pumps, the driving
force from the impeller is transferred through the liquid ring to the pumped medium. The liquid ring will accelerate on the

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_151 next page >


< previous page page_152 next page >
Page 152

Fig. 3.3
Cross section of a one-stage liquid ring vacuum pump
with canned motor (Lederle-Hermetic GmbH).

suction side and, therefore, has augmented energy. On the discharge side, the ring liquid enters the impeller
compartments, whereupon a portion of the kinetic energy is transformed into static energy of compression. On the
dissipation of the static energy, the losses in the gas and liquid streams are overcome, so that the gas is compressed and
ejected.

Assuming that the vane-impeller is arranged eccentrically in the properly adjusted housing, the moving impeller gives
impetus to the rotating liquid ring. Because of the eccentricity of the impeller to the housing, a crescent shaped cavity is
formed between the impeller hub and the liquid ring as the impeller rotates. The impeller vanes split up this cavity into
many segments of differing volume. On rotation of the impeller, the segments increase in size in the region of the intake
port (Fig. 3.3) and gas or gas vapor mixture is sucked in.

In the region of the discharge port opposite, the pumped medium in the segments (which are here becoming smaller)
compresses and is ejected. The intake and discharge ports are to be found with control plates on the side of axial-intake
machines. Liquid ring pumps are rotary machines with the characteristics of piston pumps. The liquid ring assumes the
function of the piston.

3.2.2
Single-Stage Liquid Ring Vacuum Pumps [9]

With the single-stage liquid ring vacuum pump, it is possible to attain various pressure ratios through the design and
arrangement of ports. Liquid ring vacuum pumps with control plates and venting holes are operated economically at
compression ratios of up to 1:7 for suction pressures between 1013 mbar and about 150 mbar

< previous page page_152 next page >


< previous page page_153 next page >
Page 153

Fig. 3.4
Cross section of a two-stage liquid ring vacuum pump (Sihi GmbH). 1, Shaft;
2, impellers; 3, control plates; 4, middle section of housing; 5, end cover;
6, shaft seal; 7, shaft bearing; 8, suction port; 9, discharge port.

with water at 15°C as the ring liquid. The single-stage liquid ring vacuum pump can operate at compression ratios
greater than 1:7 but at considerably diminished capacities, provided that the discharge port is reduced, the venting holes
are suitably arranged, and an automatic valve is provided. Disc valves are usually used. The compression ratio is
accommodated by means of this valve, which is contiguous with the discharge opening in the control plate. The flexible
plate in use covers or uncovers the venting holes automatically over the entire pressure range and prevents
overcompression and back flow. Using water at 15°C as the ring liquid, inlet pressures in the range between 1013 mbar
and about 33 mbar are achieved with a corresponding compression ratio between inlet and discharge pressure of 1:30.7.
If other ring liquidsfor example, oilare used, inlet pressure is limited to between about 10 mbar and 30 mbar due to
outgassing.

3.2.3
Two-Stage Liquid Ring Vacuum Pumps [8,9]

For inlet pressures which lie in the region below 150 mbar, instead of using single-stage machines with flexible
discharge ports, it is possible to use a two-stage liquid ring vacuum pump (Fig. 3.4). This construction has gas discharge
openings without valves. The discharge from the first stage becomes the suction to the second stage. The partial
compression ratio per stage is variable in regard to the inlet pressure. Through
< previous page page_153 next page >
< previous page page_154 next page >
Page 154

a fixed outlet port, a compression ratio of less than 1:7 per stage is maintained. The overall compression ratio is the
product of the two partial compression ratios. With water at 15°C as the ring liquid, two-stage liquid ring vacuum
pumps may also be used to accommodate suction pressures in the range between 1013 mbar and 33 mbar.

3.2.4
The Operating Liquid

The operating liquid serves both to transfer energy and to seal the vane-impeller and spaces between the impeller,
control plate, and housing, and it also absorbs and removes quantities of heat generated in the pump [10]. Water as well
as other ring fluids, dependent upon the nature of the process, may be usedfor example, hydrocarbons, solvents, and
synthetic light oils. In addition to the heat of compression, additional heating in liquid ring pumps may be generated by
condensation of vapors and absorption of gases, through chemical reactions occurring between process gas and the ring
liquid or through the cooling of the principal gas at elevated temperatures.

3.2.5
Operating Ranges of Liquid Ring Gas Pumps

To obtain the lowest achievable pressure with the liquid ring vacuum pump, it is necessary to prevent boiling of the
operating liquid. Using water at 15°C as the ring fluid, the lowest possible inlet pressure which is practical as well as
possible is approximately 33 mbar [11]. Lower inlet pressures may be reached through the use of a liquid having a
higher boiling point. An improvement of the inlet pressure can also be attained through the use of an additional vacuum
pump operating on a different principle connected on the inlet side of the liquid ring pump. Liquid ring vacuum pumps
are often used in combination with gas ejector pumps; and here it is possible, using water at 15°C as the ring liquid, to
accommodate inlet pressures in the range between 33 mbar and approximately 5 mbar. In the combination, the gas
ejector is placed directly at the intake of the vacuum pump (Fig. 3.5). The special curve over the operating range of a
vacuum pump with gas ejector, using water at 15°C as the ring liquid, is shown in Fig. 3.6. As the driving medium for
the gas ejector, we may use atmospheric air, the process gas itself, or any suitable noncondensable gas at atmospheric
pressure.

The liquid ring vacuum pump can also be adapted for use at inlet pressures below 5 mbar. In these cases, one or more
vacuum pumps utilizing a different pumping principle are placed at the inlet side of the liquid ring pump. The choice of
vacuum pumps is to be such that the pump combination will operate synchronously over the required suction pressure
range. Since over wide pressure ranges there are changes in specific gas densities and types of flow (viscous, Knudsen,
and molecular flow), different pumping principles are used for operation over wide pressure ranges. A condenser is
often used in a vacuum pump combination for handling vapors and often improves the overall efficiency of the pumping
equipment [12].

3.2.6
Cavitation and Protection against Cavitation

The application range of the liquid ring vacuum pump for lower inlet pressures is limited by the vapor pressure of the
operating liquid. If the physical properties of the

< previous page page_154 next page >


< previous page page_155 next page >
Page 155

Fig. 3.5
Liquid ring vacuum pump with gas
ejector. A, intake gas; B, atmospheric
air as motive gas; P0, intake pressure;
P, back pressure of the gas ejector =
intake pressure for the liquid ring
vacuum pump; C, exit of the
intake gas and motive gas.
Fig. 3.6
Operating range of a liquid ring vacuum pump with gas ejector.

< previous page page_155 next page >


< previous page page_156 next page >
Page 156

operating liquid are unsatisfactory, it will boil and the impeller vanes will fill up with vapor. The vapor transportation
by the movement of the impeller vanes toward the pump outlet side results in collapse of the vapor bubbles. This
implosion, known as cavitation, causes a banging noise and noticeable shaking of the pump. Cavitation will not occur in
the liquid ring vacuum pumps if the pump operation is always limited to a minimum inlet pressure and corresponding
volumetric flow and noncondensable gases are handled. The pump can also be provided with gas from a liquid separator
on the inlet side operating in conjunction with a regulating valve toward the inlet. If a gas ejector is used in a pump
combination and there is always a gas flow through the open driving gas inlet, cavitation will not occur [11]. To prevent
cavitation damage, a liquid ring vacuum pump can be provided with a built-in cavitation protection. This is in the form
of a gas ballast in the working space on the discharge side. This has only a small effect on the pumps operating
characteristics.

3.2.7
Types of Operation; Conveyance of Operating Liquid

Three principal methods of operation may be differentiated: operation without back flow of liquid (fresh liquid injected,
cooling circuit); operation with back flow of liquid (combination operation, economical circuit); and closed liquid
circulation (return circulation, closed circulation) [12]. The principle of operation and disposal of the operating liquid
will depend upon the process technology, the suction pressure, the available cooling medium and its temperature, the
situation with regard to corrosion of materials, and the nature of the process gas and operating liquid. The installation of
these machines and correspondingly optimized vacuum systems should be done in an economically favorable manner so
that environmental safety precautions are observed. Figure 3.7 shows the principal circuitry for operation with closed
liquid circulation.

In this method of operation, the operating liquid in the liquid separator is separated from gas as well as vapor and passes
back into the pump through a heat exchanger wherein it is indirectly cooled. In the heat exchanger, the heat is removed
from the system. The special advantage of this method of operation is that the intake gases and

Fig. 3.7
Principal circuit for operation with
closed liquid circulation.

< previous page page_156 next page >


< previous page page_157 next page >
Page 157

vapors as well as the operating liquid do not come into contact with the cooling medium. There is no problem of
disposal from the cooling liquid side. By reason of the isothermal compression at low gas temperatures, the gases or
vapors discharged into the atmosphere contain only small quantities of noxious materials.

The closed circulation method of operation is used in the chemical industry and in other situations wherein the
requirement for freedom from leakage and protection of the environment exists. The contaminated operating liquid may
require special disposal procedures.

3.2.8
Materials of Construction

In the choice of materials of construction, resistance to corrosion normally plays the major role. In process technology,
the gases to be compressed are very often chemically active or process conditions require the use of alkaline or acidic
operating liquids. Liquid ring pumps are robust machines whose moving and fixed internal parts are not subject to
failure. It is practical to use all materials in the construction of these pumps which are usual in machine building and in
process technology. Liquid ring machines may be constructed using such materials as gray cast iron, spheroidal cast
iron, cast steel, chemically resistant chrome nickel steels, high nickel alloys, titanium bronze polypropylene, polyvinyl
plastics, Noryl (polyphenylen oxide), Ryton (PPS), ceramics, and stoneware. Combinations of different materials may
also be used. Parts for these pumps may also be coated with hard rubber, plastics, enamel, or special lacquer to protect
against corrosion and erosion.

3.2.9
Sealing

In liquid ring vacuum pumps, one differentiates between static and dynamic sealing. Fluid seals based on
polyvinylacetate, polyisobutylene, and epoxyresin are used as static seals for the housings. Nitrile rubber, fluorinated
elastomers, polytetrafluorethylene, ethylenepropylene rubber and chloroprene rubber are used for O rings and flat
gaskets. Where dynamic sealing is required, such as for shafts, conventional seal elements such as stuffing boxes or
single- or double-acting mechanical seals are used. Stuffing boxes with double packing and lantern rings are usual. In
vacuum technology the demand exists that there be zero leakage when machinery is operating and when it is stopped.
Liquid ring machines in hermetic construction are presented as an alternative to vacuum pumps of the usual design with
dynamic seals.

There exist two alternative types of hermetic machines: those using a drive with permanent magnet coupling and those
with canned motors (Fig. 3.3). In choosing a drive system, the transfer of torque and the requirement for shaft sealing
exist concurrently. The hermetically encapsulated construction is recommended in cases where an absolutely leak free
and maintenance free operation is demanded. With machines of this type, a zero rate of leakage is achieved.

3.2.10
Drives

Two-, four-, or six-pole low-voltage asynchronous squirrel-cage motors are usually used as drives. For larger machines,
having drive requirements above about 150 kW, high-voltage squirrel-cage motors are used. For large pumps operating
at low speeds,

< previous page page_157 next page >


< previous page page_158 next page >
Page 158

gear reducers of V-belt reducers may be placed between motor and pump, thus allowing for speed and adjustment such
that specific operating conditions may be optimally met. Especially for large machines it is useful to place a
hydrodynamic coupling between the motor and pump in order to take care of starting resistance and frequency control
on start-up of induction motors.

Hermetic drive systems may be used as alternatives to conventional drives. The torque is transferred through either
permanent magnets or electromagnets. The complete leak tightness of these pumps is especially necessary when toxic,
carcinogenic, or radioactive gases or vapors are handled and where no leakage of the operating liquid can be tolerated.
Liquid ring vacuum pumps with hermetic drive systems are presently in use in the process industries; these pumps have
maximum capacities of about 3000 m3/h and can be provided in special construction for higher capacities.

With canned motors, the noise emissions of a pump may be reduced by an average of 810 dB(A).

3.2.11
Accessories

A heat exchanger and liquid separator are needed for the operation of a liquid ring vacuum pump with closed circulation
(Fig. 3.7). Dependent upon the operating conditions, various other accessories such as gas ejector, check valves,
emission cooler, control and surveillance instrumentation, and so on, may be required.

< previous page page_158 next page >


< previous page page_159 next page >
Page 159

Part III
Dry Vacuum Pumps

Nigel T. M. Dennis

3.3
Dry Vacuum Pumps

3.3.1
Roots Pump

Roots Pump (mechanical booster) plus oil-sealed rotary forepump

Capacities available: 75 to 30,000 m3/h


Operating pressure range: 10 to 103 mbar total pressure, inlet pressure is sensitive to fore pressure

Multistage Roots Pump

Capacities available: 25 to 1000 m3/h


Operating pressure range: (4-stage) 1000 to 5 × 102 mbar total pressure
(5-stage) 1000 to 2 × 102 mbar total pressure
(6-stage) 1000 to 103 mbar total pressure

A typical Roots pump is illustrated in Fig. 3.8. It consists of two figure-of-eight-shaped rotors, although for higher pressure duty three-lobe rotors
are sometimes used. These rotors are synchronized by external gears and rotate in opposite direction within the stator. The gears and rotor
bearings are oil-lubricated; but they are external to the pumping chamber, so the swept volume of the pump is dry. A small clearance, generally
between 0.05 and 0.25 mm, is maintained between the rotors and between each rotor and the stator wall. As the rotors run dry, back leakage
occurs through these clearances at a rate dependent on the pressure difference across the pump and

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1998 John Wiley & Sons, Inc.

< previous page page_159 next page >


< previous page page_160 next page >
Page 160

Fig. 3.8
Cross section through a Roots (mechanical booster)
pump and its operating cycle.

the gas being pumped. The pressure ratio generally achieved by this mechanism is shown in Fig. 3.11, indicating a high
compression at low pressures and a low compression at high pressure.

A mechanical booster is basically a Roots pump modified for high-vacuum applications. This generally involves
improved leak tightness, different stator/rotary seals, and the connection of the gear box to the outlet of the pump so the
same pressure exists in the gear box as in the foreline.

As indicated in Fig. 3.11, the ultimate pressure achieved by a mechanical booster is very dependent on the ultimate of
the fore pump [13], and where a liquid ring pump [14] is used it is usually necessary to use two mechanical boosters in
series to obtain pressure below 1 mbar.

Rotational speeds vary between 500 and 3440 rev/min and depend on the size of pump; the upper limit is dependent on
the supply frequency.

Forepump speed is normally between one-fourth and one-tenth the displacement of the mechanical booster so as to give
a relatively small pressure differential across the pump.
Overheating is a potential source of trouble with a mechanical booster causing loss of clearance which may result in
seizure. Operation up to about 90°C is generally permissible with cooling only applied to the shaft seal, bearings, and
gear box. The cause of overheating is that the gas in the foreline, heated as a result of compression,

< previous page page_160 next page >


< previous page page_161 next page >
Page 161

Fig. 3.9
Speed curve and power consumption of a 100-m3 · h1
multistage Roots pump at 50 Hz. The ultimate vacuum
improves when the pump is run at 60 Hz, and the
speed curve increases by about 20%. The speed
and power curves are without gas purge.

can flow back into the pump because there is no exhaust valve; it then undergoes further compression and heating [15].
This normally only becomes a problem when the inlet pressure is held between 2 and 40 mbar for long periods. With
larger-size pumps, an after cooler consisting of a water-cooled finned tube is inserted in the outlet as close as possible to
the rotor. Because the forepump is generally much smaller, start-up of the mechanical booster is normally at about 20
mbar, but with the use of a spring/gravity loaded bypass valve or a fluid coupling [16] it is possible to start the
mechanical booster at atmospheric pressure. In the case of a fluid coupling which varies the rotational speed of the
rotors, assistance is given during the roughing cycle and no overheating occurs when pumping a large volume from
atmospheric pressure.

System contamination is still possible from either the forepump or the gear box oil which is pumped via the outlet of the
mechanical booster [17]. This latter source of contamination is often overlooked. If an oil-free forepump is used to back
a mechanical booster, then to largely eliminate system contamination the gear box should be pumped separately.

The mechanical booster with a forepump provides a group capable of high pumping speed in the range 5 to 102 mbar.
In conjunction with a water ring pump it can handle fine powders, although coarse powders can cause wear.

Multistage roots pumps normally have three, four, or five stages of Roots mechanisms in series and are driven off the
same shaft. In the case of three-stage pumps it is normal to back them with a further three-stage pump, thereby giving
six in total.

Rotational speed is normally that of a two-pole electric motorthat is, 2850 rev/min (50 Hz) or 3440 rev/min (60 Hz).

Stage ratios are important to ensure efficient running at the normal inlet pressures. The inlet stage is the largest and the
outlet stage the smallest. A typical speed curve for a five-stage pump is shown in Fig. 3.9 with the typical power
requirements. The power starts low and then reaches a level which remains constant as the pressure drops.

< previous page page_161 next page >


< previous page page_162 next page >
Page 162

Overheating is a potential problem in the high-pressure stages when the pump is working at low pressure, and for this
reason the higher pressure stages are frequently fitted with water-cooled heat exchanges in their outlet port. Gas ballast
or purging is often used on a number of the stages to ensure that vapors being pumped do not condense. Frequently
there are gas purges at the shaft seals to keep them clean so as to improve their reliability.

System contamination is minimal because the gear box, which is at atmospheric pressure, is adjacent to the outlet of the
pump, or in the case of six-stage pumps it is at least three stages from the pump inlet. The bearings at the inlet stage of
the pump are normally the only source of contamination and are generally packed with a very-low-pressure
perfluoropolyether grease.

Dry pumping group controls often include thermocouples placed at critical points, outlet pressure and motor current
monitor, water flow switch, and so on. Frequently all these functions are monitored via a computer-compatible
controller so that a large number of units may be run on one in-house computer.

Dry pump outlet management [18] is an important factor in systems where aggressive toxic or explosive gases are being
pumped. The use of water scrubbers on their own is not generally advisable because water vapor and water can enter the
pump and, by combining with the pumped gases, can cause extensive damage. With the many gases used in the
semiconductor industry [e.g., halogen compounds (i.e., fluorine, chlorine, and bromine)], it is now becoming more
common to react and inert the pumped gases by the use of a number of different technologies, some of which are listed
below:

Burning/Oxidation: This needs to be combined with an absorber or wet scrubber to remove the by-products. It should
not be used for chlorinated materials due to risk of the formation of toxic by-products (e.g., dioxins).

Adsorption/Chemisorption: Possible explosive risk from adsorbed gases and hazardous waste disposal issues.

Hot Bed Reactors: Gases are passed into a heated granular bed, forming solid by-products, or are catalytically converted
by the bed into other gases or solids. By-products are nonhazardous or totally recyclable.

Application: Pumping of highly chemically active gases and the handling of fine dustthat is, semiconductor manufacture
plus where system contamination is a concern.

3.3.2
Claw Pump

Multistage Claw

Capacities Available: 25 to 500 m3/h (also including pumps with inlet Roots stage increasing to 1200 m3/h when used
with a mechanical booster)

Operating Pressure Range: 1000 to 104 mbar

With Additional Mechanical Booster: 1000 to 7 × 104 mbar

A typical claw mechanism is illustrated in Fig. 3.10. It consists of two claw-shaped rotors. These are synchronized by
external gears and rotate in opposite directions

< previous page page_162 next page >


< previous page page_163 next page >
Page 163

Fig. 3.10
The claw mechanism and its operating cycle.

within the stator. The gears and rotor bearings are oil-lubricated but are external to the pumping chamber, so the swept
volume is dry. The clearance between the rotors and the stator wall is normally 0.1 to 0.2 mm. As the rotors run dry
back leakage occurs through these clearances. The shape of the rotors and of the inlet and outlet ports is designed to act
as a valve, thereby almost eliminating back flow. This results in a much higher compression ratio at high pressure than
can be achieved with a Roots pump mechanism as indicated in Fig. 3.11. A further advantage is that the heat generation
between stages and at the outlet is much smaller than with the Roots-type mechanism, and interstage cooling on the
higher pressure stages is not required. The power consumption is also lower when the pump is operating at low
pressures.
Multistage claw pumps [19, 20] normally have three or four claw stages, although in some cases the inlet stage is of the
Roots design. The advantage of using a Roots

< previous page page_163 next page >


< previous page page_164 next page >
Page 164

Fig. 3.11
Attainable pressure ratio (for air) with a single claw and a single Roots
mechanism as a function of outlet pressure with no gas throughput.

inlet stage is that the inlet to a claw pump is more restrictive at low pressure than a Roots stage, so the use of the latter as an
inlet stage improves the speed of the pump at low pressures where the compression ratio of the Roots stage is somewhat higher
than that of a claw-type stage.

Rotational speed is normally that of a two-pole motorthat is, 2850 (50 Hz) or 3440 (60 Hz).

A typical speed curve of a multistage claw pump is shown in Fig. 3.12, along with a curve of its power consumption.

As with most other dry pumps, system contamination is minimal because the drive gears and oil box are at the atmospheric end
of the pump [21].

For gas ballast, system contamination, pumping group controls, pump outlet management, and applications, see previous
section.

3.3.3
Screw Pump

Capacities Available: 24 to 2700 m3/h


Operating Pressure Range: 1000 to 5 × 103 mbar

The mechanism of a screw pump consists of two intermeshing screw rotors enclosed in a close-fitting stator. There is a very
small clearance between the screw form of the two rotors, the rotors, and the stator wall; the stator has specially shaped ports for
the inlet and outlet. The rotors are normally coated to improve resistance to chemical attack and also to reduce the clearances
within the pumps to a minimum.

Two forms of screw are currently used; one type is of a rectangular form (see Fig. 3.13) and runs at about 10,000 rev/min, and
the other type is illustrated in Fig. 3.14 and runs at two-pole motor speed. Figure 3.15 shows a trapped volume of gas and its
transfer from the inlet to the outlet which is at atmospheric pressure. A typical speed curve is shown in Fig. 3.16 which also
indicates the power requirements.

< previous page page_164 next page >


< previous page page_165 next page >
Page 165

Fig. 3.12
Speed curve and power consumption of a four-stage
claw pump at 50 Hz. The ultimate vacuum improves when
the pump is run at 60 Hz and speed curve increases
by about 20%. The speed and power
curves are without gas purge.

Fig. 3.13
Screw pump with a square thread form.

This type of pump is particularly suitable for applications where slugs of liquid from the process vessel are likely to
reach the pump.
For gas ballast (or purging), system contamination, pumping group controls, pump outlet management, and applications,
see Section 3.3.1 on the Roots pump.

< previous page page_165 next page >


< previous page page_166 next page >
Page 166

Fig. 3.14
Alternative screw profile which is normally run at two-pole motor speed.
Fig. 3.15
Transport of gas by the square thread mechanism.

< previous page page_166 next page >


< previous page page_167 next page >
Page 167

Fig. 3.16
Speed curve and power consumption of a typical pump using a
screw form similar to that in Fig. 3.13 and without gas purge.
Obtainable pressure of pump type with the different screw form
(Fig. 3.14) and using two pole motor is about one decade higher.

3.3.4
Scroll Pump

Capacities Available: 20 to 50 m3/h


Operating Pressure Range: 1000 to 102 mbar

The two main components of this type of pump is a stationary scroll and an identical moving scroll. The moving scroll is
mounted on a spigot, or spigots, so that the rotation of the shaft produces an orbital motion of this member. The moving
scroll is constrained from rotating by an arrangement which ensures that it maintains the same angular position during the
orbiting motion. Such an arrangement is shown in Fig. 3.17, which indicates how gas is taken in at the inlet and transferred
to the outlet. The main problem in achieving a good vacuum is leakage across the edge of the moving scroll and the
stationary scroll. A spring-loaded gasket is normally fitted to ensure good sealing along the moving scroll to overcome this
problem [22]. However, this gasket can lead to the formation of a small amount of dust. So it is important to ensure that this
is not at any stage blown back into the system.

A typical speed curve and power consumption of a pump is shown in Fig. 3.18.

This pump is generally not suitable for severe or chemically active gas pumping applications but gives almost no
contamination when pumping clean systems or if it is used as a forepump (e.g., with a turbomolecular pump).

A special version of the pump has been used for many years, mainly in the nuclear industry. In this version the control of
the motion is made completely external to the system by the use of metal bellows. This ensures that there are no organic
components in the system that could be contaminated by the radioactive materials being pumped. Pumps of this type are
available in capacities ranging from 15 to 600 m3/h and with ultimate vacuum of 5 × 102 to 103 mbar.

< previous page page_167 next page >


< previous page page_168 next page >
Page 168

Fig. 3.17
Transport of gas in a scroll mechanism.
Fig. 3.18
Speed curve and power consumption of a 30-m3 · h1
scroll pump without gas purge.

< previous page page_168 next page >


< previous page page_169 next page >
Page 169

3.3.5
Piston and Diaphragm Pumps

Multistage Piston Pumps


Capacities Available: 12 m3/h
Operating Pressure Range: 1000 to 3 × 102 mbar

Fig. 3.19
Three-stage piston pump.

Two-Stage Diaphragm Pumps


Capacities Available: 0.8 to 5 m3/h
Operating Pressure Range: 1000 to 2 × 102 mbar
Fig. 3.20
Speed curve of a two-stage diaphragm pump.

< previous page page_169 next page >


< previous page page_170 next page >
Page 170

These two types of pump are basically of similar design. In the case of the diaphragm pump the diaphragm is used to
seal the piston while the piston in the piston pump either is a close fit in the piston housing or uses piston rings. This
results in the diaphragm design being more applicable to small capacities while the normal piston is used for the larger
sizes.

In vacuum applications the piston sealing is often obtained by a long path with a close clearance. The inlet valve is a
slot in the cylinder wall opened by the piston and the outlet valve is pushed open by the piston [23]. A three-stage pump
[24] is shown in Fig. 3.19. Three or four stages are used to achieve a vacuum of less than 0.1 mbar.

The diaphragm pump used in vacuum is normally a two-stage pump. The inlet and outlet valve are opened and sealed
by the diaphragm. The use of a molded diaphragm with strengthening ribs has greatly improved the reliability of this
component.

Gas ballast, pumping group controls, and pump outlet arrangements are not normally required. The main application of
the piston pump is for use on reasonably clean systems or as a forepump for a turbo pump. The diaphragm pump is
becoming more widely used, outside its normal laboratory application, as a forepump for a turbomolecular drag pump.

A typical speed curve for a two-stage pump is shown in Fig. 3.20.

References

1. F. A. Flecken, Gaede's influence on the development of mechanical vacuum pumps. Vacuum 13, 583 (1963).

2. W. Gaede, Gas Ballast Pumpen. Z. Naturforsch A. 2A, 233 (1947).

3. B. D. Power and R. A. Kenna, Vapour pumping characteristics of gas ballast pumps. Vacuum 5, 35 (1955).

4. L. Laurenson, Technology and application of pumping fluids. J. Vac. Sci. Technol. 20(4), 989 (1982).

5. L. Laurenson and G. Caporiccio, PerfluoropolyethersUniversal vacuum fluid. Proc. Int. Vac. Congr., 7th, 1977, Vol.
1, p. 263 (1977).

6. N. S. Harris and L. Budgen, Design and manufacture of modern mechanical vacuum pumps. Vacuum 26(12), 525
(1976).

7. D. Balfour, L. Budgen, P. Connock and D. Phillips, Pumping systems for corrosive and dirty duties. Vacuum 34, 771
(1984).

8. U. Segebrecht, Liquid Ring Vacuum Pumps and Liquid Ring Compressors, 1st ed. Expert-Verlag, Renningen-
Malmsheim, 1994.

9. H. Bannwarth, Liquid Ring Vacuum Pumps, Compressors and Installations, 2nd ed. VCH-Verlagsges., Weinheim.

10. H. E. Adams, Thermodynamic Characteristics of Nash Compressors. Nash Engineering Company, South Norwalk,
CT, 1953.

11. M. Wutz, H. Adam and W. Walcher, Theory and Practice of Vacuum Technology. Vieweg Verlagsges.,
Braunschweig, 1989.

12. J. L. Ryans and D. L. Roper, Process Vacuum System Design and Operation. McGraw-Hill, New York, 1986.

13. M. Bussard, La Technologie des pompes à vide système roots. Le Vide 12, 425 (1957).
< previous page page_170 next page >
< previous page page_171 next page >
Page 171

14. W. Armbatruster and A. Lorenz, Die Kombination Roots PumpeWasserring Pumpe. Vak. Tech. 7(2), 85 (1958).

15. C. M. Van Atta, Theory and performance characteristics of a positive displacement rotary compressor as a
mechanical booster pump. Trans. Natl. Symp. Vac. 3, 62 (1957).

16. H. Wycliffe and A. Salmon, The application of hydrokentic drives to high vacuum mechanical boosters (Roots
pumps). Vacuum 21, 223 (1971).

17. N. T. M. Dennis, L. J. Budgen and L. Laurenson, Mechanical boosters on clean and corrosive applications. J. Vac.
Sci. Technol. 18, 1030 (1981).

18. P. Mawle, Exhaust management. Eur. Semicond. May, p. 17 (1995).

19. H. Wycliffe, U.S. Pat. 4,504,201 (1985).

20. H. Wycliffe, Mechanical high vacuum pumps with an oil-free swept volume. J. Vac. Sci. Technol. A 5, 2608 (1987).

21. W. Wong, L. Laurenson, R. G. Livesey and A. P. Troup, An evaluation of the composition of the residual
atmosphere above a commercial dry pump. J. Vac. Sci. Technol. A 6, 1183 (1988).

22. D. Arthur, Little Inc., U.S. Pat. 1,507,254 (1976).

23. M. H. Hablanian, E. Bez and J. L. Farrant, Elimination of backstreaming from mechanical vacuum pumps. J. Vac.
Sci. Technol. A 5(4), 2612 (1987).

24. E. Bez, A compact oil-free rough vacuum pump. Solid State Technol., February (1995).

General References

L. Cummings, Liquid ring vacuum pumps optimise power generation. World pumps, January pp. 1415 (1987).

P. Duval, Will tomorrow's high vacuum pumps be universal or highly specialized. J. Vac. Sci. Technol. A 5(4), 2548
(1987).

R. G. P. Kusay, Vacuum equipment for chemical processes. Br. Chem. Eng. 16, 29 (1971).

J. F. Lennon, Vacuum process in drying of power transformer insulation. J. Vac. Sci. Technol., 19, April (1982); Proc.
Natl. Symp. of Am. Vac. Soc., Pt. 2, p. 1039 (1981).

P. A. O'Neill, Industrial Compressors: Theory & Equipment. Butterworth-Heinemann, London, 1993.

B. D. Power, High Vacuum Pumping Equipment. Chapman & Hall, London, 1966.

W. F. Ravette, Paper machine vacuum systems. TAPPI Wet End Oper., Semin. Notes, Seattle, WA, 1980, p. 97, TAPPI
Press, Atlanta, GA, 1980.

G. F. Smith, Machine vacuum systemhow to get the most out of this versatile papermaker's tool. TAPPI Eng. Conf.,
Pap., Houston, TX., 1976, Book 1, p. 243, TAPPI Press, Atlanta, GA, 1976.

A. P. Troup and N. T. M. Dennis, Six years of dry pumpinga review of experiences and issues. J. Vac. Sci. Technol. A 9
(3), 20482052 (1991).

M. Wutz, H. Adam and W. Walcher, Theory & Practice of Vacuum Technology. Vieweg Verlagsges., Braunschweig,
1989.
< previous page page_171 next page >
< previous page page_173 next page >
Page 173

4
Kinetic Vacuum Pumps

A kinetic vacuum pump is a pump that imparts momentum to the gas being pumped in such a way that the gas is
transferred continuously from the inlet of the pump to the outlet. This distinguishes it from the positive displacement
vacuum pumps described in Chapter 3 in which a volume filled with the gas being pumped is cyclically isolated from
the pump inlet and then transferred to the outlet. In the capture vacuum pumps described in Chapter 5 the gas molecules
being pumped are trapped on internal surfaces within the pump by sorption or condensation.

In the kinetic vacuum, pump momentum may be transferred to the gas being pumped in the direction of the pump outlet
by mechanical moving parts or by entrainment in a high-speed vapor stream. Both types of pumps are described in this
chapter.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1988 John Wiley & Sons, Inc.

< previous page page_173 next page >


< previous page page_175 next page >
Page 175

Part I
Diffusion and Diffusion-Ejector Pumps

Benjamin B. Dayton

When a liquid such as water, mercury, or petroleum fluid is vaporized in a boiler and the vapor is conducted through a
diverging nozzle exiting into an evacuated chamber, the vapor expands through the nozzle and continues to expand
beyond the nozzle exit, acquiring a high forward mass velocity as the random molecular motion is converted into a
directed stream of molecules known as a vapor jet. If some gas is present in the chamber, the expansion beyond the
nozzle exit is limited to a certain extent and a boundary region is formed where gas mixes with the vapor stream. This
boundary region may take the form of a turbulent layer of mixed gas and vapor when the pressure is high enough, or at
lower pressures it may be a diffuse layer of vapor mixed with penetrating gas molecules moving in laminar flow. In
either case the entrained gas is moved forward in the direction of the vapor jet; and if some means can be found to
separate the gas and vapor without allowing the gas to find its way back into the region near the nozzle, then a pumping
action has occurred.

One means of separating the entrained gas from the vapor stream is to design the pump housing or chamber wall so that
the cross section narrows gradually to a throat in front of the vapor jet and then expands rapidly beyond this throat. Such
a structure is called a Venturi tube or ''diffuser," although no gas diffusion is necessarily involved. If the cross-sectional
area of the throat of the diffuser is about equal to the exit area of the nozzle and the gas pressure beyond the diffuser in
the region known as the "fore-vacuum" is not too high, the vapor stream will converge and accelerate through the
diffuser throat at such a density and mass velocity that gas molecules cannot

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1988 John Wiley & Sons, Inc.

< previous page page_175 next page >


< previous page page_176 next page >
Page 176

readily penetrate back through the diffuser throat. The vapor expanding beyond the throat is cooled and condensed,
thereby releasing the entrained gas which must then be discharged into another (backing) pump or into the atmosphere.
This is the principle of action of vapor-jet ejector pumps. While multistage steam-jet ejector pumping systems [1] have
been used in the past to produce pressure as low as 7 Pa while discharging the gas directly into the atmosphere, they are
seldom used today because of the availability of mechanical booster or blower pumps of various sizes.

Another method of separating the entrained gas from the vapor stream which can be employed when the gas pressure in
the fore-vacuum is sufficiently low and the vapor is readily condensable, as in the case of mercury or oil vapor, is to
cool the pump housing walls in the region where the main vapor stream encounters the wall. The cross section of the
pump at this point does not necessarily have to be about equal to the nozzle exit area since at low pressures the vapor
expands freely and fills the pump cross section at the level where condensation occurs. Since the vapor must move a
certain distance along the pump axis before it is entirely condensed, a vapor barrier is formed by forward-moving vapor
molecules through which only a few gas molecules can diffuse back from the fore-vacuum. This back diffusion is not
always negligible and may limit the performance of the pump when the gas molecules have a low molecular weight and
collision cross section, as for hydrogen and helium. Vapor pumps employing a cooled casing and low-vapor-pressure
working fluids are termed "diffusion" or "condensation" pumps. They require backing pumps or ''forepumps" to produce
the necessary fore-vacuum and can only operate when the gas pressure in the chamber to be pumped has already been
reduced to less than about 70 Pa.

Ejector pumps will not be considered here, but some hybrid pumps are still in use which combine oilvapor ejector
stages with diffusion pumping stages. The latter are known as "diffusion-ejector" or "vapor booster" pumps and are
described briefly in Section 4.2.

4.1
Diffusion Pumps

4.1.1
History of Development

The invention of the diffusion pump arose out of studies made by W. Gaede [2] in Germany in 1913 on the counterflow
of mercury vapor and air in a vacuum system pumped by his rotary mercury-sealed mechanical pump with a cold trap to
condense mercury vapor. During this study to determine the lowest partial pressure of gas that could be produced
beyond the trap, he discovered the effect of the counterdiffusion of air and mercury vapor on the reading of a McCleod
gauge and conceived of the pumping effect which might be obtained by allowing gas to diffuse through a narrow slit
into a rapidly moving stream of mercury vapor. Figure 4.1 shows his first embodiment of this concept. Note that the slit
region is surrounded by a water-cooling jacket to condense mercury vapor escaping through the slit.

In 1916 Irving Langmuir [3] showed that the slit or diffusion diaphragm through which the gas was pumped in the
Gaede pump [4] could be widened if the mercury vapor was directed away from the slit by a suitable nozzle as shown in
Fig. 4.2. He emphasized the importance of adequate condensation of the mercury vapor and referred to his pumps as
condensation pumps. The pumping speed was thereby greatly increased, and Langmuir was granted U.S. Patent
1,393,550 in 1921.

< previous page page_176 next page >


< previous page page_177 next page >
Page 177
Fig. 4.1
Gaede's slit-jet mercury vapor pump. a, return
tube; b, steel cylinder with slit e; c, condenser;
d, mercury-filled gutter; f, gas inlet; f1, connection
to roughing pump; g, connection to fore-pump;
h, thermometer; V, mercury cut-off valve; Q, boiler.

In 1916 Langmuir also applied for a patent on the inverted or "mushroom cap" nozzle design which was issued in 1919
as U.S. Patent 1,320,874 as shown in Fig. 4.3.

The detailed history of the development of mercury diffusion pumps is given in earlier editions and will not be repeated
here. Mercury vapor pumps became widely

< previous page page_177 next page >


< previous page page_178 next page >
Page 178

Fig. 4.2
Langmuir's mercury condensation pump with cylindrical nozzle.

used in the electronic tube industry in the United States from 1920 to 1940, but they required the use of refrigerated
traps to keep the vapor out of the tubes (except for mercury rectifier tubes). In 1928 in connection with experiments on
the vacuum impregnation of pressboard with transformer oils to improve the dielectric strength of insulators, C. R.
Burch [5] at Metropolitan-Vickers in England became interested in high-vacuum distillation at moderate temperatures,
known as "molecular distillation." He succeeded in obtaining some very-low-vapor-pressure fractions from the
petroleum oil used in rotary mechanical pumps. It then occurred to him to try these oil fractions as the pump fluid in a
diffusion pump in place of mercury. He was able to obtain a pressure reading on an ionization gauge of 104 Pa without
a cold trap and thus began a revolution in the application of diffusion pumps (British patent, 303,078).

After visiting England, Dr. Kenneth Mees, Research Director of the Eastman Kodak Laboratories, brought the work of
Burch to the attention of Dr. K. C. D. Hickman, who had been working with low-vapor-pressure synthetic organic esters
in place of mercury in special manometers for measuring the pressure in apparatus for
< previous page page_178 next page >
< previous page page_179 next page >
Page 179

Fig. 4.3
Langmuir's metal condensation pump
with umbrella nozzle.

the drying of photographic film where mercury vapor was harmful. In 1929 Dr. Hickman found that these esters gave
good results when used as the pump fluid in small glass diffusion pumps of simplified design which he constructed
himself [6]. He applied for a patent on the use of butyl phthalate, and similar esters, in place of mercury in these pumps
(U.S. patent 1,857,506 issued in 1932). Many articles were published in the period from 1932 to 1940 on the conversion
of mercury vapor pumps to oil diffusion pumps and the use of baffles and charcoal traps to reduce oil vapor
contamination.

During the period from 1929 to 1937 the factors involved in designing diffusion pumps for use with the new synthetic
oils were studied by Dr. Hickman and his co-workers in the Research Laboratories of the Eastman Kodak Company,
resulting in the development of multiboiler "self-fractionating" oil diffusion pumps [7] which extended the lowest
pressure attainable without cold traps from 103 Pa to about 5 × 106 Pa. Figure 4.4 shows a water-cooled three-stage
fractionating pump constructed of Pyrex glass with coiled nichrome heater wire in each of the boilers which in use are
covered with glass-wool-lined insulation bags. The total length is 25 in. Two alembic heads, or catchment lobes, are
provided in the tubing leading to the forepump connection for the purpose of retaining the most volatile fluids distilled
from the boiler.
< previous page page_179 next page >
< previous page page_180 next page >
Page 180

Fig. 4.4
Hickman's three-stage glass oil fractionating diffusion pump.
Fig. 4.5
Vertical multistage oil fractionating
pump design of Malter and Markuvitz [8].

< previous page page_180 next page >


< previous page page_181 next page >
Page 181

In 1937 L. Malter [8] of the Radio Corporation of America applied for a patent on an all-metal vertical multistage oil
diffusion pump, as shown in Fig. 4.5, which incorporated the Hickman self-fractionating principle by having separate
boiler compartments feeding vapor through concentric cylindrical tubes to the various nozzles with small openings
between compartments at the bottom of the cylinders so that returning condensed oil first enters the outer compartment
feeding vapor to the lowest stage where the more volatile components are removed and the purged oil then enters the
next compartment, where more of the volatile constituents are removed, and finally passes to the innermost
compartment feeding vapor to the nozzle nearest the pump inlet, so that the oil condensed on the walls near this top
stage has a lower vapor pressure than the original unpurified oil mixture. An alembic head is also shown in the forearm
to catch the more volatile components of the pump fluid.

Much of the history of diffusion pumps involves investigation of various synthetic oils for use as diffusion pump fluids
and attempts to minimize the backstreaming of the vapor of these oils from the first stage into the vacuum chamber.

4.1.2
Diffusion Pump Design

The first practical diffusion pumps were single-stage with a boiler at the bottom and a vapor conduit leading to a
cylindrical or expanding conical nozzle, or an inverted (umbrella-type) nozzle to form a high-speed vapor jet. In early
pumps the vapor jet was directed vertically upward into a condensing region with an alembic at the bottom to catch the
distilled mercury or oil and return the fluid to the boiler. Later designs employed nozzles which directed the jet slightly
downward, or vertically downward, so that the condensed pump fluid ran down the pump wall and returned to the boiler
through a connecting tube or through a narrow passage between the pump wall and a skirt on the vapor conduit so that
the head of oil prevented the escape of vapor from inside the conduit. Single-stage pumps with parallel multiple nozzles
were tried and gave improved speed and required less power input, but multiple nozzle designs are not practical in
multistage pumps.

Multiple in-line stages were introduced when it was realized that the compression ratio between forepressure and
maximum inlet pressure before jet breakdown had to be limited for each stage of an oil diffusion pump because of the
decomposition of organic pump fluids at elevated temperatures and pressures. Since the pumping speed of any stage
depends on the nozzle clearance area (the annular gap between the rim of the nozzle exit and the wall of the pump
housing) while the forepressure at which the jet breaks down depends on the boiler pressure (or temperature) and the
ratio of the nozzle throat area (narrowest cross section) to the "body clearance area" (the annular gap between the wall
of the pump housing and the vapor conduit leading from the boiler to the nozzle), by arranging a series of nozzles along
a common axis with stepped increases in the nozzle clearance area and body clearance area from the stage nearest the
primary pump to the stage nearest the vessel to be pumped, the overall speed of the pump could be maximized while
still maintaining operation against relatively high forepressures as produced by the primary pump without requiring an
excessively high boiler temperature.

The design of the first stage (nearest the pump inlet) was particularly important since some vapor from this stage
scatters backward into the high vacuum and influences the speed and ultimate pressure obtainable. In 1937, Embree [9,
10] applied

< previous page page_181 next page >


< previous page page_182 next page >
Page 182

Fig. 4.6
Design of Embree jet (b) compared with straight-sided jet (a) [10].

for a patent on a streamlined inverted nozzle design (as shown in Fig. 4.6b) which greatly improved the pumping speed
over prior designs of the type shown in Fig. 4.6a.

Since for multistage pumps the speed depends on the nozzle clearance area of the first stage (nearest the inlet), one
method of increasing the speed without changing the size of the flange connection to the vacuum chamber or the pump
length is to widen the pump housing in the neighborhood of the first- and second-stage nozzles (as shown in Fig. 4.7) of
a pump designed by Hablanian and Maliakal [11]. Also shown in Fig. 4.7 is a cowl or cap, cooled by conduction
through the supports mounted on the water-cooled pump casing, placed over the first-stage umbrella nozzle which
serves to intercept and condense vapor scattered backwards from the outer edge of the jet or evaporating from the rim of
the hot umbrella nozzle. This "cold cap," which may also be cooled directly by a loop of water-cooled copper tubing,
invented by Power [12] of the Edwards High Vacuum Ltd., greatly reduces the back-streaming of pump fluid into the
high-vacuum region.

A similar design, including a built-in butterfly valve at the inlet, is shown in Fig. 4.8. Both Figs. 4.7 and 4.8 show a
boiler filler and drain pipe at the side of the boiler and last-stage nozzle directing the vapor jet into a water cooled side-
arm in the form of an ejector diffuser. Water cooling of metal pumps is preferably through copper tubing wrapped
around the casing and brazed to the wall since double-wall cooling jackets as shown in Fig. 4.5 are less efficient and
subject to corrosion. Coiled tubing may also be

< previous page page_182 next page >


< previous page page_183 next page >
Page 183

Fig. 4.7
Metal multistage pump with enlarged nozzle clearance
area and cold cap [Varian Vacuum Products].

placed around the boiler region through which water can be circulated during the cool-down of a pump on a system
without a high-vacuum valve.

Proper design of the pump boiler region involves several details, such as (1) the size and spacing of boiler
compartments for fractionating pumps, (2) having a "skirt" at the lower end of the outer vapor conduit close to the pump
wall and of sufficient height to maintain a head of pump fluid (from the condensed fluid returning down the wall) to
balance the vapor pressure inside the boiler, and (3) the method of conducting heat from the electric heating units to the
oil in the boiler in such a manner that superheating and "bumping" is reduced.

4.2
Diffusion-Ejector Pumps

Before the development of efficient mechanical booster (rotary blower) pumps with adequate pumping speed down to
10 Pa, there was a problem in providing an adequate fore-vacuum for the diffusion pumps used on large molecular stills.
K. C. D. Hickman decided to develop a booster pump for the 5-Pa to 200-Pa pumping range based on the design of
steam ejectors but using a stable hydrocarbon oil at high boiler pressures instead of steam. This resulted in the
production of an oilvapor ejector or 'kerosene booster" pump which was then used on the molecular stills and in the

< previous page page_183 next page >


< previous page page_184 next page >
Page 184

Fig. 4.8
Cross section of three-stage oil-vapor diffusion pump
with integral water cooled baffle and high-vacuum
valve [Edwards High Vacuum International].
drying of frozen foods and blood plasma. Hickman was unable to obtain a patent on the basic idea of an oil-ejector
because of a prior invention by W. K. Lewis of an oil-ejector pump for the petroleum industry. However, U.S. patent
2,279,436 was issued to K. C. D. Hickman and G. Kuipers for an improved oil-ejector design, and a series of single-
stage and multistage oil-ejector pumps was developed and applied to use in the vacuum metallurgy industry. The
development of mechanical boosters (of the RootsConnersville blower type with rotating lobes) eventually obviated the
need for these oil-ejector pumps. However, hybrid ejector-diffusion pumps with the high-pressure stage in the form of
an oil-ejector were developed in the United States and in England and found use in various industrial applications where
the gas load involved relatively high pressures in the fore-vacuum produced by mechanical pumps.

< previous page page_184 next page >


< previous page page_185 next page >
Page 185

Fig. 4.9
Diffusion-ejector pump [Edwards
High Vacuum International].

Figure 4.9 is a diagram of such a diffusion-ejector pump [13] with a 38.7-cm-i.d. inlet which can operate against a 200-
Pa backing pressure and which has a maximum pumping speed of 4000 liter·s1 at inlet pressures up to 1 Pa using
Apiezon A201 booster pump fluid and 6.0-kW heater input. The boiler pressure can be as high as 4000 Pa.

Theoretical and experimental studies of the ejector stages of vapor pumps were performed by Jaeckel, Nöller, and
Kutscher [14], by Kutscher [15], and by Bulgakova et al. [16].

4.3
Performance of Vapor-Jet Pumps

4.3.1
Pumping Speed

The rate at which gas flows across the inlet of a vapor-jet pump can be expressed in at least four ways: (1) the number
of molecules crossing the inlet in unit time, (2) the mass of gas in grams per unit time flowing across the inlet, (3) the
volume of gas in volumetric units (liters, m3, cubic feet, etc.) flowing through the inlet in unit time at the prevailing
pressure and temperature at the inlet, and (4) the throughput (usually
< previous page page_185 next page >
< previous page page_186 next page >
Page 186

Fig. 4.10
Typical speedpressure curve of diffusion pump
as measured with a total pressure gauge.

represented by the letter Q) of gas across the inlet area in pressurevolume units per unit time at the prevailing gas
temperature. Historically, the precedent set by W. Gaede was to express pump performance in terms of the volumetric
flow rate or ''pumping speed" (usually represented by the letter S). The flow rates (1) and (2) are almost impossible to
measure directly. Flow rates (3) and (4) can easily be measured by procedures given in the next section, but differences
arise in the method of measuring the pressure (represented in the following by the letter p) depending on the type of
vacuum gauge used, the location and orientation of the gauge tubulation relative to the pump inlet, and the method of
connecting the pump to the test system. In any case, the speed at the cross section where the pressure is measured is
defined by

for a given gas at the specified temperature.

The equation of continuity for gas flow through the diffusion pump into the forepump is

where C is the "capacity" or volumetric speed of the forepump at the prevailing forepressure F.

Measurements of the speed of diffusion pumps using the test dome recommended by the American Vacuum Society
[17] (or by the equivalent ISO standard), along with a total pressure gauge, show that the speed varies with the intake
pressure as shown in Fig. 4.10. The speedpressure curve can be divided into three regions or zones of pressure: (1) the
declining speed in the "jet breakdown region," or forepressure breakdown region, from the breakdown value of the
intake pressure, pb, (corresponding to a forepressure F equal to the limiting forepressure, or breakdown forepressure,
Fb), to the crossover pressure, pc, where the inlet pressure equals the forepressure Fc, (2) the region of nearly constant
maximum speed, Sm, from pb to pd, the lower limit of pressure at which the speed is near Sm, and (3) the declining
speed in the ultimate pressure region from pd to the ultimate pressure, pu, at which the speed S becomes zero due to
back migration of pump fluid vapor and fragments of decomposed pump fluid, and back-diffusion of gas in the fore-
vacuum through the vapor jets.

< previous page page_186 next page >


< previous page page_187 next page >
Page 187

When the forepressure is maintained at a sufficiently low value to eliminate appreciable back-diffusion of gas from the
forevacuum and the outgassing of the measuring dome is negligible, the ultimate pressure, pu, is determined mainly by
the vapor pressure of the pump fluid on the walls near the pump inlet or on a cooled baffle or trap between the pump
and the test dome. In the region from pb to pu the pump speed is given approximately by

with apparent deviations from this formula during measurement being due to the varying response of the pressure gauge
(usually a tubulated ionization gauge) to the changing mixture of air and pump fluid vapor and the change in the
outgassing rate of the ion gauge itself.

In the region from pc to pb, when no gas is leaked into the fore-vacuum while the intake pressure p is increased by
admitting gas to the test dome, the pump speed is given approximately by

where C is the volumetric speed of the forepump at the prevailing forepressure and can be assumed to be nearly
constant over the forepressures prevailing during jet breakdown.

4.3.2
Limiting Forepressure for Maximum Speed

The limiting forepressure for maximum speed depends on the amount of gas admitted through the pump inlet and the
speed of the forepump. It is convenient to distinguish between (a) static breakdown, which occurs when the forepressure
has been raised by throttling the forepump at a given gas load or by admitting gas to the fore-vacuum until the inlet
pressure (high vacuum) has risen by about 2 × 102 Pa, and (b) dynamic breakdown, which occurs when gas is admitted
only through the pump inlet until the forepressure maintained by a given forepump has risen to the limiting value, Fb,
marked by a noticeable decrease in the pump speed (knee of the pump speed curve).

When the forepressure, F, is increased by opening an adjustable leak in the fore-vacuum while the high-vacuum
pressure, p, is held constant by adjusting the rate of air admitted through a leak into the test dome, the variation of speed
as a function of forepressure for a fixed value of p can be obtained and plotted for different heater inputs. For a given
heater input the speed remains constant until the forepressure reaches the limiting value and then decreases rapidly. The
maximum speed varies with heater input because of the vapor-backstreaming factor α and the fore-vacuum gas back-
diffusion factor β as explained in Section 4.4.

When the heater input is constant and the throughput is also maintained constant by controlling the leak into the test
dome and reading the intake pressure, p, while the forepressure, F, is slowly raised by admitting gas into the fore-
vacuum, a curve of inlet pressure versus forepressure at constant throughput is obtained. Two extreme values

< previous page page_187 next page >


< previous page page_188 next page >
Page 188

Fig. 4.11
High vacuum versus forepressure with constant
throughput and constant heater input.

of throughput are of special interest: the no-load value, or Q = 0, and the maximum load value, Q = Qm. Figure 4.11
shows such curves for a two-stage pump.

For a single-stage oil-vapor diffusion pump, as explained later, it can be shown that as an approximation, when no gas is
leaked into the fore-vacuum while gas is admitted to the pump inlet until p = pb at which dynamic jet breakdown
begins, then

where At is the nozzle throat area, B is the body clearance area (the open cross section of the pump casing in front of
forward-directed nozzles, or area between the inside wall of the pump casing and the outside wall of the vapor conduit
leading to an inverted or "umbrella"-type nozzle), Kn is a constant with value in the range from 0.5 to 1.1 depending on
the design of the nozzle as explained later, and Pc is the vapor pressure inside the vapor conduit leading to the nozzle,
which is approximately equal to the vapor pressure in the boiler. Since, as shown in Section 4.3.6, the boiler pressure
increases more or less linearly with the watt input to the heater, we note that pc, Fc, pb, and the limiting forepressure,
Fb, all increase linearly with the watt input. The knee of the speed versus intake-pressure curve for a single stage pump
can be moved to higher pressures, pb, by increasing the speed C of the forepump.

The static forepressure breakdown of single-stage pumps is given by


where Kn is nearly independent of the gas load (throughput) and Pc is approximately equal to the boiler pressure, but
for multistage pumps the static limiting forepressure depends on the gas load. When curves of p versus F for a two-stage
pump are plotted

< previous page page_188 next page >


< previous page page_189 next page >
Page 189

Fig. 4.12
Intake pressure versus forepressure for each stage in
two-stage pump and combined stages at no load.

together with the corresponding curves for each stage separately, as in Fig. 4.12, using the fact that the forepressure for
the first (top) stage, F1, is the same as the intake pressure, p2, for the second stage while the inlet pressure, p, equals p1,
it can be seen from the figure that the no-load static limiting forepressure for a two-stage pump is given approximately
by

where F1b is the no-load limiting forepressure for the first stage alone, while F2b is the no-load limiting forepressure
and F2c is the crossover forepressure for the second stage alone. At the upper end of the forepressure breakdown curve
for any individual stage the crossover pressure is given by

where At/Ae is the ratio of nozzle throat area to nozzle exit area. Hence, from Eqs. (4.6) and (4.8) we obtain
< previous page page_189 next page >
< previous page page_190 next page >
Page 190

where A2e/B2 is the ratio of nozzle throat area to the body clearance area for the second stage. Usually, for the stage
nearest the forepump, Ae/B = 0.3 0.5. Thus, for a two-stage pump, Eq. (4.7) gives Fb = 0.6F1b + F2b.

Similarly, it can be shown that for a three-stage pump the no-load static limiting forepressure is

where F3b is the static limiting forepressure and F3c is the crossover forepressure for the third stage (nearest the
forepump) alone. Again, F3b/F3c = A3e/B3, the ratio of nozzle throat area to body clearance area for the third stage.

The compression ratio of a single pumping stage is defined as the ratio of forepressure to inlet pressure when pumping a
gas that is not absorbed within the pump. This ratio is a variable quantity which depends on the ratio of the pumping
speed of a given stage to the net speed of the next stage in series. However, for a given throughput there is a limiting
compression ratio which can be maintained at a given power input. At maximum throughput of air the limiting
compression ratio for single-stage ejector pumps is about 10, and for the individual stages of multistage diffusion pumps
the limiting ratio is about 4. Thus a four-stage diffusion pump can compress air from 0.2 Pa to 51 Pa, corresponding to a
total compression ratio of 44 or 256. For a single-stage diffusion pump the limiting compression ratio as obtained from
Eq. (4.5) will be

4.3.3
Influence of Nozzle and Entrance Chamber Design on Speed

In order to simulate the gas flow pattern into the pump from a pipe connecting the pump to the vessel to be evacuated,
pump speed is always measured using a test dome over the inlet with an inside diameter either equal to the inlet
diameter of the pump housing or else equal to the inside diameter of the pipe fitting which is normally used to connect
the pump to the vessel. The American Vacuum Society has established recommended configurations for the test dome
[17] as described in Chapter 12. The location and orientation of the tubulation of the total pressure gauge installed on
the dome influence the measured speed [1820]. The nozzle design of the first stage of pumping (nearest the high-
vacuum inlet) determines the pump speed. For this stage the exit of the nozzle should not be beveled to deflect the vapor
jet toward the pump wall, but rather should direct the jet along the axis of the pump as much as possible by allowing
some inward expansion as in the Embree nozzle design. The "nozzle clearance area," Anc, between the rim of the vapor
nozzle exit (or the rim of a cold cap over an umbrella nozzle) and the inside wall of the pump housing in the entrance
chamber region is the limiting factor determining pump speed. The theoretical maximum speed across this area
considered as an opening in a thin plate would be

< previous page page_190 next page >


< previous page page_191 next page >
Page 191

where fn is the "effusion law" factor in liter · s1 · cm2 as given by

in which R0 is the molar gas constant, T is the absolute temperature, and Mn is the molecular weight of the gas.
However, because the nozzle clearance area is the exit of a tube or chamber of varying cross section consisting of the
walls of the pump casing and the nozzle assembly between the nozzle clearance area, Anc, and the pump inlet area, Ap,
gas does not cross the area Anc according to the cosine law but is beamed forward across this area by scattering from
the walls of this entrance chamber. If gas at pressure p (in Pa) were to enter the pump inlet from a large chamber
directly connected to the pump so that the molecules crossed the inlet area with a cosine law distribution, then there
would be a transition probability w less than 1 that a molecule crossing the pump inlet would also cross the nozzle
clearance area into the vapor jet. The forward flow in Pa · liter · s1 across the pump inlet would then be wfnApp. For gas
at a partial pressure pj scattered back from the vapor jet and crossing the nozzle clearance area (with an approximate
cosine law distribution) and continuing on out of the pump inlet, the backward flow in Pa · liter · s1 out of the pump
would be wfnAncpj. When the back-diffusion of gas from the fore-vacuum is negligible, the pump speed as measured in
the large chamber would then be

Theoretically, w could be calculated from the known geometry of the entrance chamber by applying various formulas
due to Clausing [21], Oatley [22], Harries [23], and others, or w could be calculated by Monte Carlo methods, but pj
would not be known. However, it is not necessary to know w and pj exactly, but merely to recognize that the entrance
chamber impedance, Rc (in seconds per liter), to gas flow should be made as small as possible while the nozzle
clearance area of the first-stage nozzle should be as large as practical to obtain maximum speed. The maximum
measured speed, Sm, can then be expressed as

where H is the "speed factor" or "Ho coefficient" which has been defined by T. L. Ho [24] as

where α is Gaede's factor for the fraction of the gas molecules from the high vacuum that succeed in diffusing through
the back-scattered vapor and reaching the nozzle clearance area which depends on the ratio of the mean free path of gas
molecules through the back-scattered vapor in the nozzle clearance area and the width of the nozzle clearance, while γ is
the fraction of the gas molecules which reach the nozzle clearance area and pass on to be carried away by the vapor jet
and 1 γ is the fraction which return through the nozzle clearance area. It should be noted that for mercury vapor pumps
the vapor pressure of mercury in the entrance chamber is not negligible,

< previous page page_191 next page >


< previous page page_192 next page >
Page 192

and Ho considered the chamber impedance Rc to include a factor which increased with increasing vapor density. In
deriving Eq. (4.16), Ho used the net speed relation

which is not strictly correct unless the molecules cross the various sections involved according to the cosine law (i.e.,
with no beaming).

Gaede [2] applied kinetic theory to the interaction of gas and vapor molecules in the neighborhood of the entrance slit
(or an equivalent orifice in a thin plate separating the mercury vapor stream from the gas in the high vacuum) of his first
mercury vapor pump and derived an equation for the speed at the slit area as shown in Section 4.4.1.

Speed for Various Gases. As can be seen from Eq. (4.12) and Eq. (4.13), theoretically, the speed of a diffusion pump
varies inversely as the square root of the molecular weight of the gas flowing across the inlet. However, back-diffusion
of light gases, such as hydrogen and helium, from the fore-vacuum to the high vacuum may decrease the pumping speed
when the vapor stream does not have sufficient velocity and density to reduce this back diffusion to a negligible value.
Theoretical equations for this back-diffusion are given in Section 4.4.4.

4.3.4
Ultimate Pressure

The ultimate pressure of a diffusion pump without a cold trap, in the absence of leaks through the system walls, is
determined primarily by the vapor pressure of the pump fluid at the temperature of the walls of the baffle or connecting
tube between the first stage and the vacuum chamber. Since organic pump fluids are a mixture of molecules of varying
molecular weight, fractionating diffusion pumps were developed by Hickman [7] and by Malter and Markuvitz [8]
which circulate the pump fluid through a series of boiler compartments from a compartment feeding the last-stage
nozzle to the compartment feeding the first-stage nozzle (nearest the pump inlet) with means of separating the vapor
supply to the nozzles. The more volatile constituents tend to be vaporized first in each compartment so that the last
compartment feeding the first stage receives oil with the lowest vapor pressure. Coe [25] studied the fractionating
effectiveness of concentric cylinder vapor conduits forming separate compartments in the boiler by raising the nozzle
assembly by small distances from the base of the pump casing and also demonstrated the value of a catchment lobe in
the fore-vacuum line to prevent volatile fractions from returning to the boiler. The variation of ultimate pressure with
ambient temperature for a fractionating pump with Octoil S, DC 703, and Apiezon C was also measured by Coe and by
Latham, Power, and Dennis [26].

When the forepressure is not sufficiently low, back-diffusion of light gases from the fore-vacuum may also influence
the ultimate pressure as explained in Section 4.4.4.

Vapor Pressure of Pump Fluids. In the first diffusion pumps, mercury was employed as the working fluid. The vapor
pressure of clean mercury at 25°C is about 0.27 Pa, and at 196°C the vapor pressure of liquid nitrogen is below 1031 Pa.

< previous page page_192 next page >


< previous page page_193 next page >
Page 193

Gallium and fusible metal alloys have been suggested as pump fluids, but are not considered practical. P. Alexander [26.
a] found glycerol as the pump fluid to give advantages over mercury vapor pumps, but glycerol has a vapor pressure of
about 102 Pa at room temperature.

Table 4.1 gives vapor-pressure data for oils used as pump fluids as obtained by various authors using tensimeters over
the pressure range from 0.1 to 1000 Pa. Since most commercially available pump fluids are mixtures of organic
molecules having a range of molecular weights and isomeric forms, the measurement of the vapor pressure is
complicated by the tendency of fractionation to occur during the measuring process. Therefore it is not surprising that
different methods give slightly different results and that the data reported by different authors for the same material are
not in close agreement as shown by this table. Parameters A and B are those of the vapor-pressure relation

where P is the vapor pressure in Pa and Tc is the temperature in °C. This equation is a good approximation to most
vapor-pressure curves over the range of temperatures involved in diffusion pumps, but plots of log10(7.5 × 103P)
versus the reciprocal of the absolute temperature obtained by actual measurements usually are not perfectly straight
lines but show a slight downward curvature. Vapor pressure measurements in the range from 1 to 103 Pa by molecular
beam [27] and dew-point [28] methods give higher values of B. The latent heat of vaporization of the liquid in kcal/
mole as calculated from

is listed in the last column. The values of temperature listed under the vapor pressure equal to 67 Pa are of especial
interest since this vapor pressure is about the optimum value for the boiler pressure in oil diffusion pumps. The ultimate
pressure of fractionating pumps will be about the same as the pressure listed in the column labeled P (Pa), 25°C, but, as
proved experimentally by Blears [29], the reading of a tubulated ionization gauge may be somewhat lower than the true
vapor pressure as measured by a nude gauge because of clean-up or sorption of the vapor inside the envelope of the
tubulated gauge, resulting in a pressure drop across the tubulation. In interpreting ultimate pressure readings by nude
ionization gauges using the "nitrogen equivalent" calibration factor, it must be borne in mind that the equivalent
nitrogen pressures of organic vapors may be 5 to 10 times higher than the true vapor pressure depending on the
ionization probability for the molecules of the vapor and its decomposition products within the gauge.

Octoil® is vacuum-distilled commercial grade di-ethyl-hexyl phthalate (molecular weight 390.3). Octoil-S® is vacuum-
distilled commercial grade di-ethyl-hexyl-sebacate (molecular weight 426.3). Apiezon B and Apiezon C are molecularly
distilled petroleum fractions. DC 704 and DC 705 are silicone fluids. Chemically, DC 705 is a pentaphenyl trimethyl
trisiloxane with molecular weight 546 g/mol [30]. DC 704 is a tetraphenyl tetramethyl trisiloxane with molecular weight
484 g/mol [31]. Convalex-10® is a vacuum-distilled polyphenylether of average molecular weight 454 g/mol. Santovac-
5 is a mixed penta-phenyl ether of average molecular weight 447.

< previous page page_193 next page >


< previous page page_194 next page >
Page 194

Table 4.1. Vapor-Pressure Data for Diffusion Pump Oils


Degrees C for P in Pa P (Pa) 25°C ∆H (kcal/mol)
Name Ref. A B 133 66.7 13.3 1.33 0.133 0.013
Octoil 1 5590 188 177 153 123 4.1 × 105 25.6
12.12 97 74
2 4808 193 180 152 117 2.0 × 104 22.0
10.31 88 63
4 5540 204 192 166 134 1.5 × 105 25.3
11.62 106 82
Octoil-S 3 5514 215 204 177 142 2.7 × 106 25.2
11.26 114 89
2 4878 214 199 169 133 6.1 × 105 22.3
10.02 102 75
4 5780 213 201 175 143 4.4 × 106 26.4
11.90 115 90
Convoil-20 7 4800 210 196 166 129 9.2 × 105 22.0
9.94 98 71
Convalex-10 2 5525 281 265 230 188 3.6 × 107 25.3
9.97 156 122
Santovac 5 6 5450 290 273 237 193 3.3 × 107 24.9
9.68 157 125
Lion A 2 4950 210 196 167 131 5.7 × 105 22.6
10.24 101 74
DC 704 2 5376 215 202 174 140 1.3 × 105 24.6
11.02 110 85
5 5700 223 210 183 149 3.2 × 106 26.1
11.49 149 94
DC 705 2 6098 250 237 209 174 2.1 × 107 27.9
11.65 143 116
5 6490 254 241 214 180 4.7 × 108 29.7
12.31 151 125
Apiezon B 2 4831 214 200 170 132 6.8 × 105 22.1
9.91 101 74
Apiezon C 2 4808 246 229 195 153 1.9 × 105 22.0
9.27 119 89
Cellulube 2 5780 225 212 185 152 1.7 × 106 26.4
11.61 122 97

References

1. K. Hickman, N. Embree and J. Hecker, Ind. Eng. Chem., Anal. Ed. 9, 264 (1937).
2. K. Nakayama, J. Vac. Soc. Jpn. 8, 333337 (1965).
3. Metropolitan-Vickers, S. Dushman, personal communication (1946).
4. E. Perry and W. Weber, J. Am. Chem. Soc. 7, 3726 (1949).
5. D. Crawley, E. Tolmie and A. Huntress, Trans. 9th Natl. Vac. Symp. 9, 399, Macmillan (1962).
6. K. Hickman, Trans. 8th Natl. Vac. Symp., 2nd Intern. Congr., 1961, Vol. 1, p. 307, Pergamon (1962).
7. B. Dayton, calculated from vapor pressure data of CVC.

< previous page page_194 next page >


< previous page page_195 next page >
Page 195

The application of polyphenyl ethers as diffusion pump fluids was introduced by K. C. D. Hickman [32]. Cellulube 220
is a triarylphosphate. Lion A is an alkylnapthalene oil. Convoil-20 is a vacuum-distilled petroleum fraction and has an
average molecular weight 388 g/mol.

The stability of the various pump fluids against exposure while hot to air at atmospheric pressure has been investigated
by several authors [30, 31, 3336]. In general, the silicone fluids are more resistant to thermal decomposition and
oxidation than the phthalate and sebacate esters and the paraffin hydrocarbon oils, but have the disadvantage that
silicone vapor migrating back into a vacuum system can decompose on hot filaments and electrodes bombarded by
charged particles to form silica-like insulating deposits which disturb the operation of electrical devices. The polyphenyl
ethers have much higher thermal stability than the ester fluids in addition to being resistant to damage by radiation.

The open-cup flash point of the ester pump fluids is about 200°C, and thermal decomposition results in formation of
crystals of the acid anhydride and tarry deposits on the nozzle assembly whereas the nozzle assemblies remain relatively
clean when silicone fluid is employed. If the exposure time is brief, fractionating pumps can usually recover in a short
time of normal operation, so that the speed and ultimate pressure are not seriously affected. However, dark deposits on
the nozzle assembly increase the radiation loss to the water-cooled walls and require an increase in heater input to
maintain the normal boiler temperature. Jaeckel [34] found that thermal dissociation of organic oils is favored by metal
surfaces, so that all-glass pumps tend to give lower ultimate pressures than metal pumps. In any case, it is recommended
that for ester and hydrocarbon pump fluids the oil in the boiler be allowed to cool below 100°Cand for polyphenyl
ethers, to below 150°Cbefore exposure to the atmosphere.

Perfluoropolyether has been investigated as a diffusion pump fluid and has the advantage of comparable or greater
stability but the disadvantage of somewhat lower speeds as compared to polyphenyl ether and silicone pump fluids [37,
38]. It has been used under the trade name Fomblin as the fluid in rotary mechanical pumps [39].

4.3.5
Backstreaming and Back Migration of Pump Fluid

Vapor molecules, originating from the pump fluid and its decomposition products, can escape back through the pump-
inlet cross section into the components of the vacuum system attached to the inlet. In 1957 Ruf and Winkler [40] and
also Power and Crawley [41] showed that these vapor molecules can have four sources: (1) back migration due to
evaporation of fluid condensed on the pump wall or on any baffles in the region between the first-stage nozzle clearance
area and the pump inlet; (2) backstreaming due to backward curving streamlines and scattered vapor molecules from the
vapor jet between the rim of the nozzle exit and the pump wall; (3) backstreaming due to evaporation from hot liquid
pump fluid clinging to the outside of the first stage nozzle; (4) backstreaming due to leaks around the nut holding the
top (umbrella) cap to the nozzle assembly. The invention of the cold cap by B. Power [12] greatly reduced the
backstreaming from sources (3) and (4). Examples of a cold cap installed over the top nozzle are shown in Figs. 4.7, 4.8,
and 4.9. In general, it is common practice to place water-cooled baffles and refrigerated or liquid-nitrogen-cooled traps

< previous page page_195 next page >


< previous page page_196 next page >
Page 196

between the diffusion pump and the vacuum chamber to reduce the backstreaming and back migration to an acceptable
level.

When the backstreaming rate is high, it can be measured during pump speed measurements with a test dome by having
the top of the dome dished convex or sloping slightly to one side so that condensed oil runs down the side walls into a
gutter at the bottom of the dome from which the oil can drain into a burette as described in the AVS Standards
Committee document T.S. 4.5 (1963). When the rate is very small, other techniques must be used, such as weighing the
oil collected on a thin metal foil clamped to a liquid-nitrogen-cooled plate above the pump and any baffles employed
[42]. Measurements of the backstreaming rate, B, (in cm3/h of oil) obtained by the author on the MC series of three-
stage vertical metal diffusion pumps having a top-stage nozzle with 3-mm throat width (without cold cap) at normal
heater input and ultimate pressure with either Octoil or Convoil-20 for casing diameters 12, 14, 16, 20, 32, and 48 in.
When plotted as a loglog graph gave a straight line corresponding to

where D is the inside diameter (in inches) of the straight cylindrical pump casing. The specific gravity of Convoil-20 is
0.90, so that 1.0 cm3/h = 15 mg/min. For the 12-in. pump we obtain B = 0.52 cm3/h, which corresponds to 0.011 mg ·
cm2 · min1. Power and Crawley [41] obtained a rate of 8.2 × 103 mg · cm2 · min1 on a 6-in. pump with Apiezon C and
no cold cap, while addition of the cold cap reduced this rate by a factor of 10. Duval [43] measured the backstreaming
from various source points within a 600-mm-i.d. pump and showed that 99.8% came from the lip of the top nozzle. He
also showed that the reduction in backstreaming produced by a cold cap depended on the position of the lower lip of the
cold cap relative to the lip of the top umbrella nozzle. Baker [44] has described a cooled quartz crystal microbalance
method for measuring backstreaming and showed that the peak backstreaming rate is as much as ten times the normal
running rate during warmup of the pump and approximately four times as high during cooldown. These high rates
during start-up and shut-down can be largely avoided by using the gas purge technique [45, 46]. Hablanian, using the
foil and liquid-nitrogen-cooled plate technique, obtained 1.6 × 103 mg · cm2 · min1 for a 6-in. pump with DC-705
pump fluid and a water-cooled cold cap. Addition of a liquid-nitrogen-cooled trap with no creep barrier above the pump
reduced the rate to about 6 × 106 mg · cm2 · min1. Interposing a water-cooled baffle between the pump and the liquid
nitrogen baffle gave a rate of 2.8 × 107 mg · cm2 · min1, and adding a creep barrier to this combination reduced the rate
to about 1 × 107 mg · cm2 · min1.

Design of Baffles and Traps to Condense Backstreaming Vapor. When a baffle or trap is installed directly over the
nozzle assembly in a water-cooled pump casing, the baffle plates can be cooled by conduction through copper, brass, or
aluminum support fixtures making thermal contact with the pump wall. However, such baffles must be removable so
that the nozzle assembly can be inserted and removed for cleaning. External baffles and traps are to be preferred since
they can be cooled independently of the pump casing to any desired low temperature.

< previous page page_196 next page >


< previous page page_197 next page >
Page 197

When backstreaming is relatively high, it is important that most of the pump fluid condensed on the first baffle nearest
the pump inlet drain back into the diffusion pump. This baffle is therefore usually air- or water-cooled, and the housing
is fitted to the diffusion pump in a manner that allows the condensed fluid to return. A second refrigerated baffle can
then be installed beyond the first baffle and operated at a temperature so low that the condensate does not drain away
except when this baffle trap is warmed again. The design of an effective baffle or refrigerated trap requires that there be
no line of sight through the baffle plates or the cold-trap reservoirs and housing (which is usually not refrigerated),
while the conductance for air should be as large as possible consistent with the ''optically tight" requirement. However,
even optically tight baffles or traps may be a source of vapor contamination since condensed pump oil tends to "creep"
along the inner surface of the baffle or trap housing and may reach regions at warmer temperatures where the oil can
reevaporate directly into the vacuum chamber. Thus, anticreep barriers are sometimes incorporated [47, 48].

The efficiency, Be, of a baffle in condensing backstreaming vapor depends on the number of collisions, n, made by a
vapor molecule on the cooled baffle plates and the condensation coefficient, αc, according to the equation

where for baffles with short passages between bends in the plates or between entrance and exit cross sections the
average number of hits with the wall for molecules passing through can be estimated from

in which L is the length (in cm) of the shortest passage from entrance to exit along a path equidistant from adjacent
baffle surfaces, b is the number of bends (usually right angle) in the passage, and Re is the equivalent radius (in cm) of a
straight cylindrical tube of length L having the same probability of transmission for noncondensing molecules as the
baffle. For long straight cylindrical tubes of length L and radius R, the number of collisions with the wall in free
molecular flow is [49]

and the transmission probability is [21]

For mercury vapor molecules colliding with water-cooled steel or aluminum baffle plates the condensation coefficient,
αc, is less than 0.5, but for the large molecules of organic or silicone pump fluids αc will be nearly unity, and most oil
molecules will be condensed within the baffle. The effectiveness of the baffle then depends on the rate of reevaporation
from the baffle surfacethat is, on the sorption lifetime, ta, on the surface. Assuming ta constant for the whole baffle
surface, the mean transit time through the baffle for vapor molecules will be

< previous page page_197 next page >


< previous page page_198 next page >
Page 198

where

in which R0 is the molar gas constant, and Tn is the mean absolute temperature of the vapor in transit which depends on
the temperature in the entrance chamber, the thermal accommodation coefficient, the temperature of the baffle plates,
and the number of hits with the wall after entering the baffle. For oil molecules at the usual baffle temperatures,
When the vacuum system is operated for times longer than tn, the flow of vapor molecules through the
baffle will reach a steady state and the baffle has no effect. The quantity of vapor passing through the baffle will then be
equal to the backstreaming flux from the pump into the baffle times the probability of passage for any molecule
(without regard to the time spent in the condensed state) in molecular flow. For long narrow passages the time to reach
the steady state can be several hours, which is the principle of the Alpert copper-foil trap for oil vapor [50].

Conductance of Baffles and Cold Traps. The conductance of a baffle or cold trap can be measured directly by admitting
gas from a throughput meter into a test dome, similar to that used for measuring diffusion pump speed, over the end
which will be furthest from the diffusion pump and placing a short pipe section (collar) with inside diameter equal to
that of the exit of the baffle, and having a gauge attached to a tube joining the pipe at right angles to the axis so as to
measure static pressure between the baffle and a diffusion pump. If the maximum speed, Sd, of the diffusion pump has
already been measured without the baffle over a wide pressure range, the pipe collar can be omitted and the net speed,
Sn, of pump and baffle measured over a wide pressure range with the standard test dome, the baffle conductance, U,
then being obtained from the net speed formula

The conductance may be calculated with or without entrance and exit corrections. The conductance with these
corrections corresponds to the throughput divided by the difference in pressure between two very large vacuum
chambers placed over the entrance and exit of the baffle, as measured in the interior of the chamber far from the
entrance and exit. Since baffles are seldom used in this configuration, it is the conductance without entrance and exit
corrections, corresponding to measurement of the static pressure exactly at the entrance and exit cross sections, which is
more relevant. Because of the location of the gauge in the AVS or ISO standard test domes, an entrance correction is
included in the measurement of pump speed or net speed of pump and baffle. As an approximation, this correction term
to the impedance can be assumed to cancel out when using Eq. (4.27), so that U is the conductance without entrance and
exit corrections. The entrance correction to a straight pipe section is about equal to one-half the impedance of a hole in a
thin plate of the same area and cross section as the entrance to the pipe as measured between two large chambers. The
actual calculation is complicated by beaming of gas from one pipe section across the entrance to an adjacent section and
the scattering of gas from any baffle plates or right angle bends [18, 51].

< previous page page_198 next page >


< previous page page_199 next page >
Page 199

Among the various types of baffle design which have been tried are (1) circular disk sandwiched between two flat
annular rings or located at the center of a cylindrical housing of about twice the diameter of the disk, (2) straight
chevrons, (3) concentric circular chevrons, (4) stepped flat annular rings, (5) stepped concentric circular chevrons, (6)
alternating semicircular plates inside cylindrical housing, and (7) split-chevron or louvre baffles. When backstreaming
is high, some alteration in the design of the baffle plates is needed to avoid condensed pump fluid dripping off a baffle
plate on to the top of the hot nozzle assembly in the diffusion pump. Zinsmeister [52] has treated very thoroughly the
characteristic properties of various baffle designs, including cold caps over the top nozzle. He defines a specific
conductivity as the total conductivity of the baffle divided by the area of the cross section of the baffle housing alone
and obtains 3.6 liter · s1 · cm2 for typical single-chevron baffles. The specific conductivity of optically tight baffles is
relatively independent of the diameter of the baffle housing. Davis, Levenson, and Milleron [53] have investigated the
optimum dihedral angle for the chevrons in a straight chevron baffle for maximum conductance both by experiment and
by Monte-Carlo calculation. Their results indicate an optimum angle of 120°.

4.3.6
Throughput

In vacuum engineering calculations the quantity of gas flowing through the pumping system is usually expressed as the
throughput (symbol Q), defined as the product of pumping speed, S, and the static pressure, p, at the prevailing gas
temperature at a specified cross section. It is recommended that speed be expressed in either liters per second (liter·s1)
or cubic meters per second (m3·s1). The ISO unit of pressure is the Pascal (abbreviated Pa); but current practice also
allows use of the millibar (abbreviated mbar), which equals 100 Pa. The unit Torr is no longer recommended. One
millibar equals 0.75 Torr. Suitable units of throughput are mbar·liter·s1, Pa·liter·s1, or Pa·m3·s1 at a specified
temperature (such as 23°C).

As pointed out by Lewin [54] and by Ehrlich [55], the above definition of throughput leads to complications when the
gas temperature is not the same in all parts of the vacuum system. However, by correcting all measurements to a
standard room temperature of 23°C and using the continuity relations,

where U12 is the conductance of the flow path between the points 1 and 2 at which the static pressure is measured,
vacuum system calculations are simplified. The conductance, U12, in the molecular flow region is independent of the
wall temperature [49, 56, 57]; but the entrance and exit pressures, p1 and p2, depend on the gas temperature and, when
measured by hot-filament ionization gauges, also depend on the wall temperature of the gauge [58, 59].

Each stage of a multistage vapor-stream pump has a maximum throughput which can be handled without jet breakdown
determined by the heater input and the speed of the next pumping stage in series. From Eq. (4.5) the maximum
throughput is

< previous page page_199 next page >


< previous page page_200 next page >
Page 200

where the subscript n refers to the number of the stage (the first stage being nearest the pump inlet and the last stage
nearest the forepump connection). For the last stage of the vapor-stream pump, the next pumping stage is that of the
forepump; and Sm(n + 1) = C, the speed of the forepump. When the nozzle assembly has optimum design, the value of
Qm for a given boiler pressure, P0, should be approximately the same for each stage in series. The required speed, C, of
the forepump then depends on the characteristics of the last stage of the vapor pump as given by Eq. (4.5). When the
speed of the forepump is sufficiently high, the limiting forepressure, Fmb, at this maximum throughput can be
determined by measuring the forepressure, F, as gas is admitted to the forevacuum region until jet breakdown begins, as
indicated by a rapid rise in the high-vacuum pressure, p, when the leak rate of gas on the high-vacuum side is
maintained constant at a series of increasing values.

Heater Input and Water Cooling. Since according to Eq. (4.29) the maximum throughput is proportional to the boiler
pressure, it is of interest to determine the dependence of the boiler pressure on the watt input to the pump heaters, the
heat carried away by the cooling water, and the radiation and conduction losses from the pump casing. The heat carried
away by the cooling water per second (in watts) can be estimated from

where we have used Eq. (4.53) below and

in which P0 is the vapor pressure (in Pa) and T0 the absolute temperature in the boiler, M is the mean molecular weight
of the pump fluid, R0 = 8.314 × 107 erg · K1 · mol1, Σ Ant is the sum of all the nozzle throat areas in an n-stage pump,
Lv is the latent heat of vaporization per gram of pump fluid, s is the specific heat of the liquid pump fluid, T0 is the
stagnation temperature of the vapor from the nozzles, Tw is the lowest temperature reached by the cooled condensed oil
on its way back to the boiler, D is the casing diameter (in cm), Lbw is the distance (in cm) between the level of fluid in
the boiler and the lowest turn of the cooling coil, kc is the thermal conductivity of the casing material, and t is the
thickness of the casing wall (in cm), Wr is the rate of heat transfer (in watts) by radiation from the hot nozzle assembly
to the cooled casing in the region of the cooling coils, Aj is the area (in cm2) of the outside surface of the hot nozzle
assembly directly exposed to the casing in the region of the cooling coils, ej is the emissivity and Tj is the average
temperature (in K) of this surface which will be slightly less than the vapor temperature, T0, in the boiler, and Tac is the
average temperature (in K) of the inside of the pump casing in the region of the cooling coils which will be somewhat
higher than Tw and Wc is the rate of heat transfer by conduction along the casing wall from the boiler to the lowest turn
of the cooling coil. Heat transfer from the ambient air to the cooling coil is neglected. The

< previous page page_200 next page >


< previous page page_201 next page >
Page 201

heat exchange between the casing wall and the returning pump fluid over the distance Lbw is also neglected.

For organic pump oils, Lv is in the neighborhood of 6070 cal·g1 and s is about 0.2 cal·g1·deg1. A more exact treatment
would use the emissivity factors for concentric cylinders and consider the heat conduction through the film of pump
fluid on the casing wall; but since emissivity of the pump fluid is nearly equal to 1 and the film is thin, the above
equation is sufficient for a rough estimate. The value of ej will vary from about 0.2 to 1.0, depending on the degree of
discoloration of the surface of the nozzles assembly. It has been found that the radiation loss is somewhat less than 10%
of the heater input when the nozzle assembly is clean.

Finally, the required heater input to the boiler will be

where Wb is the rate of heat loss by radiation and air convection to the surroundings of the boiler region up to the lowest
cooling coil. The boiler region in some pumps is thermally insulated with fireproof insulating material. The electric
heating units may be installed as hot plates in direct contact with the flat bottom of the boiler or may take the form of
cartridge units inserted in cavities which extend into the pump fluid. The latter arrangement reduces the "bumping"
which sometimes occurs as the oil becomes superheated and flashes into vapor. This bumping can cause vibrations
which create problems in apparatus attached to the pump, such as electron microscopes.

The heat transferred to the cooling water has been measured to be between 55% and 75% of the heater input W for
external heaters and between 65% and 80% of W for immersion heaters. The rate of flow of cooling water through the
coils should be adjusted so that the discharge temperature is 3545°C to maintain a relatively warm temperature of the oil
descending along the wall from the last coil turn to the boiler. This helps to prevent condensation and return to the
boiler of light ends of the pump fluid which normally undergoes some decomposition during pump operation. As a rule-
of-thumb, for pumps with external heaters the cooling water should circulate through the coils from the high-vacuum
toward the fore-vacuum side at a rate (in cm3/min) about equal to the heater input in watts.

While the vapor pressure, P0, during operation of diffusion pumps may vary over the range from about 10 to 100 Pa,
depending on the heater input, W, the absolute temperature, T0, of the vapor in the boiler varies only from about 440 to
480 K for typical pump oils as shown by Table 4.1. The variation in in Eq. (4.30) is therefore only from 21.0 to
21.9, and for all practical purposes it can be assumed that the boiler pressure, P0, is a linear function of the heater input
W. This is confirmed by the experimental data.

To ensure a supply of vapor to the nozzles at a pressure near that in the boiler and to avoid overheating the fluid in the
boiler, the area of the free surface of the pump fluid in the boiler should be at least five times the total nozzle throat
area, ΣAnt. In the forepressure breakdown region from pb to pc the vapor jet is deformed and cooling is less efficient, so
that the boiler temperature tends to rise to dangerous levels. Also, the deformation of the first-stage vapor jet will
increase the backstreaming rate. Oil

< previous page page_201 next page >


< previous page page_202 next page >
Page 202

diffusion pumps should not be allowed to operate in the forepressure breakdown region for any length of time, or else
the heater power should be turned off.

4.4
Theory of Pump Performance

The theoretical analysis of the pumping action of diffusion and ejector pumps has engaged the attention of many
authors. Kinetic theory can be applied to analyze the counterdiffusion of gas from the high-vacuum side and vapor
spreading sideward and backward from the nozzle exit as well as the counterdiffusion of gas from the fore-vacuum and
vapor discharged in the forward direction from the nozzle. Hydrodynamic theory, as developed in connection with
steam turbines, can be applied to the expansion of the vapor through the nozzle and beyond the nozzle exit. Theoretical
equations have been derived for the following aspects of pump performance: speed, limiting forepressure, ultimate
pressure, and backstreaming.

4.4.1
Speed

In 1915 Gaede derived a theoretical equation for the speed of a diffusion pump by analyzing the counterdiffusion of gas
in a mercury-vapor stream as follows. First, he considered the case of a stream of mercury vapor flowing in the axial
direction (x-axis) through a cylindrical tube of radius a with average velocity v and a density high enough to be in the
viscous flow region so that entrapped gas molecules diffusing countercurrent to the vapor stream tend to be carried
along by the vapor stream. In the steady state the concentration n of gas molecules at x varies linearly from a very small
value n0 at x = 0 to a maximum nm at x = L because of a reservoir of gas beyond L from which gas molecules can
penetrate into the vapor stream and diffuse counter to the stream velocity with diffusion coefficient D. For L/a < 10 the
velocity u will be approximately constant over any cross section of the tube. Since n is proportional to the pressure p of
the gas and D varies inversely with the pressure P (in mmHg) of the mercury vapor, Gaede obtains

where D is the diffusion coefficient at 760 mmHg (101,308 Pa), and dp/dx is the pressure gradient of gas along the tube,
which is equal and opposite to the pressure gradient, dP/dx, in the vapor. Integrating Eq. (4.33) from p = p0 to p = pm
gives

The counterdiffusion of gas and vapor results in an actual pressure drop in the gas from x = L to x = 0 given by

< previous page page_202 next page >


< previous page page_203 next page >
Page 203

As Gaede pointed out, this introduces an error in the reading of gas pressure with a McLeod gauge. For a mercury vapor
stream of sufficiently high pressure P and velocity u the gas pressure at x = 0 can be reduced to a very low value, p0 =
βpm.

Gaede then considered the case of mercury vapor flowing past a small hole of diameter d in a thin wall separating the
main body of vapor from a region containing gas at the pressure p. From kinetic theory the number of vapor molecules
crossing a small surface element σ normal to the radius from the hole and located at the angle ϕ with the normal to the
plane of the hole will be

where N is the number of vapor molecules per unit volume in the vapor stream near the hole coming from the source
side and ua is the average molecular velocity at the temperature T. Since the emerging vapor molecules do not actually
start at a point at the center of the hole, Gaede introduces a spherical surface of radius x0 at which the concentration is
N. Since the total number of vapor molecules crossing the hemisphere of radius x0 per second will be equal to the
number Z crossing the hole of diameter d, it follows that the concentration N will be equal to Z divided by the volume
or

so that The number of vapor molecules per unit volume, N(x, ϕ), at σ will be

The mean free path, λ(x, ϕ), of the gas molecules through the vapor at x and ϕ can be assumed to be inversely
proportional to the concentration N(x, ϕ) of the vapor molecules since the gas pressure is small compared to the vapor
pressure, or

where λ0 is the mean free path of a gas molecule through the vapor at the hole, where x = x0. Gas molecules diffuse
through this mercury vapor toward the orifice and are scattered so that only a fraction, α, of the molecules finally reach
the orifice depending on the mean free path, λ(x, ϕ), for the gas molecules through the vapor as given by

By integrating over all angles, Gaede obtains

For λ0 = d, this gives α = 0.793.

< previous page page_203 next page >


< previous page page_204 next page >
Page 204

Then Gaede's complete formula for the speed at the entrance slit is

where A is the slit area, and β is the back-diffusion coefficient for gas from the fore-vacuum at the forepressure F which
penetrates back through the vapor jet by counterdiffusion and generates a partial pressure

of this gas at the entrance slit area, while α is the fraction of the gas molecules from the high-vacuum side that penetrate
through the backstreaming vapor and reach the slit area as given by Gaede's Eq. (4.41), and γ (using Ho's γ notation in
place of Gaede's k) is the fraction of the gas molecules which have reached the slit area that actually pass through and
are carried away by the vapor stream. The coefficient β is given by Eq. (4.34) when P, v, and L are constant, but for
vapor jets diverging from a nozzle in modern pumps these parameters are not constant and integration must be
performed over a path through the vapor jet from the fore-vacuum to the high-vacuum side. Experimentally, β
corresponds to the slope of the p versus F curve for no load in Fig. 4.11.

As explained below, Molthan [60] modified Gaede's equation for α by taking account of the effect of the stream
velocity, v, of the vapor molecules on the pattern of flow of vapor molecules emerging through the hole, or slit.

4.4.2
Limiting Forepressure

As the forepressure is increased by admitting a gas leak into the fore-vacuum with a constant gas throughput on the inlet
side, a shock boundary is formed between the supersonic vapor jet and the gas in the fore-vacuum. Increasing the fore-
pressure further pushes this shock boundary back toward the nozzle exit as shown by Dayton [61] in 1948 in
experiments with a glass diffusion pump using pump fluid colored with a red dye and observing a glow discharge in the
gas and vapor in front of the nozzle. When the shock boundary is pushed back to the point at which the jet no longer
makes contact with the cooled pump casing, the vapor jet again shoots forward as a narrow beam into the fore-vacuum
but has a side boundary with the gas at the elevated pressure and eventually bends to one side to condense on the pump
wall.

This behavior is also shown for the case of a metal pump by the isobar diagrams obtained by a pressure probe moved up
and down through the jet of a mercury vapor pump by P. Alexander [62] (see Fig. 4.13) and by the glow discharge
photographs and line drawings in a 1955 article by Kutscher [15]. Similar diagrams were prepared in 1960 by N.
Florescu [63].

The gas density distribution around the upper stage of a 6-in. vertical metal pump at an inlet pressure of 5.3 × 102 Pa
was explored by Hablanian and Landfors [64, 65] using an ionization gauge probe. They also measured the distribution
of vapor arriving at the wall from the top nozzle (both with and without a cold cap) by using a series of gutters and drain
tubes.

< previous page page_204 next page >


< previous page page_205 next page >
Page 205

Fig. 4.13
Isobars of air pressure in mercury vapor pump [62].
4.4.3
Vapor-Jet Flow Pattern

When the average stream velocity is parallel to the nozzle axis and both the throat area, At, and the exit area, Ae, are
orthogonal to this axis and no condensation occurs

< previous page page_205 next page >


< previous page page_206 next page >
Page 206

Fig. 4.14
Expanding conical nozzle showing various
defined cross sections.

in the nozzle, then the mass of vapor discharged by an expanding conical nozzle is given (in cgs units) by

where ρt is the vapor density, Pt is the static vapor pressure in dyne·cm2, vt is the stream velocity in cm/s, Tt is the
absolute vapor temperature at the nozzle throat of cross-sectional area At, and the subscript e indicates the
corresponding quantities at the nozzle exit (see Fig. 4.14). It is known that for adiabaticisentropic expansion of vapor
through the nozzle when the forepressure is below the limiting value, the velocity vt at the throat equals the acoustic
velocity,

where k is the ratio of specific heats. Since any change in downstream pressure cannot influence the throat pressure
when the flow is supersonic, the discharge rate is a maximum.

In passing adiabatically from the cross-sectional area Ac of the vapor conduit from the boiler to the throat area At the
vapor density and pressure changes according to

where ρc is the density of the vapor in the conduit from the boiler. For sufficiently large Ac relative to the throat area,
At, it can be assumed that Pc is approximately equal to the boiler pressure, P0, and ρc is approximately equal to the
vapor density, ρ0, in the boiler.

When the forepressure is below the limiting value, the discharge rate is a maximum and the static vapor pressure at the
throat cannot be increased beyond the critical value,
where Ptc will be in Pa if P0 is in Pa.

< previous page page_206 next page >


< previous page page_207 next page >
Page 207

Combining the above equations gives for the maximum discharge rate (in cgs units) of short nozzles

For adiabaticisentropic expansion through a short nozzle the static pressure, Px, the vapor density, ρx, and the absolute
vapor temperature, Tx, at any cross section x are related to the initial pressure Pc, density ρc, and temperature Tc in the
vapor conduit by

There are no experimental data on the value of k for the high-molecular-weight organic and silicone molecules of
typical diffusion pump fluids, but from the equipartition theorem as modified by quantum mechanics it can be assumed
that for all temperatures above 300 K the specific heat ratio can be estimated from

where Nv is the number of vibrational degrees of freedom which are fully excited at the temperature of the vapor in the
boiler [66]. A study of the available data on the value of k for large organic molecules reveals that in the temperature
range 30°C to 300°C the value of k will be about 1.5 times the number of atoms in the molecule [67]. The theoretical
maximum value is 3n 6, where n is the number of atoms in a molecule (with n > 3), but this value cannot be reached
except at extremely high temperatures where in most cases the molecule decomposes. The pump fluid Octoil (dioctyl
phthalate) contains 66 atoms, so that Nv = 99 and k = 1.019. For typical hydrocarbon pump fluids with molecular weight
about 390, the molecules have about 28 carbon atoms and 54 hydrogen atoms and Eq. (4.50) using the 1.5 factor gives k
= 1.015. It can be assumed, as an approximation, that k = 1.02 for organic pump fluids. The specific heat ratio for
mercury vapor is k = 1.666.

The above equations involve the assumption that the initial stream velocity, vc, in the conduit of cross-section area Ac is
practically zero. This will be true when but for some diffusion pump designs Ac is less than 5At and a factor

should be included in Eq. (4.48). Neglecting this correction factor for the initial velocity and assuming that P0 at the
nozzle entrance is about equal to the boiler pressure and using k = 1.02, Eq. (4.47) gives the critical throat pressure Ptc
= 0.60P0 and Eq. (4.48) gives

< previous page page_207 next page >


< previous page page_208 next page >
Page 208

for oil diffusion pumps with forepressure below the limiting value corresponding to a maximum discharge rate. Since
the vapor density, ρ0 (in cgs units), is not measured directly but calculated from

where M is the average molecular weight of the pump fluid, R0 is the molar gas constant, and T0 is the absolute
temperature of the vapor in the boiler region, a more convenient form of Eq. (4.51) is

As the vapor passes from the throat to the nozzle exit of area Ae > At, it expands and acquires supersonic stream
velocities. When the forepressure is below the limiting value and the initial velocity, v0, is negligible,
adiabaticisentropic flow through a short expanding conical nozzle results in an average stream velocity (parallel to the
nozzle axis) at the nozzle exit given by

where Pe is the static vapor pressure at the exit. For organic pump fluids with values of k approaching 1, it can be
shown by an application of L'Hôpital's Rule [67] that the nozzle discharge rate is given by

Differentiating G with respect to Ptc and setting the result equal to zero gives the critical pressure at which G is a
maximum,

which is approximately the value found above. Substituting in Eq. (4.55) then gives

which is the same as Eq. (4.53). Similarly, applying L'Hôpital's Rule to Eq. (4.54) as k → 1, we obtain

Then, using Eq. (4.45) and Eq. (4.63) below with k = 1, the ratio of exit to throat velocity is

< previous page page_208 next page >


< previous page page_209 next page >
Page 209

Combining Eq. (4.44), Eq. (4.56) and Eq. (4.59) using gives

Since this equation cannot be solved explicitly for Pe, but Pe/P0 < 1, then by successive approximations we obtain for
the static vapor pressure at the exit of a short expanding conical nozzle:

where e is the base of the natural logarithms and the last P0/Pe has been replaced by as an approximation. For a
cylindrical nozzle, At = Ae and the static pressure at the exit is

where Pt is the throat pressure as given by Eq. (4.56). In this case the exit velocity, ve, equals vt as given by Eq. (4.59).

When the forepressure is below the limiting value, the vapor temperature at the throat is

and for organic vapors with k = 1.02 one can assume Tt = T0, the boiler temperature, so that the acoustic velocity, vs, at
the throat is essentially the same as the acoustic velocity at the boiler temperature,

When the forepressure is below the limiting value, the vapor diverges from the nozzle exit at supersonic velocities. The
central portion of the vapor jet can be assumed to follow the hydrodynamic equations for streamline flow developed by
Prandtl and others for steam jets [68]. However, the outer regions of the vapor jet expand to such low pressures that
continuum physics cannot be applied, and the vapor expands freely according to the laws of kinetic theory so that the jet
pattern in the outer fringe region can be estimated by superimposing a downward stream velocity on the random
direction and thermal velocity distribution according to Maxwell's equation.

Figure 4.15 is a modification of a diagram first presented by Molthan [60] which shows that a hemispherical surface of
radius u, corresponding to some fixed value of the thermal molecular velocity for molecules escaping through a hole c
in a wall between the vapor source and a high vacuum with zero initial stream velocity, becomes displaced in the
direction of a superimposed vapor stream velocity, v, of the vapor on the other side of the wall. Choosing, as Gaede did,
a circular hole with

< previous page page_209 next page >


< previous page page_210 next page >
Page 210

Fig. 4.15
Displacement of vapor stream pattern
diverging from slit c for fixed thermal velocity,
u, by superimposing a stream velocity, v.

diameter d, then when the mean free path of a vapor molecule in the vapor at the hole is less than 10 times the diameter,
d, Molthan assumed that the vapor does not emerge from the hole according to the cosine law but rather emerges with
equal probability in all directions so that the density of molecules at any point on the hemispherical surface of radius u
will be a constant.

The magnitude of the thermal velocity, u, varies from 0 to ∞ with the probability, w(u), that the velocity lies between u
and u + du being given by Maxwell's distribution law:

where the unit of the velocity u is chosen as the most probable velocity

As a result of the translation of the surface of a given radius u in the direction v, the flux per unit solid angle in different
directions is no longer the same. Let a line (which will be termed the x-axis) be drawn on the surface of the wall through
the center of the hole and parallel to the stream velocity v, then any plane through this line cuts out two semicircles on
the two hemispheres as shown in Fig. 4.15. The intersection of this plane (which will be called the ''diagram plane")
with a plane through the center of the hole and perpendicular to the stream direction determines the line . To
calculate the new density distribution on the displaced hemisphere, consider a small arc on the initial semicircle
which is displaced to the position of equal arc length, on the displaced semicircle. A plane passed through the
line orthogonal to the diagram plane will determine a dihedral angle θ with the plane through perpendicular to
the stream direction, and a plane passed through orthogonal to the diagram plane will form the dihedral angle θ'
with the plane through . Now let the arc represent a small surface element, AB, cut from the hemisphere by the
planes forming the dihedral angle and planes through x-axis. The solid angle determined by the surface element AB will
now be greater than the solid angle at c determined by the surface element A'B' of area equal to that of AB. As shown in
the diagram, let the latter solid angle also determine the surface element ab on the hemisphere of radius u. The vapor

< previous page page_210 next page >


< previous page page_211 next page >
Page 211

stream flux per unit solid angle will then be greater in the direction of A'B' (or ab) than in the direction of AB by the
ratio A'B' : ab. To find this ratio, form the intersection of all the rays within the solid angle A'B' and a sphere of radius
which cuts the ray at B". Then the spherical surface element of area A'B" determines the solid angle

while the element AB determines the solid angle AB/u2 which is numerically equal to A'B'/u2. For infinitesimal surface
elements we have

where ψ is the dihedral angle B'A'B" which is numerically equal to the dihedral angle AcA' = θ' θ, where angles to the
right of cB are considered positive and angles to the left are considered negative. Then

For surface elements on the portion of the displaced hemisphere to the left of the plane through , this ratio will be
less than unity corresponding to reduced flux per unit solid angle in directions with x-components opposite to the stream
velocity vector. For θ' = 0, corresponding to a surface element of the displaced hemisphere intersecting the line , the
flux is unchanged because ab = AB. The probability that a molecule has the velocity u now depends on the direction of
flight, θ', so that Eq. (4.65) must be replaced by

in which β = θ + π/2 is the angle between v and u, and

Only those molecules having a backward component of thermal velocity greater than the downward stream velocity will
be able to travel back through the nozzle clearance area. For a cylindrical nozzle the vapor velocity at the center of the
nozzle exit will be near the acoustic velocity, which for mercury is only 0.8 times the mean thermal velocity, so that a
considerable fraction of the vapor molecules will be scattered back through the clearance area. For these molecules, θ' is
negative. To

< previous page page_211 next page >


< previous page page_212 next page >
Page 212

determine this fraction it is necessary to find the value of the integral

where θ' has been eliminated using Eq. (4.72).

For vapor molecules with positive θ' the lower limit of integration is replaced by

because then molecules with thermal velocities, u, smaller than the stream velocity, v, contribute to the flow, depending
on θ'.

Molthan evaluates the integral equations for Z(θ) by a series expansion of the integrand in different quadrants and by
use of the Error Function when appropriate. He then uses the relation

to convert the distribution Z(θ) to a function of θ', which can be denoted D(θ'), and which Molthan presents in tabular
form. For a stream velocity v = vs, the velocity of sound in the vapor, his calculated values are given in Table 4.2.

Neglecting friction at the nozzle wall, Molthan obtains the vapor jet pattern shown in Fig. 4.16.

The outer curved dashed line is orthogonal to all streamlines. Along this dashed line the distance between two
streamlines is proportional to the mean free path of gas through the vapor. In the region a where vapor travels backward
from the edge of the nozzle exit, Molthan calculates that the mean free path is 4.742/0.063 = 75 times greater than in the
region b in the center of the vapor jet. For mercury vapor at 133 Pa and 400 K the mean free path of nitrogen through
the vapor is about 7 × 103 cm, and if this vapor pressure and temperature occurs in the region b, then the mean free path

Table 4.2. Molthan Angular Distribution Function


θ' D(θ') θ' D(θ')
90° 0.063 0.422
0°10'
73°20' 0.070 0.757
15°50'
60°10' 0.081 1.309
30°40'
44°40' 0.111 2.139
45°40'
28°40' 0.169 3.177
60°30'
16°20' 0.247 4.742
90°

< previous page page_212 next page >


< previous page page_213 next page >
Page 213

Fig. 4.16
Scattering of mercury-vapor stream at
the exit of a cylindrical nozzle [60].

for nitrogen through the vapor in the region a will be about 0.53 cm, which is comparable to the slit width in Langmuir's
cylindrical nozzle pump.

Molthan did his work under Gaede's direction, and one purpose of his mathematical analysis was to confirm Gaede's
hypothesis that the speed of all diffusion pumps, including Langmuir's pump, will become small when the mean free
path of gas through the vapor at the pumping annulus (nozzle clearance area) becomes smaller than the "slit
width" (distance from nozzle exit rim to pump wall). Alexander [62] applied an analysis similar to Molthan's
(apparently without having read Molthan's papers) and came to the conclusion that Gaede's diffusion principle cannot
explain the action of vapor-stream pumps. By using a mercury vapor pump with an expanding nozzle, he showed that
the vapor stream is mainly directed downward away from the nozzle clearance area and very little vapor is scattered
backward. Using a long vertical tube to probe the vapor jet, Alexander obtained the isobars of gas pressure shown in
Fig. 4.13. However, Jaeckel [69] considering the same problem disagrees with Alexander's conclusion. Also, Avery and
Witty [70] showed that Alexander's criticism of Gaedes' theory was based on a misconception of Gaede's mean-free-
path rule.

Molthan obtained a corrected form of Gaede's α factor by dividing the original hemispherical surface centered on the
hole c into zones of area
< previous page page_213 next page >
< previous page page_214 next page >
Page 214

However, one can also use differentials,

Then, instead of Gaede's Eq. (4.38), one now has for the vapor concentration at dσ,

where D(θ') is Molthan's vapor dispersion factor. Molthan omits the factor cos ϕ for simplicity since it is difficult to
integrate this factor over a zone σ (or dσ). Gaede's α formula is then modified to

where K is of the order of unity and depends on the method of correcting for the variation over the angle ϕ with the
normal to the plane of the hole. Since

the general effect of K is to introduce a factor of 1/2. Molthan's failure to account for the angle ϕ leads him to the
incorrect result α = 2 when λ0→∞ and hence a maximum flow of 734 cm3 · s1 through a hole whose area is 0.0314
cm2.

Oyama [71] criticized Molthan's assumption of spherical symmetry for the velocity directions of the thermal motion on
the grounds that the gas is expanding after leaving the nozzle and not in thermal equilibrium and in fact may incur some
condensation into liquid droplets. He also questioned Molthan's assumption that the cosine law did not apply to the
vapor crossing the hole, although Molthan may have been justified in this assumption for the case that the mean free
path is less than d/10. However, the cosine factor should be included for the gas molecules crossing the hole since their
mean free path is greater than d, but Molthan omitted this factor. Oyama considered the problem in two dimensions and
applied the PrandtlMeyer theory of the supersonic flow around the edge of a flat wall where the streamlines curve
outward until the vapor is expanded to the point at which the mean free path was 1 mm, after which he considered the
vapor to continue in the same direction in a straight line without further change in velocity. By integrating over all
possible values of the thermal velocity u he calculates the angular distribution of the molecular flux from a
nonexpanding (two-dimensional) nozzle as a function of the angle θv between the molecular velocity w and the stream
velocity v for mercury vapor with k = 5/3 and for a vapor with k = 20/19 = 1.053 for three different boiler pressures.

The PrandtlMeyer formula [68] for the bending of the streamlines of a vapor expanding adiabatically with specific heat
ratio k around the edge of a horizontal flat plate in two dimensions into a vacuum is

< previous page page_214 next page >


< previous page page_215 next page >
Page 215

where r is the radial distance from the edge as a function of ϕ, the angle of turning measured from the starting ray
through the edge to the initial point on the streamline at which the velocity component along the ray direction is zero so
that the ray makes the angle 90° with the stream velocity. For a hypothetical nonexpanding nozzle with a rectangular
cross section of width d and length so great that the flow can be considered in two dimensions only, the vapor stream
velocity at the exit plane will equal the acoustic velocity and be directed parallel to the nozzle axis so that the radial
component toward the nozzle edge is zero and no lateral expansion occurs until the vapor has crossed this exit plane.
Then for streamlines starting near one edge of this nozzle the angle ϕ equals 0 for the ray lying in the exit plane and r0
is the vertical distance to the streamline as it crosses the exit. For mercury vapor, k = 5/3 and the expression in the
brackets becomes cos(ϕ/2); thus the maximum angle of turning is 180°, at which r = ∞. For organic vapors, such as
Octoil, k = 1.02, and theoretically for hydrodynamic flow with adiabatic expansion the maximum angle of turning

would be so that the vapor streamline closest to the wall of the flat nozzle could curl right around the edge
and flow backward on the underside. However, when the vapor has expanded to the point that the mean free path is no
longer small compared to the width (of order d) of the vapor jet, the path of the molecules is no longer influenced by a
gradient in the vapor pressure but is determined by the local stream velocity and the random thermal motion at the local
temperature as calculated by Molthan. A plot of Eq. (4.81) for the case k = 1.02 is shown in Fig. 4.17.

Fig. 4.17
Two-dimensional streamline flow pattern around edge of
flat plate into vacuum for organic pump fluid vapor.

< previous page page_215 next page >


< previous page page_216 next page >
Page 216

Fig. 4.18
Expansion ratio for k = 1.02 as a function of the angle
of turning. Nozzle exit plane horizontal.

The static vapor pressure decreases along a streamline as the vapor expands, so that the pressure P at (r, ϕ) is a certain
fraction of the initial vapor pressure, P0, before any expansion occurs as given by the PrandtlMeyer formula [68]:

where k is the specific heat ratio and P0 can be considered to be the boiler pressure. A plot of this equation for the case
k = 1.02 is given in Fig. 4.18, where P0 is the boiler pressure and Pe = 0.602 P0 is the static vapor pressure at the exit
(represented by the horizontal line), where ϕ = 0, of the above hypothetical flat rectangular nozzle or of an equivalent
nonexpanding annular nozzle formed by concentric cylinders of very large diameter. At ϕ = 180° the ratio is 3.6 × 103.

Table 4.3 lists numerical values for the PrandtlMeyer ratios r/r0 and P/P0 for Octoil (k = 1.02) and mercury (k = 5/3).

By an application of L'Hôpital's rule for the evaluation of indeterminate forms it can be shown that for organic pump
fluids with k → 1, Eq. (4.81) reduces to the very simple form

which is equivalent to the isothermal solution of the problem of steady flow around a corner as given in Stodola and
Loewenstein [68, p. 983]. Similarly, as k → 1, Eq. (4.82) reduces to
< previous page page_216 next page >
< previous page page_217 next page >
Page 217

Table 4.3. PrandtlMeyer Ratios


ϕ k = 1.02 k = 1.67 ϕ k = 1.02 k = 1.67
r/r0 P/P0 r/r0 P/P0 r/r0 P/P0 r/r0 P/P0
1.0039 0.602 1.0049 0.487 0.170
0 90 3.4507 4.000 0.086
1.0153 0.600 1.0154 0.485 0.126
5 100 4.6232 5.8582 0.053
1.0349 0.593 1.0350 0.478 0.091
10 110 6.3857 9.2421 0.030
1.0629 0.581 1.0632 0.467 0.063
15 120 9.1158 16.000 0.015
1.1000 0.567 1.1007 0.451 0.0345
20 135 16.477 46.642 0.004
1.1471 0.523 1.1488 0.410 0.0175
30 150 32.031 88.731 0.0006
1.2762 0.469 1.2824 0.357 3.6 × 103
40 180 151.15 ∞ 0
1.4641 0.408 1.4821 0.298 1.5 × 103
50 195 367.3
1.7322 0.344 1.7778 0.237 5.4 × 104
60 210 965
2.1128 0.281 2.2212 0.180 5.9 × 105
70 240 8450
2.6589 0.222 2.9036 0.129 102800 4.7 × 106
80 270

and using Eq. (4.62) this becomes

where Pe is the static vapor pressure at the exit of a short cylindrical nozzle of large diameter or of a concentric cylinder nozzle of large mean
radius so that the flow pattern is approximately two-dimensional. It is evident that the static pressure is assumed to be constant along any ray
with fixed angle ϕ. Combining Eqs. (4.83) and (4.85) gives

thus for organic vapors issuing from the type of nozzle specified above, the static vapor pressure in the jet is inversely proportional to the length
of the ray from the edge of the nozzle to the intersection with the streamline starting at r0. For concentric cylinder nozzles with finite mean
radius, one might expect the vapor pressure to decrease as some higher power of 1/r.

Using Eqs. (4.86) and (4.83) the mean free path of a vapor molecule through the vapor at (r, ϕ) will be given by

where T is the local vapor temperature, σv is the mean molecular diameter (in cm), and the pressure is expressed in Pa. As an example,
consider Octoil with σv = 1.2 × 107 cm at a local temperature of 160°C (433 K) and boiler pressure P0 = 67 Pa, for which λv = 2.33 × 103 exp
(ϕ2/2). If we assume that the boundary between hydrodynamic theory (continuum) flow and kinetic theory scattering occurs when the mean
free path is of the order of 0.1 cm, then the boundary angle will be

< previous page page_217 next page >


< previous page page_218 next page >
Page 218

ϕ = 2.742 rad = 157.1°, which from Fig. 4.17 occurs in front of the nozzle so that kinetic theory must be used for the
vapor molecules scattered back through the nozzle clearance area.

As the vapor expands the flow velocity, v(r, ϕ), becomes greater than the acoustic velocity, vs, and the Mach number is
given by v(r, ϕ)/vs. Stodola defines a Mach angle, ψ, by

where vr is the component of the stream velocity v(r, ϕ) directed downward along the ray from the turning edge. Along
the starting ray we have ϕ = 0, ψ = 0, and vr = 0, so that no lateral expansion has begun and the ray is perpendicular to
the streamline.

For Octoil with k = 1.02 this becomes

or for k → 1 by L'Hôpital's rule we have tan ψ = ϕ.

The acute angle between a ray from the nozzle rim and the tangent line to a streamline is the complement of Mach's
angle, or ψc = (π/2) ψ, as defined by Stodola, and this angle is constant for all streamlines cutting a given ray as shown
in Fig. 4.19, which is adapted from Fig. 954 in Stodola and Loewenstein.
Fig. 4.19
Prandtl's two-dimensional flow around a corner into gas at pressure p2 [68].

< previous page page_218 next page >


< previous page page_219 next page >
Page 219

Nöller [72] has modified Oyama's approach by calculating the angular distribution function from the state of mercury
vapor at the discontinuity surface between the PrandtlMeyer expansion and molecular flow and by assuming that an
observer moving with the preferred velocity will observe a normal Maxwellian velocity distribution at that point. At the
discontinuity surface the mean free path is not kept constant but rather the Knudsen number λ/x is kept constant, x being
the distance from the edge of the nozzle. He applies this approach to a nozzle for which the mercury vapor velocity at
the exit is 2.2 times the acoustic velocity; thus from Eq. (4.88) the angle ϕ in the Prandtl diagram increases from 0 at the
starting ray inside the nozzle to ϕ = 95.45° at the ray corresponding to Mach 2.2, as compared to the angle ϕ = 53.13°
for the Mach 1.0 case above. For an exit velocity of Mach 2.2 the Mach angle at the exit is ψ = 65.56° as compared to ψ
= 45° for the Mach 1.0 case. The starting ray for the Mach 2.2 case is thus rotated counterclockwise by about 21° with
respect to the starting ray for the Mach 1.0 case. This rotates the ray pattern for a nozzle with the acoustic velocity at the
exist counterclockwise by about 21°, so that the streamlines are less divergent.

In the above cases the normal to the exit plane of the nozzle was parallel to the axis of symmetry of the nozzle.
However, many diffusion pump designs involve nozzles for which this normal makes an angle with the axis of
symmetry for which we use the term beveled exit. Nozzles of this type are used in steam turbines, and Zerkowitz [73]
has analyzed the deflection of the jet pattern away from the nozzle axis. Applying his equations for the case of oil vapor
with specific heat ratio k = 1.02, we [67] obtain for a nonexpanding nozzle with exit plane making the angle ϕ = 65°
with a plane orthogonal to the axis of symmetry a jet deflection of the middle streamline by 36° as shown in Fig. 4.20.

Fig. 4.20
Oil-vapor jet deflection
from nonexpanding vertical
nozzle with beveled exit.

< previous page page_219 next page >


< previous page page_220 next page >
Page 220

In moving from the plane CD orthogonal to the nozzle axis at the outer rim to the exit plane CE the static vapor pressure
decreases from P1 to P2 and the vapor velocity increases. The ratio P2/P1 will be a function of the angle ϕ, and
Zerkowitz assumes that this ratio will be given by the PrandtlMeyer equation [Eq. (4.82)]. The velocity vector for the
vapor crossing the middle of the beveled exit CE can be shown to have the magnitude

where vs is the acoustic velocity at the throat and makes the angle with the nozzle axis as given by

For a bevel angle ϕ = 65° this gives v2 = 1.6vs and But there is a further deflection after leaving the exit
plane CE as shown by Fig. 7 in the Zerkowitz article. As the vapor crosses the exit plane CE it is free to expand on both
sides so that the pressure beyond the exit, P3 is lower than P2 and there is a driving pressure, P2 P3, acting in a
direction perpendicular to CE. The derivation of the equation [67] for this second deflection will not be given here, but
the result for oil vapor with k → 1 is the additional angle ε as given by

When the throughput is small so that the inlet pressure and forepressure are low, one can assume P3 small compared to
P2 so that ε = 16° and the total jet deflection along the middle streamline is as shown in Fig. 4.20.

Iliasova et al. [74] computed the vapor flow from the top-stage nozzle of a diffusion pump by a marching procedure in
the frames of parabolized NavierStokes equations. These authors also computed the gas flow through the pump inlet
and into the vapor stream by the Monte Carlo method and computed the net pumping speed to be expected for a given
inlet chamber design and vapor stagnation density. Sadykov and Figurov [75] investigated the vapor flow inside heated
nozzles experimentally with the help of Pitot tubes and in the free jet beyond the nozzle exit by pressure and

< previous page page_220 next page >


< previous page page_221 next page >
Page 221

temperature measuring instruments. They also calculated theoretically the vapor flow allowing for a boundary layer
using the equations of aerodynamics and rarefied gas dynamics. The mixing of the pumped gas with the vapor jet in the
backstreaming region and the forward pumping region was measured with pressure probes. Backstreaming rate from the
inlet stage was measured by weighing. Heating the nozzle to 25 K higher than the vapor conduit decreased the
backstreaming by 20%.

Attempts to explain the operation of diffusion pumps by kinetic theory considerations of the collisions of gas molecules
and vapor molecules were made by Riddiford and Coe [76], by Reichelt [77], by Florescu [63], and by Tóth [78].

4.4.4
Ultimate Pressure

Many authors have discussed the ultimate pressure obtainable with diffusion pumps which depends mainly on four
factors: (1) the back-diffusion [61] of gas from the fore-vacuum through the vapor jets to the high-vacuum side; (2) the
vapor pressure of the pump fluid [7, 79, 80] condensed on the walls of the inlet chamber or on baffles over the inlet; (3)
the evolution from the oil in the boiler of dissolved gas and low-molecular-weight (cracking) fragments [33, 34] from
decomposition of the pump oil; and (4) the outgassing [81] of the vacuum system beyond the pump. In addition, back
migration of oil vapor from a mechanical forepump may contaminate the system when the diffusion pump is not
operating and no valves are closed. This source of contamination can be reduced by the use of a purge gas during
shutdown [45].

Back-Diffusion of Gases from the Fore-Vacuum. Several authors have derived a first approximation to a more explicit
theoretical value of the back-diffusion coefficient β than that given by Gaede's simple Eq. (4.34). Matricon [82]
assumed that the vapor jet from a cylindrical nozzle in a single-stage mercury vapor pump could be simulated by
straight lines of flow originating at a point inside the nozzle exit at a distance from the exit plane of about one-third of
the exit diameter. However, to derive a formula for the β factor, as a first approximation, he assumed that in the region
of counterdiffusion of vapor and gas from the fore-vacuum the pressure gradient from the axis to the wall at any
distance x along the axis from the nozzle exit was negligible and that no condensation of mercury vapor occurred over
the distance x = 0 to x = L, where L is the useful length of the vapor jet when the forepressure is so low that the vapor
stream is not obstructed by the gas in the fore-vacuum. Then he obtains

where N is the concentration and v is the stream velocity of the vapor molecules independent of x, while D0 is the value
(in cm1 · s1) of the gas-vapor diffusion coefficient at unit vapor concentration. Using the Meyer formula for the
diffusion coefficient and the StefanMaxwell formula for the mean free path through a binary mixture, Matricon obtains

< previous page page_221 next page >


< previous page page_222 next page >
Page 222

Table 4.4. Calculated Diffusion Coefficients at 101308 Pa and 160°C


Units σ σ' M (g) M' (g) D0 D760
(108 cm) (108 cm) (1018 cm1·s1) (cm2/s)
Airmercury
3.11 5.67 29 200.6 2.88 0.17
Airmercury
3.72 4.50 29 200.6 3.30 0.19
AirOctoil
3.11 12 29 390 0.99 0.058
Hydrogenmercury
2.38 5.67 2 200.6 13.9 0.82
Hydrogenmercury
2.74 4.50 2 200.6 17.2 1.01
HydrogenOctoil
2.38 12 2 390 4.35 0.257
Heliummercury
2.2 3.5 4 200.6 19.6 1.16
HeliumOctoil
2.2 12 4 390 3.2 0.19
DeuteriumOctoil
2.38 12 4 390 3.1 0.18

D760 = D0/N760 = 5.9 × 1020D0 at 160°C

where σa = (σ + σ')/2 is the mutual collision diameter for the gas and mercury molecules, M is the molecular weight of the gas, M' is the
molecular weight of the vapor, and T is the absolute temperature of the gas within the vapor stream. The diffusion coefficient D760 (in cm2 · s1)
at 101,308 Pa (760 Torr) as used in Eq. (4.34) is obtained from

where N760 is the concentration of vapor molecules at 101,308 Pa and the specified temperature.

Values of the diffusion coefficient D760 for various gas-vapor combinations T = 160°C and 101,308 Pa as calculated by Dayton [61] from
Matricon's formula are listed in Table 4.4 together with the values of D0. The mean diameter σ' = 5.67 × 108 cm for mercury, σ = 3.11 × 108
cm for air, and σ = 2.38 × 108 cm for hydrogen are the values used by Matricon. Texts on kinetic theory give σ' = 4.50 × 108 cm for mercury
(150°C), σ = 3.72 × 108 cm for air, and σ = 2.74 × 108 cm for hydrogen. The value σ' = 12 × 108 cm for Octoil was estimated by Jacobs and
Kapff [83].

Gaede [2] measured an average value D760 = 0.18 cm2 · s1 for air through mercury vapor and an average value D760 = 0.66 cm2 · s1 for
hydrogen through mercury vapor at 140°C. It may be noted that D0 is proportional to T½ while D760 is proportional to T3/2.

Wertenstein [84] and Jaeckel [69] also arrive at Eq. (4.96) but use Enskog's [85] formula for D0,

This equation gives values of D0 about 10% lower than those from Eq. (4.97) using the same constants. Jaeckel cites σ = 2.7 × 108 cm for
hydrogen and 3.68 × 108 cm

< previous page page_222 next page >


< previous page page_223 next page >
Page 223

for air. The ratio of the diffusion coefficient for hydrogen through mercury to the coefficient for air through mercury is
about 4.8, while the ratio for hydrogen to air for diffusion through an organic vapor such as Octoil is about 4.4.

For a cylindrical nozzle with no loss from throat to exit the stream velocity, ve, of Octoil at the exit is the acoustic
velocity as given by Eq. (4.64). The molecular concentration of vapor at the exit will be the same as at the throat:

where Na is Avogadro's number and P0 is the vapor pressure in dyne · cm2. When P0 is expressed in Pa, one must use

If we now assume that the nozzle clearance is small compared to the diameter of the nozzle exit, then, as the vapor
flows from the cylindrical nozzle exit into a condenser with a cross section only slightly larger than the nozzle exit, the
vapor concentration will be only slightly decreased and the stream velocity only slightly increased, so that the product
Nv will be approximately equal to Neve = Ntvt. Then for an organic vapor, such as Octoil, Eq. (4.96) with P0 in Pa
becomes

with D0 in cgs units given by either Eq. (4.97) or Eq. (4.99) and assuming that the temperature T of the gas in these
equations equals the vapor temperature T0 and that very little condensation of the vapor jet occurs within the distance L
from the nozzle exit. As an example, at the relatively low temperature of 160°C or T0 = 433 K the vapor pressure, P0,
of Octoil is about 13.33 Pa; and using M' = 390, R0 = 8.31 × 107, Na = 6.02 × 1023, D0 = 1018 cm1 · s1, and L = 5 cm,
Eq. (4.102) gives β = 3.35 × 1029 for air through Octoil. Using D0 = 4.35 × 1018 cm1 · s1 for hydrogen through Octoil
gives β = 2.84 × 107, so that a forepressure of 1 Pa of hydrogen would limit the ultimate pressure to pu = 2.84 × 107 Pa.
Actually, there is considerable condensation of vapor over the distance L which is determined by the position of the
shock boundary created by the forepressure; and the ratio of nozzle throat area to body clearance area is about 1/2 to 1/3
for the last stage (nearest the fore-vacuum), so that the static vapor pressure at the body clearance area is about 1/2 to
1/3 of the pressure at the throat, which in turn is about 0.6 times the boiler pressure P0. Thus, even for the more
common temperature 180°C and boiler pressure P0 = 67 Pa, the β coefficient for hydrogen through Octoil can be of the
order of 107 or greater.

Theory of Pump Performance in the Forepressure Breakdown Region. A shock boundary is formed between the vapor
jet and the gas in the fore-vacuum when the forepressure rises above 1 Pa. The vapor expanding from the nozzle exit
travels with increasing supersonic velocity, and the static vapor pressure decreases as the jet expands and diverges
toward the pump walls. Any obstruction, such as a Pitot tube which has a small opening at the tip of a tapered nose,
placed in the vapor stream

< previous page page_223 next page >


< previous page page_224 next page >
Page 224

which permits the vapor to escape laterally will result in a compression shock in front of the obstruction as the flow
velocity perpendicular to the obstruction is reduced to zero. The law of conservation of momentum requires that a
pressure is exerted on the obstruction when the latter is fixed in position. According to the definition in Stodola and
Loewenstein, [68, p. 85], this pressure, known as the ''impact pressure" or "dynamic pressure," is the reading of a
manometer attached to the open channel in the Pitot tube. The magnitude of this pressure depends on the residual
momentum in the vapor stream after lateral dispersion and on the heat transferred to the Pitot tube. It has been found by
Löliger (and independently by Prandtl) that the impact pressure can be calculated by assuming that a compression shock
followed by an adiabatic compression takes place in front of the pointed end of the Pitot tube until the kinetic energy of
the flow is consumed. The shock arises from the sudden conversion of stream momentum into static pressure as the
vapor at supersonic velocity impinges on vapor moving with subsonic velocity.

The method of computing the impact pressure given in Stodola and Loewenstein involves the tracing of the "condition
loci" on the entropy diagram [68, Fig. 51]. The condition change leads from the initial condition A1 along the Fanno
and Rayleigh curves to the shock point A2, and from there adiabatically to the impact pressure Pi. The Fanno line is
given by combining the total energy per unit mass equation

where P1 is the initial static pressure and P2 is the final static pressure while E1 and E2 are the initial and final internal
energies per unit mass due to molecular rotation and vibration, with the equation of continuity

where a1 and a2 are cross sections in a tube of flow. For adiabatic expansion [86] we obtain

The Rayleigh line is obtained by combining the equation of continuity with the impulse-momentum equation

Since compression shock occurs in a very thin zone over the shock boundary, a1 and a2 will be approximately equal.
For perfect gases the algebraic solution for the intersection of the Fanno and Rayleigh lines is easily obtained, giving as
the static pressure after the compression shock the following equation:

< previous page page_224 next page >


< previous page page_225 next page >
Page 225

or for organic vapors with k → 1 we obtain

which is the total flux of stream momentum per unit area before the shock. Prandtl's solution for the stream velocity
after the shock is

For organic vapors with k → 1 we obtain

where vs1 is the value of the acoustic velocity under the initial conditions. There can be no shock unless the initial
stream velocity v1 exceeds the acoustic velocity vs1 for the initial state. From Eq. (4.110) it is obvious that if the initial
stream velocity is much larger than the acoustic velocity, the final velocity v2 after the shock will be a small fraction of
the acoustic velocity. After the shock the vapor is compressed adiabatically and the stream velocity falls from v2 to zero
while the static pressure rises to the final maximum value which represents the impact pressure, Pi.

When measuring the impact pressure at the throat of a nozzle, there is no compression shock, but the process of
adiabatic expansion is suddenly reversed to one of adiabatic compression. Then in Eq. (4.110) both v2 and v1 equal the
acoustic velocity, vs, at the throat and P2 = P1 = Pt. The measured impact pressure at the throat then equals the boiler
pressure, P0, for which the vapor stream velocity is zero. All streamlines are parallel at the throat and remain parallel to
the nozzle axis when the nozzle is cylindrical and short so that no energy is lost at the walls, but in expanding conical
nozzles of the DeLaval type the streamlines diverge and the impact pressure at the exit is less than the boiler pressure.
However, in an expanding nozzle with walls curved in a certain way, the Prandtl nozzle, all the streamlines at the exit
can be made parallel to the nozzle axis and to each other. However, for most diffusion pumps the expanding nozzles
have straight walls and the vapor acquires some momentum in directions perpendicular to the axis in passing from the
throat to the exit. In expanding from the acoustic velocity at the throat to supersonic velocity at the exit of a DeLaval
nozzle, some of the energy of random motion is converted into a flux of forward momentum parallel to the nozzle axis,
and the impact pressure at the exit is difficult to calculate. As shown by Fig. 49 in Stodola and Loewenstein [68] for the
expansion of steam through a DeLaval nozzle, the measured impact pressure at the exit is given approximately by

where At/Ae is the ratio of throat area to exit area and C is somewhat greater than 1.0.

By application of the definitions [87] of pressure in terms of momentum flux with respect to a specified surface, the
impact pressure can be related to the dynamic pressure as given by

< previous page page_225 next page >


< previous page page_226 next page >
Page 226

where n is the number of molecules in unit volume, m is the mass of each molecule,

is the average molecular velocity, and v is the stream velocity perpendicular to the surface. The impact pressure is then
obtained by omitting the last term on the right since molecules moving with velocity ua v in the negative direction with
respect to the surface normal will never strike the surface of the shock boundary. Since the vapor velocity at the throat is

one can substitute

Proceeding in this manner the author [67] obtained the following as an approximate formula for the coefficient C:

For example, for steam through the nozzle studied by Löliger [68, p. 85, Fig. 49] we have At/Ae = 0.238, k = 1.3, Pe/P0
= 0.05, Pt = 0.546P0, and C = 1.402, giving a computed impact pressure Pie = 0.33P0 close to the experimental value
0.34P0 obtained by Löliger. For oil vapor with k = 1.02, and assuming Te = Tt, the same nozzle gives ve/vt = 2.3183, Pe/
P0 = 0.0615, Pt/P0 = 0.602, C = 1.502, and Pie = 0.36P0. When Ae = At and k = 1.02, Eqs. (4.117) and (4.119) give ve/
vt = 1.1 and Eq. (4.116) gives C = 0.96 corresponding to Pie = 0.96P0.
< previous page page_226 next page >
< previous page page_227 next page >
Page 227

Fig. 4.21
Shock boundary near forepressure breakdown [15].

After the vapor leaves the nozzle the vapor diverges rapidly and the impact pressure for direct compression shock will
depend on a narrow bundle of streamlines intercepted at the shock boundary and the distance traveled from the nozzle
exit. For cylindrical or expanding conical nozzles it has been found experimentally that as a good approximation it can
be assumed that the vapor diverges in straight lines from a point source located inside the nozzle on the axis at a
distance depending on the vapor velocity at the exit. For oil vapor through cylindrical nozzles with velocity about equal
to the acoustic velocity at the exit, this distance appears to be about one-half the diameter of the nozzle exit, so that a ray
from the source point passing through the rim of the nozzle exit makes an angle of 45° with the nozzle axis and with the
wall of the pump casing. From Table 4.3 it can be seen that the static vapor pressure along this ray (ϕ = 135°) is about
0.035 times the boiler pressure, or 2.3 Pa when P0 = 67 Pa. For vapor diverging from a virtual point source the (direct)
shock boundary can be expected to form on a spherical surface perpendicular to each ray from the source. That this
actually occurs in pumps with cylindrical or conical nozzles is shown by glow discharge experiments reported by
Dayton [61] and by Kutscher [15]. The curved shock boundary shape is particularly noticeable just at the point of
forepressure breakdown as shown by Fig. 10 in Dayton [61] and by Fig. 4.21 adapted from Kutscher's article (which
includes some regions of oblique shock).

Since Eq. (4.111) indicates that the impact pressure is roughly inversely proportional to the ratio Ae/At of exit area to
throat area, as an approximation it can be assumed that the impact pressure at the wall of a pump with cylindrical nozzle
is inversely proportional to the square of the radius r from the hypothetical point source, where r is greater than the
radius to the rim of the nozzle exit. Then from Fig. 4.22, r2 = R2 + (x + xs)2, where R is the radius of the pump casing
and xs is the distance between the point source and the exit plane. The impact pressure at x will then be given by

< previous page page_227 next page >


< previous page page_228 next page >
Page 228

Fig. 4.22
Distance x of shock boundary at wall from plane of nozzle
exit with hypothetical point source of vapor flow.

where rn is the radius of the cylindrical nozzle and Pie is the impact pressure at the rim of the nozzle exit.

Assuming that at every point on the shock boundary before jet breakdown the impact pressure Pix equals the
forepressure F and that the impact pressure at the nozzle exit, Pie, is given by Eq. (4.111), then using xs = rn we obtain

where rt is the radius of the nozzle throat. From Fig. 4.22 the forepressure breakdown, F = Fb, occurs when the shock
boundary has been pushed back to x = x0 = R rn. Then for cylindrical nozzles we obtain

where B = πR2 is the body-clearance area and At is the nozzle throat area.

For most metal diffusion pumps the nozzle is constructed of concentric cylinders or cones so that the throat and exit
cross sections are annular with a width small compared to the mean radius. Then, as an approximation we can assume
that the vapor disperses in straight lines from a line source (circle of large radius) orthogonal to the plane through the
pump axis and located within the nozzle at a short distance from the exit of the order of one-half the width of the exit as
shown in Fig. 4.23.

The dashed line ti = R rB is the width of the body clearance area, B, and the angle φ gives the deflection of the
centerline of the jet caused by the presence of the streamliner (or apron) hj. For each stage, the breakdown forepressure,
Fb, at which the outer edge of the shock boundary is close to the point t, is determined by the ratio of the nozzle throat
area, Ae, to the body clearance area and the amount of the jet deflection φ. Then, as an approximation, the impact
pressure at a point on the pump casing wall located at the distance x from the plane of the nozzle exit will be given by
< previous page page_228 next page >
< previous page page_229 next page >
Page 229

Fig. 4.23
Dispersion characteristics of inverted annular nozzle.

where Pie is the impact pressure at the nozzle exit, xs is the distance of the source from the plane of the nozzle exit, rs is
the distance of the source from the pump axis, R is the radius of the pump casing, θ + φ is the angle between the vertical
line through the source and the ray from the source which passes through the rim of the nozzle exit, and n is an index
which equals 1 for annular nozzles in which the throat width w and the body clearance ti are small compared to the
casing radius R, but which approaches the value 2 as rs→ 0 corresponding to the previous case of a cylindrical nozzle
from which the vapor is assumed to issue from a point source. In general, the impact pressure at the casing wall will be
a maximum and forepressure breakdown will occur when the shock boundary reaches the point t where x = x0. At this
point, F = Fb = P1 and

so that the breakdown forepressure is


< previous page page_229 next page >
< previous page page_230 next page >
Page 230

It can be seen from the similar triangles in Fig. 4.23 that

so that, using Eq. (4.111) with C replaced by the empirical constant Kn, we obtain

which is equivalent to Eq. (4.122) when n = 2 and rs = 0. When n = 1 and R rn is small compared to rn, the ratio (rn rm)/
(R rm) will be approximately equal to the ratio of nozzle exit area, Ae, to the body clearance area, B. Then from Eqs.
(4.126) and (4.127) when n = 1 the general formula for the limiting forepressure is

Experimental data on various pumping stages has shown that the limiting forepressure is given by Eq. (4.128) with Kn
values which in general range from 0.5 to 1.1.

To calculate the β coefficient it is necessary to compute the rate of diffusion of gas from the fore-vacuum through the
vapor jet from the boundary uu' at the forepressure F to the boundary te with the high vacuum pressure p. The length of
the vertical diffusion path, Lj, between these boundaries can be calculated from the given geometry, and integration can
be performed over the upper boundary. However, it is obvious that most of the back-diffusion will occur where Lj is
smallest, or close to the casing wall through a barrier of thickness ut = x x0, where from the above equations we obtain

From Fig. 4.23 we have

Then the thickness of the diffusion barrier at the casing wall is

Substituting this for L in Eq. (4.34) together with mean values for the static vapor pressure P and the vertical component
of the streamline velocity, v, at the casing wall obtained by extrapolating Eqs. (4.117) and (4.119) in terms of the
expansion ratio Ae/B to the region x x0 would then give an approximate formula for β as a function of the forepressure
F. A more exact relationship could be obtained by extrapolating these equations from the nozzle exit to the cross section
at x to obtain the vertical component of stream velocity near the wall, vx(x), as a function of the distance x and the static
vapor pressure near the wall, P(x), as a function of x and then substituting

< previous page page_230 next page >


< previous page page_231 next page >
Page 231

these functions in place of v and P in Eq. (4.33), integrating from x x0 to x, and then using Eqs. (4.129) and (4.130).

Vapor Pressure of Fluid Condensed in the Inlet Chamber or on Baffles. The vapor pressure of the fluid condensed on
walls near the inlet of the pump (and baffle) can be reduced by proper design of fractionating pumps and baffles, as well
as the use of cold traps. For fractionating pumps with water-cooled baffles the ultimate pressure limit due to pump fluid
vapor will be slightly less than the vapor pressure in Pa at 25°C as given in the next-to-the-last column in Table 4.1 for
the given pump oil. Use of the more stable fluids, such as Convalex-10, Santovac-5, or DC-705 with the addition of a
cold trap cooled to below 40°C will usually allow ultimate pressures below 1 × 108 Pa when the forepressure is
sufficiently low. It has been reported [88,89] that using a system with an alumina or zeolite trap and gas purge
provisions to reduce contamination from the oil in mechanical forepumps and using a fractionating diffusion pump with
polyphenyl ether pump fluid, a cold cap, a cryotrap chilled with liquid nitrogen or simply by circulating methanol at 75°
C, and a high-vacuum gate valve with proper operating procedures can produce clean vacua as free of hydrocarbons and
CO as those produced with conventional (hydrocarbon lubricated) turbopumps.

< previous page page_231 next page >


< previous page page_233 next page >
Page 233

Part II
Molecular Drag and Turbomolecular Pumps

Jörgen Henning

4.5
Molecular Drag Pumps

A molecular drag pump is a vacuum pump in which gases are pumped by momentum transfer from a rapidly rotating
solid surface to the gas molecules. The rotor impulse is transmitted to the particles by superposition of the thermal
velocity of the colliding particles with the velocity component of the moving rotor surface (Fig. 4.24). The nondirected
motion of the particles is changed to a directed motion in the pumping process. When the mean free path of the particles
between the collisions with other particles is larger than the spacing between the rotating and the stationary surfaces
(molecular flow range, typically < 0.1 Pa), particles collide primarily with the rotor, resulting in an efficient pumping
process, and there is no interacting influence of the different gases.

In the laminar flow range (typically > 0.1 Pa) the action of the rotor surface is restricted by the frequent collisions
between particles. Therefore, a molecular drag pump is not capable of pumping gases against atmospheric pressure and
must be backed by an adequate roughing pump.

The first pump of this type was introduced in 1913 by Gaede [90], when he presented his "Molecular Drag Pump" using
a multistage cylindrical rotor design, with parallel slots around the circumference of the rotor, into which project
extensions from the outer casing.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5 © 1988 John Wiley & Sons, Inc.

< previous page page_233 next page >


< previous page page_234 next page >
Page 234

Fig. 4.24
Principle of operation of
molecular drag pump.

A modified form of the molecular drag pump, designed by F. Holweck [91] in 1923, used a smooth cylindrical rotor
(drum) inside a housing with spiral channels, decreasing in depth toward the exhaust side. The intake was in the middle
of the housing and the gas was dragged to both ends of the housing, making use of twice the pumping speed.

In 1940 Siegbahn [92] described his disc-type molecular drag pump using a disc-rotor. A circular disc rotates inside a
container consisting of two side plates with spiral grooves. These grooves are deeper at the periphery (inlet) and
gradually decrease in depth toward the center (exhaust).

Up to the 1970s, only few researchers were interested in these "early" molecular drag pumps because of their relatively
low pumping speed and their questionable reliability; furthermore, in those days there was no real demand for these
pumps. In order to attain low ultimate pressures with these pumps the clearances between rotating and stationary parts
were made a few hundreths of a millimeter only. Therefore, any change in temperature or intruding solid particles could
result in a failure of the pump, caused by a ceased rotor. However, recently, the basic ideas of Holweck (drum), Gaede
(slotted rotor), and Siegbahn (disc) were picked up again successfully in the design of modern pumps (molecular drag
pumps and combined molecular drag and turbomolecular pumps), in order to attain extremely low pressures and/or to
make use of simple dry roughing pumps.

4.5.1
Theoretical Considerations and Performance Data

The principal design of the pumping stage of a molecular drag pump is shown in Fig. 4.25. It does not make any
difference if the pumping channel is located in the rotating or the stationary part of the pumping stage. Different
methods have been developed to describe the pump performance of molecular drag and turbomolecular pumps. The
continuum methods [93], and methods using statistical calculations [94] agree in the basic results concerning the pump
performances. The values depend on the transmission probabilities (P) of the particles moving in and against the
pumping direction. For the molecular flow, these probabilities depend on the ratio v/u of the rotor speed v and the mean
thermal velocity u of the gas particle, but not on the pressure.

For a pumping channel, in the range of free molecular flow, the compression and the net flow rate (pumping speed) can
be calculated from the law of conservation of molecules by assuming that the velocity distribution function of the gas on
both sides

< previous page page_234 next page >


< previous page page_235 next page >
Page 235

Fig. 4.25
Principle design of a molecular
pumping stage: l, channel
length; h, channel height; b,
channel width; s, gap between
rotor and housing; FA, entrance
area; FB, exit area.

of the channel is Maxwellian. For steady-state flow, the net flow rate W (Ho-coefficient) from side A to side B can be
derived from

where PAB and PBA are the probabilities that a molecule incidenting from side A or B will be transmitted through the
channel to side B or A, and FA and FB are the entrance area of side A and the exit area of side B, respectively. K = nB/
nA is the compression (n is the number of molecules) [95].

In order to calculate the performance of a molecular pumping stage, two extremes, the maximum compression Kmax (at
zero flow) and the maximum pumping speed Smax (for equal pressure on both sides of the pumping stage) must be
known.

From Eq. (4.132) the following relations can be derived:

For maximum compression, no gas flow (W = 0):


The important problem to solve remains the calculation of PAB and PBA, which, for example, was solved by Kruger
and Shapiro [94] using Monte Carlo techniques. As a result it was found, in agreement with calculations using different
methods [93], that Kmax of a molecular drag pump is exponentially dependent on the rotor speed v,

< previous page page_235 next page >


< previous page page_236 next page >
Page 236

a pump-specific factor g (rotor and stator geometry), and the square root of the molecular weight M of the particles
pumped:

For maximum pumping speed (K = 1):

Smax is proportional to the product of a specific pump factor G (rotor and stator geometry) and the rotor speed v; it does
not depend on the pressure and the kind of gas pumped, because the molecular arrival rate at the inlet is proportional to
the thermal velocity of the gas:

An operating pumping stage works in the region between these two extremes. For real pumping conditions (gas
throughput Q = Spinlet) and different pressures on both sides of the pumping stage (K = poutlet/ pinlet) the following
relation exists between compression K and real pumping speed S of a single pumping stage:

Therefore, the pumping speed of a molecular drag pump depends on the compression and the pumping speed Sv of the
backing pump, as given in Eq. (4.138). From there it can be derived that if Kmax for any gas is small, the pumping
speed for this gas is a function of the ratio S/Sv. For practical purposes a molecular drag pump with low Kmax for H2
needs a larger backing pump to pump H2 effectively (low S/Sv ratio).

Due to the inlet conductance of the pumping channel, there is a limitation of Smax and the actual pumping speed S of a
molecular drag pump becomes a function of the ratio v/u and therefore it will depend on the molecular weight of the gas
pumped.

More exact calculations take into account the losses through gaps between the rotating and the stationary parts.
However, the main relation, Eq. (4.138), remains the same.

As for any vacuum pump, the ultimate pressure pf, that a molecular drag pump can attain can be calculated from the
pressure pv at the outlet side of the pump by dividing it by the maximum compression Kmax:

< previous page page_236 next page >


< previous page page_237 next page >
Page 237

Fig. 4.26
Principle of multistage molecular
drag pump (Balzers Pfeiffer).

The increase of Kmax with the molecular weight M means that heavy molecules are highly compressed and have a low
backflow probability; this is the reason for the ''clean" vacuum without contamination by oil vapors and hydrocarbons.

4.5.2
Design Considerations

Stand-alone molecular drag pumps are designed on the basis of the Holweck principle [91]. These pumps in general
include two different sections: (1) an entrance section with a row of turbomolecular pump blades for maximum
conductance to ensure a high pumping speed and (2) subsequent molecular pumping stages (up to 5) having in principle
a drum design with a multiribbed structure to ensure efficiency and high compression (Fig. 4.26). These pumps are
available with mechanical rotor bearings or a combination of mechanical and magnetic bearings. The lubricant used for
the mechanical bearings is either grease or oil. More details of the design of molecular drag pumps are given in Section
4.6.2.

4.5.3
Typical Performance Data of Commercial Pumps

The working range for molecular drag pumps is, depending on the manufacturer, between 103 and 103 Pa. Due to the
high admissible foreline pressure, molecular drag pumps can use small, simple, dry, and inexpensive fore-vacuum
pumps (e.g., membrane pumps) with an ultimate pressure of <103 Pa.

< previous page page_237 next page >


< previous page page_238 next page >
Page 238

4.5.3.1
Compression

Depending on the manufacturer's design, molecular drag pumps have the following compression values: H2, 2 × 102 to
1 × 103; He, 1 × 103 to 2 × 104; N2, 1 × 107 to 1 × 109.

4.5.3.2
Pumping Speed

Molecular drag pumps are available in a pumping speed range from 7 to 300 liter · s1.

Type designations of molecular drag pumps often use the pumping speed for nitrogen. Therefore the pumping speed for
N2 is used as the 100% value. This value is compared to the pumping speed data for H2 and He. From the catalogues of
different manufacturers: H2, 40% to 56%; He, 53% to 67%.

4.5.3.3
Ultimate Pressure

Depending on how many molecular pumping stages are used in series, the attainable ultimate pressure is in the range of
105 to 103 Pa.

4.6
Turbomolecular Pumps

The turbomolecular pump invented by Becker [96] in 1957 (Fig. 4.27) became commercially available in 1958. Since
then it has become very popular in every field of high- and ultrahigh-vacuum technique, due to a clean, consistent, and
predictable vacuum created, the easy operation, and the advanced degree of operating reliability. The turbomolecular
pump is the only mechanical vacuum pump which, together with a roughing pump, can attain ultimate pressures in the
range below 108 Pa. It is a bladed molecular turbine that compresses gases by momentum transfer from the rapidly
rotating blades of the rotor wheels to the gas molecules. It is working on the same principle as the molecular drag pump.

When the mean free path of the particles is larger than the spacing between rotor and stator (molecular flow range in a
turbomolecular pump, typically <101 Pa), particles collide primarily with the rotor, resulting in an efficient pumping
process, and there is no interacting influence of the different gases. In the laminar flow range (typically in a
turbomolecular pump > 101 Pa) the action of the rotor is restricted
Fig. 4.27
Principle design of turbomolecular pump [96].

< previous page page_238 next page >


< previous page page_239 next page >
Page 239

by frequent collisions between the particles. Therefore, a turbomolecular pump is not capable of pumping gases against
atmospheric pressure and must be backed by an adequate roughing pump.

The Becker design avoided the obvious disadvantages of the "early" molecular drag pumps (relative low pumping
speed, questionable reliability): The turbomolecular pump is composed of a series of wheels with coaxial blades,
alternately fixed and moving. In each wheel the sides of the blades are inclined with respect to the axis, in one direction
for the blades of the moving wheels and in the other for the blades of the fixed wheel. The moving wheels have a high
rotational speed so that the peripheral speed of the blades (up to 500 m/s) is of the same order of magnitude as the speed
of the molecules of the pumped gas.

The distances between these wheels were in the range from several tenths to a few millimeters. The channels between
the inclined blades of the wheels act like elementary molecular drag pumps, similar to the Gaede molecular drag pump.
All channels (~2050) on one wheel are connected in parallel and together can yield a high pumping speed up to several
thousand liters per second.

4.6.1
Theoretical Considerations and Performance Data

A milestone for the understanding of turbomolecular pumps was the work published by Kruger and Shapiro in 1960
[94] on blading geometry of axial-flow molecular turbines in the molecular flow range. However, most later
calculations of the pumping speed characteristics made by several authors [93, 94, 97] included complicated statistical
mathematics or, from the user's standpoint, unknown geometrical factors with different formulas for light and heavy
gases. In 1983, simple pumping speed calculations were published for gases with molecular weights between 2 and 50
[98].

The main vacuum data of a turbomolecular pump, the compression, and the pumping speed can be calculated from the
data of single rotor and stator wheels. A rotating wheel with blades pumps molecules from one side to the other, and
some particles move opposite to the primary flow direction through the blades. The structure of most common
turbomolecular pumps is shown in Fig. 4.28. In order to calculate the performance of a single wheel, the same
considerations as with the molecular drag pump can be used, by replacing the wording "molecular pumping stage" by
"single wheel." Again two extremes, the maximum compression Kmax (at zero flow) and the maximum pumping speed
Smax (for equal pressure on both sides of the wheel), must be known. For a turbomolecular pump the transmission
probabilities in the molecular flow depend on the ratio v/u of the blade speed v and the mean thermal velocity u of the
gas particle, and additionally on the blade angle α, but not on the pressure. For Kmax of a turbomolecular pump Eq.
(4.134) applies and is exponentially dependent on the blade speed v, a specific pump factor g (rotor and stator
geometry), and the square root of the molecular weight M of the particles pumped.

In agreement with Eq. (4.136), Smax is proportional to the product of a specific pump factor G (rotor and stator
geometry) and the blade speed v, and it does not depend on the pressure and the kind of gas pumped:

An operating wheel works in the region between these two extremes. For real pumping conditions (gas throughput Q =
Spinlet) and different pressures on both sides of the wheel (K = poutlet/ pinlet) the relation given in Eq. (4.138) applies.
More exact calculation takes into account the gap between the rotating wheel's outer diameter

< previous page page_239 next page >


< previous page page_240 next page >
Page 240

Fig. 4.28
Structure of turbomolecular blades [98]: α, blade angle; d, blade thickness;
h, distance between blades.

and the inner diameter of the stator housing and the variation of the blade geometry with the radius. However, the main
Eq. (4.138) remains the same.

A turbomolecular pump is assembled from several wheels in series with different blade geometries. Each wheel can be
regarded as a separate pump. If the pumping speed Sv of the next wheel downstream is known and the throughput is
constant, K in Eq. (4.138) can be replaced by S/Sv; and beginning with the pumping speed Sv of the backing pump and
using the values Smax and Kmax for the first wheel at the forevacuum side, the real pumping speed of this wheel can be
calculated. The formula can be used as a recurrence formula to calculate step by step the pumping speed of the whole
pump in the molecular flow range.

Then the pumping speed of a turbomolecular pump depends on the compression and the pumping speed Sv of the
backing pump, as given in Eq. (4.138). From there it can be derived that if Kmax for any gas is small, the pumping
speed for this gas is a function of the ratio S/Sv. For practical purposes a turbomolecular pump with low Kmax for H2
needs a larger backing pump to pump H2 effectively (low S/Sv ratio).

Due to the inlet conductance of the blade area of the wheel, there is a limitation of Smax and the actual pumping speed S
of a turbopump becomes a function of the ratio v/u [98], and therefore it will depend on the molecular weight of the gas
pumped:

where F is the pumping area of the wheel, α is the blade angle, ds is the blade thickness, h is the blade distance, and k is
the trapping probability ≈1 [98]. The pumping speed
< previous page page_240 next page >
< previous page page_241 next page >
Page 241

and the compression of a turbomolecular pump decrease at pressures of above 101 Pa, caused by the interaction of the
particles with one another, as the mean free path is no longer larger than the blade distance and the blades of the rotor
wheels no longer are in the molecular flow range. This value corresponds to a foreline pressure of 110 Pa.

Again, as for any vacuum pump, the ultimate pressure pf a turbomolecular pump can attain can be calculated from the
pressure pv at the outlet side of the pump by dividing it by the maximum compression Kmax using Eq. (4.139). The
increase of Kmax with the molecular weight M means that heavy molecules are highly compressed and have a low
backflow probability; this is the reason for the "clean" vacuum without contamination by oil vapors and hydrocarbons.
The smaller compression for light gases is the reason that the residual gas atmosphere of a turbomolecular pump
consists mainly of H2. This, however, holds true only for a "clean" system with metallic flange seals. In the case of
viton or rubber seals the ultimate pressure and the residual gas look differently [99].

4.6.2
Design Considerations

4.6.2.1
Rotor and Stator Geometry

The pumping speed and the compression of a turbomolecular pump depend strongly on the rotor geometry and speed.
Starting from the 1958 original geometry of the rotor and stator wheels, new wheel geometries have been developed.
These geometries together with increased rotor speeds allowed much smaller and lighter rotors for the supercritical
speed range.

The rotor and stator stages nearest the high-vacuum inlet are designed to serve a purpose different from those near the
outlet. The flow through each stage is constant, or, stated another way, the product of pressure times pumping speed is a
constant. The blades nearest to the inlet of a turbomolecular pump are designed to have as high a pumping speed as
possible, whereas the blades nearest to the foreline port are designed for high compression. The opening angles of the
blades are decreased from the high-vacuum side of the fore-vacuum side with the aim of optimizing the compression
and the pumping speed. For economic reasons it would be impractical to make each stage different from its neighbor. A
compromise results in groups of two to four types of blades, in which each is designed for a particular speed and
compression ratio.

The methods of manufacturing the rotors and stators have an influence on the pumping speed and compression. Rotors
can be made of individually machined wheels which are heat shrunk to the rotor shaft, by machining complete groups of
wheels from a single block of material or by manufacturing the rotors using spark erosion. The individually machined
wheels offer the advantage of making them "optically opaque", maximizing them for the compression. The other
methods of rotor production yield wheels with less opaqueness and lower compression, maximizing them for pumping
speed.

The stators are either manufactured from individually machined wheels or from stampings.

Meanwhile the first commercial turbomolecular pumps were dual-flow ("horizontal") pumps [96] having a double-
ended rotor, pumping the particles from a central inlet toward both sides and reuniting the gas flow in a common
foreline; single-flow ("vertical") turbomolecular pumps, using single-ended rotors, became available in

< previous page page_241 next page >


< previous page page_242 next page >
Page 242

1969 [100]. The double-ended rotor design allows a more stable bearing design which is advantageous for easy
balancing and lower vibration levels. The single-flow design has little conductance losses between the inlet flange and
the rotor, whereas the dual-flow design suffers losses from the inlet to both sides.

Today only a few models of commercial turbomolecular pumps still use the dual-flow design.

4.6.2.2
Rotor Suspension

The development of the turbomolecular pump was quite spectacular concerning the reduction in size. This was possible
due to (a) an increase of the circumferential (tip) speed from 150 m/s in 1958 to approximately 500 m/s today and (b)
changes of the rotor geometry. These high tip speeds relate to high rotational speeds of the rotors exerting high loads to
the rotor's suspension.

Today most of the commercial turbomolecular pumps are equipped with lubricated mechanical rotor bearings, or a
combination of permanent magnet bearings at the high-vacuum side and a lubricated mechanical bearing at the fore-
vacuum side. Depending on the wheel diameter the rotational speed of the rotor goes up to 90,000 rpm. These high rpm
have become possible because of the advances in bearing and balancing technology, without sacrificing the high
reliability of the turbomolecular pump. Today, high-precision ball bearings are available, which, when specially tuned
to a certain turbomolecular pump rotor at comparable radial and axial loads, even at much higher rpm, have a superior
lifetime compared to bearings of older design.

"Ceramic" bearings (ceramic balls) are widely used today. The ceramic balls exert lower centrifugal forces and lower
stress on the races than metal balls, are harder and more temperature-stable, and, therefore, have a stable spherical shape
and minimal wear on balls and race. Their surface is smoother, leading to less friction, and the pairing of different
materials (ceramic balls/steel races) avoids micropitting. Therefore these bearings are more reliable even under
lubrication-starved conditions.

4.6.2.3
Lubrication of Mechanical Bearings

Three main requirements have to be fullfilled by the lubricant used for the mechanical rotor bearings: (1) The lubricant
has to cool the rotor bearings since these are running under fore-vacuum conditions, and a heat transfer from the inner to
the outer race is possible only by the lubricant; (2) a low vapor pressure is required; and (3) good lubrication properties
at high speed are necessary. Today most of the lubricants used have a synthetic base.

Even at low rotor weights a minimum fluid flow through the bearings for heat transfer is necessary. Many smaller
turbomolecular pumps use a wick lubrication system or, together with "ceramic" bearings, a grease lubrication system,
while most larger turbomolecular pumps have pump systems circulating the oil.

4.6.2.4
Magnetic Rotor Suspension

After some unsuccessful experiences with commercial "gas bearing" turbomolecular pumps [101], today commercial
magnetic bearing turbomolecular pumps are available in which one, two, three, or all of the necessary five degrees of
freedom of the rotor are actively controlled either by an analog or by a digital electronic system. The position of the
rotor spindle is monitored by sensors, and the electronics corrects the spindle position to the data supplied by the
sensors. There is no mechanical friction and hence no wear. The price of these magnetic pumps is still considerably
higher as compared to standard ball-bearing turbomolecular pumps, limiting their general use.

< previous page page_242 next page >


< previous page page_243 next page >
Page 243

4.6.2.5
Balancing and Vibration

The dynamic balancing of the rotor of a turbomolecular pump is of great importance to minimize vibration and noise
levels, which are related to the mechanical bearing lifetime. Due to the high rotational speed of turbomolecular pump
rotors, the centrifugal forces associated with the residual unbalance (material inhomogeneities, radial bearing play,
geometrical imperfections) attain considerable values and transmit vibrations to the body of the pump.

The balancing process modifies the mass distribution of the rotor by adding or taking off material in order to bring the
rotational axis as close as possible to the principal axis of inertia. To further reduce the influence of the residual
unbalance, the mechanical rotor bearings within a turbomolecular pump are held in elastic "antivibration" rings which
effectively dampen the residual unbalanced forces.

Dynamic multifrequency balancing of the rotor is generally done in several balancing planes. This relative complex
procedure in the last few years has become simplified by the use of computers and dedicated software. Modern
turbomolecular pumps have very low residual vibration amplitudes, below 0.02 µm. These low values are necessary for
the use of a turbomolecular pump with vibration-sensitive instruments, such as mass spectrometers and electron
microscopes.

4.6.2.6
Rotor Materials

Most of the commercial turbomolecular pumps use high-strength Al alloys as rotor material. Compared to other high-
strength materials, such as Ti and steel alloys, these Al alloys are lighter, are much easier to machine, and have
sufficient thermal stability for the operational temperatures even at the typical turbomolecular pump bakeout cycles.
The use of ceramics (Si3N4) has been reported for the use of turbomolecular pumps in very strong magnetic fields
[102]. A typical maximum stress in the root of the rotor blades of a high-speed turbomolecular pump rotor at operating
conditions is in the range between 50 and 150 N/mm2, well below the 0.2% elongation limits of adequate Al alloys
[103].

4.6.2.7
Drive Systems

The drive rotor of a turbomolecular pump is an integral part of the turbomolecular pump rotor and together with its
drive stator can be located in the fore vacuum area. Today three different motor systems are used: DC motor, AC motor,
and hysteresis motor. The somewhat more expensive DC motor has lower energy consumption and energy losses than
the other motors. The motors are driven by solid-state frequency converters. Some of these converters can operate the
turbomolecular pump at variable rotor speeds.

For special application (e.g., radiation) motor-driven frequency converters are used.

4.6.3
Applicational Considerations

4.6.3.1
Venting

After a power shutdown a turbomolecular pump should be vented; otherwise it will be contaminated by oil vapors as a
result of pressure equalization between exhaust and inlet. By venting with a dry gas to atmospheric pressure, this effect
can be suppressed and a contamination of the vacuum system avoided. Most of the turbomolecular pumps, with the
exception of the larger pumps, can be vented at the high-vacuum side at full rotor speed. The venting air bursting into
the pumps puts a force of 1 kg on every cm2 of blade surface, which for larger areas represents a high load to the rotor
suspension. If the vacuum system cannot be vented
< previous page page_243 next page >
< previous page page_244 next page >
Page 244

directly, it is advisable to vent the turbomolecular pump by means of a special vent port which opens into the
compression stages. Venting a turbomolecular pump from the contaminated foreline should be avoided, by all means.

4.6.3.2
Baking

In order to attain low ultimate pressures in the high-vacuum and ultrahigh-vacuum range, the internal surfaces of a
turbomolecular pump (rotor, stator, housing) can be baked out. Due to the temperature sensitivity of the Al alloy used
for the rotors, there is a limit to the maximum value of the baking temperature. This maximum has to be well below the
critical temperature for the rotor alloy with respect to its strength. Typical baking temperatures are in the range of
100140°C, using the bake-out systems supplied by the pump's manufacturer.

4.6.3.3
Cooling

In order to dissipate the frictional heat from the bearing areas, the motor, and the heating by gas throughputs at high
pressures, the bearing areas of turbomolecular pumps have to be cooled. Meanwhile, for small pumps, convection
cooling is sufficient, whereas larger pumps are equipped with fans for cooling. For many pumps, water cooling is
common.

4.6.3.4
Operation in Magnetic Fields

Turbomolecular pumps with their metallic rotors in the presence of magnetic fields experience induction of eddy
currents which, due to heating, can cause serious problems concerning the material strength and the tolerances between
rotating and stationary parts in the turbomolecular pump [104]. The eddy-current loss ∆P, is the amount of energy
transformed by the eddy current into heat, and can be represented by the Eq. (4.141):

The eddy-current loss ∆P is proportional to the square of the magnetic flux B, proportional to the square of the rotor's
rotational frequency f, and inversely proportional to the resistivity δ of the rotor material. In addition ∆P is dependent on
the shape of the rotor. Since in actual turbomolecular pumps the length of the rotor greatly exceeds its diameter, the
maximum heating of the rotor results from the component of the magnetic field in a radial direction.

Therefore, turbomolecular pumps with metallic rotors can be used in magnetic fields only if certain maximum values of
the magnetic flux density will not be exceeded. These maximum values are specified by the manufacturers and, for
static magnetic fields perpendicular to the axis of rotation, typically are in the range of 1030 mT. With pulsed magnetic
fields, higher maximum values of Bmax are admissable. If the magnetic field is applied for a time t1 and off for a time
t2 and since the heating of the rotor is proportional to the square of B, the following relation for Bmax(pulse) for pulsed
magnetic fields applies:

In case of higher flux densities the turbomolecular pump will have to be shielded magnetically, which, however, can
create problems with the distribution of the magnetic field itself. An experimental turbomolecular pump with a ceramic
rotor has been reported which has been tested with a magnetic flux density of 460 mT [102].

< previous page page_244 next page >


< previous page page_245 next page >
Page 245

In this context it has to be kept in mind that the electronic drive systems and the control systems (magnetic suspension)
of turbomolecular pumps create magnetic stray fields in their vicinity. These stray fields can influence other electronic
and magnetic systems. Depending on the position relative to the turbomolecular pump, typical values for these stray
field are in the range of 150 µT.

4.6.3.5
Pumping Corrosive Gases

Modern manufacturing techniques require pumping processes where the ultimate system pressure has to be in the high-
vacuum range and corrosive gases have to be handled (e.g., below 101 Pa). At these pressures the rotor material will not
be attacked by the corrosives. However, the higher gas density at the fore-vacuum side in the rotor bearing area requires
the use of special lubricants (Fomblin®, Krytox®, etc.) which are corrosion-resistant.

In the semiconductor industry, Si and Al are etched in plasma etch machines (e.g., at pressures above 101 Pa). To avoid
etching and corrosion of the turbomolecular pump's Al rotor and other parts, all internal parts of the turbomolecular
pump having contact with the corrosives (e.g., Cl2, BCl3, CCl4) are either made of corrosion-resistant materials or are
specially coated (e.g., Ni-plated). Such coatings of the rotor can reduce the attainable ultimate pressure of a
turbomolecular pump to 5 × 106 Pa. Besides using special lubricants, for further protection the turbomolecular pumps
(for plasma etching) are equipped with a purge gas system which admits inert gas (N2, Ar, typically 30 sccm) directly
into the bearing area of the turbomolecular pump, where it creates a directed flow to the exhaust side of the pump. This
flow prevents the corrosive gases from entering the bearing area.

In some processes the reactive molecules of the etch gas from AlCl3 which is pumped off with the etch gas, and at
temperatures of 69°C solidifies inside the pump. To avoid this, the pumps will be heated to keep the temperature level
above this temperature. The use of turbomolecular pumps with magnetically suspended rotor will reduce the above-
mentioned problems.

4.6.3.6
Pumping Toxic or Radioactive Gases

An additional range of problems has to be solved in the case of a turbomolecular pump pumping toxic or radioactive
gases (e.g., tritium in plasma fusion installations). Due to the safety hazards involved, the turbomolecular pump has to
be extremely tight. With a "tritium" turbomolecular pump, all seals to the atmosphere are made of metal. The integral
leak rate is below 107 Pa · liter · s1.

4.6.3.7
Turbomolecular Pumps in Combination with Other Pumps

In order to increase the pumping speed of a turbomolecular pump for hydrogen or water vapor and make use of its
constant throughput and pumping speed, turbomolecular pumps have been combined with Ti sublimation pumps [99]
or, recently, with cryopumps [105] (see Chapter 9).

4.6.4
Performance Data of Commercial Pumps

The variations in the performance of turbomolecular pump by different manufacturers depend on their actual design
(blade design, rotor staging, etc.).

< previous page page_245 next page >


< previous page page_246 next page >
Page 246

4.6.4.1
Compression

Figure 4.29 shows typical values for the maximum compression Kmax of a turbomolecular pump for different gases as a
function of the foreline pressure.

Depending on the manufacturer, turbomolecular pumps have the following compression values: H2, 1 × 102 to 1 × 105;
He, 5 × 102 to 1 × 107; N2, 5 × 106 to 1 × 1010.

4.6.4.2
Pumping Speed

Figure 4.30 shows the pumping speed of a turbomolecular pump for different gases as a function of the inlet pressure.

Turbomolecular pumps are available in a pumping speed range from 35 to 25,000 liter · s1.

Type designations of turbomolecular pumps often use the pumping speed for nitrogen; therefore the pumping speed for
N2 is used as the 100% value. This value is
Fig. 4.29
Maximum compression of turbomolecular
pump as a function of the foreline
pressure (Balzers Pfeiffer).

< previous page page_246 next page >


< previous page page_247 next page >
Page 247

Fig. 4.30
Pumping speed of turbomolecular pump as a function
of the inlet pressure (Balzers Pfeiffer).

compared to the pumping speed data for H2 and He from the catalogues of different manufacturers: H2, 34133%; He,
60133%.

4.6.4.3
Ultimate Pressure

The ultimate pressure of a commercial turbomolecular pump is generally between 108 and 107 Pa, using metal flange
seals and a two-stage rotary backing pump.

4.7
Combined Molecular Drag and Turbomolecular Pumps

In the past several years, different kinds of combined molecular drag and turbomolecular pumps have been developed to
improve the high-pressure performance of the turbomolecular pump and to attain very low ultimate pressures.
Combinations of a turbomolecular and a molecular drag pump basically work on the same principle. Momentum is
transferred from a rapidly moving rotor surface to the particles to be pumped. Therefore the theoretical considerations
are the same.

In 1975, Schittko and Schmidt [106] tried to combine a Holweck molecular drag pump with a turbomolecular pump in
order to raise the compression for light gases. In 1973, Sawada and Tanigushi [107] proposed the concept of combined
molecular drag pumps and turbomolecular pumps in which a turbomolecular pump and a drag pump stage are mounted
in series on a single shaft (''Compound Molecular Pump") for the purpose of expanding the working pressure to a higher
pressure range. Maurice [101] developed in 1974 such a pump ("Hybrid turbomolecular pump"). However, in these
attempts the gaps between rotary and stationary parts were still extremely narrow, causing poor reliability at high-speed
rotation.

With the progress in semiconductor manufacturing there has been an increasing demand for "clean" vacuum pumps
with high throughputs and capable of being used in the process range of 1103 Pa. This demand led to the development
of combined turbomolecular pumps and molecular drag pumps in the 1980s. This combination has two effects: H2
compression is increased from typically 103 to 105, allowing ultimate pressures of below 108 Pa to be obtained and,
more significantly, allowing the

< previous page page_247 next page >


< previous page page_248 next page >
Page 248
pressure in the backing line to rise above, for example, 103 Pa and higher, thereby enabling it to be backed with a dry
membrane pump. In comparison, turbomolecular pumps, without the molecular drag pump have to be backed by a two-stage
rotary pump to ensure that the compression stages do not operate under laminar flow conditions.

With an integral molecular drag pump, oil is still present in the turbomolecular pump bearings and in the drive mechanism of
the diaphragm pump (separated from the high-vacuum space), but the elimination of the oil-sealed rotary pump is a substantial
improvement.

4.7.1
Design Considerations

This pump design combines the advantages of a multistaged turbomolecular pumps and molecular drag pumps: The
turbomolecular pump section provides a high pumping speed and low ultimate pressures, and the molecular drag pump section
provides a high compression and extends the fore-vacuum tolerance up to 102 to 5 × 103 Pa. These capabilities allow the use of
simple small, dry, and inexpensive backing pumps (e.g., membrane pumps) and therefore allow a dry evacuation from
atmospheric pressure down to 109 Pa.

These pump combinations are available in several different design configurations:

TMP stages + multistage MDP (disc type, derived from Gaede [90])

TMP stages + multistage MDP (drum type, derived from Holweck [91])

TMP + multistage MDP (drum type, derived from Holweck [91])

TMP stages + multistage MDP (disc type, derived from Siegbahn [92]).

4.7.2
Typical Performance Data for Commercial Combined Molecular Drag and Turbomolecular Pumps

4.7.2.1
Compression

H2, 2 × 103 to 2 × 107; He, 7 × 103 to 2 × 109; N2, 5 × 107 to 1 × 1012

4.7.2.2
Pumping Speed

N2, 100%; H2, 24133%; He, 42133%

4.7.2.3
Ultimate Pressure

Depending on how many turbomolecular pump and molecular drag pump stages are used in series, the attainable ultimate
pressure is in the range of 5 × 109 to 1 × 107 Pa.

4.8
Backing Pumps

Because all of the above-mentioned turbomolecular pumps and molecular drag pumps and combination of these pumps need a
backing pump, it is worthwhile to look at the existing types of backing pumps, in order to be able to evaluate which

< previous page page_248 next page >


< previous page page_249 next page >
Page 249

Table 4.5. Typical Backing Pump Data


Type Ultimate pressure (Pa) Quality of Vacuum
Membrane Absolutely dry
< 103
Rotary vane Oil vapors
< 101
Piston Technically dry
<1
Multiroots Technically dry
< 10
Claw roots Technically dry
< 10

backing pump should be used together with the high-vacuum pump (see Chapter 9). In principle, the selection of an
adequate backing pump depends on its pumping speed, its pressure range, the quality of the fore-vacuum produced, its
price, and its size. Typical backing pump data are shown in Table 4.5.

Backing pumps producing an "absolutely dry" vacuum do not have any oil or grease in possible contact with the
pumping area. "Technically dry" pumps do not have any oil or grease in the pumping area; however, they have grease or
oil (e.g, in the bearing or feed-through area) that is kept away from the pumping area by dynamic seals, because there
are labyrinths, and so on.

Together with magnetically levitated combined turbomolecular pump and molecular drag pumps, even the most
demanding applicational requirements can be met.

< previous page page_249 next page >


< previous page page_251 next page >
Page 251

Part III
Regenerative Drag Pumps

Nigel T. M. Dennis

4.9
Regenerative Drag Pumps

Capacities available: 30540 m3 h1 (25150 liter·s1)

Operating pressure range: 1000 to less than 104 mbar

4.9.1
Mechanism

The regenerative pump is a high-speed single-shaft vacuum pump capable of delivering to atmospheric pressure. It is a
dynamic machine that relies on momentum transfer from fast moving blade rows that create the pumping action. The
name "regenerative" comes from the way the mechanism circulates the gas repeatedly through the same blade row, each
time "regenerating" the forward momentum. Other names are used: side channel, vortex, or peripheral flow pump.

The advantage of this pump is that the high shaft speed results in small stage size, making the pump extremely compact.
The high rotational speed also allows other mechanisms such as a molecular drag stage (see Section 4.5) to be
effectively used on the same shaft.

A single stage of the pump consists of a row of blades mounted on a rotor; these protrude into the flow channel (see Fig.
4.31). As the blades move around they generate a spiral vortex within the channel that circulates the gas through the
blades as it moves along the channel, thus creating a helical flow path. This ensures that

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1988John Wiley & Sons,Inc.

< previous page page_251 next page >


< previous page page_252 next page >
Page 252

Fig. 4.31
One stage of a regenerative peripheral flow pump
indicating direction of gas flow and where the flow
channel is interrupted by the "stripper."

high-speed gas from the blades is continually fed into the channel, diffusing into the bulk flow to generate the pressure
rise. Within a single stage the gas may circulate many times through the blades, creating a high compression ratio.
Typical performances range from a compression of 2 to about 10, with the latter occurring in the very best machines.
The blades are shaped to catch the incident gas with the minimum of turbulence, to deflect it, and to throw it forward
into the bulk gas flow. The inlet and exhaust of a pump channel are separated by a restrictive section through which
only the blades may pass. This is generally known as the stripper. Pockets of gas between the blades are also carried
through the stripper; these form a carry-over volume, limiting the efficiency and compression of the machine. The blade
size must, therefore, be small compared to the channel size. The amount of gas carried over to the inlet can be reduced
by allowing the closed pockets within the stripper to expand through a connection to an intermediate pressure point
within the stage.

Sixsmith [108] revolutionized the understanding and design of these machines and showed that practical compressors
and vacuum pumps could be made. He developed many of the features described above and demonstrated a prototype
machine with a maximum compression of 10 and a working flow of 250 m3·h1.

Blade speeds are usually below Mach 1, and Mach 0.5 to 0.8 appears to be suitable. Speeds above Mach 1 result in
shock wave problems.

The function of the blades is to impart momentum to the gas as efficiently as possible.
< previous page page_252 next page >
< previous page page_253 next page >
Page 253

Fig. 4.32
A typical pump with inlet drag stages.
Fig. 4.33
Speed and power curve of a 240-m3·h1 regenerative
pump with two drag stages.

At the present stage of development, five or more stages of regenerative compressor are used with an inlet drag stage or
stages based on the Holweck or Siegbahn principle. The regenerative stages are typically capable of reaching 1 mbar.
The drag stages extend the ultimate to pressures to as low as 1 × 104 mbar [109]. A typical pump

< previous page page_253 next page >


< previous page page_254 next page >
Page 254

cross section is shown in Fig. 4.32. The speed curve obtained from a medium-size pump of this type is indicated in Fig.
4.33.

The power requirement of a regenerative pump is high because of the poor efficiency of the mechanism which is due to
the slip required to generate a useful compression. A typical power curve is shown in Fig. 4.33.

Pumps of this type are susceptible to dust accumulation in their mechanism and, therefore, should generally be used on
clean duties. They are, however, the only single-shaft pumps to achieve this level of vacuum when working against
atmospheric pressure.

References

1. V. V. Fondrk, Chem. Eng. Prog. 49, 3 (1953).

2. W. Gaede, Ann. Phys. (Leipzig) [4] 46, 357 (1915).

3. I. Langmuir, Gen. Electr. Rev. 19, 1060 (1916); J. Franklin Inst. 182, 719 (1916).

4. W. Gaede, Z. Tech. Phys. 4, 337 (1923); German Patent No. 286,404.

5. C. R. Burch, Nature (London) 122, 729 (1928).

6. K. C. D. Hickman and C. R. Sanford, Rev. Sci. Instrum. 1, 140 (1930).

7. K. C. D. Hickman, J. Franklin Inst. 221, 215 (1936).

8. L. Malter and N. Marcuvitz, Rev. Sci. Instrum. 9, 92 (1938).

9. N. Embree, U.S. Pat. 2,150,676 (1939).

10. K. C. D. Hickman, J. Appl. Phys. 11, 303 (1940).

11. M. H. Hablanian and J. C. Maliakal, J. Vac. Sci. Technol. 10, 58 (1973).

12. B. D. Power, High Vacuum Pumping Equipment. Reinhold, New York, 1966.

13. L. Laurenson, Vacuum 37, 609 (1987).

14. R. Jaeckel, H. G. Nöller, and H. Kutscher, Vak.-Tech. 3, 1 (1954).

15. H. Kutscher, Z. Angew. Phys. 7, 218, 229, 234 (1955).

16. N. M. Bulgakova, G. A. Khramov, O. A. Nerushev and A. K. Rebrov, Vacuum 44, 749 (1993).

17. M. H. Hablanian, J. Vac. Sci. Technol. A 5, 2552 (1987).

18. B. B. Dayton, Vacuum 15, 53 (1965).

19. W. Steckelmacher, Vacuum 15, 503 (1965).

20. W. Steckelmacher and D. Turner, J. Sci. Instrum. 43, 893 (1966).

21. P. Clausing, Ann. Phys. (Leipzig) [5] 12, 961 (1932).


22. C. W. Oatley, Br. J. Appl. Phys. 8 (15), 495 (1957).

23. W. Harries, Z. Angew. Phys. 3, 296 (1951).

24. T. L. Ho, Physica (Amsterdam) 2, 386 (1932).

25. R. F. Coe, J. Sci. Instrum. 32, 207 (1955).

26. D. Latham, B. D. Power, and N. T. M. Dennis, Vacuum 1, 97 (1951); 2, 33 (1952).

26.a. P. Alexander, J. Sci. Instrum. 25, 313 (1948).

27. A. Herlet and G. Reich, Z. Angew. Phys. 9, 14 (1957).

28. S. F. Kapff and R. B. Jacobs, Rev. Sci. Instrum. 18, 581 (1947).

29. J. Blears, Nature (London) 154, 20 (1944); Proc. R. Soc. London, Ser. A 188, 62 (1947).

30. D. J. Crawley, E. D. Tolmie, and A. R. Huntress, Trans. Natl. Vac. Symp. 9, 399403 (1962).

< previous page page_254 next page >


< previous page page_255 next page >
Page 255

31. A. R. Huntress, A. L. Smith, B. D. Power, and N. T. Dennis, Trans. Natl. Vac. Symp. 4, 104111 (1958).

32. K. C. D. Hickman, Trans. Natl. Vac. Symp., 2nd, 1961, Vol. 1, pp. 307314 (1962).

33. K. C. D. Hickman, J. Franklin Inst. 221, 383 (1936).

34. R. Jaeckel, Z. Tech. Phys. 23, 177 (1942).

35. D. Latham et al., Vacuum 2, 33 (1952).

36. G. P. Brown, Rev. Sci. Instrum. 16, 316 (1945).

37. L. Laurenson, Proc. Intn. Vac. Congr., 7th, Vienna, 1977 (1977).

38. E. H. Hirsch and T. J. McKay, Vacuum 43, 301 (1992).

39. M. A. Baker, L. Holland, and L. Laurenson, Vacuum 21, 479 (1971).

40. J. Ruf and O. Winkler, in Ergebnisse der Hochvakuumtechnik und der Physik dünner Schichten (M. Auwärter, ed.),
pp. 207224. Wissenschafttliche Velagsgesellschaft, Stuttgart, 1957.

41. B. D. Power and D. J. Crawley, Vacuum 4, 415 (1954) (published in 1957).

42. M. H. Hablanian, J. Vac. Sci. Technol. 6, 265 (1969).

43. P. Duval, Vide 135 (MayJune), 162166 (1968).

44. M. A. Baker, J. Sci. Instrum. [2] 1, 774 (1968).

45. D. Santeler, J. Vac. Sci. Technol. 8, 299 (1971).

46. D. Hoffman, J. Vac. Sci. Technol. 16, 71 (1979).

47. M. Wutz, Vak.-Tech. 5, 146 (1964).

48. N. Milleron, Trans. Natl. Vac. Symp. 5, 140 (1959).

49. R. Darbord, J. Phys. Radium 3, 345 (1932).

50. D. Alpert, Rev. Sci. Instrum. 24, 1004 (1953).

51. B. B. Dayton, Ind. Eng. Chem. 40, 795 (1948).

52. G. Zinsmeister, Adv. Vac. Sci. Technol., Proc. Int. Congr., 1st, Namur, Belg., 1958, Vol. I, p. 335 (1960).

53. D. H. Davis, L. L. Levenson and N. Milleron, Rarefied Gas Dynamics, 2nd Int. Symp., Sect. 2, pp. 99115.
Academic Press, 1961.

54. G. Lewin, J. Vac. Sci. Technol. 5, 75 (1968).

55. C. D. Ehrlich, J. Vac. Sci. Technol. A 4, 2384 (1986).

56. P. Clausing, Z. Phys. 66, 471 (1930).


57. H. L. Eschbach, R. Jaeckel, and D. Müller, Trans. Natl. Vac. Symp., 2nd Internat. Congress, Pergamon Press, New
York, 1961, pp. 11101115 (1962).

58. R. G. Herb, Rev. Sci. Instrum. 36, 367 (1965).

59. ASTM Tentative Recommended Practice E296-66T, Ionization Gauge Application to Space Simulators, p. 48,
Appendix A2. Am. Soc. Test. Mater., Philadelphia, 1966.

60. W. Molthan, Z. Tech. Phys. 7, 377, 452 (1926).

61. B. B. Dayton, Rev. Sci. Instrum. 19, 793 (1948).

62. P. Alexander, J. Sci. Instrum. 23, 11 (1946).

63. N. A. Florescu, Vacuum 10, 250 (1960).

64. M. H. Hablanian and A. A. Landfors, Trans. Am. Vac. Soc., 1967, p. 65 (1967).

65. M. H. Hablanian, Proc. Vac. Congr., 6th, Kyoto, Japan, 1974; Jpn. J. Appl. Phys., Suppl. 2 (Pt. 1), 25 (1974).

66. S. Chapman and T. G. Cowling, The Mathematical Theory of Non-uniform Gases, p. 42. Cambridge University
Press, Cambridge, UK, 1970.

< previous page page_255 next page >


< previous page page_256 next page >
Page 256

67. B. B. Dayton, MS Thesis, Dept. of Physics, University of Rochester, Rochester, NY, 1948.

68. A. Stodola and L. C. Loewenstein, Steam and Gas Turbines. McGraw-Hill, New York, 1927.

69. R. Jaeckel, Z. Naturforsch. A 2A, 666 (1947).

70. D. G. Avery and R. Witty, Proc. Phys. Soc. London 59, 1016 (1947).

71. S. Oyama, J. Phys. Soc. Jpn. 5, 192 (1950).

72. H. G. Nöller, J. Vac. Sci. Technol. 3, 202 (1966).

73. G. Zerkowitz, Z. Ver. Dtsch. Ing. 61, 869873, 889892 (1917).

74. N. V. Iliasova, S. V. Nedosekova, A. K. Rebrov, P. A. Skovorodko, and J. J. Roig, Vacuum 44, 745 (1993).

75. K. S. Sadykov and S. A. Figurov, Vacuum 41, 2061 (1990).

76. L. Riddiford and R. F. Coe, J. Sci. Instrum. 31, 33 (1954).

77. W. Reichelt, Vak.-Tech. 13, 148 (1964).

78. G. Tóth, Vak.-Tech. 16, 41, 193, 215 (1967).

79. K. Ray and N. Sengupta, Nature (London) 155, 727 (1945).

80. K. C. D. Hickman, Nature (London) 156, 635 (1945).

81. B. B. Dayton, Trans. Natl. Vac. Symp. 6, 101119 (1960).

82. M. Matricon, J. Phys. Radium 3, 127 (1932).

83. R. B. Jacobs and S. F. Kapff, Ind. Eng. Chem. 40, 842 (1948).

84. L. Wertenstein, Proc. Cambridge Philos. Soc. 23, 578 (1927).

85. D. Enskog, Phys. Z. 12, 533 (1911).

86. H. Croft, Thermodynamics, Fluid Flow and Heat Transmission. McGraw-Hill, New York, 1938.

87. E. H. Kennard, Kinetic Theory of Gases. McGraw-Hill, New York, 1938.

88. N. T. M. Dennis, B. H. Colwell, L. Laurenson, and J. R. H. Newton, Vacuum 28, 551 (1978).

89. T. J. Gay et al., J. Vac. Sci. Technol. A 12, 2903 (1994).

90. W. Gaede, Ann. Phys. (Leipzig) [4] 41, 337 (1913).

91. M. Holweck, C.R. Hebd. Seances Acad. Sci. 177, 43 (1923).

92. M. Siegbahn, Arch. Math. Astron. Fys. 30B, 17 (1944).

93. W. Becker, Vak.-Tech. 16, 625 (1966).


94. C. H. Kruger and A. H. Shapiro, Trans. Natl. Vac. Symp. 7 (1960).

95. J. F. O'Hanlon, A User's Guide to Vacuum Technology, pp. 173183. Wiley, New York, 1980.

96. W. Becker, Vak.-Tech. 7, 149 (1958).

97. I. G. Chu and Z. Y. Hua, J. Vac. Sci. Technol. 20, 1101 (1982).

98. K. H. Bernhardt, J. Vac. Sci. Technol. A 1, 136 (1983).

99. J. Henning, Vacuum 21, 523 (1971).

100. K. H. Mirgel, J. Vac. Sci. Technol. 9, 408 (1972).

101. L. Maurice, Jpn. J. Appl. Phys., Suppl. 2 (Pt. 1), 21 (1974).

102. Y. Murakami, T. Abe, S. Mori, N. Nakaishi, and S. Hata, J. Vac. Sci. Technol. A [2] 5, 2599 (1987).

103. J. Henning, J. Vac. Sci. Technol. A 6, 1196 (1988).

104. W. Becker and J. Henning, J. Vac. Sci. Technol. 15, 768 (1978).

< previous page page_256 next page >


< previous page page_257 next page >
Page 257

105. J. E. de Rijke and W. A. Klages, Jr., Solid State Technol., April, p. 63 (1994).

106. F. J. von Schittko and C. Schmidt, Vak.-Tech. 24, 110 (1975).

107. T. Sawada and O. Taniguchi, Jpn. Pat. 68/1723 (1973).

108. H. Sixsmith and H. Altmann, J. Eng. Ind., August, pp. 637647 (1977).

109. M. G. Mase, T. Nagaoka and M. Taniyama, J. Vac. Sci. Technol. A 6, 25182521 (1988).

< previous page page_257 next page >


< previous page page_259 next page >
Page 259

5
Capture Vacuum Pumps

A capture or entrapment vacuum pump is a pump usually located within the chamber being evacuated that removes gas
molecules by sorption or condensation on its internal surfaces. Three types of capture pumps are described in this
chapter; the getter pump, the sputter ion pump, and the cryopump.

Gas in the getter pump is retained principally by chemical combination with a getter material. The getter is usually a
chemically active metal or alloy, either in bulk or in the form of a freshly deposited thin film.

In the sputter ion pump the gas molecules are ionized and directed towards getter surfaces of the pump by electric and
magnetic fields. The getter surfaces are replenished in a continuous way by cathodic sputtering.

Gas molecules are condensed on refrigerated surfaces in a cryopump. These surfaces are cooled to a temperature low
enough to keep the vapor pressure of the condensate equal to or below the desired low pressue in the vacuum chamber.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_259 next page >


< previous page page_261 next page >
Page 261

Part I
Getters and Getter Pumps

Bruno Ferrario

The use of getters in vacuum technology has gained an increasing interest during recent decades, and the development
of this type of pump for new and advanced applications has required a deeper understanding of some basic scientific
concepts. These concepts are particularly related to surface and bulk characteristics of metals and gassurface interaction
phenomena. A review of some fundamental aspects of gettering will therefore be given before the description of the
working and applicative characteristics of getters. A more complete description of the basic concepts of the gassurface
interactions and diffusion phenomena can be found in Chapter 10.

5.1
Types of Gas Surface Interactions

When gas molecules interact with a solid surface [1,2], several phenomena can take place. Among them, the following
ones shown in Fig. 5.1 are of particular interest for the subject here considered: adsorption (capture of molecules by the
surface), desorption (emission of molecules from the surface), backscattering (bouncing back of molecules impinging
on the surface), diffusion (penetration of adsorbed atoms from the surface into the solid bulk or movement of dissolved
atoms from the solid bulk to the surface), displacement (displacement of an adsorbed molecule by another impinging
molecule), and surface reactions (formation of new molecules at the surface from adsorbed molecules of different
species).

Some of these phenomena tend to remove molecules from the gaseous phase, while others tend to drive molecules into
the gaseous phase. If the surface has suitable

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_261 next page >


< previous page page_262 next page >
Page 262

Fig. 5.1
Some types of gassurface interactions.

characteristics, the adsorption phenomenon can, however, prevail (together with diffusion into the bulk, under certain
conditions of temperature and pressure) in a vacuum system. As a consequence, there is a net removal of molecules
from the gaseous phase and, therefore, a decrease of the gas pressure. This phenomenon is usually known as gettering,
and the solid materials which exhibit this gettering capability are named getters. Usually the term ''getter" is adopted
when the capture of gaseous molecules is due to relatively strong forces (i.e., chemical forces), and it refers to special
metals both in pure or in alloy form. In this case, it is also common to say that "chemisorption" takes place, which
explains why getters are also called chemical pumps. Other materials such as molecular sieves, active charcoal, and so
on, exhibit adsorption of gas molecules or atoms; however, the forces involved are relatively weak and these materials
are therefore usually known as physical absorbers rather than getters. In this case, "physisorption" takes place.

Quantitatively, as will be seen in more detail later, the gettering capability is defined by a "gettering rate" and a
"gettering capacity." The gettering rate represents the number of gaseous atoms or molecules which are removed from
the gaseous phase by the getter per unit time. The gettering capacity is the number of atoms or molecules which can be
captured by the getter before it stops sorbing gas.

Chemical nature, crystal structure, and physical characteristics of pure metals or alloys used as getters are important
aspects to understand gettering; some discussion on these topics [35] is therefore presented for a better understanding of
the characteristics of getters.

5.2
Basic Concepts of Getter Materials

The lattice structure (i.e., the "bulk" parameters) can influence the gettering properties of a metal in terms of diffusivity
and solubility for gases [6]. The gettering properties are, however, first of all related to the surface characteristics of
metals. The situation in the bulk of a metal, where each atom is completely surrounded by other metal atoms and has
saturated bonds, is quite different from that at the surface. In fact, the surface atoms have a smaller coordination number
in comparison with the bulk atoms, since they are not completely surrounded by atoms; this surface coordination
number depends on the face of the crystal structure exposed at the surface. This means that the surface atoms have
unsaturated bonds which determine their reactivity versus the gas atoms or molecules colliding with the surface. The
adsorption of these molecules tends

< previous page page_262 next page >


< previous page page_263 next page >
Page 263

to saturate the free bonds and to reestablish the symmetry of the force field to which the atoms would be submitted if
they were in the bulk of the crystal. Typically the surface density of atoms is in the range of 1 × 1015 per cm2
(depending on the metal type and the exposed crystal plane).

In gettering, not only pure metals are used, but often also alloys [35] are of paramount importance to achieve special
properties. There are binary, ternary, multicomponent alloys, depending on the number of the components. The
composition and the nature of the alloys depend on the type of the alloying elements and are characterized by the
"constitutional diagram" (also called equilibrium or state or phase diagram) for the system of elements.

Most of the metals are completely miscible in the molten state, and the resulting alloys can be considered as solid
solutions. The solid solutions can be substitutional or interstitial. In the molten state, some metals can form stable
chemical bonds; when this occurs, intermetallic compounds are formed.

Alloys can exhibit more than one phase and also show complex phase diagrams. The type of phases present in a getter
material can affect not only its chemical affinity, but also other properties such as solubility, diffusivity, hardness,
thermal conductivity, and so on. Also pure metals can exhibit more than one phase, depending on the temperature
conditions. For example [6], Ti has a hexagonal close-packed (hcp) structure (α-phase) below 885°C and a body-
centered cubic (bcc) structure above 885°C. Fe exhibits an α-phase (bcc structure) below 910°C and between 1400°C
and 1540°C; between 910°C and 1400°C, Fe exhibits a γ-phase [face-centered cubic (fcc) structure].

Metals and alloys are likely more frequently polycrystalline than monocrystalline. The grains consist of atoms arranged
in a lattice with a precise space orientation. Adjacent grains have the same crystallographic structure, typical of the
element considered, but the orientation is different. Therefore, in the space between two adjacent grains, there is a
transition in the crystal orientation. This area, called grain boundary [5, 7], has an important influence on the metal
properties, also from the point of view of gettering. The grain boundary, in fact, is a zone of the metal where there is a
lower density and a smaller coordination number in comparison with the bulk. The grain boundary atoms can, therefore,
be more prone to react with foreign atoms; diffusion can also take place more easily than in the crystal bulk.
Dimensions and shape of the grains can change during time, depending on temperature.

5.3
Adsorption and Desorption

As thoroughly discussed in Chapter 10, the experimental and theoretical studies of the gassurface interaction distinguish
between two types of adsorption: physical adsorption or physisorption [8, 9] and chemical adsorption or chemisorption
[1013], which is normally involved in gettering.

It is useful to keep in mind that chemisorption and therefore gettering can be dissociative, so that the individual atoms of
the split molecules (such as O2, N2, CO, etc.) are actually bonded to the surface in the so-called adsorption sites and can
eventually diffuse into the bulk of the getter material if enough (usually thermal) energy is provided. The bond energies
involved in chemisorption normally exceed 5eV, and the so-called residence time of the molecules on the capturing
surface is so

< previous page page_263 next page >


< previous page page_264 next page >
Page 264

long that the process is practically irreversible in usual working conditions (apart from H2). In physisorption, on the
contrary, the process is nondissociative and reversible. In chemisorption, moreover, the adsorption characteristics are
specific as in all chemical reactions.

The number and position of the adsorption sites depend on the nature of the metal considered, the crystallographic
structure, the orientation of crystal faces, the presence of impurities, and so on. The fraction of adsorption sites available
at the surface, occupied by the adsorbed atoms, usually called surface coverage (ϑ), is a useful concept to describe
adsorption phenomena and capacities of getters. The surface is completely saturated by the adsorbate when ϑ = 1 (a
monolayer is formed). During chemical adsorption (as well as during physical adsorption) there is heat generation; that
is, the process is exothermic. The heats of chemisorption depend on the gasmetal system and are reported in the
literature [14].

During gettering there could be some increase of temperature which may be negligible or relevant, depending on the
amount of gas adsorbed, its nature, the area of the gettering surface and its heat dissipation characteristics, the rate of
adsorption, and so on.

In equilibrium conditions, the adsorption characteristics of a gasmetal system can be described by isotherms [1519]
which are usually of five types as seen in Section 10.1.2. It is, however, important to notice that, besides the
thermodynamic considerations, it is fundamental to describe the adsorption process and therefore the gettering process,
also from the kinetics standpoint (a process can be thermodynamically favored but too slow). Kinetically the adsorption
process can be described by the following general equation:

where S is the rate of adsoption, P1 is the probability that the particle colliding with the surface finds a free site, P2 is
the probability that the particle has sufficient energy for the adsorption to take place (i.e., has the necessary activation
energy), P3 is the probability that when the two above conditions are fulfilled the particle is actually adsorbed, and v is
the collision frequency of the particle on the surface. The product s = P1P2P3 is usually called sticking probability or
sticking coefficient and depends on the surface coverage, the size of the adsorbate, the dissociative or nondissociative
character of the adsorption, the activation energy (of the specific gas moleculemetal system), and so on. Considering the
usual mathematical expression for the collision frequency, the adsorption rate per unit time and unit area can be given
by the following equation, which also represents the basic expression for the gettering speed:

where p is the pressure of the gas, m is the mass of the gas molecule, k is the Boltzmann constant, and T is the
temperature in Kelvin. From the experimental data for S, it is then possible to derive the sticking probability values. The
sticking probability parameter is often used in the field of getters to calculate the gettering characteristics of complex
gettering structures when statistical methods are applied. The effectiveness in chemisorption of gases on metals is
related to several factors: the electronic factor,

< previous page page_264 next page >


< previous page page_265 next page >
Page 265

the geometric factor, and the effect of impurities and imperfections. Since chemisorption implies the formation of a
covalent bond, the importance of the electronic factor is obvious. In this respect, the electronic properties of the d
character metals are favorable to chemisorption for many gas molecules. As a matter of fact, getter materials are usually
selected among metals with d characters (typically among metals of the IVB, VB, and VIB groups of the periodic table).

The possibility to accommodate molecules in the chemisorbed state also depends, however, on the geometric
characteristics of the metalthat is, on the relative distances of the surface metal atoms and the atom surface density [6,
18]. Different exposed crystal planes of the same metal can exhibit different chemisorption capabilities. Impurities and
imperfections may promote or inhibit chemisorption and are sometimes responsible for unexpected changes in
chemisorption properties of the same metal.

After adsorption, molecules can be re-emitted, and this phenomenon depends on the energy imparted [810]. The
physisorbed molecules can be easily desorbed, for example, by heating, since the bond energies involved are relatively
small.

The release of chemisorbed (dissociated) molecules implies the preliminary recombination of the component atoms and
then the emission of the molecules. The energy involved in this process is large, and therefore desorption is generally
very difficult (apart for H2). Adsorption could be accompanied by desorption of gas molecules of different species
compared to the adsorbed molecules; this can be due to catalytic surface reactions [19] between the adsorbed,
dissociated molecules of different species simultaneously present on the surface or between the gas molecules and
certain surface impurities.

5.4
Bulk Phenomena

5.4.1
Diffusion

Getter materials are not only characterized, as previously mentioned, by the phenomena taking place at the surface. The
chemisorbed species can in fact diffuse [1922] from the surface into the bulk of the getter material, depending on the
nature of the diffusing species and the physicochemical characteristics of the sorbing material. The component elements
of the gas molecules chemisorbed at the surface have to break the chemical bonds with the adsorbing material to diffuse
into the bulk. The driving force for the diffusion is then the concentration gradient of the considered element, provided
that enough energy is supplied (usually thermal energy is supplied to promote diffusion). In this situation the two
processes of adsorption and diffusion proceed simultaneously. In the stationary state the diffusion rate of an element
through the x-axis of a metal sheet can be expressed by Fick's first law and is characterized by the diffusion coefficient
D (cm2·s1) and by the concentration gradient at the considered plane from the surface. D depends on temperature
according to the following equation:

where D0 is a diffusion constant depending on the diffusing element-metal system considered, E is the activation energy
for the diffusion process, T is the temperature in

< previous page page_265 next page >


< previous page page_266 next page >
Page 266

Table 5.1. Diffusion Coefficients for Various GasMetal Systems at Different Temperatures [22]
Diffusion Species Metal Crystal Structure D0(cm2·s1) ∆H (J/g·atom) D D D
25°C 400°C 1000°C
H α-Ti hcp 1.80 × 102 1.42 × 1011 1.68 × 106
51,900
3.00 × 102 4.92 × 1013 5.03 × 107
61,500
H α-Zr hcp 7.00 × 104 4.69 × 109 3.58 × 106
29,500
H β-Ti bcc 2.00 × 103 1.44 × 104
27,800
H β-Zr bcc 5.30 × 103 1.98 × 104
34,800
N α-Ti hcp 1.20 × 102 7.60 × 1036 2.39 × 1017
189,300
N β-Ti bcc 3.50 × 102 5.48 × 108
141,400
O α-Zr hcp 2.00 × 101 1.61 × 1031 9.45 × 1015
171,600
O β-Zr bcc 4.50 × 102 6.44 × 107
118,000

K, and R is the gas constant. It is often observed that D is not independent of the concentration and of the concentration gradient, particularly at
high concentration values. Typical values of D0 and E for some diffusing species-metal systems are shown in Table 5.1 [22].

The application of Fick's first law leads, for example, to the description of the permeation phenomena and can be used in first approximation
also to describe gettering when stationary diffusion is involved.

In the description of gettering phenomena, the stationary state condition, however, is not always an appropriate approximation. Time-
dependence of the phenomenon, as a matter of fact, is to be considered to better describe the process. In this case Fick's second law applies for
the linear diffusion showing the variation of the concentration as a function of time t and position along the axis of diffusion. The application of
this law for a semi-infinite slab represents a good approximation in describing gettering in some practical conditions; the derived Eq. (10.127)
of Chapter 10 indicates that the rate of removal of molecules per unit area and time is proportional to the square root of the diffusion coefficient
over time of sorption; the concentration of the diffused atoms at a certain distance from the surface is instead given by Eq. (10.126). The total
gas uptake per unit area after time t is calculated to be proportional to the square root of the diffusion coefficient multiplied by time [Eq.
(10.128)] ; therefore for a given value of the diffusion coefficient (i.e., for a given getter material), the diffused quantity is expected to be simply
proportional to the square root of the diffusion time. By plotting the sorbed quantity as a function of the square root of time, one should then
find a straight line; this is often actually found even if deviations are observed, particularly at high concentration values.

For geometries such as spherical particles of radius a, the solution of the diffusion equation can be expressed in the following form:

This is a relatively complicated equation, but can better describe the common situation where the getter material is in powder form and the
shape of the particle is

< previous page page_266 next page >


< previous page page_267 next page >
Page 267

not too different from spherical. This equation can be used to derive the mass of gas diffused during time t as follows:

where a is the radius of the sphere, M∞ is the mass of the gas in saturated conditions, and Mt is the mass of the gas at
time t.

5.4.2
Solubility

The mechanism associated with gas solution [6, 22] in a getter material is usually described as a three-step process: (a)
dissociation of the gas molecules at the gassurface interface; (b) sorption at the superficial, or near surface, sites; and (c)
dissolution in the bulk of getter material, through diffusion mechanisms.

As well as in other gasmetal systems, a solid solution is formed with the sorbed gas species usually located at the
interstitial sites of the lattice of the host metal. The structural effects caused by gas dissolution are generally limited to
some expansion of the lattice, with increase of the lattice parameter.

In the case of the dissolution of a di-atomic gas species, the formation of solid solution in a getter material can be
represented by the following reaction:

where g and s denote gas phase and solid solution, respectively.

Once the thermodynamic equilibrium has been achieved, the pressure, P, is proportional to the square of the gas species
concentration in the material, q:

More precisely, the gas concentration is related to the gas pressure and to the getter temperature according to Sieverts'
law [23], expressed by the following most common equation (for dilute gas solutions):

where A and B are constant parameters depending on the gas type and getter material, which can be determined
experimentally.

Although Eq. (5.8) is valid for any solute gas, it is used, in practice, only for hydrogen. In fact, hydrogen generally
forms a solid solution in metals with relatively low stability, enabling the reverse reaction to occur by a moderate
increase of the temperature.

The solid solution (α phase) is thermodynamically stable until the solubility limit is reached. Approaching this
concentration, the interactions between the solute atoms tend to increase, nucleation and growth of solid phases coexist,
and the isotherms

< previous page page_267 next page >


< previous page page_268 next page >
Page 268

show a plateau, which depends on temperature, according to van't Hoff's law:

where P is the pressure, T is the getter temperature in K, and ∆S and ∆H are, respectively, the entropy and enthalpy of
formation of the β phase. In the most common high-vacuum (HV) and ultrahigh-vacuum (UHV) applications of getter
materials, the amounts of gases involved in the sorption process are relatively small and therefore they typically form
dilute solid solutions. The thermodynamics of hydrogen solution and hydride formation can be described by
pressurecomposition isotherms (Fig. 5.2).

5.5
Equilibrium Pressures

In order for a getter material to work properly, particularly in ultrahigh-vacuum conditions, the equilibrium pressures of
the gettered gas should be as low as possible. The equilibrium pressures are the result of the balance between the
adsorption and desorption processes, as previously illustrated. Due to the energy involved in chemisorption, it has
already been mentioned that the residence time of the adsorbate is extremely long and therefore the desorption process
is negligible compared to the adsorption process; the equilibrium pressure of the adsorbates are therefore expected to be
very low.

If the gettering process proceeds, for example, to the formation of certain stoichiometric oxides and nitrides of certain
metals commonly used as getters, the corresponding equilibrium pressures can also be low or extremely low (e.g., well
below 1020 bar [24]).

In the case of H2, the situation is different; as discussed in the previous section, in the range of concentrations of usual
interest in gettering, Sieverts' law is found to be followed and the H2 equilibrium pressure may not be negligible
depending on the working conditions.

Fig. 5.2
Typical pressurecompositiontemperature
curves for hydrogenmetal systems.

< previous page page_268 next page >


< previous page page_269 next page >
Page 269

Of course, also the vapor pressure [25] of the getter material itself should be small or negligible compared to the
operating pressure required in the considered vacuum applications.

5.6
Getter Materials

5.6.1
Basic Characteristics of Getter Materials

The first basic characteristics for a getter material are chemical affinity with gases and bulk diffusivity; the former is
particularly important in determining the chemisorption of the gas species to be removed in a vacuum system, and the
latter allows the displacement of the adsorbate into the bulk of the getter material, thus increasing its capacity. Another
basic characteristic is related to the possibility of achieving large surface areas. The development of suitable alloys
makes it possible, however, to modulate the various relevant properties for a getter material and achieve a useful
compromise. This explains why, in practice, getter materials are often made up of alloys rather than of pure metals.

There are other important parameters to be considered in developing and/or selecting a practical getter material, which
often requires a compromise with the above chemicophysical properties. As a matter of fact, getter materials are often
manufactured on an industrial scale and have to be handled in large quantities. For these reasons, workability
(possibility to transform the original getter ingots into powder to ensure enough surface area), hardness, safety (related
to possible toxicity, pyroforicity, and high exothermic reactions with ambient or process gases and materials), stability
under usual or specific storage conditions, availability, and cost all have to be taken into consideration.

5.6.2
Sorption Speed and Sorption Capacity

The net removal rate of molecules from the gaseous phase as a result of the various possible interactions with the getter
material (i.e., the gettering rate as already mentioned in Section 5.1) is also often defined as sorption speed of that getter
material. This indicates that the removal of molecules can be due to adsorption combined or not with absorption (due to
diffusion into the bulk of the material). The sorption speed can be expressed in m3·s1, liter·s1, or other convenient
units. Multiplying the sorption speed by the pressure at which this speed is measured, one obtains the sorption
throughput, which is therefore given in Pa·m3·s1, mbar·liter·s1, Torr·liter·s1 or other convenient units.

The gettering capacity is also called, more generically, sorption capacity. It is commonly measured in Pa·m3,
mbar·liter, Torr·liter or other convenient units. Of course the maximum sorption capacity corresponds to the formation
of a stoichiometric compound. For example, in the case of O2 gettered by Ba to form BaO, the stoichiometric capacity
is 93 mbar·liter·g1.

The sorption characteristics of a getter are generally represented by the sorption speed as a function of the sorbed
quantity ; the typical trend of the curve is shown in Fig. 5.3. Q0 represents the total capacity at zero speed. This quantity
corresponds to the amount of gas saturating the surface if there is practically no bulk diffusion, which

< previous page page_269 next page >


< previous page page_270 next page >
Page 270

Fig. 5.3
Typical sorption curve for a getter material.

occurs when getters work at room temperature. Q0 corresponds to the maximum possible amount of gas sorbed if
diffusion takes places and a stoichiometric compound forms. However, the corresponding sorption speed for a certain
amount of gas sorbed can be too small and not acceptable for the application even at much smaller values than Q0. The
useful capacity of the getter is then the amount of gas sorbed until the speed reaches the minimum acceptable value for
the application. In Fig. 5.3 the capacity of the getter is Q1 if S1 is the acceptable speed; it is Q2 if the acceptable
sorption speed is S2. These values of the sorption capacity are therefore the so-called "practical capacities" and can vary
from situation to situation. ASTM F 798-82 defines a terminal gettering rate of a nonevaporable getter when the getter
has sorbed an amount of gas corresponding to a decrease of the sorption speed to 5% of its initial value; the initial value
is defined as the value measured after 3 minutes from the start of the sorption test. In the case of evaporable getters (Ba
getters) the terminal gettering rate corresponds to 1 liter·s1 and 0.1 liter·s1 for a Ba film in a large TV tube and in
receiving tubes, respectively.

It is common practice also to use the terms "specific sorption speed" and "specific sorption capacity"; they are speed
and capacity of the unit mass or volume or weight or area of the getter considered. For example, the specific sorption
speed can be measured in m3·s1·cm2 and the specific sorption capacity can be expressed in Pa·m3·g1.

The speed of a getter can decrease during time, depending on working temperature and pressure conditions. It decreases
more rapidly if sorption occurs at room temperature since bulk diffusion is not promoted (except for H2) and the
capacity is basically surface-limited. If a getter is operating at high temperatures, its speed is generally increased and
remains more constant as a function of time since the sorbed gases are diffused into the getter bulk. The behavior is
different for H2, however, because of the reversible character of its sorption. These situations are shown in Figs. 5.4 and
5.5.

Concerning pressure-dependence, it is found that, usually, the initial speed of getters is practically independent of
pressure in a very wide range: from 102104 mbar (depending on gas type) down to UHV conditions.

< previous page page_270 next page >


< previous page page_271 next page >
Page 271

Fig. 5.4
Typical sorption curves at different temperatures
for irreversible sorption.

Fig. 5.5
Typical sorption curves at different temperatures
for reversible (H2) sorption.

5.6.3
Principal Types of Getter Materials and Their General Working Conditions
A practice which has now become common identifies two different broad classes of getter materials: evaporable and
nonevaporable getters. Sometimes it is also common to say evaporated and nonevaporated getters. In principle, all
getters can be used either as evaporable or as nonevaporable; however, it is easier or more convenient to use certain
getter materials in the evaporated mode and others in the nonevaporated mode. Nonevaporated getters are also called
"bulk" or "volume" or "massive" getters.

In the case of evaporated getters, the getter material is heated to a sufficiently high temperature to evaporate and form a
film onto a surface inside the vacuum system.

< previous page page_271 next page >


< previous page page_272 next page >
Page 272

This film is immediately ready to sorb gases. The geometric area and the real area of the film determine the speed and
capacity of the getter.

In the case of nonevaporated getters, there is no evaporation or sublimation of the getter material; the gases react on the
available surface of the material and, if sufficient energy is supplied, diffuse into the bulk.

Based on the considerations made in Section 5.1, the typical materials studied as practical getters are shown in Table 5.2.

The alkaline-earth metals and particularly Ba are generally used as evaporable getters, whereas the metals of the IVB
group and particularly Zr and Ti, in pure or alloy form, are generally used as nonevoporable getters. Ti is also used as an
evaporable getter in the so-called sublimation pumps. The Zr- and Ti-based alloys can be binary, ternary, or
multicomponent, comprising elements such as Ni, Fe, Al, Co, rare earths, and so on, to obtain specific gettering
characteristics and other required physicochemical features. Rare-earth metals are often used in alloy form containing
various combinations of these elements. ''Mischmetal" is the common name of this type of alloy, rich particularly in
cerium and lanthanum. Thorium and uranium have also been found to be useable as getter materials; their radioactivity
and pyrophoricity (particularly for U), however, limit their applicability. Hafnium also exhibits some interesting
gettering capability, but availability, costs, and stability problems limit its use in real getters.

Particularly in recent years, getters with high sorption capabilities for all the residual gases found in a vacuum system
have been developed to cope with various specific requirements.

Before describing getter materials in more detail, however, the concept of "activation" has to be introduced as necessary
to understand the gettering behaviour of all these materials. The surface of the powder particles of any type of getter
material is readily covered mainly by an oxide layer when exposed to air the first time after its manufacturing. This
layer usually passivates the getter material thus preventing further pick-up of gas. The getter material, therefore, is not
immediately active and ready to work when introduced in the vacuum environment. In order for the getter to be prone to
chemisorb gases, the passivating layer has to be removed. In the case of evaporated getters, this occurs automatically
when the material is heated and evaporated because a new fresh metal surface is formed under vacuum, with oxygen
and carbon atoms being dissolved in the evaporated film mass. In the case of nonevaporated getters, the removal of the
passivating layer to get a clean, essentially metallic surface is usually necessary before the getter starts working. This
process is

Table 5.2. Typical Getter Materials


Evaporable Nonevaporable
Phosphorus (red)
Zirconium
Strontium
Titanium
Calcium
Hafnium
Barium
Thorium
Titanium
Rare earths
Alloys based on Zr and Ti

< previous page page_272 next page >


< previous page page_273 next page >
Page 273

called activation [25, 26] and is generally performed by heating the material for sufficient time to promote the diffusion
of the superficial oxygen and carbon atoms into the bulk of the material. The process is regulated by the diffusion laws
and can be performed using suitable combinations of temperature and time. Higher temperatures are more efficient in
cleaning the surface due to the exponential dependence of the diffusion coefficients on temperature; but in many
practical applications, too-high temperatures are not acceptable and the time factor can be more conveniently exploited.
In some cases, to overcome the problem of possible unacceptable heating of the surfaces surrounding the getter, a sort
of pulsed activation is adopted; it consists in successively heating to a high temperature for short times, to avoid
excessive heating of the surroundings, until the temperaturetotal-time combination for activation is reached.

The activation process can be schematically represented as shown in Fig. 5.6.

When a practically completely metallic surface is obtained, the maximum number of available adsorption sites can be
occupied and the speed is maximum: full activation has been achieved. However, in practical conditions also partial
activations can be useful. The activation conditions depend on the getter material usedthat is, on the type of oxide layer
formed (thickness, compactness) and its diffusivity characteristics. The activation of a getter requires that a reasonable
preliminary vacuum has already been achieved by conventional pumps.

It is to be added, however, that there are recent special nonevaporated getter materials [27] (based on special Ba alloys)
with particularly good sorption performances for N2, which can be used without the necessity of activation, even if they
are exposed to air (for a relatively short time such as minutes) before being used in the vacuum device.

It is to be pointed out that some getters which usually need activation can be used without being preliminary activated
when H2 is the main gas to be sorbed and its pressure is relatively high (e.g., in the range of 102 to a few millibars). The
reason is related to the mechanism described by van Vucht [28]. H2 can easily physically diffuse through the oxide
layer, particularly if it is not very compact. As a matter of

Fig. 5.6
Representation of the activation process for
a nonevaporable getter.

< previous page page_273 next page >


< previous page page_274 next page >
Page 274

fact, there may be enough porosity and/or enough cracks in the oxide layer to allow H2 to flow through it and be
absorbed by the getter metal surface immediately underneath; the local adsorption of H2 expands the lattice of the getter
material, thereby generating sufficient stress on the oxide layer to produce or increase the number of cracks and thus
further enhancing H2 absorption, and so on. There is therefore a sort of autocatalytic effect and the absorption becomes
more and more evident after some time (induction time), which may be minutes or hours, depending on hydrogen
pressure and getter temperature. It is finally to be noted that, in principle, the activation of a getter material can also be
performed either by displacing the atoms of the passivating layer away from the surface into the vacuum side or by
destroying the compactness of the layer to allow the passage of gas molecules and their contact with the metallic getter
surface, rather than promoting the diffusion of the passivating atoms into the bulk. This can be obtained by sputtering of
the surface using ion beams or a gas discharge (e.g., argon discharge) where the ions present are accelerated onto the
getter surface. These activation procedures are generally not very practical and therefore very rarely applied. The oxide
layer can also be destroyed, while new, fresh metallic surfaces are generated, by breaking the getter particles as occurs
for example during grinding in vacuum or in an inert atmosphere.

The activation process is repeated whenever the gas sorbed quantity has reduced the speed to unacceptable values and
every time the getter has been exposed to air. If no air exposure occurred, the reactivation of the getter can be made in
milder conditions compared to first activation.

After reactivation the initial speed is almost completely restored (depending on the amount of gas sorbed before starting
this process). With an increasing number of reactivations the initial speed obtainable decreases to the point that may be
unacceptable; the getter is therefore considered exhausted. Sometimes it is possible to restore somewhat better residual
gettering performances by using activation conditions that are more drastic than usual.

The typical trend of the sorption curves obtained as a function of the reactivation number is shown in Fig. 5.7.

If the getter is saturated at room temperature (RT), the possible required number of reactivations is high since the
surface capacity is generally a small fraction of the total getter capacity. If the getter is operated at high temperatures,
the number of reactivations needed is smaller the higher the temperature of operation; if the operation temperature is
sufficiently high (HT), the total capacity can be reached without need of reactivation. The number of possible
reactivations is limited and depends on the temperatures involved and on the nature and structure of the getter.

Fig. 5.7
Typical sorption curves for a nonevaporable
getter after successive reactivations.

< previous page page_274 next page >


< previous page page_275 next page >
Page 275

After saturation with hydrogen, reactivations (in this case also called regenerations), are theoretically unlimited because
of the reversible character of the absorption of this gas; in practice, the number of possible regenerations is large but
limited because of mechanical or embrittlement problems.

5.6.4
Interaction of Getters with Common Residual Gases

After activation, the getter sorbs the gases which are the residue of the preliminary pumping or which are continuously
generated in the vacuum system, thus ensuring the required vacuum level. The gases generally responsible of vacuum
deterioration in usual vacuum applications are: H2, H2O, N2, O2, CO, CO2, hydrocarbons (particularly CH4), noble
gases. The composition of the gas to be removed depends on the gas source (outgassing, leaks, permeation).

Active Gases. H2, H2O, CO, CO2, O2, and N2 are usually defined as active gases because they can chemically interact
with getter materials.

CO, CO2, O2, and N2 are chemisorbed at the surface and fixed in such a way that the adsorption is considered
permanent or irreversible; that is, no re-emission of these gases occurs in usual practical operating conditions. If the
getter material is heated, the atom components of these gases diffuse into the bulk and there they sink.

H2 is sorbed forming a solid solution, in a certain range of concentration following Sieverts' law. This is a reversible
sorption. This gas can actually be sorbed or emitted, depending on the working temperatures and pressures and on the
concentration of the gas in the getter material.

H2O is dissociated in the two components: hydrogen and oxygen. H2 is then sorbed according to the typical reversible
mechanism for this gas, whereas O2 is permanently fixed.

Hydrocarbons. These gases are not usually considered active. However, they can be sorbed by surface cracking when
the getter material is heated to sufficiently high temperatures. Again, H2 is sorbed in the typical reversible way and
carbon is permanently fixed. Heavy hydrocarbons, particularly, can be sorbed at the surface of a getter material also at
room temperature mainly by physisorption, with the capacity obviously being limited by the available surface area.

Noble Gases. Noble gases are not sorbed by a getter material. Their removal has to be achieved by other pumps. The
absence of chemical interaction of noble gases with getters can be exploited, for example, in the purification of these
gases.

5.6.5
Evaporable Getters

Evaporable getters used in practice are typically Ba and Ti. Ba getters, particularly, are by far the most commonly used
on a large industrial scale, and their application is well established and dates back several decades; their characteristics
have greatly evolved during time, and plenty of studies have been performed on the related applications.

< previous page page_275 next page >


< previous page page_276 next page >
Page 276

5.6.5.1
Ba Getters

Pure Ba is not used as such, because of its reactivity in air which would not allow its practical handling on a large scale.
In the early days of the getter developments, many attempts were made to use Ba in the most convenient forms. One of
these was the so-called reactive getter consisting in a mixture of barium strontium carbonate sprayed on a tantalum strip
(Batalum getter) [29]. Further developments led to the use of various types of Ba containing alloys and, finally, to the
stabilization of Ba by alloying with Al. The BaAl phase diagram is shown in Fig. 5.8. Among the different alloys of Ba
and Al, BaAl4 has been chosen because of its stability, allowing safe and reliable handling during manufacturing (as a
matter of fact, it is known as Stabil 2 alloy, shortened to St2). Other alkaline-earth metals have been studied [30] as
possible evaporable gettersparticularly Ca and Sr, which also form alloys with Al and have high vapor pressures (higher
than Ba) as seen in Fig. 5.9. For various reasons, however, Ba turned out to be the best compromise for different
requirements.
Fig. 5.8
Constitutional diagram of the BaAl system.

< previous page page_276 next page >


< previous page page_277 next page >
Page 277

Fig. 5.9
Vapor pressures of Ca, Sr, and Ba.

Fig. 5.10
Cross sections of a ring-shaped Ba getter:
(a) open center getter and
(b) closed center getter.

The BaAl4 alloy can be pulverized and compressed in a metallic container, usually in shape of a ring [31, 32] (with
"open" center or "closed" center) as shown in Fig. 5.10. The container can, however, also be in the shape of a disc or
wire.
This container can then be heated (e.g., inductively by a radio-frequency coil) to the dissociation temperature of the
alloy so that Ba is rapidly evaporated, thereby forming the active film. The evaporation temperature starts significantly
above 900°C (the melting point is about 1100°C). The process of rapid evaporation of Ba from a Ba getter is sometimes
called flashing of the getter. This type of getter is called endothermic [33, 34] since it needs energy for the dissociation
of the alloy. It requires high temperatures to start evaporation; and the process is not very well controlled since free Al
can also partly evaporate with Ba, thus negatively affecting the gettering

< previous page page_277 next page >


< previous page page_278 next page >
Page 278

properties of the Ba film, and partly react with the getter container (generating melted areas, holes, etc.).

To obtain getters having a lower evaporation temperature and more reliable and controllable characteristics, the BaAl4
powder is mixed with Ni powder and then compressed in the container. In this case, during heating, the following
reaction takes place between the two components: BaAl4 + 4Ni → 4AlNi + Ba. Ba evaporates and deposits onto the
available surfaces while Al is "blocked" by its reaction with Ni. The above reaction starts at approximately 800°C and
generates heat (enthalpy of reaction about 11.2 kcal/mol); the getter is therefore called exothermic [33, 34].

The time necessary to start the reaction from the beginning of heating is called start time, while the time from the
beginning of heating to the end of heating off is called total time of the flashing process. The maximum evaporated
quantity, or yield, of the getter is generally a percentage of the total quantity of Ba contained in the getter; however, in
special getters it can approach 100% (the getter is said to show total yield).

The flashing process for an exothermic getter is schematized in Fig. 5.11, where the temperature of the getter is shown
versus time. The getter heats up to the point where the above exothermic reaction starts and raises the getter temperature
up to 12001300°C. There is then a decrease corresponding to a slowdown of the reaction. During the process, together
with generation of heat there is obviously also heat dissipation mainly due to radiation losses and vaporization of Ba.

The quantity of Ba evaporated can range from few milligrams to a few hundred milligrams, depending on the getter
type and application. A typical "yield" curve for a Ba ring getter (Fig. 5.12) shows the quantity of Ba evaporated as a
function of start time for a certain total time. If the start time is too short, the getter is excessively heated and can melt;
of course, this is not acceptable and has to be avoided.

Interaction of Gases with a Ba Film. The interaction of a Ba film with residual gases in a vacuum device is a very
complex phenomenon as illustrated by Perdijk [35]

Fig. 5.11
Typical temperature trend of a Ba getter during the
"flashing" process.

< previous page page_278 next page >


< previous page page_279 next page >
Page 279

Fig. 5.12
Typical "yield curve" of a Ba getter.

in Fig. 5.13. The main reactions occurring at the surface of the Ba film with the usual main gases to be sorbed are, however, the
following primary reactions:

2Ba + O2→ 2BaO (93 mbar·liter g1)


3Ba + 2CO → 2BaO + BaC2 (107 mbar·liter g1)
5Ba + 2CO2→ 4BaO + BaC2 (67 mbar·liter g1)
3Ba + N2→ Ba3N2 (53 mbar·liter g1)
2Ba + H2O → BaO + BaH2 (80 mbar·liter g1)
Ba + H2→ BaH2 (173 mbar·liter g1)

Together with the reaction type, the stoichiometric capacity of the Ba film for the considered gas is also indicated. The actual
capacity, however, depends on the film characteristics, as will be seen later.

There could be also secondary reactions, the study of which might lead to a better understanding of the practical behavior of Ba
getters. For example, BaC2 can react with H2, and particularly with H2O, to generate hydrocarbons (especially CH4 and
C2H2). Also tertiary reactions are possible, whose practical effects are usually negligible. A comprehensive study of the
interactions of residual gases with a Ba film has been made by J. Verhoeven and H. van Doveren [36]. During sorption of O2,
following sorption of N2, a "displacement" phenomenon has been observed [37]: It consists of emission of previously adsorbed
N2 due to its replacement with O2.

Sorption Characteristics. The sorption properties (speed and practical capacity) of a Ba film depend not only on the type of the
sorbed gas and the operation temperature

< previous page page_279 next page >


< previous page page_280 next page >
Page 280

Fig. 5.13
Types of reactions which can take place on or Ba film; besides primary reactions,
also secondary and tertiary reactions can take place. The getter film is therefore a
dynamic means to keep the atmosphere in a vacuum device as low as possible and
in any case in a reducing state (for the benefit, for example, of oxide
cathodes present). (From Perdijk [35].)
Fig. 5.14
Typical sorption curves (speed
versus sorbed quantity) for various
gases on a Ba film.
(From P. della Porta and L. Michon [38].)

< previous page page_280 next page >


< previous page page_281 next page >
Page 281

but also on the physical characteristics of the Ba film itself, which are linked to the flashing conditions (pressure,
temperature of the substrate, surface characteristics of the substrate, thickness of the film, etc.).

The sorption curves for CO, N2, and H2 are typically represented in Fig. 5.14 [38]. They have been obtained by flashing
about 80 mg of Ba inside a cathode ray tube (CRT) on a surface of a few thousand of square centimeters. CO is the gas
for which the initial speed is the highest, and N2 is the gas with the lowest speed. The actual curves per unit area of Ba
film, even if they don't substantially change in the relative trend, can be modified in terms of speed and capacities,
depending on various parameters. The initial speed values can be particularly affected by flashing conditions and
sorption test conditions, as indicated by the different values shown in Table 5.3 [39,40]. This table shows the sticking
probabilities corresponding to initial speeds measured by different authors for some common gases: N2, CO, H2, CO2,
O2, and H2O.

The kinetics of the sorption process is characterized by surface and bulk diffusion, so that it can be schematically
represented as in Fig. 5.15. This is particularly

Table 5.3. Summary of Initial Sticking Probability Values for Barium Film
Gases Sticking Probabilities Authors
(Initial)
H2 Wagener
1 × 103
della Porta and Origlio
4.17 × 105
Ichimiya, Mizushima, and Oda
1 × 104
N2 Wagener
3 × 104
della Porta
5.03 × 105
1 Ichimiya, Mizushima, and Oda
Wagener
4 × 101
CO Bloomer
1.2 × 102
Morrison and Zetterstrom
2.7 × 102
della Porta and Ricca
2.8 × 102
0.3 Bloomer and Cox
O2 0.4 Verhoeven
0.6 Verhoeven
CO2 0.6 Wagener
H2O 0.6 Verhoeven
References
S. Wagener, J. Phys. Chem. 60, 567 (1956).
P. della Porta and S. Origlio, Vacuum 10, 227 (1960).
T. Ichimiya, Y. Mizushima, and Z. Oda, Proc. 1st Int. Congress on Vacuum Namur., 1958, Vol.
II, p. 641. Pergamon Press, Oxford, 1960.
P. della Porta, Vacuum Symposium, Transactions, p. 317. Pergamon Press, London, 1959.
R. N. Bloomer, Br. J. Appl. Phys. 8, 352 (1957).
J. Morrison and R. B. Zetterstrom, J. Appl. Phys. 26, 437 (1955).
P. della Porta and F. Ricca, Le Vide, 85, 1 (1960).
R. N. Bloomer and B. M. Cox, Br. J. Appl. Phys., 16, 1331 (1965).
J. Verhoeven, Vacuum, 30, 69 (1979).

< previous page page_281 next page >


< previous page page_282 next page >
Page 282

Fig. 5.15
Speed versus time during sorption according to
different mechanisms.

Fig. 5.16
Approximate structure of a Ba film.

emphasized by the fact that the actual structure of a Ba film is granular. This structure is depicted in Fig. 5.16, where the
small circles represent the Ba microparticles forming the film. The particles can be smaller or bigger and the structure
can be more or less compact, depending on the formation conditions of the film.
The model for the various stages of sorption on a Ba films is schematically shown in Fig. 5.17. At the beginning of the
process, mainly the surface adsorption on the ''external" surface takes place since this surface is readily available for
sorption. Then, surface adsorption occurs at the internal surfaces of the film; bulk diffusion takes place simultaneously,
but it is slower and therefore accounts for the part of the sorption curves with lower speed and higher capacity. The
presence of the diffusion process is evidenced by Fig. 5.18 [39]. This figure shows the amount of CO, N2, and H2
sorbed as a function of the square root of time during which the sorption has taken place. As

< previous page page_282 next page >


< previous page page_283 next page >
Page 283

Fig. 5.17
Model showing the different
stages of gas sorption on a Ba film:
(a) Surface diffusion, (b) internal
surface diffusion (intergranular
diffusion), and (c) bulk diffusion.
Fig. 5.18
Diffusion process for various gases in a Ba
film. The sorbed quantity of a gas at a given
time t, divided by the total sorbed quantity
at that time, is plotted as a function of the
square root of time. (From della Porta [39].)

illustrated previously, in a semi-infinite medium the linear dependence of the sorbed quantity versus the square root of
time is an indication of the presence of a bulk diffusion process. As a matter of fact, the plots in the figure show this
trend, apart from deviations at the beginning and at the end of the curves. The initial deviations from linearity indicates
the existence of a phenomenon different from bulk diffusion, whose

< previous page page_283 next page >


< previous page page_284 next page >
Page 284

rate decreases with time less rapidly than in the case of a diffusion process: as mentioned above, in fact, surface
adsorption prevails at the beginning of the sorption process. The deviation from linearity toward the end of the curves
indicates the influence of the limited film thickness on bulk diffusion (due to deviation from the semi-infinite medium
assumption).

Temperature-Dependence. If sorption occurs at increasing temperatures, also the sorbed quantity increases since
diffusion is enhanced, as can be seen in Fig. 5.19 with regard to CO [41]. In the case of H2, because of the reversible
character of its adsorption, the increase of temperature can eventually excessively increase the equilibrium pressure of
this gas. The H2 equilibrium pressure in the H2Ba system is shown in Fig. 5.20.

The effect of temperature is found to be different for CO, N2, and H2; this can be seen in Fig. 5.21 [4143], which shows
the temperature-dependence of nitrogen, hydrogen, and carbon monoxide diffusion in a Ba film (2 mg of Ba on an
apparent surface of 100 cm2). The figure shows that for CO and N2 there is a critical temperature indicating a change in
the sorption process, from surface diffusion (at lower temperatures) to bulk diffusion (at higher temperatures). The
curves allow one to calculate the activation energies for the processes involved, corresponding to the potential barrier
which the diffusing particle must overcome to move from one lattice point to another, when the movement is on a plane
(surface diffusion) or into the solid (bulk diffusion). On the contrary, H2 does not seem to show a critical temperature,

Fig. 5.19
Sorption throughput as a function of the sorbed quantity at different
temperatures. The apparent area of the film were 100 cm2 and the evaporated
quantities were 2 mgexcept at 423 K and 473 K, when this quantity was
5 mg and 8.5 mg, respectively. During the tests the conditions ensured that
the same number of molecules were striking the unit area per
unit time. (From Ricca and P. della Porta [41].)

< previous page page_284 next page >


< previous page page_285 next page >
Page 285

Fig. 5.20
Equilibrium isotherms for the BaH2 system.

indicating that diffusion of this gas is already important at lower temperatures and overshadows surface diffusion.

Thickness-Dependence. The sorption characteristics can change as a function of the film thickness, as can be seen in
Fig. 5.22 in the case of N2 sorption [42]. At a very small thickness, all the available Ba reacts with the gas molecules.
When the thickness increases, eventually the more internal layers of the Ba film are no longer accessible and the
quantity of gas sorbed becomes almost independent of the thickness. This trend can be modified by increasing the
sorption temperature thanks to the enhanced diffusion process.
For a compact fully sintered Ba film, it has been found that there is a maximum thickness involved in room temperature
sorption which in the case of CO is about

< previous page page_285 next page >


< previous page page_286 next page >
Page 286

Fig. 5.21
Temperature dependence of N2, H2, and CO
diffusion in barium films. The weight of the
evaporated Ba is 2 mg, and the apparent surface
is 100 cm2. (From della Porta and F. Ricca [42].)
Fig. 5.22
Total amount of sorbed N2 as a function of
the thickness of the Ba film at 298 K.
(From della Porta and F. Ricca [42].)

< previous page page_286 next page >


< previous page page_287 next page >
Page 287

90 Å [44]. This corresponds to about 0.04 Pa·liter of CO sorbed per square centimeter of a 3.6 µg Ba film as a
consequence of the reaction: 3Ba + 2CO → BaC2 + 2BaO. When the critical temperature is reached or exceeded,
however, sorption continues until all the available Ba has reacted.

Influence of Evaporation Conditions. The physical characteristics of the Ba film strongly depend on the ambient
pressure during evaporation and on the temperature and characteristics of the surface onto which it is formed [45]. As a
matter of fact, depending on these conditions the formed Ba film can be more or less porous, as already mentioned. The
structure of a Ba film can change as, for example, depicted in Fig. 5.23 at different substrate temperatures during Ba
deposition. The actual surface appearance of a Ba film deposited at room temperature and at 340°C has been studied, for
example, by Perdijk using a scanning electron microscope (SEM) [45].

Similar results are found if Ba is evaporated in the presence of a certain pressure of gases such as noble gases, methane,
or nitrogen. If the film is formed in high-vacuum conditions, it is less porous than a film deposited in low-vacuum
conditionsas was shown, for example, by Hoshimoto and Kitagawa [46]. The differences in porosity translate into a
more or less efficient sorption.

Figure 5.24 shows the amount of N2 sorbed at room temperature by a Ba film formed at different condensation
temperatures of the film [42]; the higher the temperature the more compact or sintered is the film and therefore the
lower its capacity.

The sorption performances of a Ba film can change by changing roughness and surface porosity of the substrate on
which the film is formed [47]. In general, rough
Fig. 5.23
Schematic of the structure
of a Ba film to show how
porosity of the film depends
on the formation conditions.
(a) Highly porous: Glass wall at
about 27°C and pressure (Ar,
CH4, N2) between 5 and 0.1 mbar.
(b) Partially sintered with some
porosity: Glass wall at about
100°C and some outgassing from
the getter assembly. (c) Sintered
with very low porosity: glass wall
at about 350°C and high
vacuum (about 105 mbar).
(From Perdijk [35].)

< previous page page_287 next page >


< previous page page_288 next page >
Page 288

Fig. 5.24
Total amount of N2 sorbed by Ba films as a function
temperature of condensation of the film (Ba weight,
2 mg; apparent surface area, 100 cm2; test pressure,
5.104 Torr). (From della Porta and F. Ricca [42].)

surfaces tend to produce also higher surface area films with improved sorption characteristics (unless the formed film is
relatively thick).

Ba Distribution. The distribution of a Ba film is important in determining its sorption characteristics; if the film area is
large, the total speed will also be large. The same amount of Ba in smaller areas corresponds to a thicker film with
reduced performances. On the screen of a CRT a high film thickness is to be avoided to prevent an excessive barrier to
the electron beam.

The film distribution is strongly affected by the getter position and by the actual flashing conditions. There are simple
methods developed to determine the Ba distribution [4749], for example, in a CRT. They are used to study the
parameters influencing this distribution and to choose the appropriate getter type and mounting positions.

Gas-Surface Reactions. As mentioned above, there are various possible surface reactions, among which methane
formation deserves more specific attention. In practice, it is found that the considerations made for a Ba film also hold
for a Ti film.

Specific studies have been performed to investigate the possible generation of CH4 on a getter film under sorption of
gases containing carbon and H2 [5052]. For example, results are reported by dosing CO2 and H2 or H2O with alternate
sequences on Ba and also Ti films [53]. The tests have been performed at relatively high pressures and therefore high
surface coverages. It is found that the active gases are well sorbed by the films but generate some methane in
approximately the same quantity; this quantity is in the range of 15% of the amount of CO2 admitted. It is also observed
that there is some effect of the sequence of dosing; if CO2 is admitted first, more methane is formed.

< previous page page_288 next page >


< previous page page_289 next page >
Page 289

Pumping of methane [53] can be obtained with these getter films by cracking this gas with a hot metallic surface, or by
reaction with suitable metal oxides as follows:

MeOx + CH4→ Me/MeOy + H2O + CO2.

The getter films sorb the formed active gases and force the reactions toward the right-hand side of the equation.

Getter Types. Various types of getters have been developed to cope with the evolving requirements of Ba getter
applications (particularly, color television tubes). In addition to the common types, high-performance getters are
available such as gas-doped getters, total-yield getters, high-yield getters, frittable getters, and low-argon getters.

Gas-doped getters contain, mixed with the BaAl4 and Ni powders, a small amount (usually less than 5% by weight) of a
compound which dissociates at the beginning of Ba evaporation generating a pressure peak of N2 (in the range of
102101 mbar) within the vacuum device to better distribute the Ba and allow the formation of a more porous film [52].
The commonly used compound that generates N2 is iron nitride (Fe4N) [54]. The process is shown in Fig. 5.25. The N2
is soon reabsorbed by the Ba film. The amount of N2 added is small compared to the capacity of the Ba film, so that the
overall effect is an increase of sorption performances. Figure 5.26 shows the good room temperature sorption
characteristics of Ba films for CO obtained by evaporating Ba from N2-doped getters as compared to the case when no
doping was present [55]. The curves for N2-doped getters have been obtained at different temperatures of the CCRT
bulb during flashing to also show the effect of the substrate temperature during film deposition.

Since part of the Ba evaporates when the pressure is already greatly reduced, a second N2 peak, between start time and
total time, can be generated to further improve the gettering performances of the Ba film. The getters exhibiting this
characteristics are the so-called delayed N2-doped getters [56].

Fig. 5.25
N2 evolution and barium yield during flashing of
an N2-doped getter.

< previous page page_289 next page >


< previous page page_290 next page >
Page 290

Fig. 5.26
CO sorption curves for doped and nondoped (dashed line)
Ba getters at different substrate temperatures during film
deposition (sorption test at room temperature).
(From della Porta [55].)

Fig. 5.27
Section of a "total yield" getter showing the evaporation paths of Ba.

Total yield getters have been developed to release almost all of the Ba contained in the getter. They need a special
arrangement to control evaporation. The structure of the getter (Fig. 5.27) allows evaporation from two surfaces, thus
increasing yield [57,58].
High-yield getters meet the needs for large size CCRTs requiring increased amounts of Ba to cope with the increased
gas load. Closed center getters with special indentations on the evaporating surface have been developed for this
purpose [59]. Yields of Ba in the range of 300 mg and more can be obtained.

Frittable getters have been developed to withstand the frit cycle in CCRT manufacturing processesthat is, heating in air
up to 450°C for 1 or 2 h. Under these

< previous page page_290 next page >


< previous page page_291 next page >
Page 291

conditions, they still show a controlled exothermicity during the getter flashing process [60, 61].

Low-Argon getters reduce the Ar content in a CCRT, which is partially due to a normal getter itself. Ar is not
chemically noxious to the CCRT cathodes. However, in some tubes, where impregnated cathodes are used, the Ar
partial pressure has to be minimized to avoid damaging the cathodes because of the Ar ion bombardment.

5.6.5.2
Titanium Sublimation Getter Pumps

Ti can be evaporated or sublimed onto a surface to form, as in the case of Ba, a highly active film. In case of Ti the term
sublimation is used as a synonym of evaporation (Ti vapor is produced from the solid without passing through the liquid
phase). Ba has, however, a much higher vapor pressure compared to Ti, so that its evaporation can take place very
quickly and therefore fits very well those applications (such as CRTs) where minimization of the process time is
essential. At 1000°C, for example, the vapor pressure of Ti is about 109 mbar while that of Ba is about 1 mbar. To reach
a vapor pressure of about 103 mbar, Ti needs to be heated to 1500°C, so a relatively high power is required.

In general, freshly evaporated metals exhibit more or less good chemisorption properties, as can be seen from Fig. 5.28
[62], where the sticking probabilities for O2 are given as a function of the O2 sorbed quantity. Ti is effective and
particularly good, compared to others, at relatively high O2 sorbed quantities. From various sorption studies and for
practical reasons, Ti turns out to be, in general, the most appropriate metal for use in sublimation pumps. These pumps
are applied in vacuum systems where the sublimation process can be controlled (using suitable power suppliers) so as to
adjust the sublimation rates and cycles to the needs of the

Fig. 5.28
Reaction probability of oxygen on various metal films as a function
of the gas sorbed quantity.
(From Fromm and Uchida [62].)

< previous page page_291 next page >


< previous page page_292 next page >
Page 292

application and where the surfaces available for the metal film formation are generally large.

Ti sublimation pumps have been extensively studied particularly in UHV systems. A recent comprehensive review has
been written by K. Welch [63].

Sorption Characteristics for Various Gases. For this type of film the sorption characteristics are often expressed in
terms of sticking probability or coefficient as a function of the surface coverage, expressed for example in molecules per
cm2 (2.2 × 1019 molecules correspond to about 1 mbar·liter at 20°C). Typical data are shown in Fig. 5.29 [64]. It is
seen that the sticking probabilities are widely separated for H2 and N2. H2, however, maintains a more constant sticking
probability as a function of the coverage, as expected, because of its relatively high bulk diffusivity combined with a
relatively high solubility. The capacities range from somewhat less than 1 monolayer for N2 to about 10 monolayers for
O2. When about one or a few monolayers, depending on the gas type, are formed, the sorption speed decreases very
rapidly; the film therefore has to be refreshed by a new evaporation. This continues until the available Ti is used up. Ti
can also be slowly but continuously evaporated to keep a more constant sorption speed.

Temperature-Dependence. As in the case of Ba films, Ti films sorption characteristics are influenced by the temperature
of deposition of the film. Figure 5.30a [64] shows the trend of the sticking probability for CO sorption as a function of
the surface

Fig. 5.29
Room temperature sticking probability for various gases as
a function of the surface coverage of the Ti film.
(From Gupta and Leck [64].)

< previous page page_292 next page >


< previous page page_293 next page >
Page 293

Fig. 5.30
Effect of the substrate temperatures and of the sorption
temperatures on the sticking probability (O2 and CO). For
each curve the numbers indicate first the substrate
temperature during film deposition and second the sorption
temperature. (From Gupta and Leck [64].)

coverage of a Ti film formed at different temperatures. The best results are obtained forming the Ti film at 77 K and
sorbing gases on the film kept at the same temperature. Films formed at lower temperatures are more porous and
therefore exhibit higher surface capacities. The effect is less important in the case of O2 (Fig. 5.30b) [64].
Various studies have been carried out concerning the temperature dependence of sticking coefficients [65, 66]. Harra
has summarized these data as shown in Table 5.4 [65].

Dependence on Thickness. As the Ti film thickness is increased, the sorption capacity does not increase proportionally
because the film tends to be more compact

< previous page page_293 next page >


< previous page page_294 next page >
Page 294

Table 5.4. Summary of Sticking Probability Values and Sorbed Quantities at Saturation at 300 K
and 78 K, for Various Gases with Titanium Film [65]
Initial sticking Quality Sorbed
coefficient (×1015 molecules/cm2)
300 K 78 K 300 K 78 K
H2 770
0.06 0.4 8230
D2
0.1 0.2 611
H2O
0.5 30
CO 50160
0.7 0.95 523
N2 360
0.3 0.7 0.312
O2
0.8 1.0 24
CO2
0.5 424
Fig. 5.31
Dependence of the saturation coverage for N2 on
the film thickness. (From Gupta and Leck [64].)

and its internal layers become less accessible to the gas [6466]. This can be seen in Fig. 5.31 [64], where the saturation
coverage (corresponding to practically zero sticking probability) for CO is given as a function of increasing thickness. It
is seen hat if the thickness increases about 20 times, the capacity increases about 6 times. This

< previous page page_294 next page >


< previous page page_295 next page >
Page 295

means that to better exploit the available Ti one should evaporate less Ti, but more frequently.

Displacement Phenomena and Surface Reactions. When various gases are simultaneously pumped, it is observed that
some gases can displace others [64, 67]. For example, Gupta and Leck have found that O2 can displace all other gases;
CO can displace CH4, N2, and H2; H2 can displace N2 and CH4, and N2 can displace only CH4. This can probably
occur at relatively high coverage and when no real strong chemical bond has been formed at the surface. As already
mentioned, there is the possibility of CH4 formation on a Ti film due to the reaction between surface C and the sorbed
H2O and H2. This effect is reduced if the film is at 77 K. Some authors [64], performing studies in high-vacuum
conditions, have also considered methane generation as possibly due also to the displacement of preadsorbed CH4 by
hydrogen sorption.

Peeling. It is observed that when the Ti film has grown to a certain thickness, it starts peeling off. It is reported [68] that
this effect is observed when the deposition corresponds to about 0.0230.03 g·cm2that is, about 6075 µm (if a density of
4 g·cm3 is considered). The film adhesion is poor if the substrate surface is very smooth (it is reported that sandblasted
substrate surfaces are less prone to peeling). If peeling occurs, H2 can be released and generate undesirable pressure
increases. Peeling Ti films are pyrophoric and can ignite when exposed to air. It is therefore necessary to clean the Ti-
coated surface from time to time.

Types of Sublimators and Pump Structure. The Ti vapor sources can be of different types: filament type (heated by
direct passage of current), radiation-heated type and electron-gun-heated type. Ti is often used in alloy form rather than
as a pure metal. In the filament type, Ti is commonly used as an alloy with Mo or Ta to have good mechanical
resistance and a convenient electrical resistivity. A typical ''filament" shape [69] is shown in Fig. 5.32. In the radiation
heated type a heater radiantly heats a Ti body of the shape for example shown in Fig. 5.33 [70] (known as the Ti Ball).
The evaporation rate is finely controlled by regulating the electric current used for heating. These sublimators are
immersed in a chamber and evaporate Ti onto the inner wall

Fig. 5.32
Simple Ti sublimator of the "filament" type.

< previous page page_295 next page >


< previous page page_296 next page >
Page 296

Fig. 5.33
Ti sublimator of the "ball"-type TiBall®.
(From Harra and Snouse [70].)

Fig. 5.34
Structure of a Ti sublimation pump, with
LN2-cooled walls.

surfaces. These walls are often arranged to be cooled down to LN2 temperature (Fig. 5.34). The structure of the pump is
relatively simple. Of course, in order for these pumps to exhibit high speeds the inner walls should be large. Sometimes
there are baffles to prevent the evaporated Ti from entering the vacuum system; conductance limitation could, however,
practically prevent full exploitation of the high speeds available.

Practical sorption speeds for various gases are given in Table 5.5 at two condensation temperatures [71]. The time
between regenerative sublimations depends on the pressure of the vacuum system and the gas type to be sorbed.
Sublimation can be almost continuous at relatively high gas loads (105 to 106 mbar pressure range) or periodic with
long (1015 h) intervals at low gas loads (in the range of 1011 mbar). Because of the high temperature needed for Ti
sublimation, the sublimator has to be submitted to careful degassing procedures to reduce particularly H2 generation at
onset and an excessively high steady pressure. H2 turns out not to significantly influence the sorption characteristics for
other active gases; on the contrary, H2 sorption is negatively affected by contamination of the Ti film by other gases
[64, 72].

< previous page page_296 next page >


< previous page page_297 next page >
Page 297

Table 5.5. Typical Practical Pumping Speeds per Unit Surface of a Ti Film for the Most Common Gases, at Different Sorption
Temperature [71]a
Gas Species
(Temperature) H2 N2 O2 CO CO2 H2O CH4 Inerts
+ 20°C
3 4 9 9 8 3 0 0
196°C
10 10 11 11 9 14 0 0
a Temperature is of condensing wall and shows resultant pumping speed of Ti film. Units are liter·s1·cm2.

Table 5.6. Some Physical and Electronic Characteristics of Common Base Getter Metals
Ti Zr Hf Th
Outer electron configuration
3d2 4s2 4d2 5s2 5d2 6s2 6d2 7s2
(5f) 6d 7s2
Atomic radius (6-coord.) (Å)
1.49 1.62 1.58 1.82
Differences in Al atomic radius (%)
4.03 11.66 9.49 21.43
α-phase parameters (Å)
cph cph cph bcc
a = 2.9503 a = 3.230 a = 3.1883 5.086
c = 4.6831 c = 5.133 c = 5.0422
αβ transition temperature (°C)
882 862 1950 1400
Melting point (°C)
1670 1860 ~2000 1750

5.6.6
Nonevaporable Getters

Many different kinds of nonevaporated getter materials have been developed since World War II. Metals of the IVB group and some of the
actinides and rare earths, as already mentioned, have often been used to make getters. In Table 5.6, the properties of some of these metals are
summarized. Also, in this table the percentage difference of the atomic radius of Al is compared to the other metals. Al is often alloyed with
the above-mentioned metals to obtain specific gettering characteristics. The use of Al tends to reduce the surface reaction rates with air at room
temperature, but increases the diffusion rates of the adsorbates at higher temperature.

Among the early nonevaporable getters, it is worth mentioning the Th-based getters, called Ceto getters [73] (a similar old type of getter is
called Ceralloy 400 [74]). Ceto is a quaternary alloy, made by sintering a mixture of Th powder and powdered alloy called "Ceral" at 900°C
in vacuum. Ceral consists of Cermishmetal (80% Ce and 20% La) and Al and has the following chemical composition: (Ce, La) Al2. The
final atomic composition of the getter is about 10 Th, 2.5 Ce, 0.5 La, and 6 Al. La can be completely replaced by Ce to obtain a ternary
alloy (ThCeAl) having practically the same properties as the quaternary alloy. Also, the Th2Al binary alloy has been studied and used as a
getter, particularly for H2 [75]. The ultimate amount

< previous page page_297 next page >


< previous page page_298 next page >
Page 298

of H2 in the compound is found to be approximately 16 atoms of H2 per unit cell (Th8Al4). Above this concentration
the equilibrium pressure rises rapidly to high values. An extrapolation of the curve for the Th8Al4H4 compound results
in an equilibrium pressure of approximately 1013 mbar at room temperature [73].

Zr and Ti are metals now commonly used as getters, particularly in alloy form with one or more additional elements.
Many alloys have been studied in the last few decades to meet the requirements of gettering all possible gases or some
specific gases present in various applications. Some of these alloys will be described in the following.

Zr and Ti can be used as pure elements in powder form to make getters for some applications. Ti is a somewhat better
getter at room temperature but similar or worse compared to Zr at working temperatures above 400°C [76]. A
comparison is shown in Table 5.7 [76], where the sorption characteristics at 400°C appear to be similar for the two
metals for various gases. For comparison, the characteristics for Thorium are also shown. Except for O2, this metal
exhibits poorer gettering properties than the other two metals. It is also to be pointed out that Ti is more prone to
become sintered than Zr; therefore Zr powders can be more easily activated and work at higher temperatures than Ti,
without losing too much surface area. To avoid or reduce sintering, these pure metals when used as getters are usually
mixed with powders of appropriate metals or of other substances.

The equilibrium pressures of H2 with these metals have been studied in detail [77, 78], and the related results are shown
in Figs. 5.35 and 5.36, where Sieverts' law is found to apply, respectively, in the case of α-Ti and α-Zr. These curves
can be expressed analytically as follows:

where p is in mbar, q is in mbar·liter (of H2) per gram (of getter material), and T is in Kelvin.

The Binary Zr and Ti Alloys. These alloys, particularly those with the addition of Al, have long been investigated with
very interesting results especially in the case of Zr.

The ZrAl, TiAl, and also the ThAl systems have been studied in a wide range of Al contents [76]. The gettering
properties were measured for various compositions. The results obtained in the case of N2 and at a sorption temperature
of 400°C, after

Table 5.7. Gettering Rates of Ti, Zr, and Th for Various Gases, at 400°C [76]
Gettering Rate (cm3·s1·cm2)
Gas Ti Zr Th
N2
36 30 5.6
CO
43 44 20
CO2
44 38 15
O2
85 85 122
H2
285 222 213
< previous page page_298 next page >
< previous page page_299 next page >
Page 299

Fig. 5.35
Sieverts' plots for H2 in α-Ti.
(From Giorgi and Ricca [77].)

activation at 900°C, are shown in Fig. 5.37 [76]. It is seen that the behavior of the ZrAl and ThAl systems are rather
complex, showing maxima and minima in the sorption properties as a function of the Al contents. The TiAl system, on
the contrary, shows a regular continuous decrement in gettering rates with increasing Al contents and, in any case,
poorer sorption characteristics compared to the other alloys for compositions above a few percent of Al. The ZrAl alloys
show the best sorption properties compared to the other alloys (and pure metals) in a wider range of Al contents and
have the highest peak in the gettering rates when Al is about 16% by weight. The Zr(16%) Al alloy has therefore been
extensively studied and used in many practical applications [7680]; today it is one of the most common getter materials
applied (under the trade name of St101) in vacuum technology.

Among the getter binary alloys, the Zr-based alloys containing V, Ni, Fe, Mn, or Co have also received interest for
some practical applications.

The ZrAl Getter Alloy. The sorption characteristics for H2 and CO have been studied together with those for N2 at
different compositions of the ZrAl system, still finding a maximum in the sorption properties around 16% Al. This can
be seen in Fig. 5.38 [80], where, for ease of comparison, the gettering rates are given as the ratio (G/Gmax) of each
single sample rate versus a given gas and the rate of the most active sample versus the same gas. The figure also shows
the phase compositions at various
< previous page page_299 next page >
< previous page page_300 next page >
Page 300

Fig. 5.36
Sieverts' plots for H2 in α-Zr.
(From Ricca and Giorgi [78].)

Al contents of the tested ZrAl samples (the ratio of the indicated numbers in the abscissas represents the ratio between
Zr and Al atoms). The results refer to getter samples working at 400°C, after activation at 1000°C for 30 s in the 106-
mbar range of pressures. The following main considerations can be made concerning this getter material: (a) The
maximum activity is in the biphase region containing Zr5Al3 and Zr3Al2 intermetallic compounds. (b) The activity
increases rapidly from the eutectic β-Zr-Zr5Al3 (with 11%Al) to higher Al contents with a maximum around 16% and
decreases rapidly with the Al contents greater than 17.5%. (c) The maximum activity is not exactly coincident for the
various gases. It is maximum for N2, H2, and CO, respectively, at the compositions having the following Al weight
percentages: 1415.5, 15.5, and 16. The best composition compromise is therefore Zr(15.516%) Al. (d) The relative
sorption rates found indicate that the getter alloy exhibits a particularly high rate for H2.

The sorption rates of the Zr(16%)Al composition have been investigated for various gases as a function of temperature,
with the results shown in Fig. 5.39 [76]. It is seen that great increases are observed for all the tested gases above 200300°
C and particularly in the case of CO. The diffusion rates are especially high for H2 and their dependence on temperature
can be better seen in Fig. 5.40 [76].

The characteristics of the H2ZrAl alloy system are shown by the isotherms represented in Fig. 5.41 [81, 82]. The
equilibrium pressure of H2 can be very low and
< previous page page_300 next page >
< previous page page_301 next page >
Page 301

Fig. 5.37
Relative gettering speeds of TiAl, ZrAl,
and ThAl alloys, as a function of
composition. (From della Porta [76].)
Fig. 5.38
Relative gettering speeds for different gases (at 10 min from the
beginning of the sorption test) as a function of the ZrAl alloy compositions.
(From Barosi [80].)

< previous page page_301 next page >


< previous page page_302 next page >
Page 302

Fig. 5.39
Gettering speeds for different gases on the Zr(16%)
Al getter (activated at 1000°C for 30 s) as a
function of sorption temperature.
(From della Porta [76].)
Fig. 5.40
Dependence on temperature of the diffusion rates of
various gases in the Zr(16%)Al alloy.
(From della Porta [76].)

< previous page page_302 next page >


< previous page page_303 next page >
Page 303

Fig. 5.41
H2 isotherms in the H2-Zr(16%)Al getter system.
(From Ferrario [81].)

very compatible with operation in ultra- or extra-high vacua. The correlation between H2 concentration, equilibrium
pressure, and temperature of the getter can be represented by the Sieverts' equation:

where p is in mbar, q is in mbar·liter·g1, T in Kelvin.

The activation of this binary getter alloy is performed by usually heating up to temperatures in the range of 700900°C;
the times involved range from a few hours to half a minute. Since, in some practical cases, also partial activations can
be acceptable, the various possible activation conditions for this alloy have been investigated and found to be those
shown in Fig. 5.42 [84, 85]. In this figure the 100% curve corresponds to full activationthat is, the activation conditions
giving the maximum sorption efficiency. This situation can be reached, for example, by heating at 900°C for 3040 s or
at 750°C for about 30 min. Sixty percent activation can be achieved, for example, by heating the getter at 700°C for 10
min or at 650°C for about 90 min. The activation conditions are the results of many sorption curve measurements;
however, the fundamental mechanism accounting for these results has been clarified by specific
< previous page page_303 next page >
< previous page page_304 next page >
Page 304

Fig. 5.42
Activation efficiency for the Zr(16%)Al getter as a
function of temperature and time of activation.

Fig. 5.43
Surface composition of the Zr(16%)Al getter
during heating at various temperature.
(From Ichimura et al. [86].)
surface analyses based on X-ray photoelectron spectroscopy (XPS). This type of surface analyses has been performed
on this getter material at different temperatures, and the results are shown in Fig. 5.43 [86, 87]. The surface composition
is dominated by carbon and oxygen until the temperature exceeds about 600°C, when these elements tend to disappear
and the surface finally shows a metallic and therefore reactive character by the dominant presence of Zr and Al.

< previous page page_304 next page >


< previous page page_305 next page >
Page 305

Other Zr- or Ti-Based Binary Alloys. Various other types of binary alloys based on Zr and/or Ti have been studied in
recent years for specific gettering applications. A few, so far the most common, of these alloys are briefly described in
the following.

ZrCo alloys have been studied particularly for hydrogen isotopes sorption and storage, especially in experimental
nuclear fusion machines to replace the efficient, but also unsafe, uranium. The composition having about 61% Zr has
been thoroughly investigated [88, 89].

ZrNi alloys have been studied, particularly the intermetallic compound Zr2Ni (75.7% by weight of Zr) [86, 90]. This
getter exhibits good sorption capabilities especially for H2 and H2O.

The presence of Ni in the alloy or simply mixing Ni powder with Zr powder has been shown to speed up and enhance
the sorption of H2 (when relatively high pressures are involved) even if no previous high-temperature activation is
carried out [91]. Ni is believed to act as a dissociation catalyst for H2, thus facilitating H2 sorption according to the
mechanism described by van Vucht.

ZrFe alloys, particularly the Zr2Fe composition (76.5% by weight of Zr), have been developed for some special
applications. This getter material can be activated at relatively low temperatures but shows higher hydrogen equilibrium
pressures compared to the St101 getter [82, 92, 93]. The gettering capability, very good for other active gases, has been
found to be very poor for N2 (it improves at temperatures higher than 300°C).

ZrV alloys have been extensively studied [94] among the materials to be used for H2 storage. The ZrV2 composition
appears to exhibit interesting gettering capabilities; however, it is considered to be too pyrophoric for common and
practical applications on a large scale. Ternary alloys containing Zr and V have been investigated instead with the result
that some of the best present getter materials have been produced.

ZrTi alloys have also been studied as getters showing inferior gettering capabilities compared to the fully activated Zr
(16%)Al getter alloy; this is illustrated in Fig. 5.44 [95], where Zr is shown for comparison.

5.6.6.1
Ternary Alloys

Binary and ternary alloys, particularly based on Zr, have been extensively studied as hydride materials. The ternary
alloys have been especially interesting to better modulate the characteristics of interaction between the metal and H2
[9699]. Attention was eventually focused on these alloys also from the point of view of gettering.

Pebler and Gulbransen [94] have, for example, investigated the formation of hydrides of the Laves-phase binary
compounds ZrM2, where M = V, Cr, Mn, Fe, Co, and Mo. It was found that the absorption capacity (measured up to 1
bar) decreases significantly with the increase in the 3d orbitals occupation number of the transition element, M, across
the 3d series. As an example, the capacity for the ZrV2 compound was found to be 4.8 H2 atoms per formula; when M
is Fe or Co, this capacity is instead 0.2 H2 atoms per formula. Shaltiel et al. [100] have then studied the formation of the
hydrides of ternary (pseudobinary) intermetallic compounds of the general formula Zr(A1 x, Bx)2, where A designates
V, Mn, or Cr and B designates Fe or Co, with x changing between 0.05 and 0.9. The advantage found versus previous
hydride materials was both economical and related to the fact that these Zr-based compounds did not require high
activation temperature.

< previous page page_305 next page >


< previous page page_306 next page >
Page 306

Fig. 5.44
CO gettering characteristics for different getter
materials, at 400°C. In all cases the material mass in
the same and the test pressure is 3 × 106 Torr.
(From A. Giorgi [95].)

These types of materials have been studied for storage of H2 isotopes in experimental fusion machines and as getters for
HV and UHV applications.

It is useful to point out here that a material having good sorption capabilities for a gas such as hydrogen, at relatively
high pressures, does not necessarily imply that it could be a good getter material. The sorption speed may be too small
(even for hydrogen), or the diffusion characteristics too poor, to allow diffusion of active gases different from hydrogen
when low pressures are involved. Therefore specific and sometimes complex studies must be performed to develop a
getter even starting from some known general features.

Among the various ternary alloys studied, the ZrVFe alloy in a certain range of compositions has shown excellent
gettering qualities for all active gases combined with other interesting physicochemical characteristics.

The ZrVFe Alloy. The study of ZrV2 as a getter material showed that good sorption capabilities for H2 could be
achieved. However, this material was found to be pyrophoric to a degree which may be undesirable for industrial scale
use. The introduction of Fe and the optimization of the Zr percentage in the alloy allowed the achievement of excellent
gettering properties for all active gases, after activation at relatively low temperatures, together with acceptable handling
characteristics [101]. In particular, the alloys which contain Zr ranging from 45% to 75% by weight, with the other
elements' percentages also ranging in certain intervals, allow activation temperatures not exceeding about 500°C [102].
The Zr(70%)V(24.6%)Fe(5.4%) composition (which is also known commercially as St707) is particularly interesting
[101]. The flammability point of the St707 getter alloy is found to be higher than that of the ZrV2 alloy in similar
powder size and with similar getter configuration

< previous page page_306 next page >


< previous page page_307 next page >
Page 307

characteristics [101] (the absolute values can change also as a function of particle size, configuration, compression,
storage conditions, etc.).

In Fig. 5.45 [101] the St707 and St101 sorption characteristics for H2 are compared after activation at 500°C and 900°C
for 10 min.

If the activation is performed at 900°C, the St101 getter shows a sorption curve similar (or somewhat better at small
sorbed quantities) to that of the St707 getter, but at 500°C this curve decreases well below the ZrVFe getter curve
which, on the contrary, is not too different compared to the 900°C curve.

Concerning the interaction of the ZrVFe alloy with H2, it is found that Sieverts' law is followed in a relatively wide
range of H2 concentrations as can be shown by the isotherms of the St707H2 system of Fig. 5.46 [82, 83, 101]. Sieverts'
equation in the range 108102 mbar is found to be expressed by

where p is in mbar, q in mbar·liter·g1, and T is in Kelvin. The equilibrium pressure of H2 over St707 compared to
St101, in the same temperature and concentration conditions, is about two orders of magnitude higher. However, it is to
be expected that with the gas loads usually foreseen in many vacuum applications the H2 equilibrium pressures at room
or also somewhat higher temperatures are quite compatible with the use of this ternary getter alloy in UHV applications.

As in the case of the ZrAl getter, different activation conditions have been studied for St707, and the resulting degrees
of activation are shown in Fig. 5.47. It is seen, as
Fig. 5.45
H2 gettering speed, at room temperature, for the St707
and the St101 getters, after activation temperatures of
500°C and 900°C. (From Boffito et al. [101].)

< previous page page_307 next page >


< previous page page_308 next page >
Page 308

Fig. 5.46
H2 equilibrium pressure in the H2-St707 system.
(From Boffito et al. [111].)

already anticipated, that the activation temperatures with practically reasonable activation times are in the range of
350500°C.

Full activation can be obtained, for example, at 500°C in about 10 min or at 400°C in about 100 min. The relatively low
activation temperatures mean that the bulk diffusion characteristics of this alloy are higher than those of St101.

The XPS surface analyses performed on this getter material have shown that the carbon- and oxygen-rich surface
exhibits a metallic character when the temperature is approximately above 400°C, as can be seen in Fig. 5.48 [86, 87].
This is 200300°C less than in the case of the ZrAl getter; this clearly explains the differences in the activation
temperatures found for the two getter types by performing sorption tests.
The relatively low activation temperature makes this getter very appealing in many practical applications; activation can
be performed automatically during the process if temperatures higher than about 350°C are involved during preliminary
pumping with conventional pumps and does not require a specific step. The activated getter can act

< previous page page_308 next page >


< previous page page_309 next page >
Page 309

Fig. 5.47
Activation efficiency of the St707 getter at different
activation temperatures and time combinations.
(From SAES Getters, Catalogue [111].)
Fig. 5.48
Surface composition of the St707 getter as a function
of heating temperatures. (From Ichimaru et al. [86].)

as in situ pump and can help speeding up the process besides ensuring pumping during the lifetime of the vacuum
device. Of course this also means that part of the gas load is picked up by the getter during the process, thus consuming
part of the getter capacity.

< previous page page_309 next page >


< previous page page_310 next page >
Page 310

5.6.6.2
Other Ternary and Multicomponent Alloys

Various ternary alloys with some specific characteristics have been studied and used in certain particular applications.
Among them it may be interesting to mention ZrTiNi alloys [103, 104] which have been particularly used in nuclear
fuel rods to sorb water vapor with a good retention capability for H2. A Zr(45.4 wt.%)Mn(27.3 wt.%)Fe alloy showing
a good gettering capability for various gases but relatively high equilibrium pressures for H2 (scarce retention capability
for H2) has been recently introduced to pump oxygen from water vapor while essentially releasing H2 (useful, for
example, to recover T2 from tritiated water) [105, 106]. Also, various ZrVFe alloys have been developed with different
compositions compared to St707 to obtain some specific features such as relatively high H2 equilibrium pressures to be
exploited, for example, in combination with low H2 equilibrium pressure getters in a sort of chemical H2 compressor
[107].

To further modulate the getter characteristics, alloys with four or more elements have been studied [108, 109], still
having Zr and/or Ti as the basic components. For example, AB2 intermetallic compounds with the following general
formula have been investigated [109, 110]: Ti1aZraV2cbdFebMncCrd, where 0 ≤ a ≤ 1, 0 ≤ b ≤ 2, 0 ≤ c ≤ 2, and 0 ≤ d
≤ 2. They are mainly interesting for H2 sorption and storage. The composition variations allow the modulation of the
plateau pressures, extension of the alpha phase, and modification of the H2 capacity.

5.7
Getter Configurations

To practically exploit a getter material, various configurations are developed to fit the requirements and the constraints
of the application. Therefore, the selection of the getter type also depends on the possibility to shape the getter into the
necessary physical configuration (size, shape, real surface area, total accessible mass available, possibility to be
mounted very close to the gas source, etc.).

Non-evaporable getters can be used just in loose powder or granule form but more commonly are shaped into forms
such as pills, rings, strips, and other, more complex structures. It is essential that the getter material be transformed into
powder, whose particle size, in principle, should be as small as possible. The small size is needed to maximize the
surface area and to have the necessary flexibility to obtain various shapes. It is good practice, however, to reach a
compromise between the target to maximize the surface area and therefore the surface capacity of the getter and the
possibility to handle it safely. Particles that are too fine can generate problems related to the possibility of burning when
handled in air (particularly if used in loose powder getter configuration); moreover, even if they do not burn, the oxides
formed on the surface can represent a high percentage of the getter mass, thus making the activation more difficult and
using up a large portion of the getter capacity. The practical sizes of the grains normally used are in the range of a few
micrometers to a few hundred micrometers.

The most simple and convenient way to use the getter powder is compressing it in pills, in tablets, or in ring-shaped
metal channels. These configurations are simple and safe to handle and can be easily fitted individually in the available
space of a small vacuum device or in a large number if the volume available and the gas load to cope with are large. The
shape and the confined volume occupied by the getter makes it

< previous page page_310 next page >


< previous page page_311 next page >
Page 311

easier to carry out activation, particularly when it has to be performed by supplying heat externally of the vacuum
device (for example, by a resistance heater or by a high-frequency induction coil). Typical sorption curves of St707
compressed pills for various gases are shown in Fig. 5.49 [111]. This figure also shows the effect of increasing sorption
speed and capacity by increasing sorption temperatures. For H2 this effect is seen to be less important because of the
relatively high diffusivity of this gas in the getter material already at room temperature.

Loose particles and relatively low porosity can be drawbacks of these configurations in certain applications. To address
these problems, porous configurations based on sintered powder have been developed. Depending on the type of getter
material involved, sometimes additives are used either to avoid excessive reduction of porosity or to facilitate the
formation of a sintered body in a controlled way [113, 114] and keep a high porosity and surface area. These getters can
be prepared with embedded heaters for activation. Usual shapes available are shown, for example, in Fig. 5.50 [113,
116].

The typical porous structure, indicating high accessibility of the gas, can be seen in Fig. 5.51, which is a micrograph of
a getter portion obtained by an SEM. The getter material used to manufacture these getters can be various and are
typically [113116]: (1) Ti or Zr with the addition of antisintering-nongetter materials to better control sintering. Many
other getter configurations have been developed to meet various requirements for the increasing number of applications
using nonevaporable getters. Getter layers deposited onto various substrates in the shape of strips, foils and cylinders are
currently available and widely used in vacuum technology. A typical getter strip structure consisting of a metal substrate
coated with a compressed getter

Fig. 5.49
Sorption characteristics for various gases by 100 mg St707
pellets with a 50 mm2 surface area at different sorption
temperatures (after activation at 450°C for 10 min).
(From SAES Getters, Catalogue [111].)

< previous page page_311 next page >


< previous page page_312 next page >
Page 312

layer, usually 50 to 100 micrometers thick, is shown in Fig. 5.52. The most common getter material used are the St101
and St707 alloys. The getter strip can be used in short or long pieces activated by an external heat source or by joule
heating. St101 getters of this type are used, for example, in the large electron position collider machine at CERN in
Geneva.

Thin, highly porous (over 40%), getter strips and foils (HPTF getters) have been developed recently using special
deposition and controlled sintering techniques to cope with the geometrical and physical constraints of certain very
small volume

Fig. 5.50
Typical shapes of porous sintered getters with embedded heater.
Fig. 5.51
SEM micrography of a sintered porous getter.

< previous page page_312 next page >


< previous page page_313 next page >
Page 313

Fig. 5.52
Typical structure of a getter coated metal strip.

vacuum devices such as FEDs. Getter strips can also be pleated to form ''concertina" shaped pumps as shown in Fig.
5.53. These can be used in a modulator way to build up large high speed pumping systems for particle accelerators and
other vacuum system applications.

Getter pumps have been developed recently to maximize sorption performance using highly porous sintered getter
elements. These elements contain one or more types of nonevaporable getter material in special configurations to fit the
application involved. For example, they take the form of stacked disks known as the InsiTorr pump used in
semiconductor process chambers, particle accelerators and other applications requiring vacuum systems with high speed
and pumping capacities. Extreme high vacuum in the range of 1013 mbar have been achieved using suitable
nonevaporable getters. It should also be mentioned that appropriate theoretical models based, for example, Monte Carlo
methods have been applied to optimize getter pump structures.

5.8
Getter Applications

Getters are used in a variety of applications ranging from very small sealed-off vacuum devices to very large
experimental vacuum systems. Getters (nonevaporated as well as evaporated getters) are used on a large, industrial
scale, especially in "isolated" vacuum systemsthat is, vacuum systems which are never or very seldom opened to air
after the original vacuum has been created. However, they are becoming more and more important also in the "open"
vacuum systemsthat is, vacuum systems which are more or less frequently exposed to air via valves and ports. The

< previous page page_313 next page >


< previous page page_314 next page >
Page 314

Fig. 5.53
Getter module and dependance of the sorption speed
on the ratio of the spacing d and height h of the getter
walls. Activation at 700°C for 30 minutes and operation at
400°C. (B. Ferrario and L. Rosai, 7th Int. Vac. Congr., 1977.)

recognized specific and sometimes unique features of getters and getter pumps in vacuum technology are: (a) possibility
to work with no need of power for operation (at ambient temperature), after activation or evaporation; (b) possibility of
mounting the getter close to the gas source, thus avoiding the pumping speed limitations due to the connection
conductances; (c) flexibility in structure and size to fit the geometric constraints of the application; (d) possibility of use
as in situ pumps during manufacturing processes; (e) high and practically pressure-independent speed in high and
ultrahigh vacua, particularly for H2; (f) intrinsic cleanliness (no fluids or lubricants are used); (g) no moving parts; (h)
no magnetic components. On the other hand, getters have a defined capacity and therefore they eventually become
exhausted and, where possible, are to be periodically reactivated and eventually replaced. In sealed-off devices a
determination of the gas load during manufacture and lifetime is important to that the right getter type and size can be
selected to ensure operation of the getter during all the foreseen lifetime.

5.8.1
Nonevaporable Getters Versus Evaporable Getters

In general, evaporable getters are used when large areas or a sufficiently large portion of the total internal area of the
device is available for the getter film deposition. High speeds and capacities per unit mass are thus usually obtainable at
economical convenience. There are also possible disadvantages, however. The deposited metallic film may create short
circuits between electric leads and react with internal components. It is possible that successive evaporations are needed
to cope with the gas load, and film flaking may occur. The necessity to evaporate more getter material also

< previous page page_314 next page >


< previous page page_315 next page >
Page 315

requires power which may not be available. The local high temperature required for evaporation can be unacceptable.

Nonevaporable getters are used when the above problems arise with evaporable getters and when other features are
important. Nonevaporable getters do not require large areas, and their speeds can be maximized by suitable geometries.
Nonevaporable getters can also be made in small size and various shapes, thus better fitting the space constraints. They
can be reactivated or regenerated several times and their structures allow them to work over a wide range of
temperatures.

5.8.2
Start-up and Working Conditions of Getters

To make a getter work after its introduction into a vacuum system, activation is usually performed. The vacuum at
which activation is performed is usually at least better than 102 mbar, as obtained by a trapped conventional rotary
pump, by a dry pump, or by other similar rough pumps. However, most commonly, activation or evaporation is
performed at a preliminary vacuum better than 103104 mbar, as obtained by the conventional high-vacuum pumps.

In some cases, it is now possible to use very special (nonevaporable) getters which do not require activation and can be
used at starting pressures even higher than 102 mbar. For example, the Combogetter used in vacuum insulated panels
for refrigerators. Getters are usually used to maintain or improve the vacuum produced by conventional high vacuum
pumps in the pressure range where they loose efficiency (such as UHV) or where their speeds are conductance limited.
They are also used to improve and maintain the required vacuum for the lifetime of isolated vacuum devices such as
CRTs, FEDs, X-ray tubes, metallic dewars, etc. where no other pumping means is really possible or convenient to cope
with outgassing, permeation or microleaks. Also getters are often used to speed up the manufacturing process of these
devices by in situ pumping. Further getter applications are in removing active gas impurities in devices filled with inert
gases such as fluorescent lamps and PDPs and in purifying gases down to one part per billion or one part per trillion as
required in some semiconductor processes. After start-up getters can work at room or even higher temperatures
depending on the application conditions such as the gas loads involved, required speeds, possible presence of
hydrocarbons, etc.

< previous page page_315 next page >


< previous page page_317 next page >
Page 317

Part II
Sputter-Ion Pumps

Hinrich Henning

Sputter-ion pumps (SIPs) are devices which use an electric gas discharge to create an absorption of gas. Cathode
sputtering from ion bombardment distributes getter material to form thin films which act as absorbing surfaces, and
ionized gas particles are trapped by implanting into solid material. This is briefly the principle of the SIP. Thus this type
of high-vacuum pump is a trapping pump which collects the pumped gas within itself, similar to cryopumps.

5.9
Gas Discharge Vacuum Pumps

The decrease of gas pressure in vacuum devices with an electric discharge has been known for a long time. Plücker
[117] reported in 1858 that "Certain gases react . . . with the platinum cathode and the resulting compounds are
deposited on the walls. So we approach . . . the absolute vacuum." He observed that " . . . the metal of the electrodes,
mainly of the cathode, is transmitted to the glass walls of the tube . . . " and in a magnetic field the glowing gas behaves
similar to an electric current and follows " . . . the course of the field lines" [118].

In 1916 Vegard [119] found that absorption is related to the cathode and is not a chemical bonding with the glass as
supposed by Willows [120] in 1903. Vegard observed a change in absorption when using different materials only for the
cathode but not for the anode. The absorption is measured as a pressure difference per electric charge [mbar/Coulomb]
in his discharge tube. In addition, the absorbed amount q depends on the cathode fall. Gas is absorbed only above

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_317 next page >


< previous page page_318 next page >
Page 318

a minimum voltage U > U':

q ~ (U U'), where the offset U' is a constant value.

From his data an internal volume throughput of the discharge may be deduced to a value of about s = 0.001 liter·s1 at P
= 0.4 mbar. But when the pressure continues to decrease finally the discharge extinguishes because of a low ionization
yield. At pressures below 103 mbar the particle density becomes low and the number of collisions between free
electrons and gas particles is too small to produce sufficient ionelectron pairs to compensate for the loss of charged
particles:

The conductivity of the plasma becomes too low and finally the discharge extinguishes. To sustain the discharge a
source for free electrons may be provided. This concept of a discharge device is common in high-vacuum ion gauges for
pressure measurement, and it was also used for vacuum pumps to generate clean vacuum at low pressure [121]. Herb et
al. [122] and Bills [123] show designs of electrostatic ion pumps where electrons orbit in a radial electric field.
Performance of such ion pumps is limited. To enhance the yield, a very intense electron source has to be used. To apply
gettering is only useful when continuous refreshment of the getter layers may be obtained to avoid temporary pressure
changes. Even then the getter is working in a highly selective fashion and shows no pumping action on noble and
chemically inactive gas species such as He, Ne, Ar, and methane.

A much better approach to a useable vacuum pump is to utilize the cold-cathode discharge of the Penning type [124,
125]. Here the electrode configuration leads to a high-density electron cloud by trapping the free electrons. It consists of
a cylindrical anode rather than a ring as in the original Penning discharge tube, but with similar cathode plates opposite
to both ends of the anode tube (Fig. 5.54). This configuration is placed in a homogeneous permanent magnetic field. If a
voltage of, for example, 5 kV is applied, there is a potential drop between the cathode plates and the center of the anode
axis. Free electrons from the discharge cannot move through this potential barrier.

On the other hand, they cannot move freely toward the anode cylinder because of the force from the magnetic field B on
an electric charge e moving with speed v:

m·dv/dt = (e/c)·(v·B),

where m is the electron mass and e is the electron charge. This makes the electrons rotate around a magnetic field line
with the gyromagnetic radius:

rm = (mvc)/(eB).

Thus a cloud of free electrons is trapped inside the anode cylinder of a density high enough to give sufficient ionizing
collisions to maintain the electric discharge even at very low pressures where neutral gas particle density becomes low.
Electron losses at the anode by collisions with gas particles and the generation of new electrons

< previous page page_318 next page >


< previous page page_319 next page >
Page 319

Fig. 5.54
Penning discharge. Electrode configuration in axial cross section.
ra, anode radius.

compensate over a wide pressure range to keep the electron cloud density sufficiently constant. Therefore the ionization
current is proportional to gas pressure according to Eq. (5.14).

This device serves as a source of energetic ions (the ions are not captured but are accelerated by the cathode fall to the
cathode because of the inverse electric charge) which are used to sputter the cathode made from getter material, and
which by this build up active getter deposits elsewhere, and penetrate the first atomic layers of the cathode material,
where they are trapped.

The volume throughput of a single discharge cell is low; Haefer [126] gives data which indicate about 0.0015 liter·s1 at
103 mbar to 105 mbar. In 1954 Gurewitsch and Westendorp reported [127] an "ionic pump" of the Penning type with Ti
cathodes which had a volume throughput S = 0.03 liter·s1.

Then in 1958, Hall [128] presented an "electronic ultrahigh-vacuum pump" with a multicell anode which has a pumping
speed of 10 liter·s1 at 107 mbar. He examined metals such as magnesium, iron, aluminium, molybdenum, and copper
for cathode material, " . . . but titanium is found clearly superior in pumping speed." He varied the upper limit of the
discharge current and found 140 mA as an optimal value. Lower values (e.g., 50 mA) increased the starting time
excessively, while too high a limit (e.g., 500 mA) again increased the starting time because of thermal effects. In a
thoroughly degassed vacuum system after baking the pump itself to 520°C, he obtained an ultimate pressure of 2 × 1010
mbar. This pressure was deduced from the discharge current of 3 × 109 µA.

5.10
The Penning Discharge

Penning described [124] the basic physical processes of his gas discharge configuration: "The electron . . . is prevented
by the magnetic field from hitting the anode and

< previous page page_319 next page >


< previous page page_320 next page >
Page 320

Fig. 5.55
Sputter ion pump, first design with multicell anode. View
with covering wall and one cathode removed and cross section
(partial), showing 36 discharge cells of square-shaped cross
section, the connection tubulation (left), the high-voltage feed-
through (right), and the second flat Ti cathode (below).
(From Hall [128].)

moves in helical trajectories towards the plate P2 (cathode), but is reflected because of the retarding electric field and
oscillates several times between P1 and P2" (Fig. 5.54). Furthermore, "The magnetic field acts as an enhanced gas
pressure making the ignition voltage in the presence of 0.03T and 105 mbar the same as without a magnetic field at 0.06
mbar." He recommends that at high pressure a " . . . discharge with high current should be avoided because of the
cathodic sputtering . . . ". He is concerned with a pressure gauge, not a pump (!), but "One has to take care for the fact
that the discharge absorbs gasin air about 1 liter at 0.025 mbar per Coulomb''; the gauge behaves as a pump! He
assumed the discharge may run at " . . . remarkably lower pressures than 105 mbar."

Investigations of a discharge cell very similar to those used today in vacuum pump design are made by Knauer [129].
The anode cylinder was 16 mm in diameter between cathode plates separated 23 mm. In the early development stages of
SIPs, Jepsen et al. published experimental and theoretical work [130133].

A very detailed investigation, theoretical and experimental, was made by Schuurman [134] in 1966. He included a
schematic survey of the different discharge modes, which have been observed by others [126, 135], and deduces criteria
for their appearance and for transitions from one mode to the other.

In commercial sputterion pumps, only the low magnetic field (LMF) mode, transition mode (TM) and high-pressure
(HP) mode occur. Otherwise, magnetic equipment for individual pumps becomes much too expensive and not very
effective
< previous page page_320 next page >
< previous page page_321 next page >
Page 321

Fig. 5.56
Modes of a cold-cathode Penning discharge as a
function of gas pressure P and magnetic field B. Modes:
N, no discharge; LMF, low magnetic field; HMF, high
magnetic field; TM, transition from low to high
pressure; HP, high pressure. Curved lines: . . .
extinction limit; ignition limit; Townsend limit
(T very small space charge). The crosshatched field is the
operational region for usual sputter ion pumps, and it
is extended to higher B for "build-in" pumps.
(Adapted from Schuurman [134].)

because the discharge current will not increase with higher magnetic field. But the high magnetic field (HMF) mode is
of interest, for example, for pumps which make use of the high field of bending magnets in particle accelerators ("build-
in" pumps) (Fig. 5.56).

The following discussion is about the LMF mode.

5.10.1
Pump Sensitivity

The discharge current ia and the sensitivity ia/p (A/mbar) are important properties of an SIP. The current at a given
pressure will determine the cathodic sputter yield and finally the pumping speed by the getter effect. The term
"sensitivity" is taken from the Penning vacuum gauge.
For his first sputter ionpump, Hall [128] gave a calibration function ia~pn with n somewhat bigger than 1. No attempt to
explain the deviation from linearity was made. From Jepsen [131] a linear characteristic ia = f(p) was measured between
108 mbar and 104 mbar on a multicell SIP with 250 liter·s1 nominal volume throughput. He assumed that this will
persist to much lower pressures. Knauer [129] assumed the classical cross-field mobility µr of electrons and a uniform

< previous page page_321 next page >


< previous page page_322 next page >
Page 322

space charge

ρe = (½π) (E2/V0)

in the discharge cell, and deduced an anode current density

where m is electron mass, e is electron charge, and vc is collision frequency of electron.

The measured anode current in the range of 106 mbar to 104 mbar was in very good agreement, ia = 102 mA to 1.4 mA,
with good linear dependence with pressure (see Fig. 5.59).

Measurements from Rutherford [136] show dependence of ia/p on magnetic field B and cell diameter D and in most
cases also on pressure p. He defined a "cutoff" at low pressure where the sensitivity is 0.5·(ia/p)max. This cutoff
pressure moves to lower values when B becomes stronger or D larger. A higher anode voltage increases the sensitivity
but does not change the cutoff. The change of ia/p at low pressure is visible in Fig. 5.57 below 108 mbar.

Fig. 5.57
Discharge current ia of a SIP as a function of the nitrogen
pressure p. The sensitivity ia/p is constant for p > 109 mbar.
< previous page page_322 next page >
< previous page page_323 next page >
Page 323

Because the cycloidal movement of the electrons has a cycloid height Jepsen [130] obtained

where Va is the anode voltage, L is the cell length, ra is the cell radius, q is the actual electrical charge in the cell, qm is
the maximum electrical charge in the cell, and v′ is the minimum electron velocity for ionization, but here the relation q/
qm is unknown. From numerical evaluation with usual parameter values, he used q/qm = 0.6 to arrive at an observed ia/
p = 10 mA/mbar. Schuurman [134] deduced the following equation with the same assumptions as in Knauer [129] and
Jepsen [130] for the LMF mode with a secondary electron emission yield Γ:

which is proportional to pressure by the ionization frequency vi of the electrons and to the square of magnetic field B.
Different from this the measured values showed linear dependence with B which he explained by having not included
extinction in his theory for ia = f(B). Another estimate of ia/p is given by Jepsen [133], who took the ratio of space
charge q to the product of average time t between collisions of all electrons and the pressure p:

ia/p = q/(tp).

He stated that this value is too high because only a part of the electrons has the optimum energy to make ionizing
collisions.

5.10.2
Ion Motion

At a magnetic field of 0.25 T, Knauer [129] observed a sputtering pattern with a diameter D0 between 3 mm and 5 mm
and a small nearly unsputtered center of diameter Di less than 1 mm. He explained this plateau by assuming that ion
motion is predominantly radial: " . . . ions . . . become deflected by the axial magnetic field . . . " and " . . . miss the
discharge axis. . . ."

From the cyclotron radius rc he deduced for the pattern dimensions

where ra is the anode radius, M is the ion mass, and e/V is the ion energy.

The Di corresponds to the maximum energy 3000 eV, while D0 is determined by the minimum value 25 eV where
sputtering occurs for Ar+ on Mo. Results are in good agreement with measurement by probes behind the cathode with
drilled holes. Values are given for Ti with N2, Ne, Ar, and Kr at B = 0.25 T and with N2 at B = 0.15 T to 0.3 T.
Although evidence of a radial component of ion motion is concluded, no remark

< previous page page_323 next page >


< previous page page_324 next page >
Page 324

is made that grazing incidence at the cathodes could be important to enhance the sputtering yield.

The average energy of ions striking the cathode was calculated [137] to be at least 460 eV for Ua = 2 kV anode voltage
and B = 0.2 T (gas not specified). For noble gases with Ua = 7 keV an ion energy of 3 keV is measured by Helmer and
Jepsen [138].

5.10.3
Electron Cloud

The cloud of trapped electrons is a very important feature of the Penning discharge. It provides a source for producing
new ionelectron pairs by collision ionization with gas molecules.

A description of the avalanche process building up the space charge is given by Jepsen [130] for Penning and
magnetron configurations. From Poisson's equation ∇E = ρ/ε0, and for uniform charge distribution ρ in the anode cell
he found the radial electric field strength Er to be

For the space charge q he obtained

with a maximum qm at V0 = 0. The potential V0 in the center is

A significant depression of the center potential has been verified by experiment from the energy of ions impinging on
the cathode center [138]. As long as the electron density in the cloud remains constant, the ionization rate is a linear
function of gas particle density or pressure.

In the LMF mode the negative space charge fills out the whole cell volume. On transition to the HMF mode it begins to
contract from the cell axis, leaving behind a nearly charge free plasma. On the other hand, when the pressure is rising
the central voltage drop Va V0 becomes smaller, the radial field Er decreases toward zero, and finally on transition to
the HP mode the discharge approaches the normal glow discharge. This is discussed in detail by Schuurman [134].

The motion of electrons is described [128, 129] as a rotation of the (space charge) lectron cloud with a drift velocity

v = ωra = c[(E · B)/B2].

The frequency ω = 70MHz has been measured as microwave resonance in the anode which was split parallel to the axis
in two parts for this investigation [129]. This rotation represents a current, which was measured as induced voltage in a
coil probe when the discharge was switched off and on. The current was found to be ie = 0.46 A and clearly
independent of pressure in the range 106 mbar to 104 mbar. From

< previous page page_324 next page >


< previous page page_325 next page >
Page 325

Fig. 5.58
The radial distribution of the potential V, electric field
strength Er, and electron density Ne in a Penning
discharge cell in hydrogen at a pressure
mbar in the LMF mode. There is no net space charge
in the center where the densities of ions and of
electrons are equal. rc is the radius of electron motion
("cyclotron radius"). (From Knauer and Lutz [135].)

Knauer [129] the rotational current density is


where vΘ is the rotational velocity and vr is the radial drift velocity. From this he calculated ie = 1.0 A, which is about
two times bigger than his measurements (Fig. 5.59). In 1980 Swingler repeated these measurements on a more open cell
geometry with nearly the same results [139].

Validity of classical mobility for particles in the Penning discharge at pressure of 104 mbar and lower was confirmed by
Knauer and Lutz [135], who deduced the

< previous page page_325 next page >


< previous page page_326 next page >
Page 326

Fig. 5.59
Anode current ia [mA] and electron ring current ir [A],
measured as a function of pressure p [mbar]
by Knauer [129].

electrical field strength from Stark-effect measurements on the spectral lines Hβ and Hγ in a hydrogen discharge. They
observed with this method a different behavior at higher pressures, which is the HP mode (see Section 5.10.5 below).

Redhead [140] gave a review of papers dealing with experimental and theoretical work on cold-cathode discharges
where instability of the electron sheath (e.g., diocotron instability; see Reykrudel and Smirnitskaya [141] and Knauer et
al. [142]) is made responsible for lack of linearity [134] between pressure and current at low pressures. Anomalous
electron motion toward the anode is observed which becomes more significant at lower pressures because it is pressure-
independent.

5.10.4
Secondary Electrons

Knauer made some qualitative measurements [129] to explore the contribution of secondary electrons from ion impact
at the cathode to maintain the discharge. This was done with two small hot-cathode probes behind holes in one cathode
in the center and near the edge of the cell. From this he assumed that mainly secondary electrons emitted off center from
the cathodes are trapped in the anode sheath to participate in ionizing collisions and to compensate for loss by the anode
current.

From the equation of the Townsend avalanche for a self-sustaining discharge,


where α is the number of electronion pairs made by an electron when moving the distance dr, and Γ is the number of
electrons captured from the cathode and initiating

< previous page page_326 next page >


< previous page page_327 next page >
Page 327

an avalanche, Jepsen [130] states that

where d is the typical distance of an electron's travel from leaving the cathode to the recapture on it, and λ is the mean
free path for collisions with gas particles.

Another source of secondary electrons is the emission by field effect (see, e.g., Good and Müller [143]). The electron
current depends not on pressure but on the electric field as given by the FowlerNordheim equation:

High field strength will occur when sputtered material deposits on the cathode elsewhere and produces needle-like
whiskers [144]. The ie is usually small at higher pressures but becomes comparable with the discharge current ia if the
pressure decreases and makes the measured effective current constant at low pressures.

5.10.5
Transition from HMF Mode to HP Mode

For the center space around the cell axis, Knauer [129] assumed a plasma "indicated by a weak glow which is visible at
pressures above 105 mbar." Ionization here is from fast radial moving ions, and the number of ionizations per ion is

n+ ~ (v+/v±) (L/2λi) = 1.2 × 102


at 104 mbar,

where v+ is the radial ion velocity, v± is the thermal velocity, L is the discharge length, and λi is the mean free path of
the ions for ionizing collision.

Because the density of ions throughout the cell volume is pressure-dependent, he explained that the transition of the
discharge from the HMF mode to the HP mode occurs at a pressure near 104 mbar, where the ion and electron density
become equal and the space charge is neutralized. Furthermore, at the high-pressure mode, activity ther than classical
behavior controlled by cross-field mobility of the charged particles is assumed. Schuurman [134] confirmed this result.
He deduced an expression for the particle density ntr for transition to the HP mode in two ways: First, he took account
of (not neglecting it as he did in the low-pressure mode!) the dependence of ni in Poisson's equation E/D = const·ne·ni.
At ne = ni he obtained

where D is the sheath thickness, mi is the ion mass, vz is the axial ion velocity, and σi is the ionization cross section.

< previous page page_327 next page >


< previous page page_328 next page >
Page 328

Fig. 5.60
Potential distribution in the Penning
discharge, plotted in arbitrary units against
distance on the cell axis from cathode center
K to center C, then radial to anode cylinder
wall A. From left to right there is the cathode
fall on the axis, then a plasma region around
the cell center with nearly constant potential,
followed by the sheath with linear rise of
potential (i.e., the electron cloud). For higher
B the negative charged sheath contracts, and
the cathode fall becomes lower. If the pressure
increases, the electric field in the sheath and
the space charge decrease. The dashed line
is the distribution before ignition.
(Adapted from Schuurman [134].)

Second, following Haefer [145], he compared the average transit times of ions tri and electrons tre and again at tri = tre:

where is the mean collision frequency per particle.

Comparing Eq. (5.26) with measured values, acceptable agreement is found (see Fig. 5.60). He stated that at the
transition point the negative charged sheath first constricts and then vanishes.

5.10.6
Transition from LMF Mode to HMF Mode
For the transition from the low magnetic field mode (LMF) to high magnetic field mode (HMF), Knauer assumed the
uniform space charge to contract from the center region against the anode, forming a ring-shaped sheath. Its thickness is
diminished with increasing B.

For the transition point between the LMF and the HMF, Schuurman gave

which is independent of pressure. This behavior is confirmed by his experiments, but Hartwig and Kouptsidis preferred
to use in their empirical model [146].

< previous page page_328 next page >


< previous page page_329 next page >
Page 329

The anode current

has a maximum at Btr. This is linearly dependent on pressure, although confirmation by experiment is less convincing.
The measured relation Imax/p is not constant [134].

5.10.7
Sputtering

The refreshment of the getter films is done by sputtering the cathode material. By this the efficiency of pumping
reactive gas is related to the sputter yield. Andrew [147] found from observations on a transparent and fluorescent-
coated cathode that the pattern of activity changes from a spot to a bigger ring with increasing pressure from p = 108
mbar to 104 mbar. This is in contradiction to other observations [148] where the area of sputtering will concentrate from
a large to a very small spot at 105 mbar. In flat cathodes a hole is drilled into the sheet by the discharge at high
pressureabout 2 mm deep after 5000 h at 105 mbar (author's experience). At low pressure, p < 106 mbar, the sputtered
pattern extends to an area similar to the anode cylinder cross section.

At a high magnetic field, B > 0.2 T, the cyclotron radius for ions, is small enough to produce little or no sputtering in the
center region of the cathode [129, 134] (see Section 5.10.2, "Ion Motion"). With usual values for the magnetic field B =
0.1 T to 0.15 T and anode voltage Ua = 3.5 kV to 7 kV, this is not observed. But from Knauer's probe measurements
[129] the radial component of ion velocity is obvious, which gives a striking angle for the impinging ions on the cathode
different from 90°. Thus a higher sputter yield can be expectedOechsner [149]. For Ar+ ions accelerated to 300 eV on
Ti he obtained 76% increase of yield compared to normal incidence. No value for the striking angle is given. Grazing
incident of ions on the cathode is also obtained by the electrode system design of commercial SIPs. In diode-type
pumps, "slotted" cathode plates [132] are used, and in triode pumps the elements of the cathode assembly [150] have
large areas not normal to the cell axis.

The consumption of the cathode electrode material by sputtering determines the life of an SIP. The limit is given by the
loss of mechanical rigidity of the electrode elements or decrease of the sputter yield by deformation from the original
shape. Usually the life of diode-type SIPs with compact flat cathodes is longer than for triodes, which provide a smaller
quantity of getter material because of the transparent design of the cathode. In a large SIP it becomes necessary to
change the electrode assembly, which can be easily inserted into the vacuum envelope. Sometimes it is possible to only
change the cathode elements. Smaller SIPs usually have a "one-way" design and have to be replaced completely.

5.11
SIP Characteristics

5.11.1
Gettering

In the very early stages of working on gas discharges it became clear that gas cleanup is caused by a kind of chemical
reaction between solids and gases [118, 119]. This now is called the getter effect and is treated in Part I of this chapter.

< previous page page_329 next page >


< previous page page_330 next page >
Page 330

The question here is where in a SIP will the gas be absorbed by gettering. The main area where sputtered material
deposits is the inner surface of the anode cylinders. On a virginal anode after a few hours' run a deposit can be observed
as a change in color, which in the midplane is different from that at the ends. After long use the anode surface will be
coated with thick titanium layers. If the pump is sometimes operated at high pressure, a change in temperature will
make the deposit flake off. This results in sharp pressure peaks or even short circuits in the SIP. For this reason a new
anode surface is made rough by sandblasting, etching, or plasma coating (e.g., with Mo) for better adhesion of the
sputtered layers.

Another place for deposits is the surface of the opposite cathode plate. Only in an area where sputtering is low the
deposited material can build up layers and act as an active getter. If, by design, ion-bombarded areas are restricted, the
remaining cathode surface can contribute to active gettering [132, 151, 152]. The use of cathodes made from different
material with different sputter yields have also been used to establish areas where deposition exceeds sputtering [151,
153].

Obviously the effective long-term volume throughput is related to the rate of refreshment of the sputter deposit. If an
inert gas is pumped for some time, the amount of available getter material in the layers is increasing. Then when
changing to a getterable species of gas the volumetric throughput will be clearly enhanced for a transient period [154].
After some time the original level will be reached again, where refreshment by sputtering and consumption by gettering
are in balance. This effect is known as argon treatment or argon shower (see Section 5.11.3, ''Volume Throughput").

5.11.2
Ion Burial

Ion bombardment of the cathode will not only sputter the cathode material but also capture ions in the upper atomic
layers of the solid.

This ion burial is one of the gas sorption effects known from electric discharges and is supposed to be the most
important for pumping of noble gases [128]. Different from gases which are gettered by sputtered Ti layers, the inert
gases can be pumped only by mechanically trapping the gas particles. This takes place mainly at the cathode where he
ions are driven in with kinetic energy gained from the accelerating electric field. This effect shows severe saturation as
the trapping cathode material is sputtered away at the same time. Lafferty and Vanderslice [155] and Andrew et al.
[156] showed, by autoradiography after pumping a radioactive tracer, that the surrounding border of the sputtered area
is the only place where ions are permanently captured.

Most of the captured gas is remitted, and finally the net volume throughput will be very small. Measured pumping speed
in fact is only 12% of that for air or nitrogen. To a certain degree, argon is also captured at the anode surface [157]. It is
supposed to come from instabilities of the discharge which energize the ions and allow them to reach the anode.

Later Jepsen [133] explained this with the theory of "energetic neutrals": Some of the ions striking the cathode are
neutralized by picking up an electron and are then reflected without losing all their kinetic energy. They then move free
of the influence of the electric and magnetic field. The yield of such energetic neutral particles depends on the angle of
incidence of the striking ions and the direction of the reflected particles. Grazing incidence and forward scattering give
higher yield.

< previous page page_330 next page >


< previous page page_331 next page >
Page 331

If the energy of an incident particle is E0 and that of a scattered particle is E1, then after elastic collision they are related
by

where m1 is the mass number of the particle and m2 is the mass number of the target. For m2 > m1, any scattering angle
Θ1 is possible; but when m2 < m1, only forward scattering with Θ1 < 90 is allowed. Furthermore, the scattered particle
can escape only if 90° < Θ1 < 270° (i.e., back scattering) for normal incidence, but for grazing incidence it can escape if
0° < Θ1 < 180° (i.e., forward and back scattering).

Grazing incidence and forward scattering are better realized in triodes than in diodes. So noble gas atoms are captured
as energetic neutrals on the collector surface where sputtering is prevented by the potential barrier.

This theory is able to explain the much better volume throughput of triode-type SIPs and others of similar design [132,
150] for inert gas compared to diode design here grazing incidence on the cathode surface is limited. The results of
Knauer [129] are not consistent with this because he measured a strong radial component of ion movement relative to
the cell axis of a diodetype. Perhaps the stronger magnetic field B > 0.2 T in Knauer [129] is responsible for the
difference. By calculating the scattered particle energy from Eq. (5.29) Jepsen [133] was also able to explain the good
argon volume throughput of the DI-type SIP [153]. From the mass number m2 = 181 for tantalum the energy E1 of
argon is high to provide better sticking probability for the reflected particle.

In the case of hydrogen, Rutherford et al. [157] reported that the majority of the gas following long-term pumping is
found in the cathode plates by carefully weighing them before and after the experiment. They explained this as
absorption of atoms and molecules at the clean titanium surface together with diffusion into the solid rather than ion
implantation.

5.11.3
Volume Throughput

Much effort is made to calculate and measure the volume throughput (VTP) S ("pumping speed") of SIPs because this is
the main property of interest.

Since the theory of the Penning discharge is not yet complete, the existing attempts to evaluate its properties are
semiempirical. They all are based on the close relation to the sensitivity ia/p (see Sections 5.10.7., 5.11.1.). Malev and
Trachtenberg [158] proposed an algorithm because they had to describe "built-in" pumps for which the parameters of
operation are very different from the usual values. They deduced the volume throughput S for a single cell from
experimental values of several authors and finally defined S0 as normalized to a unit of anode cross section:

where η is a filling factor which is the sum of the effective anode cross sections related to the total area covered by the
whole anode cell assembly.

< previous page page_331 next page >


< previous page page_332 next page >
Page 332

The conductance limitation of an electrode system with distance δ between anode and cathode and width 2h of the
anode assembly is given by

All values are for nitrogen (or air).

Faced with the same problem to evaluate distributed SIPs in particle accelerators, Hartwig and Kouptsidis [146]
regarded the volume throughput relation to sensitivity, S/(ia/p), as constant. From experimental work they derived a
mean value

c = S/(ia/p) = 0.075 torr·liter/(A·s) = 0.1 mbar·liter/(A·s) at p < 107 mbar.

Then they assumed that the observed decrease of this parameter at higher pressures is caused by a gas desorption from
ion and electron impact on the electrode surface. This effect compensates the VTP. Assuming a gas coverage following
the Langmuir isotherm equation, they corrected the c value to

ceff = c · c* with c* = (1 1.5 × 106p/(1 + 4 × 106 p))

and

Seff = ceffia/p.

They followed the theory of Schuurmann [134] but postulated a linear dependence of Seff on magnetic field B near the
ignition value Bi, as is obvious in their measurements.

Now using Schuurman's formalism they obtained the following for nitrogen and a single cell with
in the LMF mode:

and in the HMF mode:

Because the discharge extends into the cathodeanode gap for its effective axial length L, they added 25% of this gap δ at
both ends to the length of the anode cylinder! The results show good agreement with experiment in the LMF mode. In
the HMF mode, deviations appear to be mainly related to dependence from cell radius [159, 160]. Therefore further
corrections in Eq. (5.32c) were introduced by Suetsugu and Nakagawa, who made [161] corrections by using a factor J
for geometric distribution
< previous page page_332 next page >
< previous page page_333 next page >
Page 333

of sputtered material on the anode surface:

J = (1 + cos(2Θ))/2
with Θ = arctan(ra/(L + δ));

they also made corrections by using a second, more important factor Qn, which represents the dependence of the total
electric charge from the cell geometry. Having calculated several Qn (with FEMsee Fig. 5.61) they offered as an
analytical approach the expression

Then Eq. (5.31c) reads

S = S(5.32c) · J · Qc.

A very different approach to evaluate the performance of a SIP was given by Suetsugu in 1993 [162]. He used the finite
element method (FEM) to calculate the potential distribution in the cell, with the assumption of a constant electron
density ne throughout the volume. The choice of the maximum ne value is from the condition that the calculated
potential shall not be negative in volume elements near the cathode. Then for estimated ne the corresponding ion current
ii to cathode and electron current ie to anode are calculated (see Fig. 5.62).

The trapped electron density ntr = ne is found balancing ion current ii against electron current ie:

ie = ii,

neglecting secondary electrons from cathodes and ion density in the discharge. Finally from ii and pressure the VTP is
calculated by introducing a sputter yield for
Fig. 5.61
Volume throughput of a commercial SIP with a
nominal pumping speed of 400 liter·s1 for nitrogen.
Calculated, • measured, from Hartwig
and Kouptsidis [146].

< previous page page_333 next page >


< previous page page_334 next page >
Page 334

Fig. 5.62
Volume throughput of a distributed "built-in" SIP showing
dependence on magnetic field strength B with anode voltage
Va as a parameter. Cell diameter, Da, is 15 mm; cell length, L,
is 26 mm; gap cathodeanode, d, is 5 mm; pressure, p, is 108
mbar. Solid line: Suetsugu and Nakagawa [161].
Dotted line: Hartwig and Kouptsidis [146].

ioncathode and a sticking coefficient for gasgetter material. The calculations are compared with those from Hartwig and
Kouptsidis [146] and with measurements that Suetsugu and Nakagawa have made on "built-in" SIPs from the
TRISTAN accumulator ring [163]. Agreement is good with experiment in LMF and HMF modes for dependence on B,
Ua, and ra but not on pressure p > 107 mbar. Nevertheless, this method, after some improvement in the treatment of the
parameters, may promise better (perhaps best) approach to the real performance of the SIPs in the future.

For measurement of VTP of SIPs there are some special problems in addition to the common ones because of the
measurement of very low gas throughput ("flow") and the influence of desorbed gas from the surface of the vacuum
system. First the emission of sputtered particles in neutral, excited, and ionized state from the electric discharge has to
be encountered. Furthermore, there is also some weak electromagnetic radiation from the discharge. As is common in
UHV technology, the design of the pump is open against the pump port to get optimal effectiveness. As a consequence,
these emissions are not screened and prevented from entering the connected vacuum system, because any screening will
reduce the pumping action of the SIP. These emissions may influence the pressure measurement by BA gauges or
operation of other devices.

In a similar manner the stray field from the strong magnets shall be regarded as a source of possible malfunction.

In fact the application of SIPs has given strong impetus to the development of methods for the measurement of volume
throughput [164, 165]. This work is described in Chapter 12. Rutherford et al. [157] reported on the pumping speed and
pumping mechanism of diode-type pumps for several species of gas: H2, He, N2, O2, Ar, and air.

< previous page page_334 next page >


< previous page page_335 next page >
Page 335

An important feature of the SIP is the time-dependence of VTP. As in other trapping vacuum pumps or devices, the
accumulation of gas will influence its performance. Gas release is dependent not only on pressure but also on the history
of the pumpthat is, the species and amount of gas pumped. With a new or regenerated pump or for a new species of gas
the SIP will show a clear decrease in VTP for some time: minutes or even hours, depending on operational pressure.
This is the time needed to bring all relevant processes to an equilibrium state. After this period, called saturation time,
the VTP is generally constant for a much longer timethat is, weeks and months. Hartwig and Kouptsidis [146]
calculated the saturated speed by

with S from Eq. (5.32). Finally, behavior of the SIP becomes unsatisfactory at low pressure. The ultimate pressure
obtained will be high; and at high pressure, strong outgassing occurs when the SIP is warming up by the high-power
dissipation, so that degassing supersedes pumping and the vacuum pressure will run away [166]. It is assumed that the
cathodes partially are heated up to 500°C [148]. For these reasons the SIP now should be degassed by a regenerative
bake-out (see below).

5.11.4
Pumping Mechanism

Getterable Gases. From the discussion about the properties of the discharge and the SIP, one may easily imagine how
pumping of chemically reactive gases proceeds. Gas molecules are ionized by electron impact; then the ions strike the
cathode, they are partially captured in the solid (primary pumping effect), and they partially sputter getter material from
the cathode, which is deposited elsewhere in the electrode system, and there the getter material acts as a getter film with
a large surface, on which other gas molecules are absorbed and bound as stable chemical compounds (secondary
pumping effect). The primary effect is time-dependent because continuous sputtering will excavate earlier captured
particles, thus reducing the net pumping activity. But after operating for some time, all involved processes will be in
balance and show a stable and constant VTP. Its value depends on the ionization energy of the gas molecules, the mean
angle of incidence of the ions striking the cathode assemble, the sputter yield for the gastarget pair, the surface area of
the getter screen, and the sticking coefficient for the gasgetter material. Further parameters are gas pressure, electrode
and cell design, magnetic field strength, anode voltage (which determines the energy of the impinging ions), the
discharge current, and the resulting operational temperature. Most of these values have been taken into consideration
when calculating the VTP as described above.

A SIP with an electrode system of about 350-mm × 90-mm cathode dimension and 55-mm distance in a field of 0.12 T
with an anode voltage of 5 kV has an effective equilibrium VTP of roughly 80 liter · s1 for nitrogen. In the normal
range of operation between 105 mbar and 109 mbar · liter, it does not depend much on the type of pump: diode,
magnetron, or triode. For other gases the speed relative to N2 is given in Table 5.8 [128, 153, 167].

The described balance between several effects resulting in a net VTP can be observed on an SIP operating at low
pressure when the anode voltage is cut off. The pressure will suddenly decrease to about 6080% from equilibrium value
and only

< previous page page_335 next page >


< previous page page_336 next page >
Page 336

Table 5.8. Volume Throughput for Sputter Ion Pumps, Related to the Value for Nitrogen [128,
153, 167]
Gas Diode Triode
Nitrogen
100% 100%
Hydrogen
160200% 140180%
Oxygen
100% 100%
Carbon monoxide
100% 100%
Carbon dioxide
110% 100%
Light hydrocarbons
80110% 80100%
Air
100% 100%
Helium
3% 30%
Argon
1% 22%

a few seconds hereafter begin to rise. The reason is the immediate stop of resputtering captured gas particles from the
cathode and a residual pumping action from the continuously refreshed getter layers. Their capacity will be consumed
after some time, depending on the pressure level. Then the pressure increases. Thus it can be deduced that the
refreshment of the getter layers is more effective than consumption of the deposited getter material, giving the pump a
stable behavior.

Hydrogen. Hydrogen is a reactive getterable gas, but its behavior in a SIP with Ti cathodes is quite different from that of
others. Rutherford et al. [157] reported that the majority of the gas following long-term pumping is found in the cathode
plates by carefully weighing them before and after the experiment. They explain this as absorption of atoms and
molecules at the clean titanium surface together with diffusion into the solid rather than ion implantation. In a very
careful investigation of a single-cell SIP, Singleton [168, 169] showed that when pumping pure hydrogen the obtained
VTP is comparable with that of other gases, but only after long-term operation. After several hours of pumping at 105
mbar, the H2 speed increases up to three times the initial value. After this procedure the VTP is high even at very low
pressures (p < 108 mbar). But a small amount of impurities (e.g., N2 from 0.1% to 10%) is ufficient to reduce it
remarkably again. If more than 10% N2 is present, the VTP will now increase. The reason is that hydrogen absorption
on the Ti surface is limited by surface layers of other absorbed gas species (N2). Because sputtering by the light H2 ions
is low, the dominant process of pumping is diffusion of absorbed particles into the bulk of the bombarded cathodes. As
long as there are surface barriers from other gases as in a new cathode or from stronger absorbed gaseous impurities, the
pumping activity for H2 will be poor.

The state of the Ti surface as a limiting factor for the H2 migration between gas phase and solid is reported by
Schoenfelder and Swisher [170]. Only after sputtering the contaminated layers away, which needs less time at high
pressure, appreciable amounts of H2 will be absorbed to diffuse into the bulk Ti. With more than 10% of N2 as a
heavier gas, the sputter yield of Ti is so high as to provide enhanced refreshment of the getter films everywhere to give
higher VTP. This was verified by Singleton [168, 169], who measured the speed after cutoff of the anode voltage. The
< previous page page_336 next page >
< previous page page_337 next page >
Page 337

observed pumping action decreases very slowly, indicating some available getter capacity.

Noble Gases. For all nonreactive gas species the only effective primary pumping effect is the capture of ions at the
cathode. This was the early experience on SIPs when pumping noble gases: After some time of fairly good pumping
action the VTP decreases to only a small percentage of the initial value and, finally, to 12% of N2 speed.

Very often, when pumping argon, pumps show severe pressure fluctuations called the Argon instability [132, 157, 166,
171]. In the case of argon, when cathode material already containing pumped gas particles is sputtered, this causes a
pressure rise because no other pumping process is available. Then with the increasing pressure the sputtered area on the
cathode surface contracts and reduces the amount of resputtered gas. The pressure now begins to decrease and the pump
enters into a new cycle after some time. The pressure fluctuations produce local temperature changes, and thus
hydrogen sorption and desorption phenomena occur (see Fig. 5.63).

The main advantage of the triode-type SIP is to provide higher VTP for noble gases and stable long-term pumping for
argon. Its pumping mechanism is well understood by the hypothesis of "energetic neutrals" from Jepsen [133] (see
Section 5.11.2, "Ion Burial"). Grazing incidence of ions and forward scattering give higher yield of both nergetic
neutrals and sputtered Ti, which are deposited together on the collector surface.

For helium the behavior of a SIP is governed by the high diffusion coefficient of this light and small atom. This gas is
pumped with a relatively high VTP but is very

Fig. 5.63
Argon instability of a SIP. A record of partial
pressures for mass number 40 argon and 2
hydrogen in arbitrary units.
(From Wutz et al. [172].)
< previous page page_337 next page >
< previous page page_338 next page >
Page 338

sensitive to changes in temperature. As for all noble gases, the trapped gas is in a "forced" physical solution in the bulk
Ti. In fact no spontaneous solubility of noble gas in metal exists. An increase in temperature produced only by an
increase of pressure with higher power consumption of the pump causes remarkable desorption of the pumped gas to
occur as the He atoms diffuse quickly through the bulk. Baking an SIP at 300°C will degas it to a large extent from the
pumped He.

The absence of getter action during noble gas pumping is the reason for the enhanced pumping behavior for getterable
gases after an Argon Shower. As a heavier gas, pure Ar sputters the Ti abundantly from the cathode, but the deposit will
not be consumed by accompanied gettering because no getterable gas is present. When after this a reactive gas is
admitted, the unsaturated getter layers will show enhanced getter activity resulting in a transient high VTP.

5.11.5
Bakeout

A regenerative bakeout is made to refresh the performance of a SIP by degassing it. Trapping of the light gases H2 and
He is reversible to a high degree, but to some xtent, trapping of Ar is also reversible. The regeneration can be made by
using external heaters or by operating the SIP at a pressure of about 1 × 105 mbar to 3 × 105 mbar in a heat-insulating
shroud. It is important to bake the whole SIP to avoid gas adsorption at cooler surfaces and to ensure good degassing of
every part if the pump is to be used later at UHV pressures. Preferably it should be evacuated by an auxiliary pump
during the bakeout up to the time when the pressure has decreased below 105 mbar. If possible, the power supply of the
SIP should be switched on for about at least 10 min before the end of the bakeout to clean most of the electrode surfaces
by particle bombardment.

5.11.6
Types of SIPs

The first SIP design had two electrodes [128] like the Penning gauge. This was called the diode-type pump, and it is
currently in use (see Fig. 5.64). Its properties as a high-and ultrahigh-vacuum pump are excellent. Sensitivity and
volume throughput are high because the discharge cell volume can fill the available magnet gap to a maximum. The
simple cathode design avoids sharp edges and burrs, which can be a source of leakage currents by field emission. To
start a diode pump a starting pressure of p < 103 mbar must be provided.

A restrictive disadvantage is the very low long-term volume throughput for noble gases (mainly for Ar), which leads to
periodical pressure fluctuations (the Argon instability; see "Noble Gases" in Section 5.11.4 above). Some attempts have
been made to optimize the properties of diodes. Tom and James reported on a diode pump which has different material
for the two cathodes (DI-type pump) [153]. Using cathode plates of Ti and Ta, they measured a volume throughput of
2027% for argon compared with air and found no argon instability after pumping for 37 weeks at about 2 × 106 mbar
against a constant leak. They guessed that most of the Ar is buried on the cathodes at the border of the sputtered area
and make the higher sputtering rate of the Ta responsible for the better properties. A similar design is used by Baechler
and Henning [151]. They used different material in small pieces (pills), each for one cell on a stainless steel support
sheet. Cathodes with Ti, Ta, Mo, W, Ag, and Pb were

< previous page page_338 next page >


< previous page page_339 next page >
Page 339

Fig. 5.64
Diode-type discharge cell (axial section). A is the anode
at positive potential; C is a Ti cathode, grounded; B is
the homogeneous magnetic field. electron, moving
on cycloid path and colliding with gas particles;
neutral gas particles, partially gettered on Ti;
ionized gas molecules, accelerated toward the cathode
by electric field (not shown) and partially captured
in the bulk; • Ti particles sputtered by ions
and deposited as getter layer.
investigated. They found an inverse relation between sputtering rate and volume throughput when pumping argon: for
example, best pumping speed and lowest sputtering for cathodes of Ta. The argon instability is observed only for Ti.
They argued that the important property is to avoid reemission of earlier trapped gas particles which is found
preferentially on the lower level of the cathode between the pills. In a later commercial SIP, they used pills from Ti
together with 15% from Ta or stainless steel and obtained 1825% Ar volume throughput compared to N2. Similar
investigations were reported by Vaumoron and De Biasio [166], who found a minimum ratio of 2.5 for the heavy metal
mass number to that of the noble gas necessary for stable pumping. The relative Ar speed of the DI pump as found in
Wutz et al. [172] was confirmed by Komiya and Yagi [173], who measured also a speed of 75% for N2 compared to a
pump with only Ti cathodes. They found that both properties are significantly higher (40% and 82%) when using
cathodes which have a Ti sheet covered by a Ta sheet 1 mm thick with a lot of small punched holes. They assumed that
the cylindrical borders of the holes provide an area for grazing incident of the strong axial (!) moving ions and that the
frequent change on short distance of two different getter materials over the cathode surface is responsible for the
performance

< previous page page_339 next page >


< previous page page_340 next page >
Page 340

data. Okano et al. [174] have measured the memory effect of some gases when pumping H2, HD, and D2 in a DI pump
with Al/Zr cathodes. They found better volume throughput at 107 mbar for N2 and Ar but not for H2, which is pumped
only with 84% of the speed of an identical pump with 100% Ti cathodes. For the memory effect see Section 5.11.8.

SIPs similar to high-vacuum magnetron-type gauges have also been proposed [156]. In the center of the discharge cell a
titanium rod is placed. Slightly different is another proposal [155] with only short posts which are fixed on the cathodes
and protrude only a few millimeters into the discharge cell. Pumps of the magnetron design with a solid cell axis suffer
from the problem of proper alignment between cathode and anode.

Early in 1958, Brubaker presented the triode-type pump as a new design of SIP with greatly enhanced argon pumping
speed [150]. It has three types of electrodes and the '' . . . third electrode is of cellular structure . . . ". He called it an
auxiliary cathode. It was energized at 3 kV (anode voltage +3 kV). At 106 mbar argon pressure he observed stable
pumping over > 200 h with SAr = 0.14 liter·s1 and no instabilities. Today his "auxiliary cathode" is called the cathode;
and his "cathode" is named collector, which is now at the same potential as the anode for simpler design with no loss in
performance (see Fig. 5.65). The collector is usually identical to the vacuum housing of the SIP. The cathode is on
negative potential with respect to the two other electrodes.
Fig. 5.65
Triode-type discharge cell (axial section). A is the
anode, grounded; C is a Ti cathode, grid structured
and at negative potential; D is a collector, grounded
(= vacuum envelope); B is the homogeneous magnetic
field. electron, moving on cycloid path and colliding
with gas particles; neutral gas particles, partially
gettered on Ti; ionized gas molecules, accelerated
toward the cathode by electric field (not shown),
partially reflected as energetic neutral particles, and
captured in the bulk of the collector; • Ti particles
sputtered by ions and deposited as getter layer.

< previous page page_340 next page >


< previous page page_341 next page >
Page 341

Hall [175] described experiments with grid structured cathodes composed of strips of Ti and Zr and measured enhanced
VTP for CO2, CO, and Ar. A different design of a triode, which has a grid structure in radial symmetry to each
discharge cell, was presented by Pierini and Dolcino [152].

The starting pressure of a triode pump may be 102 mbar or lower, depending on the volume to be evacuated.

Today, SIPs are usually of the triode type (Fig. 5.66) with radial or stripe structured cathodes or of the diode-type design
with flat sheets as cathodes. Typical VTPs are shown in Fig. 5.67. Materials used are mainly titanium, together with
tantalum for noble gas pumping. The anode cells are circular tubes. Small SIPs are used as "appendage" pumps with
welded connection to devices permanently under high vacuum such as big thermionic transmitter tubes, klystrons, and
so on. This was in fact the original application of a SIP after its invention.

In particle accelerators and storage rings where suitable magnetic fields already exist, special SIPs are used as "built-in"
or "distributed" pumps. Specially designed electrode systems are arranged inside the vacuum chamber, with the
advantage of being very close to the place where the vacuum is needed, and therefore a high VTP is
Fig. 5.66
Sputter ion pump (diagram partially in cross
section) with four electrode systems and
DN150 connection flange. 1, electrode system
(triode type shown); 2, ferrite magnet system;
3, vacuum envelope; 4, high-voltage feed-through.

< previous page page_341 next page >


< previous page page_342 next page >
Page 342

Fig. 5.67
Volume throughput (VTP) of sputter ion pumps for nitrogen as a
function of pressure. IZ270 is a pump with nominal VTP, S = 270 liter·s1,
diode and triode; IZ500 is a pump with nominal VTP,
S = 500 liter·s1, diode and triode.

Fig. 5.68
"Built-in" SIP for BESSY. Cross section of the double electrode system
in the vacuum envelope. 1, Cathodes (Ti/Ta); 2, anodes (stainless steel
sheets with collinear holes, gaps for better conductance). Cell diameter,
Da is 4 mm; cell length, La is 10 mm; anode voltage, Ua is 4.8 kV;
magnetic field, B is 0.751.5 T. (From Pingel and Schulz [178].)

desired. The operational parameters of such pumps very often are unusual for SIPs, and the planning phase of these
devices is very long. From this, algorithms for the performance of such pumps had to be developed, and in fact the cited
papers for calculation of the VTP [146, 158, 162] are for this kind of application. There are pumps for extremely low
magnetic field with anode cells of 40-mm diameter (only prototype realized [176]) and another with 4-mm diameter,
designed for B = 0.75 T to 1.5 T [177], built for the Bessy electron storage ring [178], (see Fig. 5.68).

5.11.7
Starting Properties
The starting pressure of a diode SIP is lower than that of triodes for two reasons:

1. To simplify the design the cathodes are grounded and the anodes are at high potential (see Fig. 5.64). At pressures
higher than 103 mbar the glow discharge

< previous page page_342 next page >


< previous page page_343 next page >
Page 343

will spread into the vacuum housing well outside the Penning configuration; and because of this sputtering of the
cathodic material is extremely poor and no pumping action will occur. The surface desorbs gas caused by particle
bombardment, and an auxiliary pump is needed to reduce the pressure again.

2. Then if with decreasing pressure the discharge remains in the electrode system, the compact titanium cathode will be
heated by the local power dissipation and desorb previously pumped and gettered gas, mainly hydrogen. This occurs
much more than in triodes, where the cathode design is less compact and where gas trapping occurs mostly on the third
electrode (see Fig. 5.65). Here only the cathodes are at negative potential, and the discharge is forced to develop in their
close surrounding space.

On the other hand, a retarded starting procedure allows the SIP to be heated. The discharge power dissipation is of
advantage for obtaining low pressures later on. The heating will be in just the place where degassing is desiredthat is, in
the electrode assemblies. The heat transfer into the vacuum envelope is slow and is negligible into the magnets outside
the vacuum. The bakeout is effective, and a temperature of 150°C has been observed on the outer surface of the
electrode housing [148]; but consumption of getter material is also high during this procedure, thereby reducing the
lifetime of the cathodes.

5.11.8
Memory Effect

The principle of pumping by trapping gas particles implies that the history of operation is documented inside the pump.
Now when the composition of the gas to be pumped is changed, particles of gas species which are not actually present
in the gas may be desorbed for a certain time. The pump "remembers" the previously pumped gas components. This is
called memory effect. It will be prevalent for noble gases which are concentrated in the cathodesespecially for helium,
which diffuses easily through the bulk. This is also valid for hydrogen, but because it is present everywhere in an SIP or
in a vacuum system at low pressure, the effect will not be detected. For the other chemically active gases the memory
effect is clearly lower because they not only are trapped at the cathodes but are gettered elsewhere.

For the same reason, the memory effect is lower for a triode-type SIP [167, 179] (see Fig. 5.69). The amount of gas
trapped in the cathode material is less than that with diodes, even for noble gases. This may be regarded as proof of the
pumping mechanism discussed for this type of pump.

The memory effect makes it difficult to use an SIP as a pump in He leak testers.

5.11.9
Ultimate Pressure

At the lowest pressure the VTP of a SIP is in equilibrium with all sources of degassing, such as the walls of the vacuum
system and of the gauges, and all vacuum surfaces inside the SIP. The main constituent of the residual gas is hydrogen.
Singleton [168, 169] has demonstrated that well-conditioned cathode surfaces are necessary to have good VTP for this
gas in the ultrahigh-vacuum pressure range. A new electrode

< previous page page_343 next page >


< previous page page_344 next page >
Page 344

Fig. 5.69
Memory effect. The residual mass spectra of a SIP (triode)-pumped vacuum system.
1, after pumping of air, p = 9 × 1010 mbar; 2, after pumping of 400 mbar-liter argon, no
baking, p = 1 × 109 mbar; 3, same as 2, but after a several hours bake at 300°C, p = 7 × 1010
mbar. The argon partial pressure is not severly increased. (From Henning [179].)

system will operate well at low pressure only when the unavoidable contaminating surface layers have been sputtered
away. To reduce degassing, the pump itself with the vacuum system should be baked under vacuum at not less than 250°
C. It is preferable to operate the SIP during the whole procedure, but at least for the last 15 min with the magnets in
place. But pay attention to the high-voltage connector, which is normally the part that is most sensitive to elevated
temperatures!

Malev and Trachtenberg [158] obtained an expression for the lowest pressure pmin for existence of the discharge:

where Ua is in V, D is in mm, and B is in T; this pressure is not the cutoff pressure from Rutherford [136]!

At very low pressure the discharge current appears to be independent of pressure but varies strongly with anode voltage.
This is from the field emission effect as described in Section 5.10.4. The responsible tips on the cathode may be burned
by applying for 1s or 2s a short impulse of high voltage (dangerous!) preferably alternating voltage to avoid arcing.

< previous page page_344 next page >


< previous page page_345 next page >
Page 345

5.11.10
Magnets

The design of the magnet system is an economic problem. Because the SIP is a type of pump for long-time operation, its
power consumption should be minimized. Thus the use of permanent magnets is evident. Care must be taken to provide
good homogeneity of the field inside the electrode systems. Hartwig and Kouptsidis [146] noted the effect of a
misalignment angle φ between cell geometry and magnetic field which consists of a reduction of the effective cell radius
rf from ra defined by the anode dimension, and they obtained the expression

rf = ra cosφ 0.5La·sinφ.

This equation shows the disadvantage from which magnetron pumps suffer. The effective cross section is cut twice
from the outside and from the central rod.
Fig. 5.70
Module design of magnet systems. (ac) With closed
magnetic circuit; (d) With individual magnets for each
electrode unit. (From Andrew [147].)

< previous page page_345 next page >


< previous page page_346 next page >
Page 346

To obtain good performance at low pressures the pump alone or together with the vacuum system has to withstand and
even operate at elevated temperatures of 200°C to 300°C. The magnets should withstand these temperatures without
permanent loss of magnetic energy. Otherwise they have to be demounted before every bakeout; this is not very
convenient with a weight of about 60 kg for a pump having 270 liter·s1 nominal VTP. The weight can be reduced
substantially using rare-earth alloy magnets rather than ferrites. But up to now they have been expensive and have a
much lower temperature limit (180°C max). Furthermore, the yoke necessary to close the magnetic circuit reduces the
gain in weight! That brings us to another point: The field of the magnets may influence other equipment inside or
outside the vacuum. A design with a closed magnetic circuit and low stray field is preferable; otherwise, shielding may
be required [147] (Fig. 5.70).

Care must be taken when demounting the parts of a magnet system. If it has been magnetized in situ after assembly, it
will irreversibly lose some magnetic energy.

< previous page page_346 next page >


< previous page page_347 next page >
Page 347

Part III
Cryopumps

Johan E. de Rijke

Cryopumping is a means of creating a vacuum through the use of low temperatures. It occurs when gas molecules
striking a surface lose enough of their incident kinetic energy to remain absorbed on the surface by so-called dispersion
forces or van der Waals forces. Dispersion forces exist between any pair of molecules; and in the case of cryopumping,
they are the important forces of attraction that hold a pumped molecule on a surface.

The amount of molecules that can be held on a surface is dependent on a number of physical factors: the temperature of
both gas and surface, the chemical nature of gas and surface, the microscopic roughness of the surface, and the incident
flux of molecules. Typically, the dispersion forces existing between a surface and a gas molecule are greater than those
between the gas molecules themselves. We speak of cryosorption pumping when these larger forces are needed to hold
molecules on surface to the extent necessary to reach the desired vacuum level. Only several monolayers of gas can be
accrued on the surface before the effect of the surface becomes negligible and the pressure above the surface will
increase. We speak of cryocondensation pumping when the dispersion forces mutually existing between gas molecules
are sufficient to keep them on the surface to the degree necessary to maintain the desired pressure levels. In this case,
typically a very large number of monolayers can be built up. The result is that much more gas can be accumulated than
in the case of cryosorption pumping.

The majority of cryopumps presently in use on vacuum systems for high- or ultrahigh-vacuum applications have
pumping surfaces cooled by mechanical closed-loop refrigerators utilizing helium as a working fluid. The refrigeration
cycle generally

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_347 next page >


< previous page page_348 next page >
Page 348

used is the GiffordMcMahon cycle. This cycle employs two stages of refrigeration. The first refrigeration stage
normally operates between 50 K and 80 K, whereas the second stage operates between 12 K and 20 K. The temperature
of the second stage is low enough to pump all gases except neon, hydrogen, and helium by cryocondensation. These
three gases are pumped by cryosorption on a sorbent attached to this stage.

Cryopumping can also be used when clean rough pumping is required. In this case, liquid nitrogen is used to cool a
canister of sorbent material. Gas is pumped by cryosorption, and pressures of 103 Pa can be reached.

At sufficiently low temperatures, almost all gas molecules incident on a surface are captured and the speed of a
cryopump will approach theoretical limitsthat is, limits imposed by molecular velocities. Therefore, cryopumps can
have large pumping speeds as compared to other pumping mechanisms. This is especially the case for water vapor
because it is pumped at easily achievable, relatively high temperatures. Also, cryopumping is a clean pumping method;
that is, no fluids internal to the vacuum envelope are used in cryopumps. Therefore, cryopumps will be found in
applications where clean vacuum production and high water-vapor pumping speed is needed.

Because cryopumps are capture pumps, regeneration or the periodic removal of accumulated gases is required. In
principle, this is a simple process, consisting of warming all pumping surfaces to room temperature and allowing gases
to escape through a valve mounted on the pumpbody. Then the pump is evacuated to sufficiently low pressures to create
an insulating vacuum between the pumpbody and the pumping surfaces, after which the refrigerator is turned on to cool
the pump to operating temperatures. Correct regeneration is key to maintaining optimum cryopump performance.
Regeneration is typically performed by automatic controllers; and because the pump cannot be used during
regeneration, considerable effort has been expended in developing efficient and fast regeneration procedures.

5.12
AdsorptionDesorption

The operating principle of cryopumps can be best explained by the theory of adsorptiondesorption. This theory
describes (a) the interactions between gas atoms and/or molecules and a surface and (b) the resulting balance between
adsorption and desorption. For the first monolayer building up on a surface, gas molecules striking the surface are
bound by the dispersion forces existing between gas and surface. These forces are generally larger than the dispersion
forces that exist between the gas molecules themselves. This means that as monolayers of gas are built up on the surface
and the effect of the gassurface interaction diminishes, the equilibrium between adsorption and desorption will change.
The pressure above the adsorbed layer will increase as the magnitude of the dispersion forces decreases. When
approximately five monolayers have been built up, the effect of the dispersion forces from the surface will have become
negligible. Molecules are bound only by the forces mutually interacting between them, and the pressure above a
condensate will no longer increase. It will remain constant so long as the temperature of the outermost adsorbed layer
does not increase. To summarize: As a layer of gas is built up on a surface, the pressure will increase until the effect of
the surface has become negligible. Then the

< previous page page_348 next page >


< previous page page_349 next page >
Page 349

pressure will remain constant as long as the temperature remains constant. The relationship between pressure and
amount of gas accumulated on a surface at a given constant temperature is defined as the adsorption isotherm.

To numerically describe the form that an adsorption isotherm will take, we look at molecules striking a (cold) surface.
In general, a fraction of these molecules will be captured by that surface. This fraction (c) is called the capture
coefficient or the sticking coefficient. It is dependent on a number of physical factors: the temperature of both gas and
surface, the chemical nature of gas and surface, the microscopic roughness of the surface, and the incident flux of
molecules. A quantitative value for the amount of molecules captured by a surface can be derived as shown below [180]:

If the number of molecules (per square centimeter per second) of a gas with pressure p (Pa) approaching a surface is n,
then n can be expressed as

where M is the molecular weight of the gas and T is the temperature in Kelvin.

If a fraction c of the flux of molecules is captured, then nc molecules per square cm per second will be captured. For the
total surface area, ncAp molecules will be pumped. The area Ap is the projected areathat is, the area of the cryopump
seen by the approaching molecules. It is not the microscopic surface area Am, which usually is many times greater.

After a certain time, a total of N molecules, uniformly distributed over the microscopic surface Am, will reside on the
surface. If the average time they remain on the surface is τ seconds, then the mean rate of departure from the surface
will be N/(τAm) molecules per square centimeter and per second.

From the above, the net pumping rate for the cryopump can be stated:

In Eq. (5.35), the first term represents adsorption whereas the second term represents desorption. In order to achieve
effective pumping, the first term needs to be much higher than the second. Initially, as gas starts to accumulate on the
surface, N will be zero and the second term of Eq. (5.35) will be zero. The first term, cnAp, then expresses the
maximum pumping speed. An efficient cryopump will have a high value of c. Sticking coefficients for many gases have
been measured [180], and in many cases [181] they have been shown to have values approaching 1. For the case when c
is one, pumping speeds as shown in Table 5.9 are achieved. These values can also be calculated from the Kinetic
Theory of Gases.

The cryopump becomes inefficient and can be said to have reached its capacity when the total number N of molecules
residing on the surface causes the desorption term in Eq. (5.35) to becomes unacceptable. If the value of c is known, the
desorption term can be determined at any stage of pumping by halting of the inflow of gas,

< previous page page_349 next page >


< previous page page_350 next page >
Page 350

Table 5.9. Maximum Theoretical Pumping Speeds for Cryosurfaces


Atomic Mass Maximum Speed (liter·s1·cm2)
Units with Gas Temperature at:
Gas Species 295 K 77 K
H2
2 44.2 22.6
He
4 31.2 16.0
H2O
18 14.7 7.5
N2, CO
28 11.8 6.0
O2
32 11.0 5.6
A
40 9.9 5.1
CO2
44 9.4 4.8

causing dN/dt to become zero and then allowing the system to reach the equilibrium pressure:

In Eq. (5.36), σ is the concentration of molecules expressed in molecules per square centimeter. We define σm as the
number of molecules that form a monolayer, which has an approximate value of 1015 molecules per square centimeter
for adsorbates [180].

If the second term in Eq. (5.36) becomes too large while σ is less than 5 σm, it means that only pumping by
cryosorption is practical. For cryocondensation pumping, σ can become much larger than 5 σm.

From the above, it can be seen that the general shape of an adsorption isotherm will show increasing pressure as the first
several monolayers are built up. Then as the effect of the surface becomes negligible, the pressure will reach a
maximumnamely, the (saturation) vapor pressureand will no longer rise as the amount of gas accumulating increases.
Figure 5.71 [182] shows typical adsorption isotherms showing the transition from less than monolayer adsorption
through multi-monolayer adsorption to cryocondensation.

Isotherms have been extensively studied, because cryosorption is often required in order to reach the necessary vacuum
levels. There are many adsorption isotherm configurations; as many as 13 different types have been categorized [183],
and at resent no unified theoretical model exists which explains the shapes of various isotherms. Langmuir was one of
the first who attempted to model isotherms in terms of gassurface physics. His work dealt with surface coverages of less
than one monolayer. The equation that he developed for an isotherm takes the form [183]
< previous page page_350 next page >
< previous page page_351 next page >
Page 351

Fig. 5.71
Adsorption isotherms of Xe, Kr, and Ar on a porous silver adsorbent
at a temperature of 77.4 K.

where k is a constant, p is the pressure, σs is the number of sites per square centimeter, and m depicts the number of
surface sites which are occupied.

Another important adsorption model was derived by Brunauer, Emmett, and Teller [184]. They expanded on Langmuir's
theory to include its applicability to gas coverages exceeding one monolayer and derived the following equation:

where k is a constant, σm is the number of molecules in one monolayer, and pv is the vapor pressure. The above method
has proved to be very successful in characterizing certain isotherm configurations and has become an industrial standard
for specifying surface areas of porous materials used in vacuum applications. In recognition of the authors, it is known
as the BET method for determining the areas of sieve materials.

One case of particular interest is cryosorption of hydrogen by charcoal. Hydrogen commonly occurs in vacuum systems
and is generated by many vacuum processes. Charcoal is the material most used in the two-stage, high-vacuum pump
cooled by a mechanical refrigerator. Figure 5.72 shows the adsorption of hydrogen on charcoal for various temperatures
[185].

The pressures represented by the vertical portions of the isotherms shown in Fig. 5.71 represent the (saturation) vapor
pressures of the indicated gases at 77.4 K. The vapor pressure is obviously a crucial value, because it represents the
theoretical ultimate pressure that can be achieved by cryocondensation for a given gas at a given temperature. The vapor
pressure of a gas is derived from the ClausiusClapeyron equation and is usually presented in the form

< previous page page_351 next page >


< previous page page_352 next page >
Page 352

Fig. 5.72
Adsorption of hydrogen on coconut charcoal at low pressures.

Table 5.10. Vapor Pressure of Common Gases as a Function of Temperature in K [187]


Vapor Pressure (Pa)
1011 109 107 105 103 101 10 103 105
Helium
1.0 1.7 4.5
Hydrogen
2.9 3.0 3.5 4.0 4.8 6.1 8.0 12 21
Neon
5.5 6.1 6.9 7.9 9.2 11 14 18 28
Nitrogen
18 20 22 25 29 34 42 54 80
Argon
20 23 25 29 33 39 48 63 90
Carbon monoxide
21 23 25 28 33 38 46 58 84
Oxygen
22 24 27 30 34 40 48 63 93
Krypton
28 31 35 39 46 54 66 86 124
Xenon
39 43 48 54 63 74 92 119 170
Carbon dioxide
60 65 72 81 92 106 125 154 198
Water
113 124 137 153 173 199 233 284 381

where the value for A is proportional to the sublimation enthalpy (∆Hs) and the factor B contains the entropy change associated with the phase
transition. The values of A, B, and C have been summarized by Haefer [186] for various solid gas condensates. Vapor pressure curves have
been developed by Honig et al. [187] in one of the classical vacuum technology papers. Vapor pressures for some common gases are given in
Table 5.10.

5.13
Cryotrapping

Cryotrapping is defined as the concurrent or sequential cryopumping of two or more gases for the purpose of trapping a less readily pumped gas
in the sorbate of a more
< previous page page_352 next page >
< previous page page_353 next page >
Page 353

Fig. 5.73
Hydrogen speed as a function of hydrogen accumulated on an argon sorbate at 12 K.
(Courtesy of Ebara Technologies, Inc.)

readily pumped gas [183]. In other words, a gas such as hydrogen can condense on a surface of a sorbate such as argon,
when hydrogen and argon are simultaneously introduced in a cryopumped system. Or hydrogen can be pumped on
freshly condensed argon sorbate. Figure 5.73 shows the normalized pumping speed for hydrogen as a function of the
amount of hydrogen pumped for a pump used in physical vapor deposition applications where the standard second stage
array using charcoal as a sorbent has been replaced by a similar array without sorbent.

5.14
Pumping Speed and Ultimate Pressure

In general, when calculations regarding gas flow, conductance, or pumping speed are performed, it is assumed that the
all components of the vacuum system have the same temperature. This is clearly not the case when using a cryopumped
system. Usually it is not possible to directly observe pumping performance on a cold surface; instead, it has to be
derived from measurements of pressure or throughput in a second chamber held at a different temperature. Therefore,
we need to examine the flow of gas between two chambers held at different temperatures (Fig. 5.74). One chamber is
held at a temperature Tw, and the second is held at a much lower temperature Tc. The chambers are connected by an
aperture with area A. From the Kinetic Molecular Theory, the flow of particles from one chamber to the other can then
be equated to

< previous page page_353 next page >


< previous page page_354 next page >
Page 354

Fig. 5.74
Thermal transpiration.

Assume first that no pumping is taking place, so there is no net flow of gas from one chamber to the other and the
pressure in both chambers is constant. Under these conditions, nw (the flow from the warm chamber to the cold
chamber) will be equal to nc (the flow from the cold chamber to the warm chamber) and Eq. (5.40) can be reduced to

Equation (5.42) shows that even if there is no flow of gas between two chambers held at different temperatures, the
pressures in the chambers will not be the same. This effect is called thermal transpiration (see Section 1.10).

To determine the effective speed at the entrance of the cryopump or, in the above case, the speed at the aperture
between the warm chamber and the cold chamber, assume that there is a net flow of gas from the warm chamber to the
cold chamber. From Eq. (5.40) the net flow of particles through the aperture, nnet, assuming a sticking coefficient of
unity, is

Equation (5.43) may be simplified by noting that the term preceding the brackets is nw.

< previous page page_354 next page >


< previous page page_355 next page >
Page 355

The maximum particle flow into the pump occurs when no gas flows back from the pump, or nc = 0. Then nw = nmax
and Eq. (5.43) can be written as

If we now define (from Eq. (5.42))

then Eq. (5.44) can be expressed as

where pw(ult) is the pressure at the pump entrance when gas flow is halted. Equation (5.46) related the net flow of
particles to the maximum flow of particles; in other words, it relates the net speed to the maximum speed. For a
cryocondensation pump, pc is the saturated vapor pressure psat. Equation (5.45) shows that pw(ult) will remain constant
as long as psat does not change and so will the net pumping speed at the pump entrance. For a cryosorption pump, pc
can be obtained from the corresponding adsorption isotherm. Pressure pc will rise as the surface coverage increases.
This means that pw(ult) will also rise. As it approaches the operating pressure pw, the net pumping speed will decrease
and become zero when pw(ult) reaches pw. The pump can no longer accumulate gas at that pressure. However, from Eq.
(5.46) it can be seen that the pump will still retain 90% of its maximum speed if the operating pressure is raised by a
factor of 10.

To summarize: For a cryocondensation pump, the ultimate pressure will not change as long as the temperature above the
adsorbate does not change. Also, the speed will be near its maximum value as long as psat is much smaller than pw. In
the case of a cryosorption pump, the ultimate pressure will rise as the equilibrium pressure over the sorbent increases
with the amount of gas adsorbed. From Eq. (5.46) it follows that for cryosorption pumping, speed will decrease when
the equilibrium pressure approaches the working pressure pw.

5.15
Capacity

Cryopumps are capacity pumps, and thus only a finite amount of gas can be stored before the pump has to be
regenerated. The need for regeneration is typically determined by the user and will be performed when pumping
performance for a particular gas has degraded to such a point that it becomes unacceptable in the particular application.
The degradation in two pumping characteristics is used to determine when regeneration is required: (a) the decrease in
pumping speed that occurs as gas accumulates and/or (b) the increase in time needed to reach base pressure.

< previous page page_355 next page >


< previous page page_356 next page >
Page 356

In other words, capacity is defined as (a) the quantity of gas that can be stored on the arrays at a given pressure while
the pump can still maintain a pumping speed igher or equal to a defined percentage (usually 50%) of its initial value at
that pressure or (b) the quantity of gas that can be stored with the pump maintaining the ability to reach a required
pressure in a required time.

There is a third definition of capacity that also has to be consideredthat is, the capacity of the pump when a condensable
gas and an adsorbable gas are being pumped simultaneously. The geometry of the pumping surfaces is designed so that
a majority of gas molecules will strike the first stage array and/or the outside of the second stage before reaching the
sorbent. Condensable gases will be removed, and only adsorbable gases will reach the sorbent. In essence, this shielding
is a compromise between preventing condensable gas from reaching the sorbent (see Fig. 5.75) and maintaining a high
pumping speed for adsorbable gas. However, some condensable gas will reach the sorbent and occupy sites on it,
thereby decreasing the ability of the sorbent to accumulate adsorbable gas.

In this case, capacity is defined as the quantity of condensable gas that can be stored at a given pressure while
maintaining a pumping speed for the adsorbable gas at that pressure that is higher or equal to a defined percentage
(usually 50%) of its initial value after regeneration.

From the above it is obvious that the operating pressure of the pump is a key factor in determining capacity. From the
adsorption isotherm and Eq. (5.46) it can be seen that the equilibrium pressure gradually rises as gas is adsorbed on a
surface.

Fig. 5.75
Cross section of cryopump used for high-
vacuum applications.
< previous page page_356 next page >
< previous page page_357 next page >
Page 357

A cryopump can have reached capacity at, for instance, 106 Pa, while still operating efficiently if only a pressure of 104
Pa is needed.

5.16
Refrigeration Technology

Most modern cryopumps in use today are cooled by a closed-loop mechanical refrigerator using helium as a working
fluid. The cryopump system consists of a compressor and an expander on which the arrays are mounted. Compressor
and expander are connected by flexible hoses. The thermodynamic cycle generally used is based on a cycle developed
by Gifford and McMahon [188] and by Longsworth. This cycle is used because it has proven to be simple and reliable
and has a long service life, and the compressor can be remotely located from the expander and therefore the pump. A
schematic of a one-stage GiffordMcMahon (GM) refrigerator is shown in Fig. 5.76. General-purpose highvacuum
cryopumps have two stages of refrigeration in order to achieve temperatures low enough for effective use.

The schematic shows that the GM machine consists of a cylinder, which contains a cylindrical piston called a displacer.
The displacer is connected to a drive mechanism, so it can be moved up and down in the cylinder. There are two
volumes, one above and one below the displacer. They are varied from maximum size to zero during the cycle, but the
total volume remains constant. The two volumes are connected

Fig. 5.76
Refrigerator schematic.

< previous page page_357 next page >


< previous page page_358 next page >
Page 358

through a thermal regenerator (inside the displacer) and to the inlet and exhaust valve. These valves are coupled to the
drive mechanism so their operation is synchronized to the position of the displacer. At the high-temperature side, the
displacer is also equipped with a gas-tight sliding seal to prevent leakage from one side of the cylinder to the other and
to ensure that helium flows through the regenerator. With two-stage machines the first-stage regenerator consists of
tightly packed, high heat conductivity metal screens, the second stage regenerator is packed with a lead alloy shot.
Essential is that the material of the regenerator has a high heat capacity at cryogenic temperatures. The screens and/or
shot also have high surface area to volume ratios.

This construction means that the regenerator can efficiently transfer thermal energy between the screens and the
incoming or outgoing helium. Also there will be very little pressure difference between the two volumes, which
minimizes the demand on the seals. Because the pressure is essentially the same in the spaces above and below the
displacer, except for a small drop when helium is flowing through the regenerator, no work is done on the gas and the
gas does no work on the displacer. So, essentially no work is required to move the displacer in the cylinder.

The operation of the refrigerator can best be understood by reviewing a cycle with the expander at operating
temperature (see also Fig. 5.76):

1. Pressure Rise. When the displacer is at the top (low temperature) end of the stroke, the exhaust valve is closed and
the inlet valve is opened. This increases the pressure from the exhaust pressure P1 to the inlet pressure P2. Helium will
enter the inlet valve and fill the regenerator and the volume V1.

2. Intake. The inlet valve is kept open while the displacer is moved toward the bottom (high temperature) end of the
stroke. This displaces helium from the volume V1 to V2. The helium is cooled while passing through the regenerator.
This causes its pressure to decrease, and more helium will enter the system from the compressor.

3. Pressure Drop (Expansion). When the displacer has reached the bottom end of the stroke, the inlet valve is closed
and then the exhaust valve opened. The helium will expand and the pressure will drop from P2 to P1. This reduction in
pressure causes a reduction in temperature. The decrease in temperature in V2 is the useful refrigeration of the cycle.

4. Exhaust. The exhaust valve is kept open while the displacer is moved toward the top (low temperature) end of the
stroke. This displaces helium from the volume V2 to V1. The helium is heated while passing through the regenerator. In
turn, the helium cools the regenerator so as to refrigerate the helium passing through on the next pressure rise and intake
stroke.

Two stages are required to achieve low enough temperatures to pump all gases except neon, helium, and hydrogen by
cryocondensation. The added second stage is essentially providing two engines operating at two temperature levels.
This multistaging is desirable because it provides a more efficient process in achieving cryogenic temperatures. For
instance, multistaging relieves the operating regimen of the regenerators [188].

< previous page page_358 next page >


< previous page page_359 next page >
Page 359

We define a refrigeration loss Qr:

where M is the mass flow of the gas, Cp is the heat capacity of the gas, and ∆Tr is the temperature difference between
the gas entering and leaving the regenerator at the low temperature end.

Assume that the temperature of the gas decreases by ∆Te when expanding. ∆Te must be greater than ∆Tr in order to
achieve useful refrigeration. ∆Te is proportional to the absolute temperature T at which the expansion occurs, and
typically it is equal to 0.3 T. In a two-stage (GM) machine, temperatures of 10 K can be reached. This means ∆Te will
be 3 K. To achieve useful refrigeration and handle other losses, ∆Tr would need to be no greater than 1 K. This is
virtually impossible in a single-stage machine operating between 300 K and 10 K because it would require heat
exchanger efficiencies that cannot be reached in practice. However, a ∆Tr of 1 K can be achieved in a two-stage system
where the second stage operates between 50 K and 10 K.

The above briefly describes the principles of operation of the mechanical refrigerator used to cool cryopumps to
operating temperatures. They are known as closed-loop systems because the helium continually circulates between the
compressor and expander. So-called open-loop cryopumps are also used. In these, liquid nitrogen and liquid helium are
used to cool surfaces to 77 K and 4.2 K, respectively. The geometry of the arrays is similar to those described above.
The first stage consists of a bowlshaped structure and a frontal louver, both cooled by liquid nitrogen. The first stage
forms a radiation shield for the second stage, which is cooled by liquid helium. The cryogens are fed into cooling coils
attached to the arrays, and this type of pump is therefore called a boiling pool cryopump. They are called open-loop
cryopumps because the cryogen is allowed to escape into the atmosphere after use. Sometimes gas collection systems
are used in the case of rare gases in order to recycle them.

5.17
Pump Configuration

For mechanically refrigerated cryopumps, array design configurations are, out of necessity, based on the characteristics
of the refrigerator. Two stages are needed to achieve sufficiently low temperatures to pump all gases except neon,
hydrogen, and helium by cryocondensation. Because these light gases cannot be pumped by cryocondensation, a
cryosorbing surface of sufficient size for practical pumping performance must be provided on the arrays. The geometry
of the arrays must be designed so that all gases pumped must impinge on a cold surface before reaching the sorbent.
This will remove those gases being pumped by cryocondenstion before they reach the sorbent, whose full pumping
capacity will be used for removing neon, hydrogen, and helium.

The efficiency of the refrigeration cycle decreases rapidly as temperatures approach absolute zero. Therefore, it is
important that the available refrigeration power for the older second stage array be used for condensing gas and is not
dissipated through parasitic loads. A key design feature for the first-stage array, which operates at a much higher
temperature and which has much more refrigeration power available, is that its geometry is such that it shields the
second stage from radiation heat loads.

< previous page page_359 next page >


< previous page page_360 next page >
Page 360

The above considerations lead to an array design as shown schematically in Fig. 5.76. The first-stage array has a can- or
bowl-shaped configuration. Attached to the can is the frontal array which for pumps used in high vacuum applications
consists of a louver. The design has to be a compromise of providing efficient radiation shielding for the second stage,
while providing as high a conductance as possible for gases that will be pumped on that stage. The second-stage usually
consists of a set of cones as shown in Fig. 5.76. A sorbent, usually charcoal, is attached to the underside of the cones.
The second-stage design is also a compromise between ensuring efficient shielding for the sorbent and still providing
high conductance for neon, hydrogen, and helium, along with high pumping speeds.

Indium gaskets are used in connections between the frontal array and the first-stage can, as well as in connections
between the arrays and the refrigerator, so that high thermal conductance is provided over the joints. Charcoal is the
most commonly used sorbent because water vapor can be removed from it at room temperature. It has a greater capacity
than man-made molecular sieve material and is less affected by impurities. Molecular sieve has to be heated to 250°C in
order to remove water vapor. This temperature would melt the indium gaskets and would damage internal refrigerator
components, such as the seals on the displacer.

The temperature of the arrays is determined by the total heat load imposed on the pumping surfaces. The loads imposed
are from radiation, convection, and condensation. As discussed below, loads from convection and condensation can be
ignored when the pump is operating at a pressure below 101 Pa. The main load under these conditions are radiation
from the pumpbody and from the vacuum chamber. Additional radiation loads can be imposed by sources inside the
vacuum chamber, such as bake-out heaters, plasmas, and so on.

An estimate of the radiant heat flow on to the arrays can be derived from calculating the heat flow (Q) between two
concentric cylinders at different temperatures [189]:

where the subscripts c and w refer to the inner, colder cylinder and the outer, warmer cylinder, respectively, and where

σ = StefanBoltzmann constant,
e = coefficients of emissivity,
A = surface areas,
T = temperatures in Kelvin.

If both emissivity coefficients ew and ec are equal to their maximum value 1, this reduces to

< previous page page_360 next page >


< previous page page_361 next page >
Page 361

For a cryopump with an inlet diameter of 200 mm, the area of the first stage is approximately 0.130.15 m2. If the
pumpbody and vacuum chamber are at 295 K and the average first-stage temperature is 60 K, this means that the
maximum heat load the first-stage would have to absorb would be approximately 60 W. This heat load is usually much
larger than the first-stage refrigeration power of this size pump. It is clear that the radiant heat load must be reduced.
This is done by reducing the emissivity of the internal pumpbody surface and the outside of the first stage can to levels
below 0.1 through electropolishing or nickel-plating. This treatment will reduce the heat load to a value of 12 W.

Many gases, specifically water vapor, will have a high emissivity (0.80.9) when condensed. This means that the
emissivity of the first stage will rapidly increase in applications where water vapor is present, which means most
applications! In order to minimize this effect, the diameter of the first stage is designed to be only slightly maller than
that of the pumpbody, so that only a minimal amount of water vapor will condense on the outside of the first-stage can
and its emissivity will not be affected significantly while pumping for long periods.

Radiation from the first-stage to the second-stage does not play a significant role. For a 200-mm-diameter pump, the
surface area of the second stage is approximately 0.07 m2. Assuming an average first-stage temperature of 60 K and a
second-stage temperature of 12 K, from Eq. (5.53) it can be calculated that the maximum heat load imposed is less than
1 W. In many cryopumps the inside of the first stage is purposely made to have high emissivity. This will adsorb stray
radiation entering through the frontal array, minimizing the heat load on the second stage.

As mentioned above, convection heat loads can be ignored when the pump is operating at a pressure below 101 Pa. The
pump then will be operating in the molecular flow regime, and the mean free path of the molecules will be significantly
greater than the distance between the second stage array and the pump body. A high degree of insulation between the
pumpbody and the arrays is provided under these conditions. Convection loads become significant at pressures above
101 Pa, certainly if a major constituent of the gas pumped is hydrogen.

Also, the heat load imposed by condensation is negligible at high-vacuum operation. For instance, the enthalpy for
nitrogen is 15,580 J/g·mol at room temperature and 134 J/g·mol at 20 K. A total amount of 15,446 J must be removed
by the refrigerator when condensing 1 mol of nitrogen. Some of this heat will be removed when the nitrogen strikes the
first stage. Ignoring this effect and assuming a worst case where all heat is removed by the second stage, calculations
show that a second stage in a pump operating at 104 Pa with a nitrogen speed of 1500 liter·s1 will adsorb a heat load of
approximately 1 mW. The obverse is that the heat of condensation becomes significant when the pump operates at
pressures above 101 Pa (for example, when it is used in physical vapor deposition applications).

Pumps used in physical vapor deposition applications have a different first-stage can and frontal array design than those
used in high-vacuum applications (see Fig. 5.77). In the first place, the pumpbody and first-stage can are longer than
those of a high-vacuum pump, increasing the distance between first and second stage. This allows more gas, typically
argon, to be accumulated before capacity is reached. In the second place, a barrier made of an insulating material is
placed between the pumpbody and first-stage can at the pump entrance. This minimizes the amount of gas entering the
space between the first stage and the pumpbody. In addition,

< previous page page_361 next page >


< previous page page_362 next page >
Page 362

Fig. 5.77
Schematic of cryopump used in physical vapor deposition.

openings are made in the can in the area where it attaches to the refrigerator, so that any gas entering the space will be
pumped away. This geometry results in a pressure differential over the barrier. This differential is such that when the
pressure above the pump is on the order of 1 Pa, the pressure between body and can below the barrier averages a
pressure below 101 Pa, effectively removing the convection heat load.

Finally, the conductance to the second stage is deliberately reduced by changing the geometry of the frontal array.
Typically, the frontal array consists of a disk. The disk has a series of openings, sized so that the total conductance to
the second stage is on the order of several hundred liters per second. The conductance is calculated to result in the
proper process gas flow at the required operating pressure. The reduction in gas flow through the use of a small
conductance results in a smaller heat load imposed on the second stage and in longer times to reach capacity. Full water-
vapor pumping speed is maintained as the frontal array operates at temperatures below 100 K. So pumping performance
during pumpdown is not significantly affected.

One of the key parameters for a cryopump is the heat load that the pump can adsorb at the moment when the valve
isolating the pump from the chamber is opened during pumpdown. The load imposed is called the impulsive heatload. If
the amount of gas impinging on the pump is too large, the arrays will warm up to the point where previously
accumulated gas will evolve from the arrays to the extent that a runaway condition occurs. The pressure will become so
high that the resulting convection load will overwhelm the refrigerator. Hydrogen is the most critical gas in this respect
in that it evolves at lower temperatures than other gases and causes a higher convection load at a given pressure. It is
recommended that the impulsive load imposed on the pump be limited to the amount that will allow the second stage to
remain at or below

< previous page page_362 next page >


< previous page page_363 next page >
Page 363

20 K. At this temperature the hydrogen capacity is approximately 50% of the capacity existing when the sorbent is at 12
K, and the amount of hydrogen desorbing will be limited. If no hydrogen has been accumulated, the pump will still
recover when the second-stage array temperature is raised to approximately 30 K.

5.18
Regeneration

Cryopump operation entails regenerating the pump periodically. Proper regeneration is key to optimum performance of
the pump. If not done correctly, pump performance can be significantly degraded.

When the pump is operated at high vacuum, hydrogen usually will be the first gas for which capacity is reached. When
this happens, pumpdown times will become longer and the ultimate pressure will become higher. When the pump is
used in sputtering applications, argon capacity will be reached first. The pump will apparently behave normally when
argon is flowing; but when argon flow is halted, large pressure bursts will be seen during pumpdown. This is due to the
fact that the amount of argon accumulated on the second stage has become so large that the sorbate surface is no longer
adequately shielded from radiation by the first stage. Argon will sublimate irratically as it is exposed to room
temperature radiation.

The need for regeneration is best determined experimentally. First, measure the time that the pump can operate in a
given application before behavior as described above appears. Regeneration should then be performed at approximately
60% of this time limit to ensure optimum performance.

Essentially, regeneration is a simple process. The pump is warmed to room temperature. During warm-up, gas will
escape through the safety pressure relief valve attached to the pumpbody. The pump is then roughed to a pressure
between 5 and 10 Pa in order to establish an insulating vacuum inside the pump. Then the refrigerator is turned on and
the pump cooled to operating temperature.

There are several important considerations to be observed. Large amounts of sorbate might be accumulated on the
second stage when regeneration is started. The second stage is shielded from convection and radiation loads. It can take
several hours before it warms up to a temperature at which significant amounts of gas evolve and the insulating vacuum
is broken. Therefore, one of the first steps usually taken when regenerating the pump is to raise the internal pressure to
atmospheric by purging it with a dry inert gas, typically nitrogen. The purge gas is usually heated to shorten the time
needed for the arrays to reach room temperature.

The second, most important consideration is to ensure that accumulated water vapor will be adequately removed during
regeneration. Purging with dry nitrogen during pump warm-up will assist in sublimating water vapor from the first
stage. Once a temperature of 273 K has been reached, the residual water vapor will liquefy. This means it can be easily
transferred to the second-stage sorbent. The capacity of the sorbent for pumping hydrogen will be decreased if all water
vapor is not removed before subsequent pump chilldown. It is therefore necessary to determine that this residual water
vapor has been removed before operating the pump. This is done empirically by purging the pump for an extended time
with dry gas and then performing a test to determine the amount of water evolving from the sorbent by measuring the
increase in pressure with time at the end of roughing. Because it is

< previous page page_363 next page >


< previous page page_364 next page >
Page 364

uncertain how much water has been accumulated since the previous regeneration, this method can lead to errors. A
more accurate method is to attach a hygrometer to the pressure relief valve and measure the dewpoint of the escaping
gas during the purge. Dewpoint levels that do not affect hydrogen pumping at subsequent operation can be measured
during each regeneration, and more repeatable performance can be attained.

5.19
Partial Regeneration

Regenerating the pump to remove all gas as outlined above is called full regeneration. The pump has to be warmed to
room temperature, roughed, and chilled to operating temperatures. Another regeneration method, called partial
regeneration, has been developed. In this process the arrays are only warmed to temperatures between 120 K and 180
K, so that only gases accumulated in the second stage are removed. This process can be used in applications where
either hydrogen or argon capacity will be reached long before enough water vapor has accumulated to affect
performance. The process can be accomplished in approximately 45 min, instead of several hours as with full
regeneration (see Fig. 5.78).

There are potential difficulties with partial regeneration. To complete the process in 4560 min, the accumulated gas has
to be removed in less than 15 min. In many applications, especially physical vapor deposition, large amounts of gas can
easily be accumulated in the second stage. The rate of gas removal needs to be high in order to remove it rapidly. This
means that the pressure in the pump will be high, essentially atmospheric pressure or higher. Conditions of viscous flow
will exist for several minutes. Large amounts of condensable gas (argon, nitrogen, oxygen, etc.) will be able to reach the
sorbent and will be partially condensed and/or sorbed. The amount of gas accumulating on the sorbent will depend on
the type of gas, its pressure, and the temperature of the sorbent.

For optimum results, after array warm-up and gas removal at atmospheric pressures through a one-way valve, the pump
needs to be roughed to a low pressure while the sorbent is at a relatively high temperature (> 120 K). This will remove
any amount of condensable gas (nitrogen, argon) remaining on the sorbent before the array is subsequently cooled
down. Also, first- and/or second-stage array temperature should not exceed 180 K during the evaporation and roughing
process. This will exclude the possibility of water vapor sublimating off the first stage and reaching the sorbent.

The efficacy of partial regeneration on a pump can be checked by measuring the hydrogen pumping speed as a function
of the amount of hydrogen accumulated after a partial regeneration has been performed. That data can then be compared
to hydrogen pumping performance after a full regeneration, because full regeneration is the standard method by which
cryopumps are restored to their original performance.

5.20
Sorption Roughing Pumps

Sorption roughing pumps or sorption pumps are used for pumping systems from atmospheric pressure to a pressure of
approximately 101 Pa. They rely on the

< previous page page_364 next page >


< previous page page_365 next page >
Page 365

Fig. 5.78
(a) Full-regeneration timetemperature cycle (200 mm diameter pump).
(Courtesy of Ebara Technologies, Inc.) (b) Partial-regeneration timetemperature cycle
(200 mm diameter pump). (Courtesy of Ebara Technologies, Inc.)

dispersion forces existing between a gas and a surface to bind gas molecules on chilled surfaces inside the pump. In
other words, they pump by cryosorption.

Sorption pumps typically consist of a cylindrical canister that is filled with an absorbent (see Fig. 5.79). The adsorbent
is usually molecular sieve material, or zeolite, which consists of pellets made of a calcium or a sodium aluminosilicate
crystalline matrix [183]. The canister is placed in a dewar cooled by liquid nitrogen. Zeolite is a poor heat conductor, so
an array of aluminum fins inside the pump is used to improve thermal contact with the sieve material.
Sorption pumps need liquid nitrogen to operate; and, as with any capture pump, they have to be periodically
regenerated. Therefore in present-day high-throughput applications, they have been replaced by dry mechanical
roughing pumps. However,

< previous page page_365 next page >


< previous page page_366 next page >
Page 366

Fig. 5.79
Cross section of sorption roughing pump
(Courtesy of Varian Vacuum Products).

sorption pumps are very clean noncontaminating roughing pumps and are mostly used in low-throughput applications
where this feature is of prime concern. They are used in conjunction with getter pumps, ion pumps, or mechanical
cryopumps.

In a sorption pump, molecules are held on the adsorbent surface by physisorption. The number of molecules that can be
held on an adsorbent is dependent on the temperature of both gas and surface, the chemical nature of gas and surface,
the microscopic roughness of the surface, and the incident flux of molecules. There is a constant interchange between
molecules residing on the surface and molecules arriving from the gas phase. The key is to have equilibrium conditions
such that practical amounts of gas can be captured at the desired pressures.

For nitrogen, the major gas load when air is pumped, the dwell time τ of a molecule on a surface at room temperature is
approximately 5 × 1011 s [183]. At a pressure of 102 Pa, about 2 × 106 molecules per square centimeter can be
adsorbed. Considering that a monolayer of gas on a surface consists of 10141015 molecules per square centimeter, it
follows that a negligible amount of nitrogen will be pumped.

At liquid nitrogen temperatures, the dwell time will have increased to 8 × 103 s and the amount of nitrogen residing on
the surface at a pressure of 102 Pa is approximately 3 × 1014 molecules per square centimeter or half a monolayer. By
providing large surface areas, practical amounts of nitrogen can be pumped (see Fig. 5.80).

As coverage increases to above half a monolayer, the effect of the surface is rapidly lost and the equilibrium pressure
will quickly rise to the saturation vapor pressure, which for nitrogen by definition is atmospheric pressure.

A cross section of a sorption pump is shown in Fig. 5.79. Key elements of the pump are the aluminum body, the array of
fins removing heat from the zeolite, and the pressure relief valve. The adsorbent used is usually Linde 5A molecular
sieve. This material, with an internal pore diameter of 11 nm [189], has a high affinity for nitrogen and oxygen.

Zeolite also has a very high affinity for water vapor. Water vapor accumulated when repeatedly pumping down a
chamber filled with ambient air will eventually

< previous page page_366 next page >


< previous page page_367 next page >
Page 367

Fig. 5.80
Adsorption isotherms for nitrogen, hydrogen, neon, and
helium for a liquid-nitrogen-cooled sorption roughing
pump with a 1.35-kg zeolite charge.

saturate the sieve material, eliminating its capacity for adsorbing nitrogen and oxygen. The pump must then be baked
out to 250°C or higher to remove the water. The sorption pump therefore usually comes equipped with a bake-out
heater. Normally, during operation of the pump, the heater is also immersed in liquid nitrogen.

Figure 5.80 shows the adsorption isotherms for nitrogen, hydrogen, neon, and helium for a pump as shown in Fig. 5.79.
This pump has a charge of 1.35 kg of molecular sieve, which can pump approximately 107 Pa·liters of nitrogen at a
pressure of 101 Pa. Figure 5.80 shows that noble gases such as neon and helium are pumped poorly. If, for instance,
neon is pumped together with air, its capacity will be less than that shown in Fig. 5.80 because the neon will be replaced
by the active air gases, starting at pressures below 103 Pa. For this reason, sorption pumps are quite often staged. When
two pumps are staged, one pump is used to achieve a pressure of 103 Pa and is then valved off. The second pump is
then valved in and the pressure is further reduced. By this method, 99% of the air is removed by the first pump, and
noble gases are also swept into this pump. They cannot backstream into the system when pressure is further reduced.
Figure 5.81 shows a pumpdown curve for a 200-liter chamber being pumped by three-staged sorption pumps.

It is not useful to characterize sorption pumps by their pumping speed due to their batch nature [190]. However, it can
be shown (Fig. 5.81) that by using recommended sequencing and sizing, speeds approaching 300 liters per minute can
be obtained.

< previous page page_367 next page >


< previous page page_368 next page >
Page 368

Fig. 5.81
Pumpdown for staged sorption pumps.

A key safety element of the sorption pump is the pressure relief valve. When the pump is saturated with air and allowed
to warm up to room temperature, very high pressures can be built up. The operation of this valve should never be
obstructed.
References

1. W. A. Steel, The Interaction of Gases with Solid Surfaces. Pergamon, Oxford, 1974.

2. R. J. Madix, ed., Surface Reactions. Springer-Verlag, Berlin and New York, 1994.

< previous page page_368 next page >


< previous page page_369 next page >
Page 369

3. C. Kittel, Introduction to Solid State Physics. Wiley, New York, 1967.

4. B. Chalmers and R. King, eds., Progress in Metal Physics. Vols. 1 and 2. Pergamon, Oxford, 1958.

5. C. O. Smith, The Science of Engineering Materials. Prentice-Hall, Englewood Cliffs, NJ, 1969.

6. J. D. Fast, Interaction of Metals and Gases, Vol. 1. Macmillan, New York, 1965.

7. J. D. Verhoeven, Fundamentals of Physical Metallurgy, Wiley, New York, 1975.

8. J. H. DeBoer, The Dynamical Character of Adsorption. Oxford University Press, Oxford, 1953.

9. D. M. Young and A. D. Crowell, Physical Adsorption of Gases. Butterworth, London, 1962.

10. D. O. Hayward and B. H. W. Trapnel, Chemisorption. Butterworth, London, 1964.

11. V. Ponec and Z. Knor, Adsorption on Solids. Butterworth, London, 1974.

12. Z. Knor, Chem. Listy 59, 277 (1965).

13. B. M. W. Trapnel, Proc. R. Soc. London, Ser. A 218, 566 (1953).

14. E. Myazaki, J. Catal. 65, 84 (1980).

15. I. J. Langmuir, J. Am. Chem. Soc. 40, 1361 (1918).

16. H. Freundlich, Colloid and Capillary Chemistry. London, 1926.

17. S. Brunauer, K. S. Love, and R. G. Keenan, J. Am. Chem. Soc. 64. 751 (1942).

18. J. R. Anderson, Structure of Metallic Catalists. Academic Press, New York, 1975.

19. G. C. Bond, Catalysis by Metals. Academic Press, New York, 1962.

20. J. Crank, The Mathematics of Diffusion. Oxford University Press, Oxford, 1957.

21. W. Jost, Diffusion in Solid, Liquids, Gases. Academic Press, New York, 1960.

22. J. D. Fast, Interaction of Metals and Gases, Vol. 2. Macmillan, New York, 1971.

23. A. Sieverts, Z. Metallkd. 21, 37 (1929).

24. Handbook of Chemistry and Physics, 67th ed. CRC Press, Boca Raton, FL, 1987.

25. R. E. Honig and H. O. Hook, RCA Rev. 21, 360, 567 (1960).

26. P. della Porta et al., J. Vac. Sci. Technol. 6(1), 40 (1969).

27. P. della Porta, Vac. Conf., 13th, Int. Conf. Solid Surf., 9th, Yokohama (1995).

28. J. H. N. van Vucht, Proc. Int. Vac. Congr., Como, p. 170 (1959).

29. E. A. Lederer and D. H. Wamsley, RCA Rev. 2, 117 (1937).


30. J. C. Turnbull, J. Vac. Sci. Technol. 4(1), 636 (1977).

31. P. della Porta, Vacuum 6, 51 (1959).

32. P. della Porta, Le Vide 101, 484 (1962).

33. P. della Porta and S. Origlio, Le Vide 91, 3 (1961).

34. P. della Porta and G. Sormani, Nuovo Cimento, Suppl. 2(1) (1963).

35. H. J. R. Perdijk, Proc. Int. Symp. Residual Gases Electron Tubes Relat. High Vac. Syst., Roma (1967); Nuovo
Cimento, Suppl. 1(1), 73 (1967).

36. J. Verhoeven and H. van Doveren, J. Vac. Sci. Technol. 20(1), 417 (1982).

37. C. Pisani, S. J. Hellier, and T. A. Giorgi, J. Vac. Sci. Technol. 7(1) (1970).

38. P. della Porta and L. Michon, Vacuum 15(11), 535 (1965).

39. P. della Porta, Trans. Natl. Vac. Symp. 6, 317 (1959).

40. J. Verhoeven, Vacuum 30(2), 69 (1980).

41. F. Ricca and P. della Porta, Vacuum 10, 215 (1960).

< previous page page_369 next page >


< previous page page_370 next page >
Page 370

42. P. della Porta and F. Ricca, Trans. Natl. Vac. Symp. 5, 25 (1959).

43. P. della Porta and S. Origlio, Vacuum 10, 227 (1960).

44. B. Fransen and H. J. R. Perdijk, Proc. Int. Symp. Residual Gases Electron Tubes Relat. High Vac. Syst., Como, p.
199 (1959).

45. H. J. R. Perdijk, Proc. Int. Symp. Residual Gases Electron Tubes Relat. High Vac. Syst., Como, p. 204 (1959).

46. K. Hoshimoto and K. Kitagawa, Proc. Int. Symp. Residual Gases Electron Tubes Relat. High Vac. Syst., Como, p.
156 (1959).

47. T. A. Giorgi and P. de Biasio, in Residual Gas in Electron Tubes (T. A. Giorgi and P. della Porta, eds.), p. 69.
Academic Press, New York, 1972.

48. P. della Porta and S. Origlio, Vide 90, 446 (1960).

49. P. della Porta, Nuovo Cimento, Suppl. 9(1), 26 (1967).

50. P. della Porta and T. A. Giorgi, Trans. Natl. Vac. Symp. 10, 491 (1963).

51. H. J. R. Perdijk, Vaccum 20(8), 321 (1979).

52. P. della Porta, Nuovo Cimento 1(5), 26 (1967).

53. W. Kroontje, et al., J. Vac. Sci. Technol. A 4(5), 2293 (1986).

54. T. A. Giorgi, Proc. Int. Vac. Congr., 6th, Kyoto, Japan (1974); Jpn. J. Appl. Phys., Suppl. 2(1), 53 (1974).

55. P. della Porta, in Residual Gases in Electron Tubes (T. A. Giorgi and P. della Porta, eds.), p. 3. Academic Press,
New York, 1972.

56. U.K. Pat. 1,372,823 (1974).

57. U.S. Pat. 3,428,168 (1969).

58. P. della Porta, in Residual Gases in Electron Tubes (T. A. Giorgi and P. della Porta, eds.), p. 26. Academic Press,
New York, 1972.

59. U.S. Pat. 5,118,988 (1992).

60. U.S. Pat. 4,077,899 (1978).

61. U.S. Pat. 4,717,500 (1988).

62. E. Fromm and H. Uchida, J. Less-Common Met. 131, 1 (1987).

63. K. Welch, Capture Pumping Technology. Pergamon, Oxford, 1991.

64. A. K. Gupta and J. H. Leck, Vacuum 25(8), 362 (1975).

65. N. J. Harra, J. Vac. Sci. Technol. 13(1), 471 (1976).

66. G. J. Grigorov, Vacuum 34(5), 513 (1984).


67. T. Sugita and S. Ebisawa, Proc. Int. Conf. Solid Surf., 2nd, Kyoto, 1974, 113 (1974).

68. D. J. Harra, J. Vac. Sci. Technol. 12, 539 (1975).

69. A. A. Kuzmin, Prib. Tekh. Eksp. 3, 126 (1963).

70. D. J. Harra and T. W. Snouse, Trans. Natl. Vac. Symp. 9, 360 (1962).

71. Vacuum Generators, UHV Components Catalogue.

72. H. Uchida and E. Fromm, J. Less-Common Met. 95, 139 (1983).

73. W. Espe, M. Knoll, and M. P. Wilder, Electronics 23, 80 (1950).

74. E. I. Doucette, Trans. Natl. Vac. Symp. 7, 347 (1961).

75. J. H. N. van Vucht, Vacuum 10, 163 (1963).

76. P. della Porta et al., Trans. Natl. Vac. Symp. 8, 229 (1962).

77. T. A. Giorgi and F. Ricca, Nuovo Cimento, Suppl. 1(5), 472 (1967).

78. F. Ricca and T. A. Giorgi, J. Phys. Chem. 71, 3627 (1967).

79. B. Kindl, Nuovo Cimento, Suppl. 1(2), 646 (1963).

< previous page page_370 next page >


< previous page page_371 next page >
Page 371

80. A. Barosi, in Residual Gases in Electron Tubes (T. A. Giorgi and P. della Porta, eds.), p. 221. Academic Press,
London, 1972.

81. B. Ferrario et al., Proc. Symp. Fus. Technol., 13th, Vol. 1, p. 395 (1984).

82. C. Boffito, F. Doni, and L. Rosai, J. Less-Common Met. 104, 149 (1984).

83. B. Ferrario, Chem. Pump. Vac. Technol., Int. Workshop Interact. Gases Solids, 1st, Magdeburg (1995).

84. B. Ferrario, Proc. Int. Symp. Vac. Technol Nucl. Appl., Bombay, p. 175 (1983).

85. P. della Porta, Gettering, an integral part of vacuum technology, Natl. Symp., 39th (1992).

86. K. Ichimura et al., J. Vac. Sci. Technol. A 5(2), 220 (1987).

87. G. Sancrotti, G. Trezzi, and P. Manini, J. Vac. Sci. Technol. A 9(2), 187 (1991).

88. R. D. Penzhorn, M. Delvillers, and M. Sirch, J. Nucl. Mater. 179181, 863 (1991).

89. T. Nagasaki et al., Fusion Technol. 9, 506 (1986).

90. K. Ichimura et al., J. Vac. Sci. Technol. A 6(4), 2541 (1988).

91. G. Kuus and W. Martens, J. Less-Common Met. 111 (1980).

92. M. Succi and P. Manini, Proc. Semicon/East 189, 62 (1989).

93. M. Sancrotti, G. Trezzi, and P. Manini, Vuoto 20(2), 294 (1990).

94. A. Pebler and E. A. Gulbransen, Electrochem. Technol. 4(56), 211 (1967).

95. T. A. Giorgi, Jpn. J. Appl. Phys., Suppl. 2, 53 (1974).

96. P. S. Rudman and G. D. Sandrock, Annu. Rev. Mater. Sci. 271 (1982).

97. F. J. Lin, G. D. Sandrock, and S. Suda, J. Alloys Compd. 190, 57 (1992).

98. R. A. Miedema, J. Less-Common Met. 32, 117 (1973).

99. Van-Mal, K. H. Bushow, and A. R. Miedema, J. Less-Common Met. 35, 65 (1974).

100. D. Shaltiel, I. Jacob, and D. Davidov, J. Less-Common Met. 53, 117 (1977).

101. C. Boffito et al., Proc. Natl. Vac. Symp., 27th, Detroit; J. Vac. Sci. Technol. 18(3), 1117 (1981).

102. U.S. Pat. 4,312,669 (1982).

103. U.S. Pat. 3,899,392 (1975).

104. U.S. Pat. 4,126,449 (1978).

105. U.S. Pat. 5,180,568 (1993).

106. J. D. Baker et al., J. Vac. Sci. Technol. A 12(2), 548 (1994).


107. F. Doni, C. Boffito, and B. Ferrario, J. Vac. Sci. Technol. A 4, 2447 (1986).

108. J. M. Park and J. Y. Lee, J. Alloys Compd. 182, 43 (1992).

109. O. Bernauer and C. Halene, J. Less-Common Met. 131, 213 (1987).

110. O. Bernauer and K. Ziegler, Ger. Pat., DE 3151712C1 (1981).

111. SAES Getters, 707 Non Evaporable Getter Activatable at Low Temperature (Catalogue).

113. A. Barosi and I. A. Giorgi, Vacuum 23(1), 15 (1972).

114. B. Ferrario, F. Figini, and M. Borghi, Vacuum 35(1), 13 (1984).

115. E. Giorgi, C. Boffito, and M. Bolognesi, Vacuum 41(79), 1935 (1990).

116. U.S. Pat. 4,428,850 (1984).

117. Plücker, Poggendorf's Ann. 105, 84 (1858).

118. Plücker, Poggendorf's Ann. 103, 88 (1858).

119. L. Vegard, Ann. Phys. (Leipzig) 4, 769 (1916).

< previous page page_371 next page >


< previous page page_372 next page >
Page 372

120. Willows, Philos. Magn. 6, p. 502 (1902).

121. A. Klopfer and W. Ermrich, 5th Natl. Vac. Symp., 297 (1958).

122. R. G. Herb et al. J. Vac. Sci. Technol. 1, 54 (1964).

123. D. G. Bills, J. Vac. Sci. Technol. 4, 149 (1967).

124. F. M. Penning, Physica (Amsterdam) 4, 71 (1937).

125. F. M. Penning, Philips Tech. Rundsch. 2, 201 (1937).

126. R. Haefer, Acta Phys. Austriaca 8, 200 (1954).

127. A. M. Gurewitsch and W. F. Westendorp, Rev. Sci. Instrum. 25, 389 (1954).

128. L. D. Hall, Rev. Sci. Instrum. 29, 367 (1958).

129. W. Knauer, J. Appl. Phys. 33, 2093 (1962).

130. R. L. Jepsen, J. Appl. Phys. 32, 2619 (1961).

131. R. L. Jepsen, Le Vide 80, 80 (1959).

132. R. L. Jepsen et al., Trans. Natl. Vac. Symp. 7, 45 (1960).

133. R. L. Jepsen, Proc. Int. Vac. Congr., 4th, London, 1968, Vol. I, p. 317 (1968).

134. W. Schuurman, Rijnhuizen Report 66-28. FOM-Inst., Rijnhuizen, The Netherlands, 1966.

135. W. Knauer and M. A. Lutz, Appl. Phys. Lett. 2, 109 (1963).

136. S. L. Rutherford, 10th Natl. Vac. Symp., 185 (1963).

137. K. Akaischi, J. Vac. Sci. Technol. 8, 658 (1971).

138. J. C. Helmer and R. L. Jepsen, Proc. IRE 49, 1920 (1961).

139. D. L. Swingler, Proc. Int. Vac. 8th, Cannes, 1980, Vol. II, p. 238 (1980).

140. P. A. Redhead, Vacuum 38, 901 (1988).

141. E. M. Reykrudel and G. M. Smirnitskaya, Radiofizika 1, 36 (1958).

142. W. Knauer et al., Appl. Phys. Lett. 3, 111 (1963).

143. R. H. Good and E. W. Müller, Handbook of Physics, Vol. 4, Part XXI, p. 176. Springer, Berlin, 1956.

144. K. F. Poole and A. Venema, Proc. Int. Vac. Congr., 4th, London, 1968, Vol. I, p. 271 (1968).

145. R. Haefer, Acta Phys. Austriaca 8, 213 (1954).

146. H. Hartwig and J. S. Kouptsidis, J. Vac. Sci. Technol. 11, 1154 (1974).
147. D. Andrew, Proc. Int. Vac. Congr., 4th, London, 1968, Vol. I, p. 325 (1968).

148. H. Henning, Proc. Int. Vac. Congr., 4th, London, 1968, Vol. I, p. 143 (1968).

149. H. Oechsner, Z. Naturforsch. A2A, 859 (1966).

150. W. M. Brubaker, Trans. Natl. Vac. Symp. 6, 302 (1959).

151. W. Baechler and H. Henning, Proc. Int. Vac. Congr., 4th, London, 1968, Vol. I, p. 365 (1968).

152. M. Pierini and L. Dolcino, J. Vac. Sci. Technol. A 1, 140 (1983).

153. T. Tom and B. D. James, J. Vac. Sci. Technol. 6, 304 (1969).

154. D. Andrew, Vacuum 16, 653 (1966).

155. J. M. Lafferty and T. A. Vanderslice, Proc. IRE 49, 1136 (1961).

156. D. Andrew et al., Proc. Int. Vac. Congr., 4th London, 1968, Vol. I, p. 337 (1968).

157. S. L. Rutherford et al., Trans. Natl. Vac. Symp. 7, 380 (1960).

158. M. D. Malev and E. M. Trachtenberg, Vacuum 23, 403 (1973).

< previous page page_372 next page >


< previous page page_373 next page >
Page 373

159. R. J. Reid and B. A. Trickett, Proc. Int. Vac. Congr., 7th, Vienna, 1977, Vol. I, p. 89 (1977).

160. T. S. Chou et al., J. Vac. Sci. Technol. A 5, 3446 (1987).

161. Y. Suetsugu and M. Nakagawa, Vacuum 42, 761 (1991).

162. Y. Suetsugu, Vacuum 46, 105 (1995).

163. Y. Suetsugu and M. Nakagawa, Vacuum 42, 625 (1991).

164. S. L. Rutherford, Vacuum 16, 643 (1966).

165. E. Fischer and H. Mommsen, Vacuum 17, 309 (1967).

166. J. A. Vaumoron and M. P. De Biasio, Vacuum 20, 109 (1969).

167. U. R. Bance and R. D. Craig, Vacuum 16, 647 (1965).

168. J. H. Singleton, J. Vac. Sci. Technol. 6, 316 (1969).

169. J. H. Singleton, J. Vac. Sci. Technol. 8, 275 (1971).

170. C. W. Schoenfelder and J. H. Swisher, J. Vac. Sci. Technol. 5, 862 (1973).

171. P. N. Baker and L. Laurenson, J. Vac. Sci. Technol. 9, 375 (1972).

172. M. Wutz et al., Theorie und Praxis der Vakuumtechnik. Vieweg, Braunschweig/Wiesbaden, 1982.

173. S. Komiya and N. Yagi, J. Vac. Sci. Technol. 6, 54 (1969).

174. T. Okano et al., J. Vac. Sci. Technol. A 2, 191 (1984).

175. L. D. Hall, J. Vac. Sci. Technol. 6, 44 (1969).

176. J.-M. Laurent, Proc. Int. Vac. Congr., 8th, Cannes, 1980, Vol. II, p. 164 (1980).

177. D. Blechschmidt et al., Proc. Int. Vac. Congr., 8th, Cannes, 1980, Vol. II, p. 159 (1980).

178. H. Pingel and L. Schulz, Proc. Int. Vac. Congr., 8th, Cannes, 1980, Vol. II, p. 147 (1980).

179. H. Henning, Z. Vakuum Technik. 24, 37 (1975).

180. J. P. Hobson, J. Vac. Sci. Technol. 10(1) (1973).

181. M. M. Eisenstadt, J. Vac. Sci. Technol. 7(4), 479 (1970).

182. J. P. Hobson, J. Phys. Chem. 73(8), 2720 (1969).

183. K. M. Welch, Capture Pumping Technology. Pergamon, New York, 1991.

184. S. Brunauer, Emmett, and Teller, J. Am. Chem. Soc. 60(1), 309 (1938).

185. S. A. Stern, J. T. Mullhaupt, R. A. Hemstreet and F. S. Di Paulo, J. Vac. Sci. Technol. 2, 165 (1965).
186. R. A. Haefer, Cryopumping, Theory and Practice. Oxford University Press (Clarendon), Oxford, 1989.

187. R. E. Honig, RCA Rev. 567 (1962).

188. G. Walker, Cryogenic Cooling Systems. Plenum, New York, 1980.

189. J. F. O'Hanlon, A User's Guide to Vacuum Technology. Wiley, New York, 1989.

190. F. Turner, Varian Rep. VR-76 (1973).

General References

J. H. DeBoer, The Dynamical Character of Adsorption, Oxford University Press (Clarendon), London, 1953.

S. Dushman and J. M. Lafferty, Scientific Foundations of Vacuum Technique. Wiley, New York, 1962.

J. D. Fast, Interactions of Metals and Gases, Vols. 1 and 2. Macmillan, London, 1971.

< previous page page_373 next page >


< previous page page_374 next page >
Page 374

S. J. Gregg, The Surface Chemistry of Solids. Chapman & Hall, London, 1965.

R. I. Masel, Principles of Adsorption and Reactions on Solid Surfaces. Wiley, New York, 1996.

V. Ponec, Z. Knor and S. Cerny, Adsorption on Solids. Butterworth, London, 1974.

G. L. Saksagansky, Getter and Getter-Ion Vacuum Pumps. Harwood Academic Press, London, 1994.

B. M. Trapnell, Chemisorption. Butterworth, London, 1955.

K. Welch, Capture Pumping Technology. Pergamon, Oxford, 1991.

< previous page page_374 next page >


< previous page page_375 next page >
Page 375

6
Vacuum Gauges

R. Norman Peacock

In Chapter 1 the pressure of a gas was defined as the force per unit area exerted by the gas on its confining walls. It was
shown that this pressure is the result of molecular motion and is given by

where n is the number of molecules per unit volume, m their mass, and their mean square velocity. Since

for particles with a Maxwellian velocity distribution, where k is the Boltzmann constant and T the
absolute temperature, Eq. (6.1) may be written

This equation is fundamental to vacuum measurement, since it provides a relationship between pressure and molecular
density. Some gauges such as liquid manometers and capacitance diaphragm gauges actually sense force per unit area,
and thus measure the pressure ''directly." These are often called "direct gauges". Others, including ionization gauges,
viscosity gauges and thermal conductivity gauges are sensitive to the molecular density, n. These are said to be "indirect
gauges" since gas density must be converted to pressure by Eq. (6.2). It is customary to report all vacuum measurements
in pressure units even though the instrument may sense density. Direct gauges

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_375 next page >


< previous page page_376 next page >
Page 376

provide pressure information independent of the gas. The sensing methods of thermal conductivity, viscous drag, and
ionization are indirect and gas-dependent.

Only a limited number of pressure- or density-sensitive effects are available for use in vacuum measurements.
Commonly used transducers are based upon:

1. displacement of a liquid column, diaphragm, or other deformable element by force due to a pressure differential;

2. viscous drag acting upon a moving element;

3. thermal conductivity;

4. ionization by electrons, nuclear radiation, or laser photons, and sensing either ion current or emitted light.

Seldom used techniques or those of historical importance include:

1. transmission of sound;

2. scattering of neutral atoms from a molecular beam;

3. flash filament, adsorption/desorption method;

4. Brownian motion.

Vacuum measurements today cover approximately 17 decades below atmospheric pressure. No one type of gauging
principle is useful over this range. For most applications, two or more sensors must be used from atmospheric to
working pressure. Figure 6.1 shows the working ranges of several commonly used vacuum gauges.
Fig. 6.1
Ranges of commonly used vacuum gauging instruments. Solid lines represent the typical
range, while dashed lines are extensions applying only to certain examples or to regions
where, although the gauge is sometimes used, the accuracy is poor.

< previous page page_376 next page >


< previous page page_377 next page >
Page 377

6.1
Pressure Units Used in Vacuum Measurements

Mercury manometers have been used since the earliest days of vacuum technology. It is not surprising that the mmHg, or Torr (1 mmHg
= 1 Torr), is a commonly used pressure unit. However, the Torr or mmHg is not allowed under the Système International (SI), which, as a
coherent system, permits only those units which are derived from the fundamental quantities of mass (kilogram), length (meter), and time
(second) by simple multiplication, without the use of numerical factors such as the density of mercury.

Pressure is force per unit area. The SI unit of force is the Newton (symbol N), which has the dimensions kg·m·s2. Thus the dimensions of
pressure must be

(P) = (kg·m·s2/m2).

The N/m2 is named the pascal, abbreviated as Pa. It is easy to find the conversion factor from Torr to Pa. Looking ahead in this chapter to
Eq. (6.3), the pressure due to a mercury column of height h is P = hρg, where ρ is the density of mercury (1.3595 × 104 kg·m3 at 0°C)
and g is the acceleration due to gravity (9.806 m·s2). Then for h = 1 mm

P(1 Torr) = 103 m·1.3595 × 104 kg·m2 × 9.806 m·s2


= 1.333 × 102 Pa.

The bar is a pressure unit defined as 105 Pa. It is approximately equivalent to a standard atmosphere pressure (760 Torr):

1 std. atm = 760 Torr = 760 Torr × 1.333 × 102 Pa/Torr


= 1.013 × 105 Pa.

The bar and millibar (mbar) are temporarily allowed by the ISO, although they do not differ from the Pa by a factor of 103n as is usually
required. The mbar has the advantage of having the same magnitude as the Torr (1 mbar = 0.750 Torr), and it is commonly used in
Europe. Other pressure units are encountered occasionally. Table 6.1 has conversion factors for some common pressure units.

Table 6.1. Conversion Factors for Some Common Pressure Unitsa


Pa mbar Torr in. Hg atm.
Pa
1 1.00 × 102 1.33 × 102 3.39 × 103 1.01 × 105
mbar
1.00 × 102 1 1.33 3.39 × 101 1.01 × 103
Torr
7.50 × 103 7.50 × 101 1 2.54 × 101 7.60 × 102
in. Hg
2.95 × 104 2.95 × 102 3.94 × 102 1 2.99 × 101
atm.
9.87 × 106 9.87 × 104 1.32 × 103 3.34 × 102 1
a Multiply quantities given in the units shown along the top row by the factor in the table to obtain the units desired from the
vertical column on the left.

< previous page page_377 next page >


< previous page page_378 next page >
Page 378

6.2
Liquid Manometers

Figure 6.2 illustrates a simple manometer consisting of a glass U-tube filled with a liquid of density ρ. The pressure P1
in the left arm is balanced by the pressure P2 in the right (P1 > P2) plus the pressure due to the difference height, h, of
the mercury columns in the two arms, or P1 = P2 + hρg, where ρ is the density of the liquid and g is the local
acceleration due to gravity. This equation may be written in terms of the pressure difference

Mercury and diffusion pump oils are the liquids commonly used in manometers. Because the density of oil is more than
a decade less than mercury, h for a given pressure differential will be correspondingly larger. Traps are usually needed
when using mercury since the vapor pressure of mercury is 0.16 Pa at 20°C, while that of modern diffusion pump fluids
is less than 106 Pa.

The accuracy attainable with a manometer is highly dependent upon the uncertainty in measuring h. With a ruled scale
behind the arms of the "U" tube, h may be estimated to perhaps ± 0.1 mm with the naked eye. This could be improved
by a factor of 10 using a cathetometer. An electrical contact with micrometer adjustment can determine the position of a
mercury surface to ± 102 mm. Interferometric methods as used today are accurate to about ± 1 × 105 mm.

Highly developed liquid manometers are the primary pressure standards for many countries. Heydemann, Tilford, and
Hyland [1] of the (then) U.S. National Bureau of Standards reported a precision mercury manometer utilizing ultrasonic
interferometry to determine the heights of the mercury columns. The resolution of this instrument was 1.4 mPa. Ooiwa,
Ueki, and Kaneda [2] of the National Research Laboratory of Japan in Tsukuba described a mercury manometer using a
white light Michelson interferometer. This instrument, which is the primary pressure standard of Japan, has an
uncertainty at 100 kPa of about 0.4 Pa.

Fig. 6.2
The U-tube manometer. The
gas pressure in the left arm is
P1 and is greater than the
pressure in right arm, P2. The
difference in height of the
liquid in the two arms is h.
< previous page page_378 next page >
< previous page page_379 next page >
Page 379

Ueki and Ooiwa [3] also made an oil manometer with a laser heterodyne interferometer for use at lower pressures. The
uncertainty of readings from 1 Pa to 1 kPa was reported to be 0.01%, or 2 mPa.

Legras and Le Breton [4] of the Laboratoire National D'Essais in Paris reported a manometer using liquid gallium. The
liquid position was sensed with a capacitance transducer. The advantage of operation with gallium is the extremely low
vapor pressure of the liquid. The resolution was said to be about 103 Pa, with a repeatability of 0.01 Pa for pressures
less than 4 Pa.

Considerable care is needed in using liquid manometers. A mistake can cause the liquid to be violently ejected into the
system. Because the vapor pressure of mercury is about 0.16 Pa at 20°C, it is necessary to use a cold trap between a
mercury manometer and system to prevent mercury vapor from entering. Contamination of the laboratory may occur if
mercury is spilled. If mercury vapor were present in the air at equilibrium concentration at room temperature, it would
be a serious health hazard [5].

Many precautions and corrections are needed if accurate results are to be obtained with liquid manometers. The density
of the liquid must be known at the operating temperature. With mercury, surface tension effects depress the liquid
surface, and with oil they elevate it. Surface effects depend upon purity of the liquid. At low pressures, special means
such as ultrasonic or optical interferometry are needed to determine column height.

In spite of the simplicity of the manometer concept, liquid manometers are probably best used in standards laboratories
where proper care can be taken in their use and their importance as primary standards justifies the effort. Note also that
liquid manometers are not suitable for following rapidly changing pressure. For general use there are alternate gauges
without the disadvantages of liquid manometers.

6.3
McLeod Gauge

The McLeod gauge [6] makes use of Boyle's law to extend the range of the manometer to lower pressures. Figure 6.3
illustrates a simple form of the instrument. The capillary tubes (1) and (2) must have exactly the same bore diameters so
that capillary depression will be identical. The volume of the bulb plus capillary (1) and the tubing above the cutoff is V.
The open tube at the top of the gauge is connected to the pressure to be measured. A liquid nitrogen trap is required to
prevent mercury vapor contamination.

In one mode of operation the reservoir is slowly raised, and the mercury reaches the cutoff point. At this time the
pressure in the bulb is taken to be identical to that at the inlet to the system. As the reservoir is raised further, the gas in
the bulb and capillary (1) is compressed. The mercury level in capillary (2) is brought to the same height as the top of
capillary (1). If the pressure in the bulb was greater than zero at the time of cutoff, the height of the mercury in capillary
(1) will be lower than in (2) by an amount h. The pressure in the small volume at the top of capillary (1) is greater by
hρg. By Boyle's law, where P and V are the values at cutoff, Pf and Vf are as above, and A is the cross-sectional area of
the capillary (1),

< previous page page_379 next page >


< previous page page_380 next page >
Page 380

Fig. 6.3
Simple form of McLeod gauge. (a) Overall view.
(b) Detailed view defining the symbols used. The
figure is drawn for operation in the quadratic mode,
where h is the height difference between the mercury
columns when the level in capillary 2 is brought
to the level of the top of capillary 1.

Solving this equation for P gives

Typically the volume Ah is negligible compared to V, thereby giving as a good approximation

A McLeod operated according to the assumptions above is said to be used in the "quadratic mode" because P is
proportional to h2. In this mode it can cover about four decades.
If the mercury is always raised to a fixed position in capillary (1), leaving a fixed distance, d, between the mercury
surface in capillary (1) and the closed end of the capillary, while the height of the mercury in capillary (2) is allowed to
vary, and if h is the height difference between the levels in the two capillaries, then by Boyle's law,

< previous page page_380 next page >


< previous page page_381 next page >
Page 381

Again making the assumption that Ah V, the approximate result is

This is known as the linear mode of operation of a McLeod, and it is useful over a range of about two decades.

There are many sources of error when using a McLeod gauge. For a more detailed discussion see Bermann [7]. As with
other manometers, measurement of h is critical. The density of mercury, the local acceleration due to gravity, and the
bore of the capillary must be known. Use of a McLeod assumes that the gases to be measured obey Boyle's Law.
Condensible gases and vapors do not. It is assumed that the temperature of the gas in the cutoff volume is identical to
that in the system. Unless the mercury is absolutely clean, it tends to stick, especially in small capillaries.

Streaming of mercury vapor from the inlet to the cold trap produces an error due to the pumping action of this
unidirectional vapor flow. This error can be as large as 20% for a mercury temperature of 25°C when using connecting
tubing of 20-mm bore. This effect was first noted by Ishi and Nakayama [8] while studying the long-term stability of the
gauge constants for a group of ionization gauges. They observed that the gauge constants seemed to vary seasonally.
The temperature of their laboratory at the old Electrotechnical Institute in Tokyo varied from 10°C in January to 30°C
in August, causing a periodic error in the pressure as measured by a McLeod. Gaede [9], in his original paper on the
diffusion pump, published the complete theory of this effect in 1915. His apparatus for verifying the pumping of
streaming mercury vapor consisted of a McLeod with an additional mercury reservoir whose temperature could be
controlled. But it is not clear that Gaede was aware of the error caused by vapor streaming in normal operation of the
McLeod.

Operation of the McLeod is very slow, and it cannot follow changing pressure.

The cold trap can become a hazard when a calibration system is used with a condensible gas. Peacock [5] describes an
accident where argon condensed in the trap. The gauge and trap were isolated from the system by a closed valve; and
the Dewar of liquid nitrogen was removed from the trap, thereby permitting the trap to become warm. The isolated trap
and Mcleod exploded, thereby peppering the operator with flying glass and distributing many kilograms of mercury
over the laboratory.

6.4
Piston Pressure Balance Gauge

A simple form of the piston pressure balance gauge is illustrated in Fig. 6.4. This instrument is used for creating known
pressure differentials to calibrate other gauges. A small amount of leakage past the piston is of no consequence. The
piston, of mass m and area A, is a very close fit in the cylinder. This gauge is similar to the liquid manometer, since in
operation the piston is supported by the pressure differential

< previous page page_381 next page >


< previous page page_382 next page >
Page 382

Fig. 6.4
Cross section of an idealized
piston pressure balance gauge.
The piston of mass m and cross-
sectional area A is a close but
friction free fit in the cylinder.

P1 > P2. Setting the sum of the forces acting on the piston equal to zero for equilibrium,

Although useful for measuring large pressure differences, this form of the piston gauge is of little use for small
differences because of the piston mass, m.

Ooiwa [10, 11] described a variation of the piston gauge as shown in Fig. 6.5. The piston is counterbalanced by a
dummy piston on the other arm of the balance beam, and the force due to the pressure differential is sensed by the
electronic balance attached to the piston. Ooiwa was able to measure pressures in the range of 1 Pa to 10 kPa with a
sensitivity of about 5 mPa. This is adequate to permit direct calibration of some higher-range diaphragm gauges using
the piston gauge as a basic standard. With a calibration system of the expansion type, measurements can be extended to
lower pressures.

Ooiwa [10] and Solis [12] discussed sources of uncertainties in the use of the piston gauge. These include: the mass,
area, and buoyancy of the piston; vertical adjustment of the cylinder; magnetic fields; local gravity; and temperature
effects. More detail on the piston pressure balance gauge is given in Chapter 12.

6.5
Bourdon Gauge
Gauges using the deflection of an aneroid cell, Bourdon tube, or a tensioned membrane also measure true pressure
independent of the gas. The deflection of a sensing element may be measured in many ways. Mechanical coupling of a
cell to an

< previous page page_382 next page >


< previous page page_383 next page >
Page 383

Fig. 6.5
Piston pressure balance gauge by Ooiwa. With the mass of the
piston counterbalanced, the range of the piston gauge can be
extended down to 1 Pa with a sensitivity of 5 mPa. Reprinted with
permission from A. Ooiwa, Metrologia 30, 607 (1993/1994) [Ref. 10].
Copyright 1993 Springer-Verlag.

indicating needle is one means. For example, Wallace and Tiernan [13] manufacture a dial-type gauge measuring to
atmosphere with 100-Pa resolution. Simple, inexpensive Bourdon gauges are sometimes used on vacuum chambers to
monitor the progress of roughing. Bourdon gauges use a flattened tube of elastic material formed into a circular shape,
as shown in Fig. 6.6a. The Bourdon tube tends to straighten as the internal pressure increases. In Fig. 6.6a the motion is
communicated to a needle for direct reading. Greater sensitivity can be obtained with a multiturn helical Bourdon tube.
The theory of the Bourdon tube is given by Lorenz [14]. Several related gauges were discussed by Dushman [15].

The most sophisticated instrument of this type is the quartz helix Bourdon gauge (QBG). An example is that
manufactured by Ruska Instruments Corporation [16] and is shown in Fig. 6.6b. It operates on the force balance
principle with magnetic nulling. The null position is sensed optically via a mirror on the free end of the helix. Solis [17]
states that the rotation is primarily a function of the pressure differential, the cross-sectional area and geometry of the
tube, the diameter of the helix, and the number of turns. For the Ruska instruments with full-scale (FS) pressures less
than 500 psi (3 × 106 Pa), the repeatability is ±0.002% FS. The resolution is ±0.001% FS, the linearity is ±0.002% FS,
and the hysteresis is < 0.001% FS.
< previous page page_383 next page >
< previous page page_384 next page >
Page 384

Fig. 6.6
Bourdon gauges: (a) Inexpensive Bourdon gauge;
(b) quartz Bourdon gauge used for precision differential
measurements. Part b reprinted with permission from
K. Solis, "The Fused Silica Helix Bourdon Tube: Its Place
in Measurement and Control." Ruska Instrument Corp.,
Houston, TX 77063 (no date) [Ref. 17].

Since the settling time is about 90 s, the QBG is best suited for a standards laboratory or similar facility.
6.6
Capacitance Diaphragm Gauges

Commercial capacitance diaphragm gauges (CDGs) appeared around 1960. They have been improved to the extent that
they can replace liquid manometers in most applications other than as primary standards. Their use avoids messy or
hazardous

< previous page page_384 next page >


< previous page page_385 next page >
Page 385

liquids along with the possibility of contaminating the system or laboratory. Only inert materials at or near ambient
temperature are exposed to the gas. Accurate measurements over four decades using a single head are possible.

There are other diaphragm gauges since other techniques may be used for observing the deflection of a diaphragm.
Inductive transducer methods were examined early in the history of diaphragm gauges [18]. Various sensing means are
used by the manufacturers of industrial diaphragm gauges for the pressure region above 1 Pa, including strain gauges,
piezoresistive films, and inductive. However, these products are seldom seen in vacuum applications relevant to this
book.

Capacitance sensing is a simple and precise technique for observing the diaphragm position. As early as 1936, Hasse
[19] described a functional single-sided CDG. Others who published descriptions of CDGs before commercial
instruments were available include Lilly, Legallais, and Cherry [20], Alpert, Matland, and McCoubrey [21], Drawin
[22], Hecht [23], and Macdonald and King [24].

CDGs may have a sense electrode on one side of the diaphragm only, or on both. Both sides may have ports so that the
instrument can be used for differential pressure measurements, or one side may be evacuated and sealed so that the
gauge is for absolute pressure sensing. See Fig. 6.7.

6.6.1
Sensitivity of the Capacitance Method

A simple numerical example will illustrate the value of the capacitance method for sensing diaphragm deflection, and at
the same time it will make evident those characteristics of the CDG helpful in understanding its characteristics.

Figure 6.8 shows the diaphragm and sense electrode of a simple CDG sensor. Let the diameter of the diaphragm be 50
mm, and let area of the capacitance sensing electrode be 1 × 104 m2, spaced 0.1 mm from the diaphragm. The
capacitance may be estimated using the parallel plate formula

where C is the capacitance in farads (F), ε0 is a constant 8.85 × 1012 F/m, K is the dielectric constant, A is the area of
the electrode, and s is the separation.

For the numbers of this example the sense electrode to diaphragm capacitance is 8.9 × 1012 F. Allowing for some stray
capacitance, a total of 15 pF is reasonable.

Fig. 6.7
Capacitance diaphragm gauges: (a) Two-port CDG for differential pressure
measurements; (b) single-port CDG with sealed and gettered vacuum
reference for absolute pressure measurements.
< previous page page_385 next page >
< previous page page_386 next page >
Page 386

Fig. 6.8
The diaphragm and capacitance sensing electrode
of a simple CDG sensor. The area of the
capacitance probe is A, and the separation
from the diaphragm is s. The thin membrane
is deflected a distance Y by the pressure P.

Assume that a capacitance change of 3 × 1017 F is a practical limit of resolution for an inexpensive instrument. Note
that if 15-pF capacitance were used in an LC circuit resonant at 10 MHz, an incremental change of 3 × 1017 F would
change the resonant frequency by 10 Hz. Solving Eq. (6.12) for s and taking the derivative, the change in spacing for an
incremental change of capacitance is

For δC = 3 × 1017 F and the numbers of the present example, a resolvable deflection is 3 × 1010 m, or about the size of
an atom.

6.6.2
Deflection of a Thin Tensioned Membrane

Referring to Fig. 6.8, the center deflection, Y, of a circular thin flexible diaphragm of radius a tensioned at the edge by a
radial tension, T, per unit length of the circumference is, for small deflections, given by Rocard [25] as

where P is a uniformly distributed pressure. For a δY just resolvable by a capacitance measurement, the resolvable δP
by Eq. (6.14) is 2.2 × 103 Pa. This example is similar to a 100-Pa full-scale sensor. That is, for δp = 100 Pa the
deflection would be about 2 × 102 mm. This is one-fifth of the 0.01-mm diaphragm to sense electrode separation that
was assumed for zero pressure differential.
The capacitance method senses not only movement of the diaphragm caused by pressure change, but also dimensional
changes due to temperature effects, gravity, and vibration. Hysteresis, or return to zero after deflection, must also be
considered. All of these problems can cause shift of the zero pressure setting. Zero stability is often a problem with
CDGs in the lowest decade of their range.

< previous page page_386 next page >


< previous page page_387 next page >
Page 387

Fig. 6.9
Schematic of a valve arrangement for verifying the zero on
(a) a differential CDG and (b) an absolute CDG.

In designing a sensor, attention must be given to choice of materials to avoid differential motion and zero shift caused
by temperature effects. When using a CDG, it is imperative that means be provided for verifying the zero. For a two-
port differential CDG, this may simply necessitate valving means to make possible equalizing the pressure at the two
ports. For a single-port, sealed reference CDG, verifying the zero requires a pumping system capable of reducing the
pressure to about 1% of the lowest resolvable pressure. Valving arrangements for the two-port and single-port gauges
are shown schematically in Fig. 6.9.

Residual temperature effects are dealt with in two ways: A temperature sensing element may be used in a circuit to
compensate for temperature variations, and/or the sensor may be placed in a constant temperature enclosure maintained
somewhat above the highest expected ambient. This temperature is usually between 45°C and 50°C.

6.6.3
Accuracy of Commercial Gauges

Present commercial instruments cover a wide range, are convenient to use, and are remarkably stable and accurate.
Circuits attached to the pressure sensing cell process the capacitance information and give analog or digital output.
Inexpensive single-port heads with sealed vacuum reference are available with full-scale ranges from 133 Pa (1 Torr) to
3.3 × 106 Pa (25,000 Torr). Claimed accuracy after subtracting zero drift is ± 0.5% of reading, and optional ± 0.15%
accuracy is available. These uncertainties are due to causes such as non linearity and hysteresis. These sensor heads
need only be supplied with + 15/0/ 15 V dc, and the output is 0 to 10 V dc [26].

For critical laboratory use, precision temperature-controlled sensor heads with sealed reference are available in full-
scale ranges from 13.3 Pa (0.1 Torr) to 3.3 × 106 Pa (25,000 Torr). For heads 133 Pa to 105 Pa full scale, accuracy of ±
0.05% of reading is available [27].

Sullivan [28] and Sullivan and Uttaro [29] discussed the various sources of uncertainty in the CDG. They assumed an
instrument with stated uncertainty of 0.08% of reading due to nonlinearity and hysteresis, resolution of 0.0001% of FS,
zero drift coefficient of 0.0004% of FS per °C, and span temperature coefficient of 0.002% of reading. The relative
importance of these errors is apparent if they are applied to

< previous page page_387 next page >


< previous page page_388 next page >
Page 388

the example of a 1-Torr (133-Pa) head used at 1 × 104 Torr (1.33 × 102 Pa). Then the uncertainty due to nonlinearity
and hysteresis is 8 × 108 Torr; due to resolution, 1 × 106 Torr; due to zero temperature drift, 4 × 106 per °C; and due to
span temperature coefficient, 2 × 109 per °C. The total uncertainty is 5.08 × 106 per °C; and at this pressure, 1 × 104 is
5.08% of reading per °C. The significance of temperature zero drift is clear.

Several international standards laboratories have studied the long-term stability of CDGs for use as transfer standards.
Hyland and Tilford [30] examined 14 different temperature controlled gauges at the US National Bureau of Standards
over long times of up to four years. Percentage shift at midrange pressure for the various CDGs varied from 0.05% to
2.02%, with most shifts well under 1%. Grosse and Messer [31] of the Physikalisch-Technische Bundesanstalt in Berlin,
calibrated two CDGs in many gases several times between 1980 and 1985. The deviation of the pressure indication from
the calibration pressure never exceeded 1%. M. Bergoglio and A. Calcatelli of the Istituto Metrologia G. Colonnetti,
Torino, found similar stabilities [32].

The result of these studies has been the acceptance of CDGs as secondary or transfer standards within their pressure
range. It should be noted that the CDG has improved in stability, the variety of products available, and convenience of
use, since the instruments were purchased for the tests described above.

6.6.4
Thermal Transpiration

When temperature-controlled sensor heads are used, the temperature of the sensor usually is greater than that of the
chamber whose pressure is to be measured. The gauge reading will be a few percent high under some conditions. This
error becomes significant when the pressure is lower than that where the mean free path of the gas is similar to the
diameter of the tubing connecting chamber and sensor (see Section 1.10).

A plot of an experimental transpiration error measurement by Jitschin and Röhl [33] is reproduced as Fig. 6.10. The
temperature of the chamber was about 26.5°C, and the sensor was 40.5°C. The diameter of the tubing where the
temperature change occurred was about 4.6 mm. The curves show the error to be negligible above 100 Pa. It increases
to about 2% at 102 Pa. In the low-pressure limit the ratio of the gauge to chamber pressure should approach

where p2 is the pressure in the heated gauge at temperature T2, and p1 is the chamber pressure at ambient temperature
T1. Similar results were found by Poulter et al. [34].

6.6.5
Conclusions

Measurements made with a CDG are independent of the gas except for the small transpiration correction. With single-
sided gauges, only noncorrosive metals such as stainless steel or Inconel are exposed to the vacuum, making the CDG
inert for most gases, not decomposing or otherwise altering the gas. Its volume is small, and its response is fast. The
long-term accuracy, when used properly, can exceed 1%, justifying its use as a secondary standard or transfer gauge. It
may be purchased with

< previous page page_388 next page >


< previous page page_389 next page >
Page 389

Fig. 6.10
Plot of the thermal transpiration error for two sets of experimental
data (triangles and inverted triangles). Reprinted with permission
from W. Jitschin and P. Röhl, J. Vac. Sci. Technol. A5, 372 (1987)
[Ref. 33]. Copyright 1987 American Vacuum Society.

two ports for differential measurements, or with an evacuated reference for absolute pressure measurements. The most
serious disadvantage of the CDG is the lack of an absolute zero. It is usually necessary to provide a pumping or valving
system to check the zero.

The American Vacuum Society ''Recommended Practices for the Calibration and Use of Capacitance Diaphragm
Gauges as Transfer Standards" [35] is a valuable source of information.

Some Suggestions to Help Obtain Good Results with CDGs

1. Avoid over pressurizing the gauge. If the chamber is frequently vented to the atmosphere, an isolation valve at the
CDG inlet will prolong life.

2. For units with heated heads allow a warm-up time of several hours.

3. Maintain a stable ambient temperature.

4. Prevent mechanical stress on the sensor cell by proper installation, using a bellows when necessary.

5. Avoid vibration.

6. Avoid particulate contamination that could inhibit diaphragm motion.

7. Recalibrate frequently.
6.7
Viscosity Gauges

Viscosity gauges utilize the drag effect observed when gas molecules act upon a moving object or surface. Gas
molecules leaving a surface moving with a velocity

< previous page page_389 next page >


< previous page page_390 next page >
Page 390

parallel to the surface plane gain momentum in the tangential direction, while that of the moving body is decreased.
Figure 6.11 illustrates several types of viscosity gauges: (a) those based upon observing the decrement of an oscillating
vane; (b) those of an scillating disc in torsion; (c) those of an oscillating fiber; (d) those based upon the coupling of a
rotating disc to a stationary disc; (e) those using the pressure dependence of the impedance of an oscillating object such
as a quartz crystal tuning fork; and (f) those measuring the decrease of the angular velocity of a freely suspended
spinning sphere.

Dushman [36] covered the early history and application of the first two of these varieties of viscosity gauges. In 1972,
Steckelmacher [37] reviewed the theory of viscous drag for planes and cylinders moving in a gas, just before the
spinning rotor gauge (SRG) of Fig. 6.11f became important.

Although investigators from Edwards High Vacuum International and the University of York have revived the coupled
disc viscosity gauge of Fig. 6.11d in recent publications [38, 39], the viscosity gauges of Fig. 6.11ad are mostly of
historical importance.

Two gauges of Fig. 6.11 are worth detailed consideration. Since 1984 the vacuum group of the Electrotechnical
Laboratory in Japan has published a series of papers concerning the quartz oscillator gauge. It may have some
importance in the future and is discussed later in this section. The spinning rotor gauge is one of the newest instruments
for vacuum measurement. Although mentioned briefly in the 1962 edition of this book, it was not practical at that time.
Advances in digital electronics now make accurate timing of rotational speed possible, and the SRG has become
important as a secondary, or transfer, standard.

Fig. 6.11
Viscosity gauges: (a) Oscillating vane;
(b) oscillating disc in torsion; (c) oscillating
fiber; (d) gas coupling of a rotating disc to a
torsion disc; (e) oscillating quartz crystal, here in
the shape of a tuning fork; (f) spinning rotor gauge.

< previous page page_390 next page >


< previous page page_391 next page >
Page 391

6.7.1
Spinning Rotor Gauge

The SRG determines pressure from the decrement of the angular velocity of a small spherical rotor suspended in the
vacuum. The frictional loss due to gas at a pressure of 104 Pa is small. At this pressure, a steel ball 4.5 mm in diameter
spinning at 410 rps would require about 18 h for a decrease of 1 rps. A nearly frictionless support is necessary if the
decrement in rotational speed of a small sphere is to be used as a measure of high vacuum. Beams and associates at the
University of Virginia were the first to apply a magnetic "bearing" to an SRG [40]. They compared it to an ionization
gauge over the range 1.3 × 102 to 6.5 × 106 Pa and found good agreement. However, they stated that their SRG was
very sensitive to shock.

Electrostatic suspension can provide an even lower loss bearing for a rotating sphere. An electrostatically suspended
vacuum gyroscope suggested by Nordsieck [41] was developed during the 1950s at the University of Illinois. Viscous
drag due to gas was observed to pressures as low as 107 Pa. The feasibility of using electrostatic suspension for a
vacuum gauge was discussed by Nuttall and Witt [42]. They concluded that it might be possible to build such a gauge
enabling measurements at pressures lower than existing instruments with magnetic levitation.

However, electrostatic levitation of a massive rotor is possible only at low pressures because the high electric fields
required for suspension would cause a gas discharge. This probably makes a wide-range electrostatically suspended
SRG impossible. Today, the only commercially available SRG uses magnetic suspension.

6.7.1.1
Theory

A spherical rotor freely spinning in a gas at low pressure is slowed by interaction with the gas. An equation for this
deceleration may be derived if three reasonable assumptions are made about the gassurface interaction:

1. The mean free path for the gas must be greater than rotor-to-wall separation. This ensures that gas arriving at the
rotor has a proper Maxwellian velocity distribution and does not come from a gas cloud spinning with the rotor. The
situation at higher pressures is more complicated and not covered here.

2. Most of the gas molecules arriving at a surface dwell there briefly, and they leave with a symmetrical cosine spatial
distribution and with a velocity distribution corresponding to the temperature of the surface. The alternate is specular
reflection, where the molecule "remembers" the tangential component of its momentum. These cases are illustrated in
Fig. 6.12. The effective accommodation coefficient, σ, is defined as the
Fig. 6.12
Molecules departing fixed (a) and
moving (b) surfaces. The molecules
transfer momentum from the moving
surface since, as seen in the rest frame,
the velocities of the departing molecules
are the vector sum of the usual
cosine velocity distribution and the
velocity of the moving surface. Part b
shows the resultant velocity for
one molecule.

< previous page page_391 next page >


< previous page page_392 next page >
Page 392

Fig. 6.13
Coordinates and element of area for integrating
the loss of tangential angular momentum
over the surface of a spherical rotor.

fraction of the molecules that are scattered diffusely. Typically, σ is near unity. It may exceed unity for rough surfaces.

3. Angular momentum is conserved in the gas/rotor system. To make use of the conservation of angular momentum,
expressions must be found for the angular momentum of the rotating sphere and for the component of the momentum
transverse to the spherical surface lost by the gas. Figure 6.13 shows the first quadrant surface of a spherical rotor of
radius a spinning with angular velocity ω. An element of area at azimuthal angle φ is

The surface element dA, having radius r, has a tangential velocity vT (dA):

A molecule of mass m striking the surface element dA and accommodating will depart with increased angular
momentum mvTr, while the angular momentum of a specularly scattered molecule will not change. The incremental
change of tangential angular momentum, δLT, of a gas molecule striking dA is

From Chapter 1, the number of molecules, ν, arriving at a surface per unit time and area is
< previous page page_392 next page >
< previous page page_393 next page >
Page 393

where n is the number of molecules per unit volume, and is the mean molecular velocity, . Then,
assuming perfect accommodation, the time rate of change of LT for all gas molecules colliding with the surface element
dA per second Substituting for dA from Eq. (6.16) gives

This may be integrated over the surface of the sphere to give the total rate of change of transverse momentum, dT/dt, for
the colliding gas

The angular momentum of the sphere is Ls = Iω. If Ls changes because of change of I caused either by thermal
expansion or by change of ω, then the rate of change of angular momentum for the spherical rotor may be written

The moment of inertia, I, of a solid sphere of radius a made of a material of density ρ is

Taking the derivative of Eq. (6.23) with respect to a, where a may change because the temperature is time-dependent,
and noting that the mass of the sphere is constant while ρ and a vary gives dI/dt = I·2αdT/dt. Substituting in Eq. (6.22)
and simplifying yields

By conservation of angular momentum, we obtain

Substituting Eqs. (6.21) and (6.24) into Eq. (6.25), introducing the accommodation coefficient, σ, to permit less-than-
perfect accommodation, and using Eq. (6.23) for I gives

< previous page page_393 next page >


< previous page page_394 next page >
Page 394

From Chapter 1, n = P/kT, where P is the gas pressure in Pa, and k is the Boltzmann constant 1.38 × 1023 J/K.
Substituting for n and solving for P, Eq. (6.26) may be written in terms of v as

or, substituting for v,

The negative signs of the terms in the parentheses are proper since dω/dt due to viscous drag will be negative, and a
positive dT/dt will cause a negative dω/dt at P = 0. If the residual drag (RD) is defined as the relative deceleration at
zero pressure, it may be included in a term with the pressure-dependent deceleration for the spinning rotor:

It is clear from Eq. (6.29) that since the molecular weight, and possibly σ, depend upon the gas, then measurements
made with a SRG are gas-dependent.

Fremerey [43] suggested extending the range of the SRG from the limit of about 104 Pa assumed in deriving Eq. (6.29)
to higher pressures approaching an atmosphere. This requires that the viscosity of the particular gas be included in the
pressure calculation made by the SRG controller.

6.7.1.2
Commercial Gauges

The present commercial magnetically levitated SRG is largely the result of the work of Fremerey and colleagues, first at
the University of Bonn and then later at the KFA, Jülich. In 1971, Fremerey [44] described an early SRG utilizing
electronic damping in the suspension and optical sensing of rotational speed. He examined the decay ratio (dω/dt)/ω for
several spheres of differing diameter for rotational frequencies from 2 × 104 to 1 × 105 Hz. The residual drag at zero
pressure, mostly due to eddy currents, corresponded to a few times 104 Pa nitrogen equivalent. In 1972 he introduced
inductive sensing of rotation [45].

Figure 6.14 shows the construction of an SRG head. A thimble (T) attached to the vacuum chamber contains a small
spherical rotor (R) of ferromagnetic material such as SAE 52100 carbon steel or 440C stainless steel. The outside
diameter of the thimble is approximately 8.5 mm, and that of the ball is 4.5 mm. Two permanent magnets (M), with
positioning coils (A) and lateral damping coils (L), are used with associated circuits to suspend the ball. Rotation of the
ball is sensed inductively by two pickup coils (P). There are four drive coils (D) to drive the ball up to speed
inductively. Figure 6.15 illustrates how the rotation of the ball is sensed. To induce a signal in the pickup

< previous page page_394 next page >


< previous page page_395 next page >
Page 395

Fig. 6.14
View of the internal structure of a spinning
rotor gauge head. The components are
explained in the text.

Fig. 6.15
Inductive sensing of the rotation of a
magnetized spherical rotor in an early
SRG head with two pickup coils.
coils the rotor must be slightly magnetized, and the magnetic moment must not coincide with the spin axis.

When turned on, the controller quickly brings the rotor up to about 410 rps. The ball is allowed to coast, and the
decrement of the angular velocity is measured by a digital timer system in the controller. When it slows to about 400
rps, the drive

< previous page page_395 next page >


< previous page page_396 next page >
Page 396

actuates briefly to bring it back to 410 rps. A good discussion of the operation of the SRG is given by Fremerey [43],
along with consideration of an algorithm used to obtain the average decay rate.

The commercial SRG controllers provide statistical information on the measurements. A pressure determination may
consist of 1 to 99 independent measurements, each between 0.5 and 30 s long [46]. The controller averages the
individual results and prints the average pressure, the signal scatter, and the standard deviation of the measurements.
Scatter of values is caused by timing errors, pressure drift during the measurement period, and vibration.

The fluctuations due to timing errors are unavoidable. Figure 6.16 from Redgrave and Downes [47] is a plot of the
standard deviation versus the sampling interval for a second-generation controller. Note that the random noise is
equivalent to a pressure of 104 Pa for a sampling interval of 10 s.

Looney et al. [48] placed all of the control electronics for an SRG on an accessory plug-in board for a personal
computer (PC). The PC then performs all of the functions of the present controllers, including levitation, spinning,
timing, and the statistical calculations with the pressure data. Greater control of some parameters such as the
Fig. 6.16
Standard deviation of SRG measurements versus
the duration of the sampling interval in seconds. Data
were obtained using a SRG-2. Reprinted from
F. J. Redgrave and S. P. Downes, "Some Comments
on the Stability of Spinning Rotor Gauges," Vacuum,
Vol. 38, 839842 [Ref. 47]. Crown Copyright 1988, with
kind permission from Elsevier Science Ltd., The
Boulevard, Langford Lane, Kidlington OX5 1GB.

< previous page page_396 next page >


< previous page page_397 next page >
Page 397

measurement time and rotor frequency is permitted. By using long measurement times the useful range can be extended
below 105 Pa. These authors also derived an algorithm for measuring the deceleration rate. Another advantage of the
plug-in board is that one board will operate up to four gauge heads.

6.7.1.3
Stability

The accommodation coefficient, σ, of a rotor must be constant if the SRG is to be used as a secondary standard.
Numerous papers have examined the value of σ and its stability over periods as long as several years [47, 4952].

The range of values of σ found with uncalibrated balls is quite limited. Dittmann, Lindenau, and Tilford [52] measured
σ for 68 smooth balls. The results are shown in Fig. 6.17. For this group we have 0.96 < σ < 1.06. From references 4953
it appears that 0.95 < σ < 1.05 for new ball bearings. For Grade 5 bearings the surface roughness is about 0.8 µin. (0.02
µm), and the departure from sphericity is 5 µin. (0.13 µm).

Redgrave and Downes [47] of the National Physical Laboratory (NPL), Teddington, calibrated one rotor several times
over a two-year period with an SRG-2. Table 6.2 shows that σ can be very stable under laboratory conditions. The
accommodation coefficient does not change unless the surface of the ball is damaged. Significant changes in σ result
when the ball suspension fails while it is rotating ("crashing"), or when a head is shipped with the ball free in the
thimble [31, 52].

Figure 6.18 reproduces a plot from Messer et al. [53] showing repeated determinations of σ for four balls during
19811986. The sudden changes were presumably caused by damage to the ball surfaces. The value of σ may be as high
as 1.2 for intentionally or accidentally roughened balls [50]. Surface contamination, or damage by chemical etching,
will change σ also. Subject to the possible ± 5% error resulting when a rotor is not calibrated, the SRG may be used as
an absolute gauge as proposed by Fremerey [43].
Fig. 6.17
Histogram showing the distribution of effective accommodation
coefficients for 68 smooth steel bearing balls used as rotors.
Note that the accommodation coefficients of all fall in the
interval 0.961.06. Reprinted with permission from S. Dittmann,
B. E. Lindenau, and C. R. Tilford, J. Vac. Sci. Technol., A 7,
3356 (1989) [Ref. 52]. Copyright 1989 American Vacuum Society.

< previous page page_397 next page >


< previous page page_398 next page >
Page 398

Table 6.2. Repeated Calibrations Over a Two-Year Period of the Accommodation Coefficient, σ, for One Spherical Rotora
Pressure (Pa)
Date 3 × 104 3 × 103 2 × 102 3 × 101 3 × 100
Nov. 1987 1.041 1.035 1.035 1.030 0.973
July 1987 1.029 1.030 1.035 1.032 0.974
Jan. 1987 1.034 1.034 1.034 1.030 0.972
Sept. 1986 1.037 1.028 1.030 1.030 0.974
Jan. 1986 1.042 1.038 1.035 1.029 0.973
aFrom Redgrave and Downes [47].

According to Eq. (6.16), the temperature dependence in pressure determination by SRG enters in two ways: (1) through and the
temperature of the gas and (2) through thermal change of the moment of inertia of the rotor. The equilibrium temperature of the gas at low
pressure is the same as the thimble. Transient inductive heating of the ball occurs during "spinning up," causing a temperature rise of
several degrees. In the thermally isolated rotor, it requires many hours to reach thermal equilibrium [5458].

Figure 6.19 is a curve by McCulloh, Wood, and Tilford [58] showing that the measured residual drag changes for five hours after spin-up.
To avoid this problem it is advisable to suspend and spin-up the ball the day before an SRG is to be used as a reference gauge.

Although the frictional loss resulting from the use of a magnetic bearing is small, it is not zero. It results primarily from eddy currents in
the ball, as well as from those induced in the thimble by the magnetic field of the ball [5961]. It can be dependent upon the ball material
[62]. Normally, the residual drag at high vacuum corresponds to a nitrogen pressure of 104 to 103 Pa.

This residual drag should be measured every time the ball is spun-up, and occasionally during a long sequence of measurements. The
thimble must be evacuated to a pressure negligible compared to the residual drag. The indicated " pressure" at P = 0 is the residual drag.
The residual drag is then entered into the controller, where it is automatically subtracted from each pressure measurement.

The effect of vibration on SRG measurements has also been studied [54]. The suspended ball tends to remain fixed in space while
vibration causes the head with its pickup coils to move. The result is a timing error of the induced signal in the pickup coils.

Recent heads use four coils for rotation sensing as sketched in Fig. 6.20. They are not sensitive to vibration as the older head illustrated in
Fig. 6.14. However, it is beneficial to minimize vibration due to causes such as mechanical pumps, air conditioning, and walking. The
effect of vibration is to cause an apparent increase in the residual drag. But the contribution due to vibration may not be stable. The
presence of periodic vibration may be seen with an oscilloscope used to view the induced signal output.

< previous page page_398 next page >


< previous page page_399 next page >
Page 399

Fig. 6.18
Repeated argon calibrations of four rotor balls at the PTB (Berlin) over
a period of several years. The SRG thimbles were transported to the
international laboratories shown on the top line with the balls free to move
in their thimbles. Reprinted with permission from G. Messer, P. Röhl,
G. Grosse, and W. Jitschin, J. Vac. Sci. Technol. A 5, 2440 (1987) [Ref. 53].
Copyright 1987 American Vacuum Society.
6.7.1.4
Secondary or Transfer Standard

The reproducibility of measurements made with an SRG can be within 1%. Factors influencing long-term stability have
been discussed above. Tests of SRGs as transfer standards in international comparisons have verified their long-term
stability over periods of several years (see Section 12.3.2). The long times and shipping hazards make international
intercomparisons demanding tests.

< previous page page_399 next page >


< previous page page_400 next page >
Page 400

Fig. 6.19
Change in residual drag due to cooling of ball after spin-up. The
curve shows that about five hours is required for the residual drift to
approach a stability of 1 × 106 Pa nitrogen equivalent. Reprinted with
permission from K. E. McCulloh, S. D. Wood, and C. R. Tilford, J. Vac.
Sci. Technol. A 3, 1738 (1985) [Ref. 58]. Copyright 1985
American Vacuum Society.
Fig. 6.20
Four rotation sensing coils used in
improved SRG head provide reduced
sensitivity to vibration. If the ball
changes position with respect to the
head because of vibration, the output
signal of the four phased coils tends
to remain constant.

Messer et al. [53] of the Physikalisch-Technische Bundesanstalt (PTB) in Berlin reported an intercomparison with
laboratories of eight countries. They found that σ changed during shipping if balls were free to move about in their
thimbles. Even so, they concluded that calibration data for most of the participating labs were within

< previous page page_400 next page >


< previous page page_401 next page >
Page 401

± 1%. The following year, using a spring-loaded device to prevent motion during transport, another intercomparison
was made between the PTB and NIST without change of the rotor calibrations [63]. The authors suggested that
intercomparisons might be possible to an accuracy of ± 0.2%. Sharma and Mohan [64, 65] of the National Physical
Laboratory (India) also participated in these intercomparisons.

6.7.1.5
Use Precautions

Most of the precautions recommended when using an SRG are evident from the above discussion. They are collected
below:

1. To prevent changes of the accommodation coefficient of calibrated balls, the rotor surface must not be damaged by
handling, either out of the thimble or during transport while free to move within the thimble. It is especially important
that the rotor suspension not be interrupted while the ball is spinning. Damage could result from power interruptions if
the storage battery within the controller is not maintained. An uninterruptable supply is desirable

2. Install the SRG head using a level to ensure that its axis is vertical. This helps to minimize the residual drag.

3. Verify the ability to suspend the ball immediately after pump down. Some balls will not suspend and spin up
properly. It is disappointing to learn this after processing the system and it is time to make a measurement.

4. Avoid magnetic fields from external sources.

5. A stable ambient temperature is desirable. Avoid locating where sunlight can strike the head or near air-conditioning
vents.

6. Allow at least several hours for temperature equilibrium after spin-up.

7. Use only in clean systems which will not contaminate or damage the rotor surface.

8. Know the gas in the system. SRG measurements are dependent upon the gas.

9. Verify that the parameters entered into the controller are correct for the gas, rotor, and current temperature.

10. Check the residual drag frequently while the SRG is pumped to a pressure not higher than 1% of the lowest pressure
to be measured.

11. Avoid locations with significant vibration. Look for effects of vibration as evidenced by an unstable residual drag,
or as vibration frequencies superimposed on the induced signal output.

12. Experiment with rotor magnetization. It may be changed by bringing a small permanent magnet near the thimble
with the head assembly removed. An excessive magnetic moment of the ball increases the residual drag. Too little gives
a weak induced signal that does not permit accurate timing. Viewing the signal with an oscilloscope is necessary to
verify the induced signal voltage.

13. Verify occasionally that the accommodation coefficient, σ, of the rotor is nchanged. Comparison to a capacitance
diaphragm gauge at about 102 Pa may be adequate. If two SRG systems are available, they should be compared with
one another.

6.7.1.6
Advantages and Disadvantages

A careful worker can make gauge comparisons with an accuracy of about 2% over the range 104 to 103 Pa. The range
< previous page page_401 next page >
< previous page page_402 next page >
Page 402

overlaps with that of the ionization gauge, so that the calibration factor for ionization gauges may be found. The
necessary investment, although significant, is not unreasonable. The SRG is inert; it does not alter or contaminate the
gas to be measured.

Proper use of the SRG requires a suitable system, an expensive instrument, and trained personnel. The system must be
capable of pressures low enough to determine the residual drag. Determining a single pressure may require several
minutes, during which the pressure must be stable. The SRG is unsuited for following changing pressures.

In situ comparison must be carefully evaluated in each case. Process systems would often contaminate or damage the
rotor surface. The SRG is best suited for use in separate calibration systems.

6.7.2
Oscillating Quartz Crystal Viscosity Gauge

In 1959, Pacey [66] was the first to report vacuum measurements with an oscillating quartz crystal viscosity gauge. He
used as the sensor a 200-KHz DT cut crystal operating in a face shear mode. It was driven by a triode circuit. Pacey
showed calibrations for the gases argon, hydrogen, and air over the range 10 to 105 Pa. The pressure corresponding to a
given output current differed by about a factor of 100 for the gases argon and hydrogen.

The oscillating quartz crystal gauge was revived about 1985 by the vacuum group at the Electrotechnical Laboratory,
Tsukuba, Japan, and has been the subject of a number of papers. In one of their early papers, Ono, Hirata, Kokubun, and
Murakami of the Electrotechnical Laboratory, with Tamura, Hojo, Kawashima, and Kyogoku of Seiko Instruments,
published results for an instrument based upon a miniature quartz crystal tuning fork as used in wristwatches [67]. The
resonant frequency was 32,768 Hz. The circuit drove the resonator at a constant ac voltage, and it measured the
impedance of the crystal fork as a function of pressure. Calibration data extended from 7 × 101 to 1 × 106 Pa. Meter
reading as a function of pressure was linear over this interval.

Figure 6.21 reproduces a set of curves showing the output meter reading as a function of pressure for several gases
when the controller was direct reading for nitrogen. One of the disadvantages of the quartz oscillator gauge is the
approximately two-decade uncertainty in pressure when the gas is not known.

In another paper, Hirata et al. [68] explored quartz oscillators of differing sizes and operating modes and proposed a
''string of beads" model to analyze the results. Low-frequency tuning forks were best, and some showed pressure
response from 102 to 105 Pa.

Temperature sensitivity of the crystal oscillator was found to limit the lowest useful pressure for a practical gauge to 1
Pa. The shift of resonant frequency of the quartz oscillator with temperature was then used to make a self-compensated
sensor useful to 0.01 Pa [69]. In a 1995 paper given at the Thirteenth International Vacuum Congress in Yokohama,
Kobayashi et al. [70] proposed using the differing sensitivity of the oscillator gauge for different gases as a means of gas
analysis. A quartz oscillator gauge based upon this research at the Electrotechnical Laboratory is manufactured by
Vacuum Products Corp [71].

6.7.2.1
Advantages and Disadvantages

The sensor is very small, and it is inert. The effective measuring range is 101 to 105 Pa. The differing calibration for
various

< previous page page_402 next page >


< previous page page_403 next page >
Page 403

Fig. 6.21
Calibration of the quartz oscillator gauge in H2, He, N2, Ar, and Kr. The
plot shows indicated pressures when the instrument is calibrated for nitrogen.
Reprinted with permission from M. Ono, M. Hirata, K. Kokubun, H. Murakami,
F. Tamura, H. Hojo, H. Kawashima, and H. Kyogoku, J. Vac. Sci. Technol. A 3,
1746 (1985) [Ref. 67]. Copyright 1985 American Vacuum Society.

gases is a problem. Referring to Fig. 6.21, assume a gauge calibrated for nitrogen, but unknown to the operator the
system gas is hydrogen. At an indicated pressure 3 × 103 Pa, the actual pressure would be 105 Pa (one atmosphere).
Attempting to back fill to an indicated pressure greater than 3 × 103 Pa would cause an overpressure condition.

6.8
Thermal Conductivity Gauges

A hot wire in a gaseous environment loses heat (thermal energy) in three ways: (1) radiation, (2) conduction to supports,
and (3) transfer by the gas. Let the rates of energy loss be WR, WC, and WG, respectively. These energy transfer
mechanisms are illustrated in Fig. 6.22. Let WT be the sum of the rates of energy loss by these means, or

Energy transfer by the gas is pressure-dependent. It is this effect that is used to make a thermal-conductivity-based
pressure sensor. Ignoring the change of WR and WC with pressure caused by the pressure dependence of the
temperature distribution along the length of the wire, WR + WC, establishes a constant background loss. The magnitude
and stability of this background determine the lowest useful pressure of the gauge.

Two types of thermal conductivity gauge are in general use. In the Pirani gauge the temperature of the hot wire is found
from its resistance. In the thermocouple gauge,

< previous page page_403 next page >


< previous page page_404 next page >
Page 404

Fig. 6.22
Energy loss mechanisms for a heated
wire in a gas at reduced pressure.

a thermo-junction of dissimilar metals provides a temperature-dependent output voltage. Descriptions of both gauges
were first published in 1906. Manfred von Pirani was the first to describe the gauge based on wire resistance [72]. This
was followed in a few months by Voege's paper on the thermocouple gauge [73].

In the introduction to his paper, von Pirani explains why he undertook the development of a new gauge. It is worth
repeating, since his reasons are as valid today as in 1906. He was employed in the incandescent lamp factory of Siemens
and Halske. He writes [74]: "I was assigned the task of developing a simple, inexpensive, vacuum measuring instrument
which could replace the McLeod gauge. The project had not only the technical goal of making possible rapid
recognition of small pressure changes at high vacuum, but it also had significant importance from the health standpoint.
As is well known, there is an increasing effort to remove harmful mercury from the workplace."

Voege [73] was constructing thermocouple based ac measuring instruments when he noticed that the output voltage of a
junction at a constant ac input was pressure dependent. Both the Pirani and thermocouple gauges are in widespread use.
Perhaps there are more of them than any other vacuum gauge. Figure 6.23 illustrates a simple form of each of these
gauges.

6.8.1
Theory

From a fundamental standpoint, Pirani and thermocouple gauges differ only in the means of observing the wire
temperature. The discussion of the three modes of energy transfer given below applies equally to both. Although a
sensor based upon a hot wire is assumed, it is clear that the reasoning could be extended to other geometries where the
sensor element was in the form of a disc or a spherical bead. For simplicity, the equations are written for an element of
wire of length dl at position l with temperature Tl. This makes it possible to examine the functional dependences while
neglecting the actual temperature distribution.

< previous page page_404 next page >


< previous page page_405 next page >
Page 405

Fig. 6.23
Construction of a Pirani gauge
(a) and thermocouple gauges (b).

To examine the loss by radiation, consider an incremental length dl of a long wire of diameter r1, emissivity ε1, and
temperature T1 centrally located in a long cylinder of radius r2, emissivity ε2, and temperature T2. Assuming T1 > T2,
the rate of energy transfer by radiation, WR, is [75]

where σ is the StefanBoltzmann constant 5.673 × 108 W·m2·K4. In the limiting case where Eq. (6.31)
becomes

The emissivity of a clean bright metal surface such as platinum is about 0.05, while for a surface coated with soot it can
approach unity. In designing a gauge it is desirable to choose a sensor wire with a low and stable emissivity. Low
thermal conductivity and a large temperature coefficient of resistance are also important. Commonly used materials
include platinum, nickel, and tungsten.

Heat transfer by conduction from the wire to the cooler support is also shown in Fig. 6.22. Assuming the same situation
at each end, the local rate of energy transfer by conduction along the wire, ½Wc(l), for an element dl at position l is
where G is the thermal conductivity of the material of the wire, and dTl/dl is the temperature gradient. Roberts [76]
derived an equation for the temperature distribution along an electrically heated wire in vacuum. The temperature
distribution

< previous page page_405 next page >


< previous page page_406 next page >
Page 406

changes somewhat with pressure as heat transfer by the gas becomes the dominant loss mechanisms. For many purposes
an adequate model is a temperature distribution constant over the central region, with a linear decrease at the ends. The
fraction of the heat loss due to conduction can be reduced by using a long wire of small diameter.

Heat transfer by gas molecules is simple to analyze for the case when the mean free path is greater than r2 r1. Gas
molecules arriving at the hot wire will have a Maxwellian energy distribution corresponding to T2. These molecules
usually dwell on the surface for a short time and depart with an energy distribution corresponding to T1. An
accommodation coefficient, α, is defined as the probability of this process.

Kennard's text [77] on the kinetic theory of gases has an equation for the heat transfer between coaxial cylinders for the
case where the molecular mean free path is long. Specializing to the present case where his result becomes

In this equation, γ is the ratio of specific heats of the gas, Cp/Cv, m is the mass of a gas molecule in kilograms, and k is
the Boltzmann constant, 1.38 × 1023 W·s·K1. The numerical value of γ is dependent upon the number of degrees of
freedom of the molecule. The classical value of γ for a point mass is 5/3, for a rigid dumbbell 7/5, and for a diatomic
molecule with vibration 9/7. Values of γ according to both classical theory and experiment are tabulated in Kennard
[78]. Substituting numerical values for the constants and letting M be the molecular weight gives the result

The region of linear behavior of WG with P extends to about 10 Pa for nitrogen. At higher pressures, energetic
molecules departing from the wire collide with others before getting far from the wire, and they deposit their energy to
form a sheath of hot gas near the wire. This prevents further effective heat transfer until the onset of convection.

6.8.2
Calibration

Figure 6.24 is a plot of data from the calibration of a constant temperature Pirani gauge with the gases He, N2, and Ar.
The plot extends to atmospheric pressure for the two heavier gases. The sensor wire was 150 mm of 0.0254-mm-
diameter platinum wire. A control circuit similar to that used in taking the data is shown in Fig. 6.25. The bridge
voltage, Vbr, used as the pressure-dependent output signal plotted on the ordinate in Fig. 6.24. If R1 = R2 in the bridge
circuit of Fig. 6.25, the sensor resistance, Rs, is equal to Rc when the bridge is balanced. Rc is chosen to provide the
desired operating temperature of the sensor wire using known temperature versus resistance data. One-half of Vbr is
applied to the sensor when R1 = R2. The total power supplied to the sensor wire is then This must equal WT
of Eq. (6.30).

< previous page page_406 next page >


< previous page page_407 next page >
Page 407

Fig. 6.24
Experimental calibration curves in helium, nitrogen, and argon for a
constant-temperature Pirani gauge.

Fig. 6.25
Bridge amplifier for constant
temperature operation of a Pirani gauge.

As an example, the experimental sum of WR and WC may be found for the data of Fig. 6.24. The bridge voltage is
0.3907 V as the pressure approaches zero. The value of Rc (and therefore Rs) was 43.61 Ω. Then
The experimental values for the rate of energy transfer by the gas may be found by subtracting this background from the
power supplied to the sensor for each data point. The resulting values of WG are plotted against pressure in Fig. 6.26.
The plot is

< previous page page_407 next page >


< previous page page_408 next page >
Page 408

Fig. 6.26
Energy transported by the gas in a Pirani gauge. A plot of WG versus
pressure for the nitrogen data of Fig. 6.24.

a straight line until the long mean free path assumption fails above 10 Pa. The value of WR + WC from Eq. (5.36) is also
plotted on Fig. 6.26.

It is interesting to calculate a point according to Eq. (6.35) and compare it to the experimental curve for WG plotted as
Fig. 6.26. Assume the same gauge parameters, molecular weight of 28 for nitrogen, a wire temperature of 400 K, and an
envelope temperature of 300 K. Let α and ¼[(γ + 1)/(γ 1)] both be unity. Then, for P = 0.13 Pa the result is WG = 1.26 ×
104 W, in good agreement with the experimental value of 1.3 × 104 W from Fig. 6.26. The fact that the background WR
+ WC is almost an order of magnitude greater than WG at this pressure will be used below in discussing the low
pressure limit.

6.8.3
Lowest Useful Pressure

The radiation component of the background, given by Eq. (6.31), may not be stable. The emissivity may vary from 0.05
for a clean wire to unity when contaminated. Choice of wire material and temperature is important for stability of the
emissivity. The stability of the background in actual use, and therefore the lower useful pressure limit, is largely a
matter of experience. It is reasonable to use a lower pressure limit for the thermal conductivity gauge determined by the
condition
For the gauge used in Fig. 6.26, which was optimized for low-pressure performance, it as found above that this occurs at
about 0.1 Pa. The fact that the gas term can be

< previous page page_408 next page >


< previous page page_409 next page >
Page 409

Table 6.3. Comparison of Accommodation Coefficients as Measured by Dickins [80] and by


Thomas and Olmer [81]
Gas Dickens Thomas and Olmer
H2 0.35 0.22
He 0.51 0.24
CO 0.91 0.75
O2 0.90 0.74
N2 0.90
A 0.88 0.89
CO2 0.92 0.76

followed below 103 Pa in the laboratory under ideal short-term conditions is of little practical value. Ellett and Zabel
[79] decreased the importance of radiation in a Pirani by cooling the envelope to 90 K and using a moderate sensor wire
temperature. Measurements to 105 Pa were possible. However, there is no inexpensive and energy efficient way at the
present time to cool a gauge for continuous operation.

The accommodation coefficient, α, in Eq. (6.35), is a function of the gas, the metal surface, contamination, and the
temperature. Table 6.3 compares experimental values of α by Dickins [80] and Thomas and Olmer [81] for several
gases on platinum at about 20°C.

6.8.4
Constant Temperature Pirani

The benefit of operating the sensor wire at constant temperature is an extended high pressure range. If, in making a
calibration similar to that of Fig. 6.24, the bridge voltage had been fixed at 0.3907 V, and the bridge-out-of balance
voltage used as the signal, the resulting nitrogen curve would have negligible slope above 100 Pa. With constant input it
is possible to choose a wire temperature to optimize performance for a limited pressure range. Leck [82] investigated
this for a Pirani, while Teledyne-Brown Hastings Engineering [83] offers a number of thermocouple gauge sensors,
each optimized for a different pressure range. With this approach the total useful range of each is between one and two
pressure decades less than for a thermal conductivity gauge operated at a constant temperature.

Before electronic feedback control circuits became feasible, operating a Pirani sensor wire at constant temperature
required manual adjustment at each pressure. Von Ubisch [84] published the first feedback circuit for constant
temperature operation in 1948. Today most commercial Pirani control circuits hold the wire at constant temperature,
although many low-cost thermocouple gauge controllers continue to have constant voltage circuits. The constant
temperature mode is possible with a thermocouple gauge, although it has been used less. Zettler and Sud [85] described
a control circuit for a thermocuple gauge which programmed the junction temperature for better results in each pressure
region.

Another advantage of the constant temperature mode is the faster response to pressure transients. Since the wire
temperature is constant within the error of the

< previous page page_409 next page >


< previous page page_410 next page >
Page 410

feedback circuit, the response time is largely a function of the amplifier gain and any limitations on high-frequency
response of the circuit that may be needed for high-frequency stability.

6.8.5
Calibration Dependence Upon the Gas

Equation (6.35) for WG has explicit M1/2 dependence upon the molecular weight, and γ is also a function of the
molecule. Thus the calibration of a thermal conductivity gauge is dependent upon the gas. The curves in Fig. 6.24 are
typical examples. The departure of the curves from one another is particularly serious at the higher pressures. With
nitrogen the output is insensitive to pressure change above 104 Pa. With argon, it would never reach an indicated
atmosphere, even with a large overpressure. With helium, the gauge would indicate a pressure greater than atmospheric
for any pressure above 300 Pa.

The differing sensitivity of the thermal conductivity gauge to various gases can be used to make a simple leak detector.
Blears and Leck [86] analyzed the Pirani leak detector in terms of the minimum detectable pressure change.
Steckelmacher and Tinsley [87] reported the construction of a "sniffer" Pirani leak detector which gave one-tenth FSD
(full-scale deflection) on the most sensitive range with hydrogen search gas and a leak rate of 5 × 104 std cm3/min.
Sharma and Gupta [88], Minter [89], and Aleksandrovich, Sokovishin, and Sazanov [90] also published variations of
the Pirani leak detector.

Today, commercial Pirani controllers often incorporate leak detector circuits that provide a large leak signal when the
pressure change is too small to see. Depending upon the pumping system, leaks as small as 106 std cm3/s may be
detectable.

6.8.6
Upper Pressure Limit

At higher pressures where the mean free path becomes comparable to the diameter of the hot wire, molecules departing
from the wire lose their thermal energy within a few wire diameters. This forms a hot sheath of gas that inhibits further
heat transport. There are at least two means of increasing high pressure gas heat transfer described in the literature: (1)
move the wire or (2) allow convection to move the gas. The moving wire technique was described by Birshert [91]. His
experimental curve shows good sensitivity extending to one atmosphere without the inflection near 104 Pa common to
convection gauges.

The inside diameter of the envelope of the Pirani sensor used in recording the data plotted in Fig. 6.24 was intentionally
small (9.7 mm) to inhibit convection. When convection is encouraged by making the tube diameter larger and mounting
it horizontally, a calibration as in Fig. 6.27 results. With nitrogen, convection starts near a pressure of 5 × 104 Pa.
Johnson [92] was the first to use convection to extend the range of an experimental thermocouple gauge. The large
change of output with position caused him to suggest possible application of the gauge as an inclinometer. McMillan
and Buch [93] described a convection-enhanced Pirani gauge the following year. These papers were ignored for 20
years before commercial convection enhanced Piranis appeared.

Just as with a normal Pirani, the calibrations of a convection enhanced Pirani for gases of differing molecular weight
differ widely at higher pressures. It is important

< previous page page_410 next page >


< previous page page_411 next page >
Page 411

Fig. 6.27
Calibration of a convection enhanced Pirani gauge with nitrogen.

that the gas be known if a convection gauge is to be used at higher pressures. This is discussed in Section 6.8.11.

The reader interested in the theory of heat transfer by convection is referred to texts on heat transfer, such as those by
Kreith [94] or Eckert [95].

6.8.7
Ambient Temperature Compensation

Equation (6.35) for WG contains explicit dependence upon the envelope temperature, T2, of the form

. The accommodation coefficient and the ratio of specific heats are also temperature-dependent.
Thus stable operation of a thermal conductivity gauge requires either temperature control or temperature compensation.

Two types of temperature compensation are used for Pirani gauges. An early method, found in Pirani's original paper
[72], uses an identical sealed-off gauge, or temperature-sensitive coil of wire, as Rc. The intent is to keep the difference
temperature (T1 T2) constant. In 1952, von Dardel [96] published a complete circuit for a controller using a
compensating coil wound on the outside of the sensor tube. With proper choice of compensating coil wire and its length,
good compensation is possible over the normal ambient range.

In the other approach [97], all component values within the bridge remain fixed. A network of resistors containing a
thermistor in contact with the gauge envelope is used as a variable gain voltage divider operating on the bridge output
voltage. The design of the network requires temperature behavior data for the gauge and the thermistor. The resulting
compensation is effective for the ambient temperature range of 1050°C.
< previous page page_411 next page >
< previous page page_412 next page >
Page 412

6.8.8
Comparison of Pirani and Thermocouple Gauges

More depends upon whether a particular gauge operates in the constant temperature or constant input mode than upon
the way the wire temperature is measured. However, one fundamental difference is the sensor output signal level. The
output of a thermocouple gauge is a few millivolts, while the bridge voltage output signal of a Pirani with feedback is
several volts. Noise immunity near sources of electrical noise, as well as compatibility with long leads, may differ
because of this. The bridge can be located in the head assembly of a Pirani; and by using potential leads, the voltage
reaching the controller can be made independent of cable length. Pirani gauges have operated well with 150-m cables.
The Pirani systems are usually more expensive, better corrected for ambient temperature, and more accurate than most
thermocouple based systems.

Leak detector capability is a feature included in some Pirani controllers.

6.8.9
Stability

Only limited information exists in the literature about the accuracy or repeatability of measurements made with thermal
conductivity gauges. This may be because the usual applications in monitoring fore and roughing pressures do not
require high accuracy. Poulter, Rodgers, and Ashcroft [98] undertook a six month's evaluation of several gauge heads
from one manufacturer. The maximum change in the mean nitrogen calibrations did not exceed 1.6% for any gauge.
They concluded that pressure measurements made with thermal conductivity gauges should have a total uncertainty not
exceeding ± 4.4%. Jitschin and Ruschitzka [99] found accuracy to be a function of the pressure, and they attributed
much of the drift at low pressure to change of emissivity. The power loss at zero pressure drifted several percent over a
few months.

6.8.10
Thermistor Pirani Gauges and Integrated Transducers

The preceding discussion has assumed sensors using metallic wires. The large temperature coefficient of thermistors
made with semiconducting materials suggests their use in place of wire. Becker, Green, and Pearson [100] were the first
to build and test a gauge using a thermistor. Used in a bridge, with the out of balance voltage as the output, they found
the output to be a nearly linear function of pressure from 102 to 100 Pa. Varicak * and Saftic* [101] mounted
thermistors on discs of tin foil 30 mm in diameter by 0.01 mm thick to obtain a large surface with low emissivity, and
they supported the discs on 0.05-mm manganin wire. Their circuit covered 104 to 100 Pa in three ranges.

Shioyama et al. [102] made a Pirani using semiconducting thin films of TaN. They were able to calibrate their gauge
from 102 Pa to atmosphere.

These papers using high-temperature-coefficient sensors appear to describe gauges with improved low pressure
capability as compared to those using metallic wires. However, gauges using thermistors are seldom found in
commercial instruments designed since 1975.

Integrated silicon microtransducers are available and will increasingly compete with the hot wire types in the future. M.
Esashi [103] of Tohoku University wrote a review paper describing silicon micromachining of integrated pressure
transducers.

< previous page page_412 next page >


< previous page page_413 next page >
Page 413

Teledyne-Brown Hastings Engineering [104] manufactures an instrument using a microelectronic sensor. The specified
range is 103 to 105 Pa. Van Herwaarden and Sarro [105] reported a silicon integrated thermal sensor 6 × 6 mm with
good sensitivity from 0.1 to 100 Pa. Weng and Shie [106] using a transducer with a largest dimension of only about 150
µm were able to measure from 102 to 8 × 102 Pa in the constant temperature mode.

6.8.11
Commercial Gauges and Applications

The above discussion covers most of the properties of these gauges that determine whether an application is appropriate.
However, there is another important characteristic of the thermal conductivity gauge. Their output is a single-valued
function of pressure over the range from high vacuum to atmosphere. This fact, combined with a response time of a few
milliseconds, makes them ideally suited for protective functions, as in determining when hot-cathode ionization gauges
should be activated. Their range makes them the usual choice for measuring backing and roughing pressures. Because
their calibration is highly gas-dependent, they are not well suited for most backfilling applications. They are not
recommended for use in contaminating environments because of their sensitivity to surface conditions.

Controller/gauge systems are available from numerous manufacturers. Figure 6.28 illustrates a controller with range 101
to 104 Pa, leak detector capability, LCD display, and RS232 digital output option [97]. ''Transducer" units comprising
sensor and a simple controller as part of the head are popular. The user need only supply them with low-voltage dc to
obtain a logarithmic dc output signal.

A hazard can arise when using a thermal conductivity gauge because of the multidecade difference in calibrations for
different gases at higher pressures. Consider

Fig. 6.28
Photograph of HPS Series 315 Pirani gauge controller
and gauge head. Photograph courtesy of MKS
Instruments, Inc. HPS Division, 5330 Sterling
Drive, Boulder, Colorado, 80301.
< previous page page_413 next page >
< previous page page_414 next page >
Page 414

a gauge calibrated for nitrogen. Then, according to Fig. 6.24, the gauge would indicate an atmosphere at an actual
pressure of 300 Pa when used in helium or other light gas. With a heavy gas such as argon, the gauge will not indicate a
pressure greater than a few hundred pascals, even for several atmospheres overpressure. The first situation can lead to
implosion, whereas the second one can lead to explosion. Peacock [5] discusses an accident of each type. To avoid
accidents, the following precautions are recommended:

1. Include an inexpensive gauge that is not gas-dependent, such as a Bourdon gauge, in applications where a large
measurement error would be hazardous.

2. Provide overpressure protection by a relief valve or burst disc on any system connected to a source of pressurized gas.

3. If the principal use of a gauge is to be at pressures greater than 100 Pa, it is advisable to use a gauge whose output
signal is not gas-dependent.

Thermal conductivity gauges using platinum or platinum alloy wires should not be used with gases capable of forming
explosive mixtures with air. Fine wires of these metals heat autocatalytically in explosive gas mixtures and may ignite
them. It is not necessary that the gauge be turned on. Any gauge made with any wire could initiate an explosion if the
circuit failed and overheated the sensor wire.

6.9
Ionization Gauges

Virtually every high-vacuum system uses some form of ionization gauge for pressure measurement below 102 or 103
Pa. Below 104 Pa there are no realistic alternates. Ionization gauges are used to the lowest attainable pressures around
1012 Pa.

Although Buckley [107] has long been credited with the invention of the hot cathode ionization gauge (HCG), Redhead
[108] has pointed out the prior work of von Baeyer [109]. Until 1950, HCGs were of normal triode design. Many HCGs
were simply standard triode vacuum tubes with a tubulation. A major improvement took place in 1950 when Bayard and
Alpert [110] revealed a new electrode configuration allowing pressures measurements two to three decades lower than
were possible with the old standard triode geometry.

Another frequently encountered ionization gauge is the crossed field magnetic discharge gauge, which was invented by
Penning [111]. It is often called the coldcathode gauge (CCG). The electron trapping was improved by Beck and
Brisbane [112] using a central wire anode, a concentric cylindrical cathode, and an axial magnetic field. This is called
an "inverted magnetron" arrangement. Hobson and Redhead [113] developed it into a practical gauge usable from 102
to 1010 Pa. It has come to occupy an important position among ionization gauges because of its relative freedom from
outgassing, its ruggedness, and wide range. Cold cathode ion gauges are discussed in Section 6.9.9.

6.9.1
Hot-Cathode Gauge Equation

Figure 6.29 represents a generalized ionization gauge. Ionizing particles (arrows) arrive from the left. They are usually
electrons from a thermionic source with an

< previous page page_414 next page >


< previous page page_415 next page >
Page 415

Fig. 6.29
Generalized ionization gauge.

energy of 100180 eV. However, α or β particles from nuclear decay have been used, and laser excitation is a subject of
current research. The electrons enter the ionizing space containing gas molecules (circles) at reduced pressure. Inelastic
collisions of the electrons with gas molecules produce ion/electron pairs (shown as circles with + or ). Ions are collected
by the suitably biased lower electrode. The electrometer in the external circuit measures the ion current as an indirect
measure of gas pressure. Figure 7.2, page 450, is a plot of ionization probability versus electron energy for several gases.

The number of ions formed, and therefore the current in the circuit, is a function of the number of gas molecules per
unit volume, the ionization cross section energy, arrival rate, and path length of the electrons.

The ionization gauge equation provides the relationship of these quantities to one another. To derive it, let σi be the
total ionization cross section for a gas molecule, L the length of the ionizing space, and A the cross-sectional area of the
electron beam. The number of molecules in this volume is nLA, where the number density, n, is related to the gas
pressure by n = P/kT. The projected area, Aσ, of the gas molecules within this volume is nLAσi = σiLAP/kT. The
fraction of the incoming electrons which participate in ionizing collisions will be Aσ/A. Let N be the number of
electrons arriving per unit time, each of charge q. The number of ionizing collisions per unit time is σiLP/kTN. Or,
converting to current, the ion current, i+, is

The incident electron current is Nq = i, and σiL/kT can be defined as the gauge constant, K. Then Eq. (6.38) becomes the
usual ionization gauge equation

< previous page page_415 next page >


< previous page page_416 next page >
Page 416

It is preferable to use the term "gauge constant" or "gauge coefficient" for K. Then "sensitivity" (symbolized by S) can
be reserved for the product K·i, which is also an important parameter of an ionization gauge.

It is known from experiments that the collector current, ic, as measured by the electrometer is the sum of i+ from Eq.
(6.38) and a residual current, ir, which may have contributions from different sources. A more general gauge equation is
then

The gauge constant, K, as defined above, is a function of the gas, the geometry of the gauge, the electron energy, and
the absolute temperature of the gas. These dependencies are inherent in ionization gauge measurements.

Until 1950 the ionization gauge was similar to the triode vacuum tube with its thermionic electron source, fine wire grid
surrounding the cathode, and outside that an electrode of sheet metal (anode or plate in vacuum tube terminology). A
good description of a 1930 triode gauge was given by Jaycox and Weinhart [114]. Examination of their data suggests
that they reached pressures at least a decade lower than their triode gauge could measure, or about 107 Pa.

A schematic circuit for use with a HCG is illustrated in Fig. 6.30. It is drawn with a triode gauge. A grid voltage of 180
V, and cathode bias of 30 V are typical. The effective energy of electrons in the ionization space will range between
zero and 150 V. The electrons have enough energy for ionization over a portion of their path in the space between grid
and ion collector.

In the late 1930s it was general knowledge that the triode gauge never indicated a pressure lower than about 106 Pa.
This was hard to understand since vacuum techniques had improved a great deal, and surface contamination times
extended to hours or even days. Anderson [115], in his studies of contact potentials, had good evidence that the true
pressures were much lower. Apparently the collector current consisted of a pressure-dependent ion current and a
residual current as anticipated in Eq. (6.40). In 1938, Bell, Davies, and Gossling [116] correctly recognized that a similar

Fig. 6.30
Basic circuit for operating a hot-cathode
ionization gauge.

< previous page page_416 next page >


< previous page page_417 next page >
Page 417

current in vacuum tubes was caused by soft x-rays. Perhaps because of World War II, their paper went unnoticed. At the
Seventh Physical Electronics Conference at the Massachusetts Institute of Technology in 1947, Nottingham [117]
suggested independently that the residual current is the result of a two-step process: (1) Electrons striking the grid
produce soft x-rays, and (2) these x-ray photons then cause the emission of photoelectrons from the surrounding ion
collector. To the electrometer this current is indistinguishable from an ion current. This was discussed again at the 1948
meeting. In the summer of 1948, D. Alpert of Westinghouse Research Laboratories, who had attended these
conferences, asked Robert Bayard to investigate the validity of the x-ray hypothesis [118]. The immediate result was the
invention of the BayardAlpert gauge (BAG), illustrated in Fig. 6.31, an ingeniously simple solution to the problem.
Whereas in the triode gauge the ion collector was a metal cylinder surrounding the grid and cathode, in the BAG the
grid was cylindrical, the cathode was external to the grid, and the ion collector was a fine wire centered in the grid.
Operating voltages were similar to those of the triode gauge. The cross section for interception of the x-rays was
reduced 1001000 times. The x-ray limit (the pressure where the x-ray current equals the ion current) decreased to about
5 × 109 Pa. Bayard and Alpert [110] compared the x-ray limits of the new gauge and the old
Fig. 6.31
BayardAlpert ionization
gauge. The figure shows a
gauge as made at the
Westinghouse Research
Laboratories in the early
1950s. The grid is
molybdenum and is intended
for electron bombardment
degassing. Reprinted with
permission from D. Alpert, in
Handbuch der Physik
(S. Flügge, ed.), Vol. 12, p. 609.
Springer-Verlag, Berlin, 1957
[Ref. 125]. Copyright
1957 Springer-Verlag.

< previous page page_417 next page >


< previous page page_418 next page >
Page 418

triode using plots of collector current versus grid voltage made at pressures as low as the 109 Pa decade.

A production BAG was sold by Westinghouse as the WL 5966. It had a Nonex glass envelope, and a molybdenum grid
intended for electron bombardment degassing. The grid was 19 mm in diameter by 38 mm long. There were two
tungsten filaments. The ion collector was made of 0.13-mm tungsten wire. The gauge constant of the Alpert BAG was
about 0.09/Pa for nitrogen, and the x-ray limit was about 6 × 109 Pa. Its simplicity made it easy to degas and
inexpensive to manufacture. Figure 6.32 shows a nude BAG for use in metal systems.

A period of rapid exploration of variations of the BAG followed, although it was difficult to improve upon the original.
However, it will be helpful to discuss two additional contributions to the residual current before beginning the subject of
BAG modifications and other new gauges. BAG users noted that the residual current changed with the presence of
absorbed active gases on the grid. It was the custom in Alpert's laboratory in the mid-1950s to operate BAGs at 10-mA
emission current to keep the grid clean. When the grid current of a gauge operating at a pressure below 1010 was
switched from 10 to 1 mA the initial pressure reading was the same, but

Fig. 6.32
Nude BayardAlpert gauge, for use in metal systems.
< previous page page_418 next page >
< previous page page_419 next page >
Page 419

Fig. 6.33
Section through a BayardAlpert gauge illustrating the origin of the
"normal" and the "reverse" x-ray effects. Reprinted with permission from
W. H. Hayward, R. L. Jepsen, and P. A. Redhead, in 10th Natl. Vac.
Symp. 228 (1963) [Ref. 123]. Copyright 1963 American Vacuum Society.

a slow upward drift started, eventually stopping at a new value. Moore [119] investigated the ion emission from
surfaces of Mo and W covered with adsorbed CO and bombarded by electrons. The ion yield from the surface was
50100 times that from the volume. This surface ionization process is often called electron-stimulated desorption (ESD).
Ackley, Lathrop, and Wheeler [120], examining the effects of different grid currents on the collector current, proposed
that an ESD ion current as found by Moore was one possible explanation of their results. Redhead [121], using a
modulated BAG, then showed that ESD ions can make a large contribution to the residual current. With a clean grid the
residual current was 6 × 1012 A. However, adsorption of a monolayer of oxygen caused the residual current to increase
by 400 times. He also noted that the modulated BAG measured true pressure in the presence of absorbed layers. Later,
Huber and Rettinghaus [122] analyzed ESD ions using a quadrupole analyzer, and they gave ion yields for several
adsorbed gases from Pt, PtIr, and Mo surfaces. It was proposed that Pt was a desirable construction material for
ionization gauges.

Another component of the residual current was explained by Hayward, Jepsen, and Redhead [123]. It is called the
reverse x-ray effect. Both the normal and reverse x-ray effects are shown in Fig. 6.33. X-ray photons striking the interior
wall of the gauge can free electrons. Since both wall and collector may be at ground potential, some of these may reach
the collector, causing a current of sign opposite to the normal x-ray current. Thus the name reverse x-ray effect.

6.9.2
Geometric Variations in the BayardAlpert Gauge

In conceiving new gauges intended to function at lower pressures, it is necessary to improve the ratio of i+ to ir. Either
increasing the gauge constant K or decreasing ir is

< previous page page_419 next page >


< previous page page_420 next page >
Page 420

helpful. The original Westinghouse BAG had a grid in the form of an open cylinder. Electrons orbiting the collector and
having an axial velocity component could escape through the open ends. Nottingham [124] found that the gauge
constant could be increased by about a factor of two by adding grid structures to close the ends of the grid. He also used
a shield grid enclosing the other electrodes to prevent charges on the glass walls from influencing the electron
trajectories.

Alpert had tried earlier to reduce the x-ray current with smaller ion collectors made by electrochemically etching the
collector wires to a taper [125]. However, the grid ends were open in his experimental gauges. The smaller collector
combined with the open grid decreased the probability of ion capture. The x-ray current and the ion current both
decreased, so there was little improvement in the x-ray limit.

At Philips Laboratories, van Oostrom [126] made BAGs with closed grid ends and collectors as small as 4 µm. The
gauge constant with the 4-µm collector, although reduced, was 0.03/Pa for nitrogen, and the x-ray limit was estimated to
be 2.1 × 109 Pa. However, with this small collector the grid and collector biases had to be considerably higher than
usual.

A longer ionizing path length and larger K would be expected with increased grid diameter. It is less evident that the
filament to grid separation is very important in determining the gauge constant. Redhead [127] and Pittaway [128]
examined electron path length within the grid as a function of launch angle. Near-diametric entry gives deeper
penetration and longer paths. The divergence of entering electrons can be decreased with greater filament to grid
spacing. However, the separation cannot be greater than a few millimeters because the cathode must be emission-
limited with moderate cathode-to-grid potential differences. Redhead found that an auxiliary electrode behind the
cathode served the same purpose and increased the gauge constant. One interesting item in Pittaway's paper is his
calculation of the number of passes of an electron through the grid as a function of grid transparency. For 90%
transparency, an electron will average four passes.

Ohsako [129] also designed a BAG using an auxiliary electrode behind the cathode at ground potential. His purpose was
to extend the linear range to higher pressures. With a small grid of 12-mm diameter, an emission current of 1 × 105 A
gave good linearity to 27 Pa for nitrogen or argon, while retaining a sensitivity of 0.03/Pa.

Comsa [130, 131], through the calculations and experiments in these and other papers, contributed to a better
understanding of the factors influencing ion collection efficiency in the BAG. In the 1972 paper he calculated the
collection efficiency as a function of collector diameter. The fall-off is extremely rapid as the collector diameter
decreases to 0.1 mm.

A large body of literature exists on BAGs with collector wires just entering, or perhaps not quite entering, the grid.
These are often called "buried collector" gauges. The intent is to reduce the cross section of the collector to x-rays,
while maintaining ion collection efficiency. The success in achieving these aims varies. Groszkowski's 1966 paper
[132] is one of many that he published on this subject. He was able to maintain K = 0.15/Pa for nitrogen with the "x-ray
limit an order of magnitude lower than for a BAG". Melfi [133] describes similar results, and Beitel and Gosselin [134]
describe somewhat better results. Fletcher [135] found a lower-gauge constant of 0.075/Pa for argon with a residual
current of 2 × 1012 A. Watanabe [136] described a "point collector gauge" shown in Fig. 6.34 that belongs in this group
by its design. It used a spherical grid 26 mm in diameter, and the collector was 0.03 mm in diameter by

< previous page page_420 next page >


< previous page page_421 next page >
Page 421

Fig. 6.34
Watanabe point collector gauge. The operating
potentials are: filament, +200 V; grid, +310 V;
shield, +200 V; modulator, +310 V, 0 V. Reprinted
with permission from F. Watanabe, J. Vac.
Sci. Technol. A 5, 242 (1987) [Ref. 136].
Copyright 1987 American Vacuum Society.

only 0.05 mm long. The gauge constant was 0.4/Pa, and the calculated x-ray limit was 2 × 1012 Pa. This gauge is
difficult to fabricate because of the precision required.

6.9.3
Modulated BayardAlpert Gauge

Redhead [137] was the first to use modulation to overcome the residual current problem. It can be used with many
HCGs, but it is especially associated with the BAG. It requires only a very simple modification of a BAG: the insertion
of a wire probe, similar to the collector, within the grid space. The use of modulation to determine the residual current
and true ion current is explained in Section 11.7. When using a modulated BAG, the ion current is not masked by ESD
current. In this paper, Redhead describes switching the potential of the modulator electrode between grid and ground
potentials. This is known as Mode 1 modulation. Lange and Singleton [138] describe problems encountered with Mode
1, including modulation of ir. An assumption of the modulation method is that the x-ray photocurrent is not changed by
altering the modulator potential. The authors reported better results switching the modulator potential from grid to a
value corresponding to maximum ion current collection, or about 20 V below grid potential (Mode 2). Lange and
Singleton state that Mode 2 operation was possible only with open grid ends. Redhead [139] defines five modes of
modulation.

Perhaps the highest development of the modulated BAG is that of Benvenuti and Hauer [140]. Several hundred of these
gauges were installed at CERN, Geneva to monitor XHV pressures in the ISR (intersecting storage rings). This gauge,
manufactured by Thompson/CSF, had a grid volume that was larger than usual: 35 mm diameter × 40 mm long. With a
50-µm collector, ir was 4 × 1013 A, the gauge
< previous page page_421 next page >
< previous page page_422 next page >
Page 422

constant was 0.32/Pa, and x-ray limit was 1.3 × 1010 Pa. This paper also reports measurement of the background due to
evaporation of the tungsten filament, as well as selection of operating parameters.

6.9.4
Extractor Gauge

The extractor gauge by Redhead [141] is shown in Fig. 6.35. It is probably the most widely used XHV gauge. Its
construction is relatively simple, which facilitates degassing. The extractor gauge discriminates against ESD ions. ESD
ions leave the grid with about 6 eV energy, while gas phase ions originate within the grid in a region here the potential
is depressed by the electron space charge. ESD ions are unlikely to reach the ion collector. Because of the location of
the ion collector the x-ray effect is inherently small, but the extractor gauge can also be modulated. Redhead reports a
gauge constant of 0.1/Pa for nitrogen, with the gauge useful to 7 × 1011 Pa. The commercial extractor gauge Type IE-
511 from Leybold Inficon, Inc. [142] has a gauge constant [143] of 0.067/Pa for nitrogen and an x-ray limit below 1010
Pa equivalent. There are many papers concerning variations of the extractor gauge. One interesting one is that of Fujii,
Inoue, and Kanematsu [144].
Fig. 6.35
Extractor gauge. The operating potentials are:
filament, +200 V; grid, +305 V; ion collector, 0 V;
modulator, +305 V, 0 V. Reprinted with permission
from P. A. Redhead, J. Vac. Sci. Technol. 3, 173
(1966) [Ref. 141]. Copyright 1966 American
Vacuum Society.

< previous page page_422 next page >


< previous page page_423 next page >
Page 423

6.9.5
Helmer Gauge

The gauge described by Helmer and Hayward [145] and illustrated in Fig. 6.36 was manufactured by Varian, although
very few were made. It is important both because of its use as an XHV reference gauge and because it has been the
basis of other XHV gauges with good characteristics. There is the possibility that an energy analyzer gauge may become
a commonly used UHV/XHV reference gauge in the future.

In the Helmer gauge the ions are extracted from the end of a cylindrical grid and introduced into a 90° electrostatic
energy analyzer. Since, as stated above, the energy of ESD ions differs from that of gas phase ions, the analyzer
provides a means of distinguishing them. The energy resolution of the original Helmer gauge is limited, and full
separation of gas phase and ESD ions is not possible. Helmer and Hayward gave the gauge constant as 0.13/Pa, the
emission 6 mA, and the residual current 1.5 × 1014 A without suppressor and 1.5 × 1015 A with it. This corresponds to
an x-ray limit of 2 × 1011 Pa.

Han et al. [146] also studied the performance of the standard Varian Helmer gauge. They found a smaller gauge
constant of 0.067/Pa for argon and found an x-ray limit of 2 ± 2 × 1011 Pa nitrogen equivalent. This is reasonable
agreement since the gauge constant of a Helmer gauge will depend upon the entrance and exit apertures of the energy
analyzer.

Benvenuti and Hauer [147] made some modifications to a Varian Helmer gauge and were able to reduce ir to 1012 Pa
equivalent. They found it a valuable reference gauge in the XHV region.

Watanabe [148] has published results for an improved ion energy analyzer gauge. Using a spherical grid and
hemispherical energy analyzer, he obtained good separation of gas phase and ESD ions. The sensitivity with
molybdenum grid was

Fig. 6.36
Helmer ion energy analyzing gauge. The operating potentials are
shown in the figure. Reprinted with permission from J. C. Helmer
and W. D. Hayward, Rev. Sci. Instrum. 37, 1652 (1966) [Ref. 145].
Copyright 1966 American Institute of Physics.
< previous page page_423 next page >
< previous page page_424 next page >
Page 424

Fig. 6.37
Ion spectrum made with the Otuka and Oshima ion
energy analyzing gauge. The gas phase and ESD
ion peaks are resolved. Reprinted with permission
from C. Oshima and A. Otuka, J. Vac. Sci. Technol.
A 12, 3233 (1994) [Ref. 150]. Copyright 1994
American Vacuum Society.

4.5 × 104 A/Pa for nitrogen at 5 mA emission, giving a gauge constant of 0.09/Pa. The lowest useful pressure was given
as 1010 Pa.

Otuka and Oshima [149] investigated a variation of the ion energy analyzer gauge, also with good results. Oshima and
Otuka [150] gave the limit of their gauge as 4 × 1013 Pa and reported the gauge constant 0.018/Pa for hydrogen. Figure
6.37 illustrates its energy resolution.

6.9.6
Long Electron Path Length Gauges

Increasing the ionizing electron path length at a given emission current has the effect of improving the ratio ic/ir. With
the magnetron geometry shown in Fig. 6.38 Lafferty [151, 152] was able to measure pressures to 1011 Pa. To prevent
instabilities and space charge effects it was necessary to use small emission currents in the range 109 to 106 A. Using
information from Fig. 7 of Lafferty [152], for 107 A emission in a field of 0.025 T, the sensitivity was 5.3 × 104 A/Pa
and the gauge constant was 5.3 × 103/Pa. The test gas was air [153]. An example of a recent related gauge is the axial
emission magnetron suppressor gauge of Chen, Suen, and Kuo [154]. With an emission current of 1 × 106 A, a field of
0.03 T, and pressure of 1.3 × 106 Pa they obtained a sensitivity of 6.75 × 104 A/Pa, and a gauge constant of 6.75 × 102/
Pa. The x-ray limit was estimated to be 6 × 1013 Pa, although the gauge was not tested below 1010 Pa.
< previous page page_424 next page >
< previous page page_425 next page >
Page 425

Fig. 6.38
Lafferty magnetron gauge. The operating
potentials are: filament, 0 V; ion collector,
45 V; shield, 10 V; anode, +300 V,
magnetic field, 0.025 T. Reprinted
with permission from J. M. Lafferty, J. Appl.
Phys. 32, 424 (1961) [Ref. 151]. Copyright
1961 American Institute of Physics.

The Orbitron gauge of Mourad, Pauly, and Herb [155, 156] uses the electrostatic field created by a wire anode on the
axis of a cylindrical cathode to trap electrons, giving very long electron path lengths. At a pressure of 1 × 107 Pa of
nitrogen, a sensitivity of 5.3 × 105 A/Pa was found, giving a gauge constant of 5.3 × 102/Pa. This is a rather low
sensitivity: The sensitivity of a typical BAG is 1 × 104 A/Pa. The stability of the Orbitron is poor. Slight changes of
electron injection conditions, caused, for example, by warping of the filament, cause large changes in sensitivity.

6.9.7
Secondary Standard Hot-Cathode Gauges

Although the stability of the BAG (discussed later in this chapter) is good, special gauges with precise geometry and
well-defined electron paths can be better (see Section 12.3.3). The need for a transfer standard ionization gauge has
decreased as spinning rotor gauges have come into general use. Hirata et al. [157] of the Electrotechnical Laboratory in
Japan described a normal triode gauge, Type VS-1, used as a standard. The standard deviation of the gauge constant for
258 VS-1 gauges was ± 6.5%
Gentsch, Tewes, and Messer [158], in association with the Physikalisch-Technische Bundesanstalt in Berlin (PTB),
built and tested a gauge for use as a standard.

< previous page page_425 next page >


< previous page page_426 next page >
Page 426

Many features were incorporated for stability: well-defined electron paths; gold coatings for constant surface potentials;
and a shielded ionizing volume. The x-ray limit was 2 × 107 Pa hydrogen equivalent.

6.9.8
High-Pressure Ionization Gauges

The high-pressure limit of BAGs is often thought to be 102 Pa. In fact, for tungsten cathodes in active gases, 102 Pa is
excessive. But B-A gauges with suitable cathodes can be used to 1 Pa. Special high-pressure gauges can be used to
more than 100 Pa. Although there are other gauges better suited for accurate work at high pressures, sometimes an
ionization gauge is convenient. There is a paper by Wang [159] on the theory of the high-pressure ionization gauge.

Schulz [160] tested the performance of standard WL 5966 BAGs from 103 to 101 Pa in argon, helium, neon, hydrogen,
nitrogen, and SF6. He found that linearity extended to 5 × 101 Pa with electron emission currents of 104 A or less and
using argon gas. The high-pressure linearity of wide-range gauges was the concern of Ohsako [129], who found that it
was improved by an electrode behind the cathode. Peacock and Peacock [161] were interested in the overlap of the
ranges of the SRG and typical BAGs. Operation above 103 Pa required reduced emission current. Peacock and Peacock
also found the high-pressure linearity to be poorer when grid end closures were used.

Another approach to high-pressure operation of a standard BAG is given in a patent by Paitich and Briglia [162]. This
patent describes a method of measuring up to 100 Pa with a BAG by modulating the grid voltage. Controllers based
upon this concept are manufactured by Terranova Scientific, Inc. [163]. Gauges with very small electrode structures
would be preferable for use at pressures above 0.1 Pa. In a later paper, Schulz and Phelps [164] described two special
high-pressure ionization gauges with small dimensions. The one that was more linear at high pressure was the planar
type, shown in Fig. 6.39. The electron and ion collector were parallel plates 12.8 × 9.5 mm spaced 3.2 mm, and the
filament was 0.13-mm-diameter wire × 12.8 mm long spaced midway between the plates. The linearity extended to
between 13 and 130 Pa depending upon the gas. This gauge was manufactured for many years as the WL 7903.

Figure 6.40 illustrates the JHP (jauge haut pression), or Choumoff gauge, which was intended primarily as a high-
pressure secondary or transfer standard. Electrons pass through a box containing two ion collector rings. The length of
the box was about 6 mm. Poulter et al. [165] gave the results from the circulation of several of these gauges among
three European national standards laboratories. The stability found during these comparisons was good. The gauge
constants found for gauge No. 20 by the three laboratories were 0.0212/Pa, 0.01987/Pa, and 0.0202/Pa. In 1989,
Peacock [166] compared gauge No. 20 to a spinning rotor gauge. The gauge constant after 10 years in storage was
found to be 0.019/Pa. The JHP appears to be one of the more successful high-pressure ionization gauges, but it is not
available commercially.

There are many more high-pressure ionization gauge papers in the literature. Kudzia and Stôwko [167] used a very
small, spherical, ionization volume. This gauge was linear to 103 Pa. Edelmann [168] investigated a SchulzPhelps
gauge. Y. H. Kuo [169] described a high-pressure ionization gauge of simple construction. The linearity

< previous page page_426 next page >


< previous page page_427 next page >
Page 427

Fig. 6.39
SchulzPhelps gauge type WL 7903.
The operating potentials are: filament, +60 V;
anode, +120 V; ion collector, 0 V. Reprinted
with permission from G. J. Schulz and A. V.
Phelps, Rev. Sci. Instrum. 28, 1051 (1957)
[Ref. 164]. Copyright 1957 American
Institute of Physics.

was excellent up to 133 Pa. Depending upon the mode of operation, the gauge constant ranged from 1.1 × 103 to 3.5 ×
103.

6.9.9
Cold-Cathode Gauges

In the HCG the ionizing electrons are supplied by a thermionic cathode. In the CCG they are part of a self-sustaining
gas discharge. Several differences result. The emission current in the HCG can be adjusted by control of the cathode
temperature. It is not changeable in a CCG. Since the emission current is held fixed in the HCG, the x-ray current is
constant, determining the lowest measurable pressure. Although there is always a dense space charge in the CCG, the
anode current, and therefore x-ray production, decreases with pressure. There is no background current to mask the ion
current. The total power into filament and grid of a HCG may exceed 30 W, independent of pressure, and the resulting
heat is a cause of outgassing within the gauge. The power into a CCG is only about 0.1 W at high pressure and
decreases with pressure.
The CCG was invented by Penning [111]. Figure 6.41 illustrates the electrode and magnetic field arrangement of a
gauge of Penning type. From 1937 until about 1960, gauges like this were the only CCGs available. They were used
primarily in

< previous page page_427 next page >


< previous page page_428 next page >
Page 428

Fig. 6.40
Choumoff gauge. The operating
potentials are: filament reflector, 0 V;
filament, +50 V; anode box, +150 V;
electron collector, +350 V; ion collector,
0 V. Reprinted with permission from
K. F. Poulter, A. Calcatelli, P. S.
Choumoff, B. Iapteff, G. Messer, and
G. Grosse, J. Vac. Sci. Technol.
17, 679 (1980) [Ref. 165].

applications where the pressure was not below 104 Pa and where cost was more important than accuracy. These gauges
provided only crude pressure measurement.

The curves given by Conn and Daglish [170] are an excellent illustration of the discontinuities which can occur in the
pressure versus current curve of a Penning gauge. These authors found sudden shifts between 101 and 102 Pa as large as
50% of reading. The other problem with commercial gauges of original Penning design was that the discharge often
extinguished at pressures below 103 or 104 Pa, and the gauge then gave no reading. The triggered discharge gauge of
Young and Hession [171] shown in Fig. 6.42 was of Penning design. It contained a filament used momentarily to
provide initial electrons to start the discharge. Lange, Singleton, and Eriksen [172] calibrated a gauge of the type
described by Young and Hession. To examine the detailed structure of the pressure versus current curve, they used a
valving system providing a linear rate of pressure rise combined with a strip chart recorder for current. Their result,
reproduced as Fig. 6.43, consisted of a series of line segments separated by abrupt changes. The average slope with
nitrogen was 1.19.
Beck and Brisbane [112] published one of the early papers describing a CCG with cylindrical geometry. Their gauge
had an axial wire surrounded by a coaxial cylinder, along with a magnetic field in the axial direction. After trying the
wire as the negative electrode, they ''soon found experimentally that the current was much increased when

< previous page page_428 next page >


< previous page page_429 next page >
Page 429

Fig. 6.41
Electrode arrangement, fields, and trajectories
in the Penning gauge.
Fig. 6.42
The Young and Hession triggered
discharge gauge. Reprinted
with permission from J. R. Young
and F. P. Hession, in Trans. Nat.
Vac. Symp. 10, 234 (1963) [Ref.
171]. Copyright 1963 American
Vacuum Society.

the wire was made the anode." This arrangement, commonly known as the inverted magnetron, has been the key to
obtaining good results with the CCG gauge. Their plots of gauge current versus pressure follow a straight line from 102
to 106 Pa. The acceptance of the inverted magnetron owes a great deal to P. A. Redhead and colleagues at the National
Research Council of Canada in Ottawa. In 1958, Hobson and Redhead [113] described an inverted magnetron gauge,
sketched in Fig. 6.44,

< previous page page_429 next page >


< previous page page_430 next page >
Page 430

Fig. 6.43
Calibration of a Young and Hession triggered discharge
gauge by Lange, Singleton, and Eriksen showing breaks
in the gauge current versus pressure curve. Reprinted
with permission from W. J. Lange, J. H. Singleton, and
D. P. Eriksen, J. Vac. Sci. Technol. 3, 338 (1966) [Ref. 172].
Copyright 1966 American Vacuum Society.

designed for UHV with features such as guard rings to prevent measurement of field emission currents. They verified
well-behaved operation of the gauge from 101 to 1010 Pa. The slope of their current versus pressure plot was 1.10.

Peacock and Peacock [173] studied the inverted magnetron shown in Fig. 6.45. The gauge had feedthroughs for both the
anode and cathode, making the circuit of Fig. 6.46 convenient. The electrometer and high-voltage supply are both
referenced to ground; and should there be leakage currents in the anode cable or feedthrough, the leakage currents are
not measured by the electrometer. Figure 6.47 is a plot of current versus pressure data obtained with this gauge at a
magnetic field of 0.12 T [173]. The slope of the linear portion is 1.09. Another typical feature evident in CCG
calibrations is the increased slope below about 107 Pa.

The plot of Fig. 6.47 is typical for a CCG. The straight-line portion of a CCG characteristic may be fitted to a power law
equation

where ig is the gauge current, and n and K are constants. The departure from linearity is not great; values of n found in
the literature usually fall between 1.05 and 1.2. Peacock and Peacock [174], investigating the nonlinearity, found n to be
a function of the magnetic field. It decreased to 1.02 at 0.22 T for pressures in the straight-line portion above 106 Pa.
< previous page page_430 next page >
< previous page page_431 next page >
Page 431

Fig. 6.44
UHV inverted magnetron gauge. Reprinted with
permission from J. P. Hobson and P. A. Redhead,
Can. J. Phys. 36, 271 (1958) [Ref. 113].
Copyright 1958 NRC Research Press.

The magnetron gauge [175] is similar to the inverted-magnetron gauge but has the positions of the anode and cathode
interchanged; a schematic diagram of the original gauge design is shown in Fig. 6.48. The magnetron gauge has been
operated in the pressure range 105 to 1011 Pa; above about 108 Pa the currentpressure characteristic is linear [n = 1 in
Eq. (6.41)], and below that pressure the value of n increases. The magnetron gauge sensitivity (A/Torr) is generally
higher than that of the inverted-magnetron gauge by a factor of about 10 at pressures above 108 Torr.

There is a large body of literature on the theory of the crossed field discharge. Redhead [176] briefly summarized its
status. Existing theory explains some aspects of the discharge; but as pointed out by Redhead, classical theory for the
mobility of electrons cannot explain the nonlinearity of ion current with pressure. In a very helpful paper, Knauer [177]
discussed some of the complex phenomena of the crossed field discharge, including dynamic effects that generate radio
frequencies. Figure 6.49, from Peacock, Peacock, and Hauschulz [143], shows, greatly simplified, electron and ion
trajectories in the inverted magnetron discharge. In uniform crossed electric and magnetic fields, charged particles move
in cycloidal jumps. In the cylindrical geometry of the inverted magnetron the electrons circle the anode in a series of
small jumps,
< previous page page_431 next page >
< previous page page_432 next page >
Page 432

Fig. 6.45
HPS inverted magnetron gauge. Reprinted with permission
from R. N. Peacock, N. T. Peacock, and D. S. Hauschulz, J.
Vac. Sci. Technol. A 9, 1977 (1991) [Ref. 143].
Copyright 1991 American Vacuum Society.
Fig. 6.46
Circuit for operation of a cold-cathode gauge.
Reprinted with permission from N. T. Peacock
and R. N. Peacock, J. Vac. Sci. Technol. A 6,
1141 (1988) [Ref 173]. Copyright 1988
American Vacuum Society.

< previous page page_432 next page >


< previous page page_433 next page >
Page 433

Fig. 6.47
A calibration of the inverted magnetron gauge of Fig. 6.45. The
anode supply was 4.0 kV, and the field was 0.12 T. Reprinted with
permission from N. T. Peacock and R. N. Peacock, J. Vac. Sci.
Technol. A 6, 1141 (1988) [Ref. 173]. Copyright 1988
American Vacuum Society.
Fig. 6.48
Schematic diagram of a cold-cathode magnetron gauge.
Reprinted with permission from P. A. Redhead, Can.
J. Phys. 37, 1260 (1959) [Ref. 175].
Copyright 1959 Canadian Journal of Physics.

< previous page page_433 next page >


< previous page page_434 next page >
Page 434

Fig. 6.49
Ion and electron trajectories in the inverted magnetron.
Reprinted with permission from R. N. Peacock, N. T.
Peacock, and D. S. Hauschulz, J. Vac. Sci. Technol.
A 9, 1977 (1991) [Ref. 143]. Copyright 1991 American
Vacuum Society.

Table 6.4. Electron transit and Collision Parameters for a Penning Dischargea
Pressure (Pa) τe τi v vi N Ni
101 25 2
1.2 µs 0.6 µs 21 MHz 1.7 MHz
104 25 2
1.2 ms 0.6 ms 21 kHz 1.7 kHz
107 25 2
1.2 s 0.6 s 21 Hz 1.7 Hz
1010 25 2
20 min 10 min 21 mHz 1.7 mHz
aThe columns are: average electron transit times, τe; average time between ionizing collisions, τi; electron collision frequency, v;
electron ionizing-collision frequency, vi; total number of collisions during transit, N; total number of ionizing collisions during transit,
Ni. Calculated for a Penning cell with anode voltage 5 kV, cell diameter 30 mm, and magnetic field 0.1 T. From Redhead [175].

repetitively stopping, accelerating, and returning to rest. To maintain a discharge the peak energy must be adequate to create ions during
inelastic collisions. As the magnetic field increases, the peak energy will decrease until ionization is not possible. In this simple model the
electrons can move inward only as a result of inelastic collisions. Not only do electrons enter the discharge from ionizing collisions, but
secondary electrons ejected from the cathode by ions may also participate. At low pressures it can take a long time for an electron to travel
from cathode to anode. Table 6.4, from Redhead [176], gives some interesting numbers for a CCG, including transit times.

< previous page page_434 next page >


< previous page page_435 next page >
Page 435

The resulting circulating current can reach about 0.1 A. The space charge is large enough to depress the electrostatic
field. With sufficient depression of the potential, electrons from the cathode are no longer able to enter into the
circulating current. It is this effect that stabilizes the current, and causes the gauge to have a constant sensitivity. Lange
[178] measured the space charge in a Penning cell using a microwave probe method and found a space charge density
nearly independent of pressure between 104 and 108 Pa. From their curves, for a potential of 2500 V and magnetic field
of 0.19 T the electron density was 4 × 1015/m3. Their plots usually show a sudden change in density at about 3 × 108
Pa.

Hobson and Redhead [113] and Feakes and Torney [179] proved the ability of the cold-cathode gauge to measure to the
lowest laboratory pressures then available. Special Redhead CCGs were left on the lunar surface to monitor the lunar
atmosphere. Some accelerator storage rings operating at XHV use CCGs. Gauges similar to that shown in Fig. 6.45 are
available commercially [180], and there are other manufacturers [181, 182] also.

When the high voltage is applied to a CCG, there is a delay before the discharge starts. This is called the "striking time."
It is only a few seconds at pressures greater than 104 Pa and is not noticeable. At pressures of 108 Pa and less when no
source of initial ionization is present, it can be hours or days [143]. If the gauge cannot be turned on at a pressure where
the striking time is short, then supplying a source of initial ionization will overcome the problem. Including a
radioactive source within the gauge has been suggested and tried with varying success [183187]. Another means of
initiating the discharge is a source of UV light. Peacock and Peacock [188] showed a convenient way to introduce this
light source.

6.9.10
Ionization Gauge Accuracy

Accuracy of measurement with ionization gauges is a complex topic. Calibration of either HCGs or CCGs is normally a
matter of determining the gauge constant by comparison with a standard in the range 104 to 103 Pa. Many users simply
accept the manufacturer's catalog value. Whether calibrated or not, the application of the gauge may be at pressures
several decades lower than the direct calibration interval of 104 to 103 Pa. Linearity must be assumed if the gauge is to
be used at lower pressures.

Linearity of ion current with pressure was first verified for the BAG by Alpert and Buritz [189] using a valving
arrangement designed to produce a quadratic change of pressure with time. To the accuracy of the method, the BAG
was linear from 107 to 101 Pa. Redhead [127], in a paper on factors influencing the gauge constant, plotted the gauge
constant against pressure from approximately 104 to 101 Pa. It peaked sharply near 101 Pa. Peacock and Peacock [161],
comparing a BAG to a SRG, found the same peaking at emission currents of 1 mA, but at low emission currents the
peak in the gauge constant disappeared and it simply decreased above the linear region. Filippelli and Dittmann [190]
searched for pressure dependence of the gauge constant between 108 and 105 Pa and concluded that there was no
evidence for pressure dependence down to 5 × 108 Pa. Thus, as long as the calibration to determine the gauge constant
is done at pressures verified to be in the linear region for an individual gauge, the extrapolation is acceptable.

< previous page page_435 next page >


< previous page page_436 next page >
Page 436

Scatter of calibrations for new gauges has been examined at NIST (National Institute for Standards and Technology,
Gaithersburg, MD). The premise, as stated by Tilford [191], is "that gauge types that show small unit to unit variations,
in spite of manufacturing tolerances, would be more stable and predictable with time and use." Tilford [191] compared
his results for the calibrations of the gauge constant for six types of HCG to the manufacturers' specified values. The
result is reproduced as Fig. 6.50. The range of the gauge constants for 24 filaments of 12 conventional BAGs with two
opposed (tungsten) filaments was + 20 to 10%. For 10 filaments of conventional BAGs with thoria cathodes, it was
greater: + 13 to 38%. The UHV nude gauges had the greatest dispersion: For the sample of 15 cathodes the range was +
22 to 65%. The same figure also lists results for conventional triodes, BAGs with side-by-side filaments, and wide-
range BAGs with small grids.

Stability of calibration over long operating times is very important. Filippelli and Abbot [192], comparing repeated
calibrations of 20 gauges returned to NIST, concluded that for gauges with tungsten cathodes the standard deviation of
the maximum difference between successive calibrations was 3.1%, and for gauges with thoria cathodes it was 5.7%.
Poulter and Sutton [193] at the National Physical Laboratory, Teddington, United Kingdom, operated five triode-type
gauges and six BAGs for periods of about 1000 hours, making up to 100 calibrations on each. The BAGs had a greater
scatter of initial gauge constants, and for one gauge they decreased at an average rate of 1.4% per 100 hours. For two
triode gauges the changes were 0.08% and 0.45% per hundred hours. BAGs exposed to atmosphere, but not triode
gauges, showed changes as large as 25%.

Fig. 6.50
Average offset from the specified nitrogen gauge constants,
standard deviation, and range of gauge constants for six
different types of ionization gauge. The gauge types are
designated at the bottom of the figure and explained in the text.
The numbers at the bottom are the number of filaments tested.
The mean offset is indicated by the asterisk (*), ± one standard
deviation by the wider box, and the range of gauge constants is
indicated by the narrower box. Reprinted with permission from
C. R. Tilford, J. Vac. Sci. Technol. A 3, 546 (1985) [Ref. 191].
Copyright 1985 American Vacuum Society.
< previous page page_436 next page >
< previous page page_437 next page >
Page 437

There are many more papers on the reliability of HCG measurements made under the conditions of a standards
laboratory. An excellent summary was given by Tilford, Filippelli, and Abbott [194]. The papers cited above have many
references, and they provide access to earlier literature.

Less is known about the accuracy, repeatibility, and stability of CCGs. Peacock, Peacock, and Hauschulz [143] state
that early production records for one group of 159 gauges as in Fig. 6.45 show that 80% were within ± 20% of the mean
calibration at 103 Pa. An individual gauge, on the calibration system for more than a year with several exposures to
atmosphere, was compared frequently to an SRG. The scatter of the calibrations was about ± 5%. Filippelli [195]
reported that NIST had examined a group of several commercial CCGs and that they exhibited reproducible behavior
within ± 10%. There may, however, be many situations in the normal use of high vacuum gauging where CCGs deliver
far more accurate results. Peacock, Peacock, and Hauschulz [143] describe an experiment comparing readings of hot-
and cold-cathode gauges during and after a pumpdown. They noted that when their UHV comparison system, equipped
with both HCGs and CCGs, was pumped down following a calibration experiment the CCGs consistently indicated
lower pressures. Measurements showed that the pumping speed of the CCGs was about 3 × 105 m3·s1, and that of the
BAGs was about 2 × 105 m3·s1. These numbers are similar, and in any case much too small to explain the difference
during pumpdown. This suggested comparing both types of ionization gauge during a simulated pumpdown. The results
are shown in Fig. 6.51. At t = 0 the system was at base pressure, and all gauges had been calibrated previously. At t = 12
min the system pressure was increased stepwise by admitting nitrogen. The gauges were compared to the SRG up as far
as 103 Pa. With all ionization gauges off, the pressure was increased to 1 Pa. The pumpdown started at t = 107 min. The
curves show the nitrogen equivalent pressures measured by the various gauges during the pumpdown. The calibrations
were also verified at the end of the experiment. At 200 min, there is a difference of more than a decade in the nitrogen
equivalent pressures given by BAGs and CCGs. The difference is due to higher pressures within the HCGs caused by
degassing of material adsorbed during the exposure to nitrogen at 1.3 Pa. The extractor gauge, which has the lowest
power input of the HCGs, shows the best agreement with the CCG. This experiment was repeated many times with the
same results. If the system is simply allowed to pump, the readings of the HCGs drift downward toward the CCG very
slowly, further indicating that the HCGs are in error. Degassing the HCGs reduces their outgassing, and they come into
better agreement with the CCG.

Gauge constants change with time under the conditions of a standards laboratory. Industrial systems with residual gases
and vapors that contaminate the electrodes are much worse. However, there is little written on the effects of
contamination. Young [196] described the results of operating a gauge in methane. An insulating film, removable only
by electron bombardment, formed on the ion collector. Traces of oil in industrial systems frequently cause the collection
efficiency to go to zero. The situation with CCGs and contamination is identical to that of HCGs.

External magnetic fields have large effects upon ionization gauge calibration. Unfortunately, small fringing fields from
ion pumps or cold-cathode gauges are often present. Gauges used around nuclear accelerators may experience large
fields of the

< previous page page_437 next page >


< previous page page_438 next page >
Page 438

Fig. 6.51
Differing behavior of a nude BAG ( ) a glass BAG , an extractor gauge (×),
and a CCG ( ) in monitoring a pumpdown. The BAGs were at 1 mA emission.
At t = 0 the system was at a base pressure of about 2×1010 Torr. Reprinted
with permission from R. N. Peacock, N. T. Peacock, and D. S. Hauschulz, J. Vac. Sci.
Technol. A 9, 1977 (1991) [Ref. 143]. Copyright 1991 American Vacuum Society.

order of 1 T, and frequently pulsed fields. There are a number of papers on the effect of fields on gauges [197201].

Hseuh [197] studied a BAG in fields up to 6 × 103 T. For a field of 6 × 103 T perpendicular to the axis of the gauge,
changes of gauge constant were as large as 140% depending upon the angle of rotation of the field about the axis of the
gauge. Effects of about 10% occurred at fields as small as 1 × 103 T. Filippelli [198] used fields up to 0.16 T. His paper
has much detail regarding direction of the field and also regarding ionizing electron, collector, and wall currents. Pickles
and Hung [199] were interested in operating a gauge in fields up to 0.7 T. With proper orientation they succeeded. Dylla
[200] considered both hot- and cold-cathode gauges used near magnetic fusion devices. Martin [201] had similar
interests. These papers confirm that operation of a BAG in a magnetic field is possible with suitable orientation and
altered gauge constant.

6.9.11
Gauge Constant Ratios for Different Gases

It would be desirable to have ionization gauges calibrated for all the gases of interest. The user would then have a set of
constants, Kg, for use as needed. There is an imperfect alternate. In principle, if the ratios of the gauge constants, Rg,

< previous page page_438 next page >


< previous page page_439 next page >
Page 439

defined by

were known for various gases, unknown gauge constants could be estimated by multiplying the known constant for a
reference gas (here nitrogen) by the value of Rg from a table. There are tables of these ratios, such as Holanda's [202].
The problem is that the ratios as measured at different times with different gauges and operating conditions scatter
badly. This technique for estimating a gauge constant is at best only approximate. Figure 6.52 from Tilford [203] shows
gauge constant ratios plotted against pressure for a number of common gases. Curves are given for both conventional
triode and for BAGs.

6.9.12
Ionization Gauge Controllers

The HCG controller supplies the grid and cathode bias voltages and the filament heating power, and it usually includes
an electrometer for measuring the collector current. It is customary to operate with the cathode off ground so that the ion
current can be measured to ground.

The grid and cathode supplies should be well regulated, and particularly for laboratory use these voltages should be
adjustable over the full range of commonly used values. Adjustable grid current from 10 µA to 10 mA is almost a
necessity. Filament heating can be by dc with series transistors, or ac with SCR control. The bias situation differs
somewhat with these methods, and dc is less likely to cause problems. Abbott and Looney [204] found that SCR
emission control with glass gauges without internal shield caused erratic gauge calibration. This effect is due to
changing wall potentials.

Failure modes of a controller should be considered. An ionization gauge is a vacuum tube operated at vacuum tube
voltages. Interfacing with solid-state circuits that may be damaged by transients is not simple. If the grid circuit goes
open loop, the filament will go full on, perhaps burning it out. A control to set the maximum filament current can
prevent this. Most modern controllers turn the filament off if the pressure reaches some maximum value such as 101 Pa.
This may help to prevent filament burnout or damage and electrode contamination.

Long gauge cables can be a problem with HCGs. Filament currents can be as large as 5 A, so voltage drop along the
cable may be significant.

Accurate measurements with a BAG are never possible until the gauge has been processed by some combination of
baking and degassing until normal operation of the gauge does not cause a pressure increase. The type of degassing,
either by electron bombardment or resistive heating, demands careful thought. Electron bombardment degassing is
effective and can be used on all gauges. Resistive heating requires that the grid be a continuous wire in the form of a
helix or double helix. Electron bombardment degassing requires a supply providing 600700 V, with current adjustable
to 70 mA. A supply of this type can be lethal. Morrison [205] describes the problems and grounding precautions to
observe when using an electron bombardment degas supply. Resistive heating requires voltages less than 10 V and
currents of several amperes. The hazard is negligible. The pressure rises in the gauge during degas with either type of
heating. But if it rises to pressures above 102 Pa with electron bombardment

< previous page page_439 next page >


< previous page page_440 next page >
Page 440

Fig. 6.52
Gauge constant ratios for several gases as a function
of pressure. Reprinted with permission from C. R. Tilford,
J. Vac. Sci. Technol. A 1, 152 (1983) [Ref.203].
Copyright 1983 American Vacuum Society.
degas when using oxide cathodes, the cathode emitting material can be stripped in minutes by ion bombardment.

Electrometer stability and range must be compatible with the requirements. Electrometers are usually direct reading in
pressure units. However, in the laboratory it is important to have the choice of reading the current directly.

< previous page page_440 next page >


< previous page page_441 next page >
Page 441

The CCG controller supplies only the high voltage required by the gauge head, and it measures the current in the same
loop. Although CCGs require several kilovolts for operation, the current is limited to about 0.1 mA so the danger of
serious electric shock is reduced. Because the power requirements of the CCG are small (< 1 W), controllers can be
compact.

A wide range of vacuum equipment is available commercially. Little is fundamentally new, but it is smaller and more
convenient to use. The trends toward miniaturization, lower pressures, and standard off-the-shelf equipment will
probably continue.

References

1. P. L. M. Heydemann, C. R. Tilford, and R. W. Hyland, J. Vac. Sci. Technol. 14, 597 (1977).

2. A. Ooiwa, M. Ueki, and R. Kaneda, Metrologia 30, 565 (1993).

3. M. Ueki and A. Ooiwa, Metrologia 30, 579 (1993).

4. J. C. Legras and D. Le Breton, Vide 252 (Suppl.), 127 (1990).

5. R. N. Peacock, J. Vac. Sci. Technol. A 11, 1627 (1993).

6. H. McLeod, Philos. Mag. 48, 110 (1874).

7. A. Berman, Total Pressure Measurements in Vacuum Technology, p. 75. Academic Press, New York, 1985.

8. H. Ishi and K. Nakayama, 8th Nat. Vac. Symp., 519 (1962).

9. W. Gaede, Ann. Phys. (Leipzig) 46, 357 (1915).

10. A. Ooiwa, Metrologia 30, 607 (1993/1994).

11. A. Ooiwa and M. Ueki, Vacuum 44, 603 (1993).

12. K. Solis, Review of the Fundamentals of the Piston Pressure Balance and its Application. Ruska Instrument Corp.
Houston, TX (no date).

13. Wallace and Tiernan, Division of Penwalt, Belleville, NJ.

14. H. Lorenz, Phys. Z. 18, 117 (1917).

15. S. Dushman, in Scientific Foundations of Vacuum Technique (J. M. Lafferty, ed.), 2nd ed., p. 237. Wiley, New
York, 1962.

16. Ruska Instrument Corp., 3601 Dunvale Rd., Houston, TX.

17. K. Solis, The Fused Silica Helix Bourdon Tube: Its Place in Measurement and Control. Ruska Instrument Corp.,
Houston, TX (no date).

18. C. Brinkmann, Arch. Elektrotech. (Berlin) 32, 59 (1938).

19. A. Hasse, Z, Ver. Dtsch. Ing. 80, 563 (1936).

20. J. C. Lilly, V. Legallais, and R. Cherry, J. Appl. Phys. 18, 613 (1947).
21. D. Alpert, C. G. Matland, and A. O. McCoubrey, Rev. Sci. Instru. 22, 370 (1951).

22. H. W. Drawin, Adv. Vac. Sci. Technol., Proc. 1st Int. Vac. Congr., Namur, Belg., 274 (1960).

23. R. Hecht, U. S. Pat. 3,446,075 (1958).

24. W. R. Macdonald and C. King, An Electrostatic Feedback Transducer for Measuring Low Differential Pressures,
Tech. Rep. 71022. Royal Aircraft Estab., 1971.

25. Y. Rocard, General Dynamics of Vibrations, p. 47. Crosby Lockwood, London, 1960.

26. MKS Type 122A/B. MKS Instruments, Inc., Andover, MA.

27. MKS Types 690/698. MKS Instruments, Inc., Andover, MA.

< previous page page_441 next page >


< previous page page_442 next page >
Page 442

28. J. J. Sullivan, J. Vac. Sci. Technol. A 3, 1721 (1985).

29. J. J. Sullivan and F. Uttaro, Performance and Calibration of Capacitance Diaphragm Gauges. MKS Instruments,
Inc., Andover, MA.

30. R. W. Hyland and C. R. Tilford, J. Vac. Sci. Technol. A 3, 1736 (1985).

31. G. Grosse and G. Messer, J. Vac. Sci. Technol. A 5, 2463 (1987).

32. M. Bergoglio and A. Calcatelli, in Atti 7th Congr. Naz. Sci. Technol. Vuoto, Bressanone, p. 193. (1981).

33. W. Jitschin and P. Röhl, J. Vac. Sci. Technol. A 5, 372 (1987).

34. K. F. Poulter, M. J. Rodgers, P. J. Nash, T. J. Thompson, and M. P. Perkin, Vacuum 33, 311 (1983).

35. R. W. Hyland and R. L. Shaffer, J. Vac. Sci. Technol. A 9, 2843 (1991).

36. S. Dushman, in Scientific Foundations of Vacuum Technique (J. M. Lafferty, ed.), 2nd ed., Wiley, New York, pp.
244259, 1962.

37. W. Steckelmacher, Vacuum 23, 165 (1972).

38. A. Chambers, A. D. Chew, and A. P. Troup, Vacuum 43, 9 (1992).

39. A. D. Chew, A. Chambers, G. J. Pert, and A. P. Troup, J. Vac. Sci. Technol. A 13, 2271 (1995).

40. J. W. Beams, D. M. Spitzer, Jr., and J. P. Wade, Jr., Rev. Sci. Instrum. 33, 151 (1962).

41. A. T. Nordsieck, U. S. Pat. 3,003,356 (1961).

42. J. D. Nuttall and D. C. Witt, Vacuum 37, 347 (1987).

43. J. K. Fremerey, J. Vac. Sci. Technol. A 3, 1715 (1985).

44. J. K. Fremerey, Rev. Sci. Instrum. 42, 753 (1971).

45. J. K. Fremerey, J. Vac. Sci. Technol. 9, 108 (1972).

46. Instruction Manual for SRG-2. MKS Instruments, Inc., Andover, MA.

47. F. J. Redgrave and S. P. Downes, Vacuum 38, 839 (1988).

48. J. P. Looney, F. G. Long, D. F. Browning, and C. R. Tilford, Rev. Sci. Instrum. 65, 3012 (1994).

49. J. K. Fremerey, Vacuum 32, 685 (1982).

50. K. E. McCulloh, J. Vac. Sci. Technol. A 1, 168 (1983).

51. G. Comsa, J. K. Fremerey, and B. Lindenau, in Proc. 7th Int. Vac. Congr., Vienna, 1977, Vol. 1, p. 157 (1977).

52. S. Dittmann, B. E. Lindenau, and C. R. Tilford, J. Vac. Sci. Technol. A 7, 3356 (1989).

53. G. Messer, P. Röhl, G. Grosse, and W. Jitschin, J. Vac. Sci. Technol. A 5, 2440 (1987).
54. M. Hirata, H. Isogai, and M. Ono, J. Vac. Sci. Technol. A 4, 1724 (1986).

55. B. E. Lindenau, Vacuum 38, 893 (1988).

56. J. Setina, Vacuum 40, 51 (1990).

57. H. Isogai, Vacuum 44, 1181 (1993).

58. K. E. McCulloh, S. D. Wood, and C. R. Tilford, J. Vac. Sci. Technol. A 3, 1738 (1985).

59. J. K. Fremerey, Rev. Sci. Instrum. 43, 1413 (1972).

60. G. Comsa, J. K. Fremerey, and B. Lindenau, Proc. 8th Int. Vac. Congr., Cannes, 1980, Vol. 2, p. 218 (1980).

61. S. -H. Choi, S. Dittmann, and C. R. Tilford, J. Vac. Sci. Technol. A 8, 4079 (1990).

62. S. -H. Choi and D. Son, Vacuum 42, 897 (1991).

63. P. Röhl and W. Jitschin, Vacuum 38, 507 (1988).

64. J. K. N. Sharma, P. Mohan, W. Jitschin, and P. Röhl, J. Vac. Sci. Technol. A 7, 2589 (1989).

< previous page page_442 next page >


< previous page page_443 next page >
Page 443

65. J. K. N. Sharma, P. Mohan, and D. R. Sharma, J. Vac. Sci. Technol. A 8, 941 (1990).

66. D. J. Pacey, Vacuum 9, 261 (1959).

67. M. Ono, M. Hirata, K. Kokubun, H. Murakami, F. Tamura, H. Hojo, H. Kawashima, and H. Kyogoku, J. Vac. Sci.
Technol. A 3, 1746 (1985).

68. M. Hirata, K. Kokubun, M. Ono, and K. Nakayama, J. Vac. Sci. Technol. A 3, 1742 (1985).

69. T. Kobayashi, H. Hojo, and M. Ono, Vacuum 44, 613 (1993).

70. T. Kobayashi, H. Hojo, and M. Ono, Vacuum 47, 479 (1996).

71. Vacuum Products Corp., 4-1-1 Honcho, Koganei-shi, Tokyo 184, Japan.

72. M. von Pirani, Ver. Dtsch. Phys. Ges. 8, 686 (1906).

73. W. Voege, Phys. Z. 7, 498 (1906).

74. Translated by R. N. Peacock.

75. E. U. Condon, in Handbook of Physics (E. U. Condon and H. Odishaw, eds.), Sect. 5, p. 76. McGraw-Hill, New
York, 1958.

76. J. K. Roberts, Proc. R. Soc. London, Ser. A 135, 192 (1932).

77. E. H. Kennard, Kinetic Theory of Gases, p. 319320. McGraw-Hill, New York, 1938.

78. E. H. Kennard, Kinetic Theory of Gases p. 252. McGraw-Hill, New York, 1938.

79. A. Ellett and R. M. Zabel, Phys. Rev. 37, 1102 (1931).

80. B. G. Dickins, Proc. R. Soc. London, Ser. A 143, 517 (1933).

81. L. B. Thomas and F. Olmer, J. Am. Chem. Soc. 65, 1036 (1943).

82. J. H. Leck, J. Sci. Instrum. 29, 258 (1952).

83. Teledyne-Brown Hastings Engineering, Hampton, VA.

84. H. von Ubisch, Nature (London) 161, 927 (1948).

85. J. Zettler and R. Sud, J. Vac. Sci. Technol. A 6, 1153 (1988).

86. J. Blears and J. H. Leck, J. Sci. Instrum. 28, Suppl. 1, 20 (1951).

87. W. Steckelmacher and D. M. Tinsley, Vacuum 12, 153 (1962).

88. J. K. N. Sharma and A. C. Gupta, Vacuum 36, 279 (1986).

89. C. C. Minter, Rev. Sci. Instrum. 29, 793 (1958).

90. E. -G. V. Aleksandrovich, V. A. Sokovishin, and A. I. Sazanov, Instrum. Exp. Tech. (Engl. Transl.) 5, 939 (1964);
Pribo. Tekh. Eksp. 5, 162 (1962).
91. A. A. Birshert, Instrum. Exp. Tech. (Engl. Transl.) 3 (MayJune), 712 (1964); Prib. Tekh. Exsp. 3 (MayJune), 216
(1964).

92. J. B. Johnson, Rev. Sci. Instrum. 27, 303 (1956).

93. J. A. McMillan and T. Buch, Rev. Sci. Instrum. 28, 881 (1957).

94. F. Kreith, Principles of Heat Transfer, 3rd ed., pp. 398400. Harper & Row, New York, 1973.

95. E. R. G. Eckert, Heat and Mass Transfer, pp. 312319. McGraw-Hill, New York, 1959.

96. G. von Dardel, J. Sci. Instrum. 30, 114 (1953).

97. Series 315 Pirani Gauge System. MKS Instruments, Inc., HPS Division, Boulder, CO.

98. K. F. Poulter, M. -J. Rodgers, and K. W. Ashcroft, J. Vac. Sci. Technol. 17, 638 (1980).

99. W. Jitschin and M. Ruschitzka, Vacuum 44, 607 (1993).

100. J. A. Becker, C. B. Green, and G. L. Pearson, Bell Syst. Tech. J. 26, 170 (1947).

101. M. Varicak * and B. Saftic*, Rev. Sci. Instrum. 30, 891 (1959).

102. T. Shioyama, S. Ogawa, K. Takiguchi, T. Yotsuya, and T. Hasegawa, J. Vac. Soc. J. 22, 26 (1979) (in Japanese).

103. M. Esashi, Vacuum 47, 459 (1996).

< previous page page_443 next page >


< previous page page_444 next page >
Page 444

104. Model 2000 Solid State Vacuum Sensor. Teledyne-Brown Hastings Engineering, Hampton, VA.

105. A. W. van Herwaarden and P. M. Sarro, J. Vac. Sci. Technol. A 5, 2454 (1987).

106. P. K. Weng and J.-S. Shie, Rev. Sci. Instrum. 65 (2), 492 (1994).

107. O. E. Buckley, Proc. Natl. Acad. Sci. U.S.A. 2, 683 (1916).

108. P. A. Redhead, J. Vac. Sci. Technol. A 2, 132 (1984).

109. O. von Baeyer, Phys. Z. 10, 168 (1909).

110. R. T. Bayard and D. Alpert, Rev. Sci. Instrum. 21, 571 (1950).

111. F. M. Penning, Physica (Amsterdam) 4, 71 (1937).

112. A. H. Beck and A. D. Brisbane, Vacuum 2, 137 (1952).

113. J. P. Hobson and P. A. Redhead, Can. J. Phys. 36, 271 (1958).

114. E. K. Jaycox and H. W. Weinhart, Rev. Sci. Instrum. 2, 401 (1931).

115. P. A. Anderson, Phys. Rev. 57, 122 (1940).

116. J. Bell, J. W. Davies, and B. S. Gossling, J. Inst. Electr. Eng. 83, 176 (1938).

117. W. B. Nottingham, Report on the Seventh Conference on Physical Electronics, p. 14. MIT, Cambridge, MA, 1947.

118. D. Alpert, in Vacuum Science and Technology Pioneers of the 20th Century (P. A. Redhead, ed.), p. 144. Am. Vac.
Soc., 1993.

119. G. E. Moore, J. Appl. Phys. 32, 1241 (1961).

120. J. W. Ackley, C. F. Lathrop, and W. R. Wheeler, 9th Natl. Vac. Symp., 452 (1962).

121. P. A. Redhead, Vacuum 13, 253 (1963).

122. W. K. Huber and G. Rettinghaus, J. Vac. Sci. Technol. 6, 89 (1969).

123. W. H. Hayward, R. L. Jepsen, and P. A. Redhead, 10th Natl. Vac. Symp., 228 (1963).

124. W. B. Nottingham, 1st and 2nd Natl. Vac. Symp. Trans., 76 (1955).

125. D. Alpert, in Handbuch der Physik (S. Flügge, ed.), Vol. 12, p. 609. Springer-Verlag, Berlin, 1957.

126. A. van Oostrom, 8th Natl. Vac. Symp., 443 (1962).

127. P. A. Redhead, J. Vac. Sci. Technol. 6, 848 (1969).

128. L. G. Pittaway, J. Phys. D. 3, 1113 (1970).

129. N. Ohsako, J. Vac. Sci. Technol. 20, 1153 (1982).

130. G. Comsa, J. Vac. Sci. Technol. 8, 582 (1971).


131. G. Comsa, J. Vac. Sci. Technol. 9, 117 (1972).

132. J. Groszkowski, Bull. Acad. Pol. Sci. 14, 169 (1966).

133. L. T. Melfi, Jr., J. Vac. Sci. Technol. 6, 322 (1969).

134. G. A. Beitel and C. M. Gosselin, J. Vac. Sci. Technol. 7, 580 (1970).

135. B. Fletcher, Vacuum 20, 381 (1970).

136. F. Watanabe, J. Vac. Sci. Technol. A 5, 242 (1987).

137. P. A. Redhead, Rev. Sci. Instrum. 31, 343 (1960).

138. W. J. Lange and J. H. Singleton, J. Vac. Sci. Technol. 3, 319 (1966).

139. P. A. Redhead, J. Vac. Sci. Technol. A 10, 2665 (1992).

140. C. Benvenuti and M. Hauer, Nucl. Instrum. Methods 140, 453 (1977).

141. P. A. Redhead, J. Vac. Sci. Technol. 3, 173 (1966).

142. Leybold Inficon, Inc., E. Syracuse, New York.

143. R. N. Peacock, N. T. Peacock, and D. S. Hauschulz, J. Vac. Sci. Technol. A 9, 1977 (1991).

< previous page page_444 next page >


< previous page page_445 next page >
Page 445

144. Y. Fujii, H. Inoue, and F. Kanematsu, J. Vac. Sci. Technol. A 1, 90 (1983).

145. J. C. Helmer and W. D. Hayward, Rev. Sci. Instrum. 37, 1652 (1966).

146. S.-W. Han, W. Jitschin, P. Röhl, and G. Grosse, Vacuum 38, 1079 (1988).

147. C. Benvenuti and M. Hauer, in Proc. 8th Int. Vac. Congr., Cannes, 1980, Vol. 2, p. 199 (1980).

148. F. Watanabe, J. Vac. Sci. Technol. A 10, 3333 (1992).

149. A. Otuka and C. Oshima, J. Vac. Sci. Technol. A 11, 240 (1993).

150. C. Oshima and A. Otuka, J. Vac. Sci. Technol. A 12, 3233 (1994).

151. J. M. Lafferty, J. Appl. Phys. 32, 424 (1961).

152. J. M. Lafferty, 7th Natl. Vac. Symp., Trans., 97 (1961).

153. Personal communication with Dr. Lafferty, 19 August 1996.

154. J. Z. Chen, C. D. Suen, and Y. H. Kuo, J. Vac. Sci. Technol., A 5, 2373 (1987).

155. W. G. Mourad, T. Pauly, and R. G. Herb, Rev. Sci. Instrum. 35, 661 (1964).

156. E. A. Meyer and R. G. Herb, J. Vac. Sci. Technol. 4, 63 (1967).

157. M. Hirata, M. Ono, H. Hojo, and K. Nakayama, J. Vac. Sci. Technol. 20, 1159 (1982).

158. H. Gentsch, J. Tewes, and G. Messer, Vacuum 35, 137 (1985).

159. Y.-Z. Wang, Vacuum 34, 775 (1984).

160. G. J. Schulz, J. Appl. Phys. 28, 1149 (1957).

161. R. N. Peacock and N. T. Peacock, J. Vac. Sci. Technol. A 8, 3341 (1990).

162. U.S. Pat. 4, 314, 205 (1982) to R. M. Paitich and D.D. Briglia.

163. Terranova Scientific, Inc., Auburn, CA.

164. G. J. Schulz and A. V. Phelps, Rev. Sci. Instrum. 28, 1051 (1957).

165. K. F. Poulter, A. Calcatelli, P. S. Choumoff, B. Iapteff, G. Messer, and G. Grosse, J. Vac. Sci. Technol. 17, 679
(1980).

166. R. N. Peacock, Vide 252, Suppl. (May, June, July), 122 (1990).

167. J. Kudzia and W. Stôwko, Vacuum 31, 9 (1981).

168. C. Edelmann, Vacuum 36, 503 (1986).

169. Y. H. Kuo, Proc. 8th Int. Vac. Congr., Cannes, 1980, Vol. 2, p. 207 (1980).

170. G. K. T. Conn and H. N. Daglish, Vacuum 4, 136 (1954).


171. J. R. Young and F. P. Hession, 10th Natl. Vac. Symp., 234 (1963).

172. W. J. Lange, J. H. Singleton, and D. P. Eriksen, J. Vac. Sci. Technol. 3, 338 (1966).

173. N. T. Peacock and R. N. Peacock, J. Vac. Sci. Technol. A 6, 1141 (1988).

174. N. T. Peacock and R. N. Peacock, J. Vac. Sci. Technol. A 8, 2806 (1990).

175. P. A. Redhead, Can. J. Phys. 37, 1260 (1959).

176. P. A. Redhead, Vacuum 38, 901 (1988).

177. W. Knauer, J. Appl. Phys. 33, 2093 (1962).

178. W. J. Lange, J. Vac. Sci. Technol. 7, 228 (1970).

179. F. Feakes and F. L. Torney, Jr., Trans. 10th Natl. Vac. Symp., 257 (1963).

180. Series 421 Cold Cathode Gauge. MKS Instruments, Inc., HPS Division, Boulder, CO. 80301.

181. Balzers ag, Balzers FL-9496, Liechtenstein.

182. Televac Division of Fredericks Co., Huntingdon Valley, PA.

183. O. E. H. Klemperer, Br. Pat. 555, 134 (1943).

184. C. Hayashi, J. Vac. Sci. Technol. 3, 286 (1966).

< previous page page_445 next page >


< previous page page_446 next page >
Page 446

185. H. Mennenga and W. Schaedler, Proc. 4th Int. Vac. Congr., Manchester, 1968, p. 656 (1968).

186. B. D. Power and C. R. D. Priestland, Br. Pat. 1, 535, 314 (1978).

187. K. M. Welch, L. A. Smart, and R. J. Todd, J. Vac. Sci. Technol. A 14, 1288 (1996).

188. R. N. Peacock and N. T. Peacock, U.S. Pat. 5, 198, 772 (1993).

189. D. Alpert and R. S. Buritz, J. Appl. Phys. 25, 202 (1953).

190. A. R. Filippelli and S. Dittmann, J. Vac. Sci. Technol. A 9, 2757 (1991).

191. C. R. Tilford, J. Vac. Sci. Technol. A 3, 546 (1985).

192. A. R. Filippelli and P. J. Abbott, J. Vac. Sci. Technol. A 13, 2582 (1995).

193. K. F. Poulter and C. M. Sutton, Vacuum 31, 147 (1981).

194. C. R. Tilford, A. R. Filippelli, and P. J. Abbott, J. Vac. Sci. Technol. A 13, 485 (1995).

195. A. R. Filippelli, Pap., 38th Natl. Symp. Am. Vac. Soc., Seattle, WA, 1991, unpublished.

196. J. R. Young, J. Vac. Sci. Technol. 10, 212 (1973).

197. H. C. Hseuh, J. Vac. Sci. Technol. 20, 237 (1982).

198. A. R. Filippelli, J. Vac. Sci. Technol. A 5, 249 (1987).

199. W. L. Pickles and A. L. Hunt, J. Vac. Sci. Technol. A 4, 1732 (1986).

200. H. F. Dylla, J. Vac. Sci. Technol. 20, 119 (1982).

201. G. D. Martin, Jr., Trans. 8th Natl. Vac. Symp., 476 (1962).

202. R. Holanda, J. Vac. Sci. Technol. 10, 1133 (1973).

203. C. R. Tilford, J. Vac. Sci. Technol. A 1, 152 (1983).

204. P. J. Abbott and J. P. Looney, J. Vac. Sci. Technol. A 12, 2911 (1994).

205. C. F. Morrison, J. Vac. Sci. Technol. A 3, 2032 (1985).

< previous page page_446 next page >


< previous page page_447 next page >
Page 447

7
Partial Pressure Analysis

Robert E. Ellefson

Measurement of total pressure in a vacuum system is often not sufficient to characterize the vacuum for processing or
experiments. The measurement of partial pressures of the components making up the total pressure is useful for vacuum
diagnostics or process monitoring. A partial pressure analyzer consists of an ion source, a mass analyzer, an ion current
detection system, and the control electronics. A partial pressure analyzer used to measure the composition of the
residual gas in the vacuum system is usually a quadrupole mass spectrometer (QMS) with its ion source immersed in the
vacuum system. Such a QMS is referred to as a residual gas analyzer (RGA). The measured ion currents from gas
species present are converted into partial pressures using a sensitivity for each gas species determined by a calibration
process. When the total pressure of a vacuum processing system exceeds the operating pressure of the RGA, pressure
reduction methods are used which allow indirect measurement of partial pressures in the process or composition of the
gas present. When pressure reduction is needed, a closed-ion-source mass spectrometer can be used which offers
improved detection limits for small component partial pressures. The ions sources for partial pressure analyzers
typically use electron impact ionization. Partial pressure analysis methods which measure optical absorption or emission
can monitor certain components in a vacuum that have strong absorption or emission signatures.

7.1
Ion Sources

The choice of ion source for a partial pressure analyzer depends on the pressure regime for the gases to be analyzed and
the accuracy and precision required for the

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_447 next page >


< previous page page_448 next page >
Page 448

measurement application. To measure the residual gases in the medium to high vacuum pressure regime, an ion source
with good conductance to the vacuum system is needed. If the vacuum system is evacuated to ultrahigh vacuum (UHV)
or extreme high vacuum (XHV), a high conductivity ion source similar to a total pressure ionization gauge is needed.
The residual gas composition measured by the RGA can include a significant contribution from the outgassing and
electron stimulated desorption from a hot-cathode ionization source. Ion sources for UHV and XHV are designed to
minimize these effects. Process gas in the vacuum system is often analyzed for minor impurities. Low-level detection of
impurities can be achieved by using a closed ion source. Most ion sources for vacuum applications use a hot cathode for
the formation of ions by electron impact ionization. This common ionization process and the different types of ion
sources are described in the next paragraphs.

7.1.1
Electron-Impact Ionization Process

The electron-impact ionization process is an inelastic scattering process where the incident electron transfers energy to
an electron in the target molecule or atom; the energy transferred is typically less than the kinetic energy (KE) of the
incident electron. The reaction can have a number of products with various probabilities (cross sections) for formation:

where M+ is termed the parent molecular ion where one electron is removed from the molecule. The product M++ is a
doubly charged molecular ion formed when two electrons are removed from the molecule. The other examples of the
process are the formation of fragment ions, F+, and associated neutral fragment, N. The ionization process is shown in
Fig. 7.1 as a potential energy diagram for the singly charged reaction products of Eq. (7.1). Electron-impact energy
transfer to an electron in the molecule is described by the FranckCondon principle [1], which states that ionization is a
''vertical transition" where little change in internuclear separation(s) occurs because the energy transfer occurs in a time
much shorter than a period of vibration of the molecule. Ionization proceeds from the ground state of the molecule
vertically to a coordinate of the potential energy curve describing the resulting products such as parent molecular ion
plus escaping electrons or fragment ion with a neutral and electrons. The amount of energy just sufficient to produce a
parent molecular ion, M+, and free an electron is defined as the ionization potential (IP) of the molecule. The minimum
energy to create a fragment ion or higher charge state of the ion is called an appearance potential (AP), which is a
characteristic of the molecule related to its bond strengths. It is possible to produce an ion in an excited state which can,
after a few molecular vibrations, fall apart into other fragment ion and neutral species represented in Fig. 7.1 as
Energy deposited in the molecule can also lead to a repulsive state where the fragment ion and
neutral species have excess kinetic energy at the appearance potential, AP2. The energy dependence for the cross
section for production of parent molecular ions [2] is indicated in Fig. 7.2. Most electron-impact ionization mass
spectrometers use 70-eV electrons, which provide ample energy for

< previous page page_448 next page >


< previous page page_449 next page >
Page 449

Fig. 7.1
Potential curves for a molecule, M, and electron-impact
ionization products producing M+ parent ion and two fragment
ions, when the incident electron has sufficient energy.

production of the parent molecular ion and fragment ions characteristic of the molecule. Ionization potentials of some
common molecules and atoms are given in Table 7.1 from a detailed compilation [3] by the National Institute of
Science and Technology. Characteristic fragment ions and their abundance for 70-eV energy ionization are given in
Table 7.2 for common molecules [4]. Additional spectra and tables of characteristic fragment ions and their abundances
are given in collections of mass spectral data [46].

7.1.2
Open Ion Source

The basic ion source for a residual gas analyzer is an open ion source similar to the structure of a BayardAlpert gauge.
Figure 7.3 shows various styles of open ion sources. The term "open" refers to a high conductance of the ion formation
region to the surrounding vacuum region to ensure that the gas composition being analyzed reflects the composition of
the vacuum [7]. The conductance of some styles is restricted by solid support structures with the filament mounted
within the structure (Fig. 7.3d,e). Such styles can be used in medium- and high-vacuum applications. In a partially
enclosed ion source, the hot cathode can induce local formation of various reaction products [8] (see Chapter 10). In a
more open source, reaction surface areas are reduced and surface temperatures are lower so the filament reaction effect
is reduced (Fig. 7.3b,c,f). For UHV and XHV applications, Watanabe [9, 10] has shown that outgassing of the ion
source due to the hot filament can be further reduced

< previous page page_449 next page >


< previous page page_450 next page >
Page 450

Fig. 7.2
Ion pairs/mbar-cm for various gases versus kinetic energy of incident electrons which
is directly proportional to the cross section for ionization for each species shown [2].

Table 7.1. Selected Ionization Potentials and Appearance Potentials of Common


Molecules and Atoms [3]
< previous page page_450 next page >
< previous page page_451 next page >
Page 451

Table 7.2. Fragmentation Patterns for Common Molecules with 70-eV Electron Ionization [4]a

a Fragmentation abundances in quadrupole mass spectrometers differ depending resolution and ion energy settings
used.
Fig. 7.3
Five variations on the design of open ion sources. Note the similarity of filament and grid
structures between an open ion source and a BA gauge.

< previous page page_451 next page >


< previous page page_452 next page >
Page 452

by modifications of the ion source (Fig. 7.3f). A berylliumcopper ring around the source was used to reflect heat back to
the filament and to shield the analyzer from filament heat by conducting the heat to the outside of the housing.

In open ion sources, different filaments geometries are used. Some ion sources have the filament located as a ring
outside the central cylindrical grid structure [7]. Another design has the filament located parallel to the axis on the side
of the central grid, and a third design has the filament on the end of the ionization region in an axial ion source. An
electron emission current from the heated filament is established with the filament at a negative voltage (typically 70 V)
with respect to the grid, thus defining the energy of ionizing electrons. The repeller (when present) is held at a few volts
more negative than the filament to repel electrons emitted outward back toward the grid. The grid is held at a positive
potential with respect to the analyzer axis potential; the difference between the grid potential and analyzer axis potential
defines the energy of ions that enter the analyzer. In a quadrupole mass spectrometer, this potential difference (ion
energy) is 515 V. In a magnetic sector mass spectrometer, this potential is 10010,000 V depending on the application
and design of the mass spectrometer. The other electrodes labeled "focus plates" in Fig. 7.3 are adjusted to potentials
that focus the ion beam for maximum transmission through the mass analyzer. The operating pressure regime for the
open ion source is from UHV to 102 Pa. Higher pressures are possible, but nonlinear response of the ion source occurs
due to three competing effects: (1) space charge (e and I+) buildup in the ionization region which alters the ion
transmission out of the ion source into the analyzer compared with lower pressure operation [1114]; (2) spreading of ion
beam due to Coulomb repulsion of intense ion beams during transit from source to analyzer and for intense ion beams in
the analyzer [15]; (3) collision losses during transit through the mass analyzer.

The operational parameters that control the production of ions in the ion source of a mass spectrometer are identified in
Eq. (7.2). The ion current, , for the ith component is given by

where Ie is the electron emission current, σi(E) is the cross section for ionization at electron energy, E, and
F is the ion transmission factor dependent on ion energy, mass, geometry of the grid structure, and
the electrostatic potentials extracting and focusing the ions from the source and the total pressure, Pi, and gas
composition, Xi. The ratio of Pi, the source partial pressure, to Tsource(K), the source temperature, gives the gas density
of the target component in the ionization region.

7.1.3
Closed Ion Source

A closed ion source encloses the ionization region; the gas to be analyzed enters through a supply tube and exits through
the electron beam hole and the ion exit slit or hole. Figure 7.4 shows a closed ion source used in quadrupole mass
spectrometers [16, 17] and the Nier-type ion source [1821] used in magnetic sector mass spectrometers. An advantage
of the closed ion source is that pressure in the ionization region is higher than in the analyzer region by a ratio of the
analyzer pumping speed to the closed ion source pumping speed. Typically, this ratio is 10100, which produces ion

< previous page page_452 next page >


< previous page page_453 next page >
Page 453

Fig. 7.4
Closed ion source designs for quadrupole and magnetic sector mass spectrometers.

source pressures up to 101 Pa while the analyzer operates at 103 Pa. High source pressure allows measurement of low-
level impurities while the analyzer pressure is maintained at a low enough pressure to minimize ion losses in the
analyzer section. Gas throughput needed for the analysis is also reduced by the ratio of the analyzer pumping speed to
the closed ion source pumping speed. Background components from outgassing or sample-induced desorption come
mainly from the small surface area inside the ionization region and are small contributions to the gas sample. These
small additions allows the measurement of low-level impurities in gas samples. The filament is located external to the
ionization region, which minimizes filament surface reaction products from diffusing back into the ionization region
and contributing to the composition of the gas sample being analyzed.

The QMS closed ion source uses a simple design with the filament located in front of a slit in the ionization chamber.
The electrons are accelerated by applying 70 V (or other voltage to define electron energy) to the filament with respect
to the ionization chamber. A significant fraction (1020%) of the electrons from the filament go through the electron
beam hole or slit terminating on the inside wall of the ionization chamber. In a Nier-type ion source, the electron beam
is focused with a grid and collimated by a small magnetic field parallel to the electron beam. The electron beam exits
through a hole on the opposite side of the ionization chamber into an electron trap electrode maintained at + 100 V. The
electron trap current is measured and used in an emission regulator circuit to produce a constant electron current.
Electrons execute a helical motion in the source magnetic field creating a higher current density than without the
magnet. The narrow collimated beam also minimizes the energy spread in ions extracted and accelerated into the mass
analyzer. This is important for magnetic sector mass spectrometers because the resolving power is inversely
proportional to the energy spread of ions.

Variations on these basic designs for ion sources are used in commercial mass spectrometers. Evolution of ion sources
for specific applications has occurred especially with quadrupole mass spectrometers in recent years and will likely
continue.

< previous page page_453 next page >


< previous page page_454 next page >
Page 454

7.2
Ion Detection

Early detection of positive ions was by observing the fluorescence of natural phosphors like glass, willemite (zinc
silicate), and zincblende (a zinc sulfide) when positive rays were incident on the scintillators. By 1910, fluorescent
screens were replaced by photographic plates which gave a permanent record of spectra and intensity measurement
from the opacity of the ion exposure [22].

7.2.1
Faraday Cup Ion Detection

By 1930, ion currents could be measured directly from a mass analyzer with an electrometer as ions strike a plate or
Faraday cup [23]. This detection method is still preferred using modern electrometers and is generically termed Faraday
detection in honor of Michael Faraday, a nineteenth-century physicist who in his experiments collected electrical
charges in a metal cup. Ion detection occurs when electron current flows to the plate to neutralize the charge of the ions
that arrive; the current is measured by an appropriate electrometer circuit. A "deep cup" detector is used in magnetic
sector mass spectrometers to ensure that secondary electrons released by energetic incident ions do not escape from the
cup and thereby generate an apparent additional ion current when the electrons leave the plate. Other designs use
suppressor plates to repel secondary electrons back to the Faraday plate. Most quadrupole mass spectrometers do not
generate ions with enough energy to release secondary electrons at the detector plate. The time constant for a Faraday
type of ion current measurement is dictated by the input resistance and the distributed capacitance of the current
measurement circuit. A time constant τ = RC ≈ 0.1 s are typical for Faraday cup/electrometer systems. The detection
limit of a Faraday detector is typically the noise limit of 3 × 1016 A [104 ions/s] in modern field effect transistor
electrometers. Faraday detection is the simplest and least expensive ion detection device and is often used in low-cost
quadrupole mass spectrometers.

7.2.2
Secondary Electron Multiplier Detection

Small partial pressures produce small ion currents that may be below the detection limit of a Faraday detection system.
The use of a secondary electron multiplier (SEM) can convert the electrons released from a single ion incident on the
detector surface into a larger electron current with a multistage amplification of the electron current. Early electron
multipliers used discrete dynodes (1012) with a 50- to 100-V accelerating potential between each stage to increase the
electron current. Later designs use a continuous dynode like that shown in Fig. 7.5. The gain of the multiplier, G, is due
to two factors: The first is the conversion of the incident ion to electrons at the first dynode or point of incidence on a
continuous dynode device. The number of electrons released in this primary event, p, is one to five depending on ion
energy, mass, molecular structure, and even ionization potential. The second factor is the electron gain per stage
because the secondary electrons are accelerated between dynodes or through the continuous dynode structure. This is
summarized with the relation

< previous page page_454 next page >


< previous page page_455 next page >
Page 455

Fig. 7.5
A continuous dynode detector for ion detection by secondary
electron multiplication (Galileo Electro-Optics Corporation).

where n is the number of discrete dynodes or qn can be the overall electron amplification in a continuous dynode
structure. In Fig. 7.5, the continuous dynode film is formed on the inside of a glass horn-like structure with the resistive
film (~ 108Ω) extending to the outside on both ends to provide connection points for the high voltage and ground. The
nonconductive zone indicated in Fig. 7.5 is on the outside of the glass structure and electrically separates the ends. The
applied voltage on the horn end is typically 2000 V with 20 µA of current through the film establishing the potential
gradient inside the channel to transport and amplify secondary electrons from the first ion-to-electron conversion event
to the capture of the electrons by a collector plate connected to the electrometer/preamplifier. The output electron
current of the SEM is 103 to 107 times the ion current at the first dynode, which is easily measured with an electrometer
with smaller input resistors than the Faraday detector and shorter time constant. Manufacturers recommend SEMs be
used for ion currents less than 10 pA, which means less than 104 Pa (N2 equivalent) gas pressure in the ion source of a
typical mass spectrometer. Higher currents are reported to degrade gain. SEM gain loss by a factor of 10 can occur from
total transported electron charge of 4 mA·hr [24]; gain loss can occur from deposits, especially hydrocarbons, from an
ion beam or from contamination. It is preferable to keep the SEM operating in a clean vacuum of 105 Pa or lower to
minimize contamination. Discrete dynodes are typically made of CuBe(2%) or AgMg(24%) "activated" in air to form a
stable Be oxide (or Mg oxide) film on the surface which controls the secondary electron yield. This layer can become
chemically altered or contaminated during use in a mass spectrometer, giving rise to changes in multiplier gain. A
continuous dynode 104 SEM replaces the discrete dynodes with a resistive layer on an insulator. The semiconductive
film on a continuous dynode SEM is chemically inert and adsorbed gases are desorbed by continued use, thus producing
a more stable gain over time than a discrete dynode SEM. Lifetime of an SEM is related to the total

< previous page page_455 next page >


< previous page page_456 next page >
Page 456

electron dose to the electrode surface. Some SEMs are designed specifically for use in pulse counting as an alternative
to analog measurement of low-level ion current. Typically, pulse counting detectors are limited to less than 106 ions per
second, which corresponds to an ion current of 2 × 1013 A.

The use of a SEM to measure small ion currents introduces a new calibration parameter due to the gain in the ion
current signal from a multiplier. The gain is adjustable by changing the voltage applied to the SEM. The gain at a given
applied voltage can degrade due to contamination of the multiplier surfaces and/or ion beam damage. The practical
matter of cross-calibrating the output of a Faraday detector with an SEM detector to give useful combined data is
resolved by measuring an ion current from a peak where the intensity is on scale for both detectors. A measurement of
the SEM gain factor G can be defined:

Ion current data from the SEM can be divided by G to scale the value for use with Faraday detector data. When
sufficient ion current is present, the Faraday detector should be used for quantitative measurements to avoid monitoring
the stability of the gain factor, G.

7.2.3
Microchannel Plate Detector

The gain of a continuous dynode SEM depends on the length-to-diameter ratio, which allows for miniaturization of this
type of SEM [25]. The combination of a large number of small channels into a planar array results in a microchannel
plate (MCP) where miniature continuous dynode SEMs in parallel form a compact multiplier-detector for ions. A
representation of the MCP array and the operation of an individual channel is shown in Fig. 7.6. The short length of the
microchannels limits the applied voltage to about 1000 V producing a gain up to 25,000.

The multichannel plate detector output can be resolved spatially, which offers the possibility for ion current dispersion
on the array giving a "mass spectrograph"-like output. An example of a hybrid detector with MCP/phosphor/photodiode
arrays to convert dispersed ions into a spatial image is reported in the literature [26].

7.3
Mass Analysis

Mass analysis of ions is achieved by various methods of acceleration and deflection by electric and magnetic fields. In
all the mass analysis methods, it is the mass-to-charge ratio of the ion, M/e, that is analyzed by the physical separation
of ions in space or time. Thus ions like 40Ar++ and 20Ne+ will be analyzed with the same nominal M/e of 20.

7.3.1
Quadrupole Mass Analyzer

The quadrupole mass spectrometer (QMS) is the most popular mass analyzer used for vacuum partial pressure
measurements because of its rapid scanning capability,

< previous page page_456 next page >


< previous page page_457 next page >
Page 457

Fig. 7.6
A microchannel plate detector array and electron
multiplication in a single channel
(Galileo Electro-Optics Corporation).

compact size, linear mass scale and relatively low cost. The QMS uses a mass filter that consists of four parallel
conductive rods arranged in a square array with opposite rods connected electrically in parallel [27] as shown in Fig.
7.7. Ions from a source enter an end of the quadrupole mass filter near the axis drifting parallel to the rods (defined as
the z-axis) with a kinetic energy of 315 eV. The combination of a direct-current (dc) potential and radio-frequency (rf)
potential applied to the rods accelerates the ions perpendicular to the z-axis; transmission through the quadrupole
potential field of the rod assembly occurs for ions in a narrow mass range. Low M/e ions move nearly in phase with the
applied RF voltage and are accelerated to large x and y displacements. These light ions collide with the rods and are
neutralized and lost from the beam. High M/e ions do not gather sufficient x or y velocity during the rf cycle to achieve
much displacement, but the dc potentials give a constant acceleration that centers the ions between the rods with
positive potential and attracts the ions to the rods with negative potential where the ions collide with the rods and are
neutralized. Between the high mass and low mass extremes, there is a range of M/e ions that can oscillate with small
amplitudes and drift through the rod structure without striking the rods. These ions are transmitted to the detector [28].
The value of M/e
< previous page page_457 next page >
< previous page page_458 next page >
Page 458

Fig. 7.7
Quadrupole mass filter structures. Parabolic rods
generate a true quadrupole field; cylindrical rods
approximate the parabolic shape and are less costly to
manufacture. The rf circuit is tuned to resonate at the
driving frequency for optimum coupling to
the rod assembly.

that is transmitted depends on the amplitude of the rf voltage; and the range of M/e values, ∆M/e, depends on the
number of oscillatory cycles the ion spends drifting through the rod assembly. The equations of ion motion in the
quadrupole field were first given by Paul et al. [29] and are summarized as follows.
Figure 7.7 shows the quadrupole rod assembly and applied constant potential and rf potentials. The potentials applied to
the rods is

where U is the dc amplitude, V is the amplitude of the rf applied potential with angular frequency ω, x and y are
coordinates perpendicular to the axis of the rod

< previous page page_458 next page >


< previous page page_459 next page >
Page 459

assembly, and r0 is the radius of a circle tangent to the inside of the rods. The equations of motion for ions in this
potential are given by the Mathieu equations [29, 30]:

Equations (7.6) and (7.7) together describe an oscillating system with restoring force that is periodic in time with
angular frequency ω as the ion drifts through the rod assembly in the z direction. Simplification of these equations
occurs by defining parameters a and q:

The practical operation of a QMS chooses values of a and q within the lowest region of stable oscillatory solutions of
the Mathieu equations. Normal operation is chosen for highest resolving power, M/∆M. (In much of the QMS literature,
M/∆M is referred to as resolution.) Choices like q = 0.706 and a = 0.233 to a = 0.236 produce 50 < M/∆M < 500 [30].
The actual adjustment of resolving power is done electronically by adjusting the ratio of U/V, which is related to a and q
by the ratio of Eq. (7.9) to (7.10),

However, as resolving power increases, ion transmission decreases due to a reduction in radius of ion beam accepted by
the mass filter as U/V increases. Mass scanning relations are predicted by Eq. (7.10) for a given choice of q and for
fixed values of r0 and ω, which shows that M/e is directly proportional to the rf amplitude V:

or for a practical quadrupole application with q = 0.706, the mass (in amu) transmitted is

where V is the rf voltage, r0 is the quadrupole radius in meters, and f is the rf frequency in Hz. The resolving power can
also be limited by the length of the rod assembly and the ion energy, eVz from the ion source:

< previous page page_459 next page >


< previous page page_460 next page >
Page 460

The coupling of the ion source to the rod assembly also affects transmission through the mass analyzer. Ions with small
radial displacements and small radial velocity components have the greatest chance of traversing the mass analyzer and
being detected. Also, off-axis location of the source or misalignment of the source or rods can degrade the peak shape
with notches or shoulders, making algorithms for quantitation of ion current more difficult. A common mode of QMS
operation is to produce mass peaks with nearly constant ∆M of less than 1 amu for all M/e over the mass range of the
instrument. This enables measurement of small peaks at a mass adjacent to an intense peak even for high mass. Constant
∆M can be produced by maintaining a relationship between the dc voltage, U, applied to the rods and the rf maximum
voltage, V, given by U = KV + Uoffset, where K is a constant fraction and Uoffset is a negative offset voltage indicated
in Fig. 7.8. In Fig. 7.8, the shaded areas describing stable trajectories for two particular masses come from the allowed
values of a and q of the Mathieu equations that produce stable ion trajectories [30]. By solving Eq. (7.9) and (7.10) for
U and V, respectively, the scaling of the a and q parameters by mass M is indicated for the applied voltages U and V
shown on the axes of Fig. 7.8. Useful mass separation is indicated when Uoffset lowers the operating line to allow a
narrow range of rf voltages (dotted lines) centered around VM1 and VM2 that transmits masses M1 and M2,
respectively, producing the observed near-constant width, ∆M, for each peak [31]. The constant ∆M mode of operation
results in ion transmission decreasing with mass as M1, which limits useful sensitivity at high mass. However, the near-
independence of ion transmission from ion energy means that changes in ion energy by collisions does not result in
large loss in transmission, so a QMS can operate at a relatively high analyzer pressure [32]. This makes the QMS
especially useful for residual gas analysis applications.

7.3.2
Magnetic Sector Analyzer

A magnetic sector analyzer achieves mass separation by deflection of ions of the same energy by a uniform magnetic
field oriented perpendicular to the motion of the ion.

Fig. 7.8
QMS operating line for producing a constant ∆M peak width over the
mass range of the instrument [31]. The mass scale is directly
proportional to the applied RF peak voltage.

< previous page page_460 next page >


< previous page page_461 next page >
Page 461

Ions formed in the source are accelerated by applying a voltage, V, to the source. Each ion achieves an energy given by

where z is the charge state on the ion of mass, M. When the ion enters a uniform magnetic field, it experiences a
centripetal force

which traces an arc with radius, R. By eliminating v from these equations, the radius of the path of an ion of mass, M, is

or, by rearrangement,

The mass separation that occurs is shown in Fig. 7.9, which is a 60° magnetic analyzer used for residual gas analysis
[33]. Light M/e ions (small circles) have the smallest radius path in the magnet and exit the magnet with greatest
deflection from the original ion path before entering the magnet. The heavy M/e ions (large circles) have the largest
radius within the magnet, as predicted by Eq. (7.18), and experience the smallest deflection. In a practical instrument, a
fixed radius of curvature is defined by the source slit, analyzer baffle, and collector slit. This allows a narrow range of
mass to

Fig. 7.9
A small commercial 60° magnetic sector mass spectrometer used for
residual gas analysis. Non-normal entry of ions into a 92° magnet
shortens the focal length of magnet making a more compact analyzer [33].
< previous page page_461 next page >
< previous page page_462 next page >
Page 462

be measured for a given magnetic field, B, and accelerating voltage, V. By scanning V or B, different values of M/z can
be brought into focus.

An early method for measuring a mass spectrum was to measure the different radii [Eq. (7.18)] separating the ions of
different mass on a focal plane exiting the magnetic field. Placing detectors or a photographic emulsion at the focal
plane records the ion currents as a mass spectrograph. The first measurement of 20Ne and 22Ne isotopes and
abundances was done by Aston [34] by building and using a mass spectrograph similar to this one. Modern versions of
the mass spectrograph locates detectors for selected masses to continuously monitor specific processes [34, 35].

An example of a 90° sector mass spectrometer is shown in Fig. 7.10 with ions entering and exiting the magnetic field
normal to the pole face. The focusing of this geometry is simple: The distance from pole face to source and pole face to
collector slit is equal to the radius of magnetic deflection. Small adjustments to focal length is accomplished by moving
the magnet as needed radially to accomplish the best peak shape. The resolving power of a 90° magnetic sector mass
spectrometer with ion beam entering and exiting the magnet normal to the face is given approximately by

where R is the magnet radius, Wcollector is the collector slit width, and Wsource is the source slit width. To achieve
unattenuated ion transmission indicated by flat-topped
Fig. 7.10
A 90° magnetic sector mass spectrometer. Note that the distance from
the source and collector slits to magnet pole face is equal to the radius
in a 90° deflection sector, which is characteristic of a 90°
magnetic sector with ion entry normal to magnetic field.

< previous page page_462 next page >


< previous page page_463 next page >
Page 463

peaks in the mass spectrum, the collector slit width must be greater than the source slit, and often it is twice the source
slit width to cover broadening or rotation of the ion beam before detection. The most common way to operate a
magnetic sector mass spectrometer is to fix the accelerating voltage and control the magnetic field to scan over the M/e's
of interest. This gives nearly constant sensitivity for ions as a function of mass, but the control of the magnetic field is
slow compared with the rapid changes possible with a quadrupole mass spectrometer or with voltage scanning a
magnetic sector mass spectrometer. Equation (7.19) shows that M/e depends on magnetic field squared. This
compresses the range of magnetic field needed; however, it creates a nonlinear mass scale if the magnetic field is
scanned linearly. Figure 7.11 shows a voltage-scanned mass spectrum for the mass spectrometer in Fig. 7.9 which uses
a permanent magnet. Note that the fixed resolving power of the magnetic sector analyzer produces peak widths, ∆M/e,
that are narrow at low mass and get wide at high mass.

Magnetic analyzers offer the advantage of a simple, fixed geometry that controls the resolving power and ion
transmission of the analyzer. This results in very stable sensitivity over time which is needed for quantitative
measurements. Magnetic sector analyzers are not as popular as quadrupoles because the necessary magnetic field can
interfere with the application; to achieve useful resolving power, the size of the analyzer can be large. Also, any ion that
scatters with a gas molecule in the analyzer loses enough energy to usually be lost from the mass resolved beam, so the
analyzer pressure needs to be kept low for quantitative measurements.

Fig. 7.11
A computer-controlled analog mass spectrum made by voltage scanning a
magnetic sector analyzer. The mass scale is linearized; evidence of near-constant
resolving power m/∆m of a magnetic sector analyzer is seen with the narrow peak
width (∆m) at low mass compared with the peak width at higher mass [33].

< previous page page_463 next page >


< previous page page_464 next page >
Page 464

7.3.3
Time-of-Flight Mass Analyzer

The concept of time-of-flight (TOF) of ions as the basis for mass separation is attributed to Stephens [36], with a
prototype TOFMS developed by Wiley and McLaren [37]. A schematic of the operational features of a linear TOFMS is
shown in Fig. 7.12. Ions are formed in an open source region continuously or by (pulsed) electron impact followed by
extraction of ions with a drawout pulse and acceleration with an applied voltage, V, resulting in a drift velocity of the
ions given by

The drift time for each ion M/z (referenced to the time that the drawout potential is applied) is

where D is the length of the flight tube. Arrival times at the electron multiplier detector vary with the square root of M/z;
for a 1-m tube and 3-kV acceleration, the arrival times range from 1.9 µs to 13 µs for M/z = 2 and 100, respectively.
Measurement of a mass spectra is accomplished by recording the ion current in a time window ∆t recorded at a variable
time delay, tdelay. By varying the time delay linearly, tdelay = At + t0, from 1 to 15 µs and recording the ion current
arriving in a 10-ns time window, a mass spectrum covering mass 2100 is recorded (for a TOFMS with D = 1 m; V = 3
kV). The mass scale of this linear time base will be proportional to the square root of mass according to Eq. (7.22).
Multiple ion monitoring is achieved by setting multiple time delays with measurement windows, ∆t, adjusted to capture
the whole peak width. The time delays are set to measure selected masses. With the multiple delays, it is possible to
measure ion currents of many species formed at a single ionization event. Repetition time for replicating the ionization
process is limited by the longest drift time of ions being measured (clearing time) and the duration of the ionization
process. With continuous ionization, ions are formed during the mass analysis of the previous pulse. With pulsed
electron impact, the electron beam is turned on for a desired time period followed by ion extraction. Repetition rates
vary from 10 kHz to 100 kHz depending on mode of operation, mass

Fig. 7.12
A linear time-of-flight analyzer mass spectrometer.

< previous page page_464 next page >


< previous page page_465 next page >
Page 465

range, and TOFMS design. The resolving power for a TOFMS is measurable on a time spectrum as

where tM is the time of flight for an ion M and ∆tFWHM is the full-width at half maximum spread in arrival times for
the ions. The origin of the peak width, ∆tFWHM, in the mass spectrum is mainly due to the variation in ion energy, ∆V,
from the finite width of the electron beam forming the ions that are drawn out by the extracting field, E. An alternate
expression for resolving power is

where for ion acceleration V = 3 kV, extraction field, E = 150 V/cm, and the electron beam width in the source, ∆d = 0.1
cm, a resolving power of 200 is predicted. The development of the TOFMS has been limited by this modest resolving
power and the cost of fast measurement circuits to record and process microsecond time frame signals. Time-of-flight
mass spectrometers were popular in the 1960s for many applications including monitoring of fast reaction kinetics.
Recent developments have resulted in improvements in both aspects. Resolving power can be improved by employing
energy focusing methods like the ''reflectron" [38], where an electrostatic mirror is used to focus ions with slightly
different energies to arrive nearly simultaneously at the detector, thereby reducing the width of the peak, ∆tFWHM.
Availability of relatively inexpensive electronics with fast (ns) response and data storage electronics has revived an
interest in time-of-flight mass spectrometry for a variety of applications including partial pressure analysis.

7.3.4
Trochoidal (Cycloid) Mass Analyzer

The double-focusing properties of crossed electric and magnetic fields were first described by Bleakney and Hipple
[39], with possible ion motions described as trochoidal paths. A commercial instrument shown schematically in Fig.
7.13 was developed by Robinson and Hall [40] where the path traced is a prolate cycloid with the spacing between
source slit and detector, D, equal to 2.7 cm. Ions are formed and analyzed within a region of electrostatic field, E, with a
magnetic field perpendicular to the plane of ion motion generating a mass-dependent cycloidal motion given by

Mass analysis can occur by scanning the magnetic field, B, or by changing the electric field, E, due to the applied
voltage V2 V1 shown in Fig. 7.13 on the electric field plates in the figure. The excellent focusing of this design gives
stability and high sensitivity for ion detection with resolving power up to 400 in a very short ion path, making it useful
for quantitative measurement of partial pressures [41]. Permanent magnet versions of this instrument were made with a
(relatively) open ion source for residual gas analysis and in closed ion source versions for analytical measurements of
gas samples. For RGA applications the source within the magnetic field region limits use to an appendage mounting.
Also, the electrode structure is complex and has a lot

< previous page page_465 next page >


< previous page page_466 next page >
Page 466

Fig. 7.13
A cycloidal mass spectrometer.

of surface area for outgassing, which limits its usefulness for UHV applications. The location of the ion detector does
not allow a traditional electron multipler detector to be used, although a microchannel plate detector might now be used.
Cycloidal mass spectrometers have not been made since the early 1970s because of some of these limitations;
instrument companies have instead focused on quadrupole mass spectrometers as a more flexible design.

7.3.5
Omegatron

The omegatron was developed by Sommer, Thomas, and Hipple for measuring atomic constants [42] and has been used
by Alpert and Buritz [43] as a mass spectrometer for measurement of partial pressures in an ultrahigh-vacuum system.
A schematic diagram of the omegatron developed by Alpert and Buritz is shown in Fig. 7.14. The operation of an
omegatron is similar to a cyclotron. Ions are formed by electron impact ionization in the center of a uniform magnetic
field. A resonant ion traces a spiral path as the ion of mass-to-charge ratio M/e gains energy from an rf field with each
orbit within the magnetic field. The condition for cyclotron resonance of an ion is

where B is in gauss and M/e is in amu. The resolving power of the omegatron is
< previous page page_466 next page >
< previous page page_467 next page >
Page 467

Fig. 7.14
Schematic diagram of a simplified version of the omegatron
developed by Alpert and Buritz [43].

where R0 is the distance from the ionization center to the detector (cm), and E0 is the magnitude of the rf field (V/cm).
In the Alpert and Buritz omegatron, the mass range from 1 to 40 amu was covered with a frequency range of 3.2 MHz
to 81 KHz, respectively, with a magnetic field of 2100 gauss. With an rf electric field of E0 = 1 V/cm and R0 = 1 cm, a
maximum resolving power of 212 at M/e = 1 is achieved and 5.3 at M/e = 40. The resolving power of an omegatron
clearly favors measurement of low mass ions. Recent use of an omegatron is seen in the measurement of compositions
of hydrogen and helium isotopes for fusion-related applications.

7.4
Optical Measurement of Partial Pressures

Optical measurements of partial pressures of selected gases in a vacuum system have been employed in specific
applications where a noninvasive measurement is needed. The advantages of optical measurement methods are that the
measurement is speciesspecific and does not introduce reaction species like hot-filament measurement devices.
Advances in laser technology that have produced high-intensity laser beams with a wide range of wavelengths have
allowed the development of multiphoton photoionization processes to produce ions and resonant absorption methods for
measurement of impurity species in process gases and in vacuum systems.

< previous page page_467 next page >


< previous page page_468 next page >
Page 468

7.4.1
Photoionization Measurement of Partial Pressure

The use of laser multiphoton ionization with mass analysis to detect low-pressure gas species had been reported by
many researchers [44, 45]. A quantitative study on measurement of CO partial pressures using resonance-enhanced
multiphoton ionization with a time-of-flight mass spectrometer (Fig. 7.15) has been reported by Looney [46, 47]. The
CO partial pressure is measured by resonance excitation of the CO molecule with two 230-nm photons followed by
single-photon ionization of excited-state CO molecules. The ionization process is shown schematically in Fig. 7.16. A
linear formation of CO+ ions with CO partial pressure has been demonstrated over the pressure range 107 to 104 Pa
with uncertainties of ± 1015% limited by repeatability of laser pulse energy [47]. This method has also directly
measured the production of CO by hot filaments of ionization gauges and RGAs. Reduction of CO in the vacuum was
measured when gauges were sequentially turned off. The photoionization-based measurements where the beam passes
through the chamber do not induce reaction products but are currently limited to selected molecules with a dipole
moment (CO, NO). Nonresonant photoionization with multiphoton absorption using 248-nm photons from a KrF
eximer laser has been demonstrated for numerous species [45] with demonstration of linear production of ions for Ar,
Kr, and Xe over the pressure range 106 to 103 Pa for fixed laser power [44, 45]. Data indicate an ion current production
relation of the form

where is the ion current of the ith component formed with cross section σi from a laser intensity ILaser for
component partial pressure, Pi. The exponent of laser

Fig. 7.15
Schematic for a resonance-enhanced multiphoton ionization apparatus.
< previous page page_468 next page >
< previous page page_469 next page >
Page 469

Fig. 7.16
A schematic energy diagram for three-photon
ionization of CO. Here λ is the two-photon
resonance absorption probability, β is the
probability for single photon ionization from
the n* excited state, and A is the probability
for radiative decay from the n* state.

intensity, n, is 2 or 3 depending on laser power and accidental occurrence of wavelength combinations, giving a
resonant excitation. As high-power UV lasers become more cost effective, the use of nonresonant photoionization for
partial pressure measurement will probably increase.

7.4.2
Infrared Absorption Measurement of Partial Pressure

Infrared (IR) absorption has been extensively used to measure part-per-billion levels of H2O, CO, CO2, CH4, NO,
N2O, NO2, NH3, and SO2 impurities in N2, Ar, He, and H2 process gases and in air [4850]. The method requires that
the impurity to be measured must have IR absorption at a different wavelength than any absorption by the matrix gas. A
variety of gas-phase IR spectroscopies have been used for measuring impurities in matrix gases [48]. The ppb detection
limits of these methods suggest at least 104 Pa detection capability for partial pressures in a vacuum by IR absorption.
The infrared transmitted through an absorber is described by a Beer's law expression:

where I0 is the incident intensity, ai(v) is the absorbance (Pa1·cm1) at frequency v, L is the path length in cm, Pi is the
partial pressure of the component being measured. Measurement of partial pressure of a species has been done by
modulating the frequency of a tunable diode laser and phase detecting the Ii(v). From the signal and a calibrated
absorbance and cell path length, component partial pressures are measured [49, 50].
< previous page page_469 next page >
< previous page page_470 next page >
Page 470

Fig. 7.17
Schematic of a cavity ring-down measurement system for measurement of partial
pressure of H2O.

Application of IR absorbance measurements to vacuum systems has produced commercial instruments for measurement
of H2O, organics, and metal organic vapors for chemical vapor deposition for thin films [51, 52]. The measurement of
water vapor partial pressure in a vacuum system with an RGA has uncertainties caused by induced desorption by the hot
filament or H2O production from filament reactions [8]. Water vapor has a strong IR absorption band such that IR
spectroscopy is a inviting technique to use. Standard IR absorption spectroscopy for gases uses a gas cell and reference
beam to measure absorbance. However, with the high level of water vapor in the atmosphere, fluctuations of water
vapor in the reference beam and sample beam paths can cause measurement errors that are difficult to compensate. A
method termed cavity ringdown spectroscopy (CRDS) has been developed which measures the rate of IR absorption by
water vapor in the sample chamber only [5355].

A schematic of CRDS apparatus for measuring partial pressure in a vacuum system is shown in Fig. 7.17. The key
element is the mirrored cavity with reflectance ≥ 0.9999 for each mirror. A pulsed laser source of light is used to initiate
the measurement process. A fraction of the incident IR light enters the cavity and is multiply reflected. A small (104)
fraction exits to the detector, which allows monitoring of the decrease in the internally reflected signal. By
differentiating Beer's law [Eq. (7.29)] with respect to path length (number of round trip reflections), a relationship can
be defined between the decrease of monitored signal with time (number of round trips) to the partial pressure of water
vapor present. The beauty of the method is that the measurement of water vapor in the cavity (and vacuum system) is
independent of the initial laser pulse intensity and external water vapor and depends only on absorption within the
cavity.

7.5
Computer Control, Data Acquisition, and Presentation

Most mass spectrometers used for residual gas analysis or process monitoring are controlled by a computer with
software provided by the manufacturer or developed by the user. Methods for data acquisition, intensity measurement
algorithms, and display of mass spectra are often proprietary by the manufacturer or, if developed by an instrument user,
reflect the particular needs of the user. The fundamental data

< previous page page_470 next page >


< previous page page_471 next page >
Page 471

provided by a mass spectrometer is an analog scan of the mass spectrum. This records the raw ion current versus mass
and displays it on a computer screen or chart recorder. An example of the analog mode is shown in Fig. 7.11 for a
magnetic sector RGA. The bar graph display mode gives a software-interpreted intensity of ions at an M/e which is
displayed at the mass as a vertical line whose height represents the ion current as shown in Fig. 7.18. Data for selected
peaks in a mass range can also be displayed in a table mode. The single intensity number for each M/e detected
represents the ion current correctly only if the mass scale is well calibrated, and the algorithm for measuring peak ion
current covers all shapes of peaks that can be encountered in the mass spectrum that the software interprets. This
intensity versus M/e data can also be printed out or used in spreadsheets. Many RGA manufacturers produce control
point outputs or switches where the switch status is controlled by a user-selected ion current level or range for a chosen
M/e. This allows process changes to be made based on the measured ion current of the selected M/e. The third common
software mode of software control for RGAs is multiple-ion monitoring. This is an extension of the bar graph mode
where the ion current of selected M/e components are monitored periodically and their ion current displayed as a
function of time to record trends in composition. This multiple ion monitoring mode is especially useful for process
monitoring and control.

7.6
Residual Gas Analysis

The dominant use of partial pressure analyzers (PPAs) is for measuring residual gas composition in a vacuum system. In
this mode, a PPA with an open ion source is used

Fig. 7.18
Multiple modes of computer-controlled data display: (1) Multiple ion detection showing
selected ion currents as a function of time. (2) Bar-graph display of mass and associated
maximum ion current determined by peak measurement algorithm. (3) Table of data giving
mass and associated maximum ion currents for all peaks found in a mass
range (Leybold-Inficon, Inc.).

< previous page page_471 next page >


< previous page page_472 next page >
Page 472

with the ion source immersed directly into the vacuum system in a manner and location similar to the use of an
ionization gauge.

In mounting the PPA on a vacuum system, the port used must have good conductance to the vacuum system being
measured, and its location should minimize interference with other gauges. Simple on/off tests with other gauges
identifies any interference problems. A primary function of this mode of operation is to diagnose the condition of the
vacuum system by identifying the components of the residual gas as a pumpdown progresses or to diagnose leaks and
contamination problems. In Fig. 7.19, there are three spectra, measured with a QMS using a Faraday detector, that can
be used to determine the components of the total pressure. Note that the display is logarithmic in ion current from 1014
to 109 A to more visually display all components detected. The top spectrum shows a pattern of peak groups with high
mass of each group spaced 14 amu from the next group. This is characteristic of hydrocarbon oils where fragment
groups differ by a mass corresponding to the mass of CH2. The source of this hydrocarbon contamination is
backstreaming of forepump oil perhaps during a long period of foreline pumping without adequate traps. A second
spectrum in the middle of Fig. 7.19 shows evidence of an air leak in the vacuum system. The largest peak is at mass 28
and is probably N2 since there are other air components like O2 at mass 32, argon at mass 40, and carbon dioxide at
mass 44. The significant isotopes 15N and 18O are evident as 15N14N+(29), 16O18O+(34); however, the species at
mass 30 has an abundance too large to be and is probably NO+ generated as an artifact of the ionization process.
There is also evidence of water vapor at mass 18 with many fragment ions nearby and at lower masses:
Ar2+(20), F+(19), OH+(17), O+(16), 15N+(15), N+(14), C+(12), O2+(8), N2+(7), He+(4), H+(1). The lower
spectrum in Fig. 7.19 shows a clean vacuum system with a base pressure of 2 × 107 Pa. This spectrum is dominated by
H2O+(18) with an associated OH+ fragment ion at mass 17 and CO+ at mass 28. The remaining peaks are
which are characteristic of a clean stainless steel vacuum system.

Selection of a filament for an RGA depends on the application and dominant gas present during use. Common filament
materials for PPAs are thoria- or yttria-coated iridium (ThO2/Ir or Y2O3/Ir), tungsten (W), and rhenium (Re). All of
these filament materials are compatible with inert gases and N2. The thoria (yttria)-coated iridium are all "burnout-
resistant" because they are already oxidized and produce electrons while operating at low temperatures (see Chapter
11). Because of the low operating temperature of these oxide coated filaments, reactions with gases are minimized and
the filaments have a long lifetime. However, the thoria (yttria)-coated iridium can become a source of oxygen for
production of water vapor, CO, and CO2 in the presence of hydrogen [8]. The oxide layer may also entrain water.
Tungsten works best in a reducing atmosphere or for UHV applications. When oxygen is present, the tungsten oxides
formed are volatile compounds which leave the filament, causing thinning and eventual burnout of the filament.
Rhenium at

< previous page page_472 next page >


< previous page page_473 next page >
Page 473
Fig. 7.19
Mass spectra showing (top) hydrocarbon contamination of a vacuum system, (middle)
an air leak into a vacuum system, and (bottom) a clean vacuum system operating at
5 × 107 Pa total pressure (Balzers-Pfeiffer, Leybold Inficon).

< previous page page_473 next page >


< previous page page_474 next page >
Page 474

its normal electron-emission operating temperature evaporates Re atoms and ions, which causes thinning of the
filament. Rhenium can also catalyze chemical reactions on its surface. The lifetime of an Re filament in continuous use
is a few months, whereas a W filament in a high vacuum or reducing atmosphere can last for years.

7.7
Pressure Reduction Sampling Methods for Vacuum Process Analysis

A common use of RGA and closed-source mass spectrometers is measurement of partial pressures or gas compositions
in a vacuum process where the process gas pressure exceeds the normal operating pressure of the mass spectrometer.
For these applications, a pressure reduction apparatus is needed to accomplish the transition. For convenience, this
transition is often done with a variable leak valve. This is suitable for pressure reduction to allow qualitative viewing of
the composition of the gas in the higher-pressure region. However, the mass dependence of the flow through the
variable leak valve depends on the smallest "critical" dimension of the leak. Figure 7.20 shows the type of flow
(molecular, transition, or viscous) for a given critical dimension as a function of the upstream pressure [56]. For a fixed
dimension, the flow can be molecular, transition, or viscous depending on the upstream pressure. It is desirable to
establish molecular flow into the mass spectrometer ionization region where exit flow is normally molecular. This
results in a simple, mass independent pressure reduction expression between process pressure and ion source pressure:

Fig. 7.20
Diagram to indicate the flow regime of gas in a process by position of the
coordinate of the pressure and critical dimension that limits the flow rate
in the process.

< previous page page_474 next page >


< previous page page_475 next page >
Page 475

where Ain is the area of the inlet orifice. For a closed ion source, Asource is the sum of the area of the electron entrance
port and are of the ion exit hole. For an RGA whose housing is pumped to reduce the ion source operating pressure,
Asource is an effective area related to the pumping speed of a RGA housing. By calibration of the ion source to measure
partial pressures, process pressures can be inferred knowing the ratio of conductances into and out of the ion source for
gas in molecular flow. Equation (7.30) becomes more complicated if the flow into the ion source is not molecular and
requires process-specific calibration. The use of fixed conductances defines a fixed factor relating the pressures while
the use of a variable leak requires remeasurement with each setting. Some common fixed-conductance reduction
methods include: (1) molecular flow orifice reduction used for the pressure regime from 103 Pa to 100 Pa for most
gases, (2) long, narrow-bore capillary tubes used as a viscous flow restrictive element for pressures near an atmosphere
or greater, and (3) a capillary tube for viscous flow to a pumped interstage region where the reduced-pressure gas is
sampled by molecular flow through an orifice into the mass spectrometer. In the latter method, the composition at the
low-pressure interstage reflects the composition of the process gas, and the introduction into the mass spectrometer
gives a mass independence flow for species as predicted by Eq. (7.30).

7.8
Calibration of Partial Pressure Analyzers

Calibration of a mass spectrometer consists of establishing a correspondence between the change in a representative ion
current due to a change in partial pressure of the gas which produces the ion [57]. Also see Section 12.3.4, which
discusses calibration of mass spectrometers. It is useful to define a sensitivity for each gas species to quantify the
calibration. Sensitivity is defined as the ratio of the change in ion current (Ii Ii0) due to an addition of a known partial
pressure (Pi) for the particular gas species, where Ii0 is background ion current at the mass of interest:

Linear response of the mass spectrometer occurs if Si is a constant. A measure of deviation from linearity is defined as
the largest percent deviation from the average sensitivity over a specified range. In Fig. 7.21, the sensitivity of a QMS is
plotted for measurements from 105 Pa to nearly 1 Pa [57]. The linearity is less than 30% for 3 eV ion energy and 15%
for 8 eV ion energy for source pressures less than or equal to 3 × 102 Pa. Above this pressure, deviations in sensitivity
occur due to changes in ion transmission efficiency from the ion source through the rod assembly to the detector. Using
the definition of sensitivity in Eq. (7.31) and the relation for production of ions in Eq. (7.2), the sensitivity of a mass
spectrometer depends on the following physical parameters:

Emission current, Ie, electron energy, E, and source temperature, Tsource (K) can be kept constant as the source
pressure changes and the ionization cross section, σi(E), is a constant for the species being ionized. This leaves the
expression for ion

< previous page page_475 next page >


< previous page page_476 next page >
Page 476

Fig. 7.21
Relative sensitivity versus ion source pressure for a QMS with differing nonlinear
response depending on ion energy [12]. Linearity for 3-eV ion energy ( ) is 30%,
and for 8-eV ion energy (×) it is 15%, for pressure less than 3 × 102 Pa.

transmission, F , as the term that varies above 102 Pa in Fig. 7.21. For the example of low ion
energy, Eion = 3 eV, space charge (ions and electrons) builds in the ion source region at higher source pressures; this
leads to an increase in sensitivity, reaching a maximum at about 2 × 102 Pa followed by a sharp decrease in sensitivity
as the pressure is increased. This variation of ion transmission with pressure has been the subject of numerous
investigations [1115]. At the higher gas pressures (1021 Pa), ion space charge can exceed the electron space charge and
alter the electric fields in the ion source and in the entrance to the mass analyzer, as well as within the mass analyzer.
Changes in effective ion energy can change the extraction of ions from the source and coupling to the analyzer .
Repulsion of ions and gas scattering can also occur at high ion current densities and at high gas pressures where mean
free paths of ions become of the order of the dimensions of the analyzer structure. In Fig. 7.21, the sensitivity for
operation with 8-eV ion energy, a larger fraction of ions are extracted from the source than for the 3-eV ion energy. Yet
as pressure is increased, there is a gradual decrease in sensitivity. For quantitative measurements, the nonlinear response
regions should be avoided by keeping the ion source pressure low, electron emission low, and ion energy relatively high
consistent with good resolution of peaks [1215].

Sensitivities for each gas of interest should be determined by introducing a known partial pressure of the gas measured
by a calibrated gauge. For example, pure gas introduced and measured by a calibrated total pressure gauge and
corresponding

< previous page page_476 next page >


< previous page page_477 next page >
Page 477

measurement of ion current can be used to establish the sensitivity [Eq. (7.31)]. Gas introduction needs to be low
enough in pressure to be in a linear operating range yet high enough pressure to get good measurements of both ion
current and partial pressure gas addition. Once a set of sensitivities, Si, are determined for the gases of interest, partial
pressures, Pi, present in the ion source can be calculated from the ion currents measured:

The correction term, ΣjRijIj, subtracts fragmentation contributions to the ith ion of interest from all higher massed ions,
Ij, that contribute a peak. Often there are no fragment interferences; however, the formalism raises the question of
interferences at each mass. The fragmentation ratio, Rij, is the ratio of the observed fragment ion abundance to the most
abundant peak measured from a mass spectrum of the pure gas that produces the pattern. Examples of abundances are
given in Table 7.2; however, the values are typically different for each mass spectrometer. In complicated spectra such
as hydrocarbon mixtures, the interferences become significant and the calculation of partial pressures is best handled
with a matrix formalism. For closed ion source mass spectrometers, it is difficult to measure the pressure in the
ionization region directly such that sensitivities can be referenced to the higher process pressure typically measured with
a capacitance diaphragm gauge. This type of calibration calculates directly the partial pressure in the process from ion
currents. Care must be taken to differentiate between ion currents generated from the background of the mass
spectrometer and ion currents from species in the process gas especially when measuring small impurity components.
This is done by modifying Eq. (7.33) to subtract the background ion currents, Ii0 (when no process gas is flowing), and
using sensitivities Si(process) referenced to process pressures for gas addition in Eq. (7.31):

From the partial pressure measurements, gas composition can be calculated relative to the partial pressures measured:

This definition of composition ignores peaks that are not in the summation, so it is important to include appropriate
peaks into the summation. A check for a reasonable inclusion of all components is made by comparing the sum of
calculated partial pressures (ΣiPi) with a total pressure gauge reading of the sample.

References

1. F. H. Field and J. L. Franklin, Electron Impact Phenomena and the Properties of Gaseous Ions, pp. 5762. Academic
Press, New York, 1957.

2. A. von Engel, Ionized Gases, AVS Classics Ser., p. 63. AIP Press, Woodbury, NY, 1994.

3. R. D. Levin and S. G. Lias, Ionization Potential and Appearance Potential Measurements, 19711981, NSRDS-NBS
71. U.S. Department of Commerce, Washington, DC, 1982;

< previous page page_477 next page >


< previous page page_478 next page >
Page 478

and/or S.G. Lias, J. F. Liebman, R. D. Levin, and S. A. Kafafi, Positive Ion Energetics, Version 2.0 (Software
Files). National Institute of Science & Technology (NIST) 1993.

4. NIH Mass Spectral Data Base, Version 1 (Software File). NIST, EPA, Washington, DC, 1995.

5. M. J. Drinkwine and D. Lichtman, Partial Pressure Analysis and Analyzers. AVS Monogr. Ser. AIP Press,
Woodbury, NY, 1979.

6. J. F. O'Hanlon, A User's Guide to Vacuum Technology, 2nd ed. Wiley, New York, 1980.

7. E. V. Kornelsen, J. Vac. Sci. Technol. 13, 716 (1976).

8. P. A. Redhead, J. P. Hobson, and E. V. Kornelsen, The Physical Basis of Ultrahigh Vacuum, AVS Classics Ser., pp.
275280. AIP Press, Woodbury, NY, 1993.

9. F. Watanabe, J. Vac. Sci. Technol. A 8, 3890 (1990).

10. F. Watanabe and A. Kasai, J. Vac. Sci. Technol. A 13, 497 (1995).

11. W. E. Austin, J. H. Leck, and J. H. Batey, J. Vac. Sci. Technol. A 10, 3563 (1992).

12. L. Lieszkovszky and A. R. Filippelli, J. Vac. Sci. Technol. A 8, 3838 (1990).

13. M. C. Cowen, W. Allison, and J. H. Batey, Meas. Sci. Technol. 4, 72 (1993).

14. M. C. Cowen, W. Allison, and J. H. Batey, J. Vac. Sci. Technol. A 12, 228 (1994).

15. M. Li and H. F. Dylla, J. Vac. Sci. Technol. A (in press).

16. J. Blessing, Res. Dev., September (1987).

17. Closed Ion Source, Bull. BR31D38K. Leybold Inficon, Syracuse, NY, 1992.

18. A. O. C. Nier, Rev. Sci. Instrum. 11, 212 (1940).

19. A. O. C. Nier, Rev. Sci. Instrum. 18, 398 (1947).

20. M. G. Ingrham and R. J. Hayden, A Handbook on Mass Spectrometry, Nucl. Sci. Ser., Rep. No. 14. National
Academy of Science, Washington, DC, 1954.

21. H. E. Duckworth, R. C. Barber, and V. S. Venkatasubramanian, Mass Spectrometry, p. 43. Cambridge University
Press, Cambridge, UK, 1990.

22. H. E. Duckworth, R. C. Barber, and V. S. Venkatasubramanian, Mass Spectrometry, p. 70. Cambridge University
Press, Cambridge, UK, 1990.

23. G. F. Metcalf and B. J. Thomson, Phys. Rev. 36, 1489 (1930).

24. E. A. Kurz, Am. Labo., March (1979).

25. J. L. Wiza, Nucl. Instrum. Methods 162, 587 (1979).

26. C. Giffin, R. Britten, H. Boettger, J. Conley, and D. Norris, Ann. Conf. Mass Spectrom. Applied Top., Seattle, WA
(1979).
27. F. A. White, Mass Spectrometry: Applications in Science and Engineering, p. 67, Wiley, New York, 1986.

28. Transpector Gas Analysis System, Manual PN 074-201, Sect. 3. Leybold Inficon, Syracuse, NY, 1993.

29. W. Paul, H. P. Reinhart and U. von Zhan, Z. Phys. 152, 143 (1958).

30. P. H. Dawson, Quadrupole Mass Spectrometry and Its Applications, AVS Classics Ser., pp. 1336. AIP Press,
Woodbury, NY, 1995.

31. D. H. Holkeboer, private communication.

32. W. E. Austin, A. E. Holme, and J. H. Leck, in Quadrupole Mass Spectrometry and Its Applications (J. H. Dawson,
ed.), pp. 121125. AIP Press, Woodbury, NY, 1995.

33. Aero Vac 1000 Brochure. Vacuum Technology Incoporated, Oak Ridge, TN, 1989.

34. F. W. Aston, Philos. Mag. 38, 709 (1919).

35. A. O. Nier and M. B. McElroy, J. Geophys. Res. 82, 4341 (1977).

36. W. E. Stephens, Phys. Rev. 69, 691 (1946).

< previous page page_478 next page >


< previous page page_479 next page >
Page 479

37. W. C. Wiley and I. H. McLaren, Rev. Sci. Instrum. 26, 1150 (1955).

38. B. A. Mamyrin, V. J. Karatajev, D. V. Shmikk, and V. A. Zagulin, Sov. Phys.JETP (Engl. Transl.) 37, 45 (1973).

39. W. Bleakney and J. A. Hipple, Phys. Rev. 53, 521 (1938).

40. C. E. Robinson and L. G. Hall, Rev. Sci. Instrum. 27, 504 (1956).

41. R. E. Ellefson, D. Cain, and C. N. Lindsay, J. Vac. Sci. Technol. A 5, 134 (1987).

42. H. Sommer, H. A. Thomas, and J. A. Hipple, Phys. Rev. 82, 697 (1951).

43. D. Alpert and R. S. Buritz, J. Appl. Phys. 25, 202 (1954).

44. R.-L. Chien and M. R. Sogard, J. Vac. Sci. Technol. A 8, 2814 (1990).

45. K. Kokobun, S. Ichimura, H. Hashizume, H. Shimizu, Y. Oowadano, Y. Matsumoto, and K. Enda, J. Vac. Sci.
Technol. A 8, 3310 (1990).

46. J. P. Looney, J. Vac. Sci. Technol. A 11, 3111 (1993).

47. J. P. Looney, J. Vac. Soc. Jpn. 37, 703 (1994).

48. Y. Ogawara, A. Bruneau, and T. Kimura, Anal. Chem. 66, 4354 (1994).

49. J. A. Mucha, Appl. Spectrosc. 36, 393 (1982).

50. R. S. Inman and J. J. F. McAndrew, Anal. Chem. 66, 2471 (1994).

51. J. A. O'Neill, M. L. Passow and T. J. Cotler, J. Vac. Sci. Technol. A 12, 839 (1994).

52. A. E. Kaloyeros, J. Loan, B. Zheng, I. Lou, J. Lou, and J. W. Hellgeth, Thin Solid Films 262, 20 (1995).

53. A. O'Keefe and D. A. G. Deacon, Rev. Sci. Instrum. 59, 2544 (1988).

54. R. T. Jongma, M. G. H. Boogaarts, I. Holleman and G. Meijer, Rev. Sci. Instrum. 66, 2821 (1995).

55. P. Zaliski and R. N. Zare, J. Chem. Phys. 102, 2708 (1995).

56. D. H. Holkeboer, D. W. Jones, F. Pagano, and D. A. Santeler, Vacuum Technology and Space Simulation, AVS
Classics Ser., p. 4. AIP Press, Woodbury, NY, 1993.

57. J. A. Basford, M. D. Boeckmann, R. E. Ellefson, A. R. Filippelli, D. H. Holkeboer, L. Lieszkovszky and C. M.


Stupak, J. Vac. Sci. Technol. A 11, A22 (1993).

< previous page page_479 next page >


< previous page page_481 next page >
Page 481

8
Leak Detection and Leak Detectors

Werner Grosse Bley

In setting up vacuum systems, one of the most important requirements to fulfill is the leak tightness of the vessel.
Depending on the intended experiment or process, different levels of leak testing have to be applied. These range from
the simple observation of the achievable pressure to the use of tracer gases detected with highly sophisticated mass
spectrometer leak detectors.

Many of these methods, though developed for vacuum science, have now been extended to the nondestructive testing of
a variety of components reaching from big barrels with leakage rates of about 103 mbar·liter·s1 to cardiac pacemakers
with less than 109 mbar·liter·s1. An overview of all relevant leak testing methods can be found in handbooks of
nondestructive testing [1].

The following sections deal with all the different aspects of modern leak detection methods of vacuum systems and
components. It is important to distinguish between vacuum systems, consisting of a vacuum vessel with appropriate
high-vacuum pumps and components such as tubes, valves and fittings without pumping equipment. In systems, leak
detection is often possible with the built-in pressure sensors, whereas for component testing a complete leak detector
with a pumping system is needed. Different gases and pressure ranges have to be dealt with; thus even with modern
automatic test equipment, basic knowledge of the vacuum physics involved is indispensable for reliable results.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_481 next page >


< previous page page_482 next page >
Page 482

8.1
Principles of Vacuum Leak Detection

8.1.1
Types of Leaks and Leak Rate Units

At first sight a leak is always a kind of hole in the wall of a vacuum vessel. However, there can also be porous or gas-
permeable areas permitting an undesired flow of gas, usually air, into a vacuum chamber. Depending on the size of the
leak, the amount of gas flow ranges from less than 1011 mbar·liter·s1 up to more than 1 mbar·liter·s1, which represents
more than 11 orders of magnitude. The gas species and mode of flow influence the amount of gas flowing through a
given leak, so the characteristics of leaks should be known for a prediction of leakage rates under different operating
conditions of a component or system. The pressure-volume units used above were meant to indicate the quantity of gas
flow. This is most convenient in vacuum leak detection, but other units are also used [2].

Gas flow is normally given in units of particle flow (s1), mass flow (kg·s1), or molar flow (mole·s1). All these units
describe the amount of substance moving per unit time through a given cross section. For leak detection in vacuum
systems, the pV throughput of a leak, given in pressure times volume per unit time, is a more convenient unit. Looking
at the equation of state of the ideal gas, written for molar gas flow dn/dt,

where n is the number of moles, R the universal gas constant, T the absolute temperature, p the pressure, and V the
volume, it is clear that the gas flow described by pV throughput is only defined for a given absolute reference
temperature T, often chosen as 296 K (23°C). In practical cases, when leaks have to be found and repaired (not to be
measured), temperature effects can be neglected. (There is, however, an additional temperature effect in practical leak
detection when dealing with permeation gas flow; see below.)

If required, pV throughput can be readily converted into the amount of particles, mass, or moles passing through a leak
by using the above gas equation of state, Eq. (8.1). Different units of pV throughput are given in Table 8.1.

Like gas flow in general, leak gas flow can be viscous or molecular in nature. As has been described before in Chapter
2, the Knudsen number, being the ratio of mean free path of the gas particles and the typical diameter of the leak
channel, determines the type of gas flow. For viscous flow the dynamic viscosity and mean pressure of the gas or
mixture determine the flow conductance whereas in molecular flow molar mass alone is important. Gas flow is always
given by the product of flow conductance and pressure differential. This leads to a relation quadratic in pressure for the
viscous case and linear for molecular flow:

< previous page page_482 next page >


< previous page page_483 next page >
Page 483

Table 8.1. Conversion Factors n for Leakage Rate Units

< previous page page_483 next page >


< previous page page_484 next page >
Page 484

where Q is the gas flow, η is the dynamic viscosity, M is the molar mass, and p1 and p2 are the pressures at either side
of the leak.

Normally, in vacuum leak detection, the pressures on either side of a leak are atmospheric pressure and ''zero" pressure,
respectively. Leakage rates under such conditions are called normalized leakage rates. The above formulae are used if
the acceptance level of leakage in a test is given as a normalized leakage rate but different pressure conditions are used
during the actual test. In this case, one has to make an assumption about the mode of flow. If in doubt, both formulae
should be tried and, to stay on the safe side, the bigger leakage rate taken as the result of the test.

A special kind of leakage can occur through permeable materials such as rubber, plastics or glues present in gaskets or
joints. These are often not recognized as leakage but may contribute a considerable amount of gas flow into the vacuum
chamber. Permeation gas flow is similar to molecular flow in that it is following a linear pressure law:

where A is the cross-sectional area and d is the thickness of the permeable wall.

The permeability is described by a constant P depending on the material and gas species involved. P is the product of
the solubility and diffusion constant of the gas in the permeable material. It is therefore exponentially temperature-
dependent. Numbers for P for various gases and elastomers may be found in Yasuda and Stannett [3] and Laurenson
and Dennis [4].

8.2
Total Pressure Measurements

The simplest way of finding out whether a vacuum system is leaky is by using the built-in total pressure gauges. Three
kinds of tests can be distinguished:

• Ultimate pressure test (lowest achievable pressure in the system)

• Isolated pressure test (pressure rise without pumping)

• Gas spray test with gas-dependent gauge (for leak localization).

All these tests are only useful for large leaks to be found in systems with pumps and gauges, not for the test of
components. During the setup or service of a vacuum system, they are most helpful, especially when no specific leak
detection equipment is at hand.

If the achievable ultimate pressure of a vacuum system is not reached, one can assume a leak. It should be considered,
however, that increased gas desorption from the walls of the vessel (especially from newly inserted components) can
also produce an ultimate pressure higher than normal. If the achievable pressure pe is known from experience and if the
pumping speed S has not changed, the gas flow through the leak can be estimated by the following equation:

< previous page page_484 next page >


< previous page page_485 next page >
Page 485

where p is the ultimate pressure actually reached in the system, pe is the achievable pressure, Qleak is the leakage rate,
and S is the pumping speed at the position of the pressure gauge.

If the achievable pressure or the pumping speed is not known, isolation of the vacuum chamber from the pumping
system can help decide whether a leak exists. In a system with volume V and leakage rate Qleak, the pressure rise is
given by

After closing the pump valve, the rise in pressure should be recorded for a suitable period of time. Because there will
always be an initial pressure rise due to gas desorption, one will have to wait for some time until desorption has come to
an equilibrium with the vapor pressure in the gas phase [5]. Then a further pressure rise due to leaks can be
distinguished. Figure 8.1 shows the three possible cases. Only leaks with gas flows much higher than desorption can be
found in this way.

To find gross leaks in a vacuum system, one may take advantage of the fact that the indication of many vacuum gauges
depends on the gas species present. This is true for heat conduction gauges as well as ionization gauges because it is not
pressure itself but a gas-dependent quantity that is measured in these gauges. For historic reasons the halogen sensor
(alkali ion sensor) should be mentioned which is able to detect very small amounts of gaseous halogen compounds
(refrigerant gases such as Freon, Frigen, etc.).

If a leak is sprayed with a gas different from air, the air flowing through the leak is replaced by this gas. This results in a
change of gauge indication according to the different gas heat conductance or ionization probability or, in the case of
the halogen sensor, an increased amount of alkali ions emitted from a heated platinum wire. Helium or methane (natural
gas) are very suitable for use with heat conductance or ionization gauges because (a) they have much higher heat
conductivity than N2 and (b) their ionization probability is much different from N2 (for helium lower by a factor of 7,
for methane higher by a factor of about 2).

Fig. 8.1
Pressure rise versus time in a closed vessel
after pumpdown. (a) Pure gas desorption,
(b) pure leakage, (c) leakage and gas desorption.

< previous page page_485 next page >


< previous page page_486 next page >
Page 486

In practice, the change of gauge indication is rather small since only the leak gas flow is replaced. For systems with high
desorption flow and high pumping speed the resulting change of indication is often difficult to see. A zero suppression
function or differential indication is provided by some gauge controllers to support leak detection.

8.3
Partial Pressure Measurements

If a vacuum system is equipped with a partial pressure gauge, leak detection is possible in a much more elegant and
more reliable way than with total pressure gauges alone. In vacuum systems, normally a quadrupole mass spectrometer
head is used as a partial pressure gauge. The halogen sensor with alkali ion emission is also a type of partial pressure
gauge, though a very unstable and nonlinear one. For overpressure tests on refrigeration systems ("sniffing"), it has
established itself as a cost-effective instrument.

The type of instrument used most in vacuum systems is a quadrupole mass spectrometer with a mass range starting with
1 or 2 amu and ending at 100 amu. This enables the experimenter to use nearly every available tracer gas from hydrogen
over helium, methane, or argon to krypton or halogen compounds [6]. Each of these gases has a prominent peak in the
range between 1 and 100 amu. If the gas is applied to the outer surface of the vacuum chamber with a concentration
cTG, atm, a certain partial pressure depending on the leak size and the pumping speed for that gas (at the position of the
quadrupole) can be measured inside. A quantitative measure of the total gas leakage rate Qleak is given similar to Eq.
(8.5) for the achievable total pressure:

with the index TG characterizing the tracer gas used. (The tracer gas partial pressure pTG, meas in the vacuum chamber
is measured by the quadrupole sensor head according to pTG, meas = sM TG is the measured
ion current on mass M for the tracer gas and sM, TG is the sensitivity for that gas on mass M).

In most cases, simply the ambient air can be used as the tracer gas: In case of a leak, two peaks on masses 28 and 32,
respectively, show up with the characteristic height ratio of approximately 4:1 corresponding to the concentrations of
nitrogen and oxygen in natural air. A typical spectrum of a leaky vessel is shown in Fig. 8.2. The partial pressure gauge
is also a very powerful instrument for further residual gas analysis. For example, a clear distinction between "virtual
leaks" (i.e., water vapor desorption) and real leaks can be made which would not be possible by watching the total
pressure alone.

8.4
Measurement of Leakage Rates with Helium Leak Detectors

The most convenient way of detecting and measuring leaks is with dedicated instruments such as helium leak detectors.
In principle, these are stand-alone vacuum

< previous page page_486 next page >


< previous page page_487 next page >
Page 487

Fig. 8.2
Mass spectrum of a leaky vacuum chamber.

systems with a mass spectrometer, of either magnetic sector or quadrupole type measuring the partial pressure of helium
or hydrogen as a tracer gas.

Since mass spectrometer leak detectors include a pumping system of definite pumping speed, a calibration of the
incoming helium gas flow in units of pV throughput is possible with suitable reference leaks. Basically, the
measurement of a leakage then follows Eq. (8.7). (In practice, there are different types of pumping and detection
systems in modern helium leak detectors leading to slightly different expressions; see Section 8.8). Helium leak
detectors are simply coupled to a vacuum chamber with their inlet port (for the best measurement position see Section
8.6). Leak detection of components is also possible because a pumping system is provided in the leak detector. This is
described in more detail in the following paragraph.

8.5
Helium Leak Detection of Vacuum Components

Generally, leak detection of components should start with an overall (or integral) test in order to minimize the testing
time. Localization and repair can be very time-consuming and difficult, especially if leaks are not directly accessible
because of complicated design of the component. Only if the integral test has shown that the component is leaky,
localization of leaks can be tried if it seems worthwhile.

In an integral test a definite concentration of helium is applied on the outside surface of the component to be tested
while it is coupled to the leak detector. Except for industrial tests, where a great number of identical parts have to be
tested, in the laboratory a simple plastic bag is most suitable as a hood for such a test.

The test is started by pumping down the component and noticing the background indication of the leak detector (this
may be set to zero, if possible). Then the hood is put over the component, filled with pure helium, and fixed with
adhesive tape in such
< previous page page_487 next page >
< previous page page_488 next page >
Page 488

a way that no gas can escape in short terms. (It is important to keep the coupling flange to the leak detector free of the
hood because it is not to be tested). By keeping the hood slack, the applied helium pressure is rather precisely kept at 1
bar, which is important for determining normalized leakage rates (see Section 8.1). For smaller objects a complete
enclosure is made. On big components it may be convenient to cover only those areas where leaks are suspected,
especially welds and flanges.

The leakage rate indication of the leak detector is watched for a short period of time (usually some seconds). The stable
value is then taken as the overall leakage rate. An increasing indication over some minutes is usually due to helium
permeation through rubber gaskets or other permeable elements. Helium permeation gas flow is normally not regarded
as a leakage rate. Because the permeation of air is much lower than that of helium, the resulting leakage rate under
operating conditions can normally be neglected in rubber-sealed systems.

Once a component has been shown to be leaky in the integral test, repair of the leaks is normally necessary except for
very inexpensive parts. Hence, leak localization is the next step. To find the leaks, after removal of the hood a fine jet of
pure helium is directed on the object's walls while its inner volume is still coupled to the leak detector. If a leak is hit by
the helium jet, the leak gas flow is replaced by helium which is pumped away and detected by the leak detector.

Because flow conditions in the component's volume are normally molecular, a very quick response within a few
seconds or even less can be expected. This enables the tester to get a definite correlation between his spraying action
and the leak response, so the position of the spray gun will be very close to the leak when an indication is noticed. The
finer the helium jet, the more precise the leak localization.

To find small leaks in the vicinity of big ones, it is sometimes necessary to mask or close the big leaks. Masking can be
done by covering leaky areas with some foil or sticky tape. Temporary closing of leaks is best done by applying some
alcohol with a syringe that clogs the leaks. After the evaporation of the alcohol the leak opens again without any further
damage to the system.

To perform a leak test within reasonable time, some understanding of the time aspects of a test is necessary. Pumpdown
time and response time are the crucial parameters.

In component testing, a certain pumpdown time is necessary until the pressure in the test object is low enough and the
leak detector is ready for measurement. This time depends very much on the surface conditions of the test object and on
the type of leak detector involved. In general, the maximum inlet pressure of the leak detector has to be reached by
pumping the component down before any leak indication is possible. As gas desorption from the test object's walls
starts below about 0.1 mbar, the pumpdown curve flattens appreciably below that pressure. The time law of pressure
decrease for volume gas only (i.e., without desorption) is

where p0 is the starting pressure, V is the object's volume, and S is the pumping speed of the leak detector pumping the
test object. If the pressure is only governed by gas desorption from metal walls, the pressure decrease after desorption
has started can

< previous page page_488 next page >


< previous page page_489 next page >
Page 489

approximately be described by the expression

where Q0, des is the initial desorption gas flow at time t0 and S is the leak detector pumping speed.

Although a fast pumping speed of the leak detector accelerates the pump process, the slope of the pumpdown curve is
much less steep with the same pump once the desorption regime has started below 0.1 mbar (see Fig. 8.3). In industrial
leak detection pumping time is lost time. Leak detectors with maximum inlet pressures above 0.1 mbar have an
advantage when rapid cycle testing is required.

When trying to localize leaks by spraying helium, a rapid signal response is important for unambiguous results. For
simple exponential processes, response time is expressed as a time constant (rise to 63% of the equilibrium level or
respective decay of signal). The overall response time is dependent on both (a) the properties of the leak detector in use
and (b) the test object. This means that in most cases it cannot simply be expressed by one time constant alone. Time
behavior can be described by three parameters:

• The electrical response time needed to measure and average the signal, described by a time constant te of the leak
detector's signal processing unit
Fig. 8.3
Pumpdown curves for three different pumping speeds.

< previous page page_489 next page >


< previous page page_490 next page >
Page 490

• The volume response time needed to pump away gas from the object's volume, described by a vacuum time constant
τV

• The diffusion response time needed by the tracer gas to traverse regions of high total pressure, described by a
diffusion time constant τD.

The first time constant is dependent on the leak detector's preamplifier and the averaging by the internal software. The
second time constant depends on both (a) the basic principle and actual mode of operation of the leak detector (see
Section 8.9) and (b) the test object's internal volume. The third is mainly a property of the vessel's dimensions and its
total pressure conditions under test. The electrical and volume response time can be shortened by sacrificing lowest
detection limit. Avoiding the most sensitive amplifier range by looking only at leakage rates above a certain level, a
much shorter electrical signal response can be achieved.

If a big test volume is involved, an additional roughing pump will reduce the vacuum time constant, but it will also
reduce the detection sensitivity for the tracer gas, because of operation in a partial flow mode (see Section 8.6).

The diffusion time constant is important mainly in big vessels with high total pressure. In such cases there is an
optimum total pressure for minimum response time [7]. It should be noted that in such cases helium is a very good
tracer gas because of its high diffusion velocity.

8.6
Helium Leak Detection of Vacuum Systems

Vacuum systems basically differ from components in that they incorporate a roughing or high-vacuum pumping system.
Sometimes big components have to be treated as a system because an additional vacuum pump is necessary to pump
down the volume in a reasonable time. The principal tests, integral tightness and leak localization, are quite the same as
for simple components, but the sensitivity of the test will be appreciably lower because a portion of the incoming tracer
gas is pumped away through the auxiliary pumping system without producing a signal in the leak detector. This can be
favorable when rather big leaks are expected and the leak detector would be driven into overflow nearly all the time. A
careful consideration of the partial-flow sensitivity is necessary, and the connection position of the leak detector to the
vacuum system is of great importance for both sensitivity and response time.

The general partial-flow arrangement found in every vacuum system leak test situation is shown in Fig. 8.4. Because a
leak detector always shows a signal proportional to the incoming flow of tracer gas, the ratio of the total leakage gas
flow QL to the gas flow portion into the leak detector QLD has to be calculated. With a system pumping speed Ssyst
and a leak detector with inlet pumping speed SLD, this is given by

If a partial-flow factor γ = Ssyst/SLD is defined, the indicated leakage rate QLD is correlated to the true leakage rate QL

< previous page page_490 next page >


< previous page page_491 next page >
Page 491

Fig. 8.4
Partial-flow configuration in the leak test of a vacuum system.

To estimate an order of magnitude in practical cases, 1 + γ can be replaced by γ with good accuracy. The reduction of
sensitivity can be several orders of magnitude if the leak detector is connected to a vessel pumped by a high-vacuum
pump. This problem can be avoided by choosing a more suitable position to connect the leak detector.

In principle, a leak detector can be connected to a vacuum system in two ways: in place of the pumping system (full
flow through leak detector) or in parallel with it (partial flow). The full-flow connection gives the best sensitivity and a
true quantitative measurement because all the gas must pass through the leak detector. However, this is only possible if
the total gas flow out of the system can be pumped by the leak detector's built-in pumps. Often this is only possible after
some time is spent in degassing. In addition, some valving is necessary to shut off the pumps of the system under test. In
all other cases, the leak detector has to be connected in a partial flow manner.

Three different locations for the partial flow connection of a leak detector to a high vacuum system are possible: directly
to the high vacuum chamber, between the high vacuum and backing pump, and at the exhaust of the backing pump.
These are shown in Fig. 8.5. In principle, true leakage rate measurements can be made at all three positions after
appropriate calibration. There are, however, certain peculiarities of the system that have to be known to achieve
reasonable results.

Although the most obvious, a direct connection to the high-vacuum chamber is normally not the most effective way to
use a leak detector. Because the pumping speed of the high-vacuum pump is very high compared to the leak detector's
inlet pumping speed, the sensitivity is reduced by some orders of magnitude (large partial-flow factor γ). This means
that the detection limit for leaks is less by the same amount. This is normally unacceptable for leak detection on a high-
vacuum chamber.

There are situations where one has no choice but to connect the leak detector directly to the high vacuum, usually when
adsorbing or gettering pumps such as cryo or sputter ion pumps are installed in an ultrahigh-vacuum (UHV) chamber. In
these cases, however, there is the risk of contaminating a clean UHV chamber by hydrocarbons diffusing backwards
from the leak detector inlet port into the vacuum

< previous page page_491 next page >


< previous page page_492 next page >
Page 492

Fig. 8.5
Positions for the connection of a leak detector to a high-vacuum system:
(a) directly to high-vacuum chamber, (b) between high vacuum and
backing pump, (c) at exhaust of backing pump.

under test. Only "oil-free" or "dry" leak detectors (see Section 8.8.5) can avoid this problem. The best solution in such
cases is to install a quadrupole mass spectrometer on the system and use it not only for residual gas analysis but also for
leak detection (see Section 8.3).

Normally the pumping system of a vacuum chamber consists of several pumping stages or separate pumps. This makes
it possible to use the optimum connecting position for the leak detector. With a two-stage pumping system, the optimum
position is between the high vacuum and the backing pump. Here, the pumping speed of the backing pump is moderate
(moderate partial flow factor γ) and the intermediate volume between the pumps is small enough to keep sufficient
sensitivity at a reasonable time constant. Additionally, the high-vacuum system cannot be contaminated by
backstreaming oil because the high-vacuum pump's compression ratio will prevent that.

If there is no appropriate access either to the high-vacuum chamber or between the high-vacuum pump and the backing
pump, the leak detector has to be connected at the exhaust of the backing pump. Because the exhaust pressure is
atmospheric, a sniffing device has to be used for that purpose. In the simplest case, this is a capillary of appropriate
length and diameter in order to reduce the pressure from atmospheric to the maximum inlet pressure of the leak
detector. The sensitivity is greatly reduced by the partial-flow factor given by the ratio of the flow of the sniffer tip to
the total gas flow.

< previous page page_492 next page >


< previous page page_493 next page >
Page 493

8.7
Special Methods and Other Tracer Gases

Tracer gas leak detection with mass spectrometer leak detectors is such a versatile and sensitive method that it is used
not only for vacuum systems or components but also in nondestructive testing of gas pressurized or hermetically sealed
parts in modern production processes. A short description of such test methods will be given in the following.

Hermetically sealed parts such as electronic packages or relays cannot be connected to a leak detector and tested by
spraying them with tracer gas. In some cases they can be filled with tracer gas (especially helium) during the
manufacturing process, but even this is not possible for integrated circuit packages. To test such devices for tightness,
the so-called "bombing test" was developed. The parts are placed in a chamber where they are pressurized with helium
for several hours. If a leak exists, a certain amount of helium will enter the inner volume of the test object. Afterwards,
the part is taken out of the pressure chamber; and after some waiting time, necessary for desorbing helium from the
outer surfaces, the part is placed in a vacuum chamber connected to a leak detector. Now the escaping helium can be
detected, and even a quantitative measurement is possible with some computation and calibration [8]. The detection
limit is somewhere in the 108 mbar·liter·s1 region.

Many industrial systems have to be leak-tight under pressure. Often these parts cannot be evacuated or have no
appropriate connection flanges for a leak detector. To test such objects, sniffing devices were developed for helium leak
detectors. As mentioned earlier, the simplest way to make a sniffer for a helium leak detector is to use a capillary or a
hose with a fine opening at one end and connect it to the inlet port of a leak detector. To achieve maximum sensitivity,
the gas flow through the hose should produce the maximum tolerable inlet pressure for the leak detector. If the tip of
this device is slowly moved across the surface of the pressurized object, escaping tracer gas will be detected and
indicated by the leak detector. Detection limits of 1 × 106 mbar·liter·s1 and even lower can be achieved.

Because pressurized objects are often filled with gases different from helium, one would like to perform leak tests just
by using the filling gas itself. Recently, the introduction of new refrigerant fluids has accelerated the development of
mass spectrometer sniffing leak detectors for those gases. In principle, they do not differ from the helium sniffing leak
detectors except that gases with higher masses can be detected only with quadrupole mass spectrometers. There may be
severe problems in sensitivity or mutual interaction of peaks, so each gas species has to be treated separately.

For vacuum leak detection, tracer gases different from helium are not very convenient because either they are present in
the ambient atmosphere in rather high concentrations or their properties (toxicity, inflammability, etc.) do not make
them very suitable for spraying into the environment. Gases such as methane, CO2, argon, or other noble gases may be
used for leak detection of vacuum systems if a quadrupole mass spectrometer is present and no helium is available.

8.8
Mass Spectrometer Leak Detectors

Mass spectrometer leak detectors are units containing a mass spectrometer and a high-vacuum pumping system
producing the necessary vacuum for the mass

< previous page page_493 next page >


< previous page page_494 next page >
Page 494

spectrometer (below 104 mbar) and pumping the tracer gas in a stable and reproducible way. Although mass
spectrometers have been known since the beginning of this century, the early ones were big and expensive systems
requiring a vacuum pressure below 106 mbar.

The solution to many problems came about with the invention of the helium leak detector based on a small mass
spectrometer. The need for the detection of really small leaks in an industrial plant was motivated by the requirements
of the uranium enrichment technique necessary for producing the atomic bomb in the United States in the 1940s [9].

The next sections dealing with different types and specifications of helium leak detectors will discuss problems in the
design of a leak detector vacuum system and the state of the art today.

8.8.1
Mass Spectrometer System for Helium Leak Detection

Dedicated magnetic sector mass spectrometers with a mass range from 2 to 4 amu are used in helium leak detectors
most frequently. These spectrometers are designed to be of high sensitivity and good resolution even at high pressure
(up to 103 mbar) and under rather dirty vacuum conditions (oils and vapors from pumps and test objects).

The first helium mass spectrometer for industrial use was designed by Nier [10]. It was an all-metal magnetic sector unit
of reasonable small size that could really be used under industrial conditions. The mass spectrometer itself had all the
basic features known about modern helium spectrometers. It was connected to the test object by means of a leak valve,
while the object was pumped by an auxiliary pump. The response time in such a topology was rather long because there
was nearly no pumping speed at the inlet port of the arrangement. Nevertheless, a response ''within a few seconds" at
that time was judged to be "a rapid response."

8.8.2
Direct-Flow Helium Leak Detectors

For some time, helium leak detectors were operated in a partial flow mode. In this mode a portion of the tracer gas does
not reach the detection system (the mass spectrometer plus its pump system) and the sensitivity is limited. On the other
hand, the total pressure in the test object can be rather high, resulting in short pumpdown times.

The introduction of a liquid nitrogen (LN2) trap made it possible to open the inlet valve wider without having to wait
too long for water vapor desorption to become low. Now a "crossover" from roughing to full-flow measuring was
possible. This means that the total gas flow including all tracer gas from the test object flows through the leak detector's
high-vacuum system and is measured by the mass spectrometer. The unit is a "direct-flow" leak detector. The vacuum
schematic of a modern direct-flow leak detector is shown in Fig. 8.6. In such an arrangement the partial pressure of
helium in the mass spectrometer, pMS, He, generated by a given leakage rate QHe flowing into the leak detector is
given by

where SHV, He, is the high-vacuum pump's speed for helium.

< previous page page_494 next page >


< previous page page_495 next page >
Page 495

Fig. 8.6
Vacuum schematic of a modern direct-flow leak detector (Leybold model UL400).
To find out the crossover point, first the bypass valve is opened to check for pressure
rise in the mass spectrometer. Only if the pressure rise keeps below a certain level,
the inlet valve is opened for direct-flow operation.

The intrinsic sensitivity sDF of a direct-flow leak detector, defined as the ratio of the partial pressure of tracer gas
(generated in the mass spectrometer) to a given leakage rate [11], can therefore be expressed by the equation

In order to achieve enough sensitivity in a direct-flow unit, the high-vacuum pumping speed for helium, SHV, He, is
throttled to typically 1050 liter·s1 and can even be throttled by another factor of 10 to achieve ultimate sensitivity
("throttle valve" in Fig. 8.6).

The overall sensitivity (defined as the ratio of output voltage to input leakage rate) is the product of several terms. sDF
is only one of these (see Section 8.9).
The "crossover"that is, the change in pumping conditions from roughing to measurementis always the most critical
problem to be dealt with in a direct-flow leak detector. That is why numerous solutions were developed to make it as
easy as possible for the user. In early units, the user had to manually open an inlet throttling valve while carefully
observing the pressure in the mass spectrometer to keep it below 104 mbar. This procedure works best with LN2 in the
trap, but there is no protection for the filament if a sudden air inrush takes place. That is why solenoid valves were

< previous page page_495 next page >


< previous page page_496 next page >
Page 496

introduced which are automatically opened at the crossover point (expressed either as pumping time or as inlet pressure)
and closed rapidly if the pressure rises. Now a new problem arises: Depending on the amount of water vapor on the
inner surfaces of the test object, the crossover point is not known in advance. For production testing of the same parts in
quick cycles, the crossover point can be found and set by trial and error. For a variety of objects exceeding a certain
volume, this method is not practicable either with or without LN2 because the crossover pressure will vary with
different rates of water vapor desorption.

To overcome those crossover problems, pressure-controlled piezo- or motor-driven inlet valves allowing continuous
flow control are used in the latest direct-flow leak detectors. The leak detector in Fig. 8.6 uses the bypass valve for
checking the pressure before crossover. Although, after these developments, direct-flow leak detectors are reliable units,
their reliability is very much dependent on the treatment they get from the operator. Frequent use without LN2, testing
of "dirty" (i.e., oily) objects, and switch-off without obeying the proper warming-up procedure of the LN2 trap can lead
to high service expenses. So there was more and more demand for leak detectors that can really operate without LN2.

8.8.3
Simple Counterflow Helium Leak Detectors

The solution of LN2 problems came about when the so-called "counterflow" of helium was developed [1214].
"Counterflow" is based on the property of all molecular pumps, diffusion and turbomolecular pumps, to have a limited
compression ratio for the gases pumped. This means that if a certain gas species with a given partial pressure pFV is
present on the fore-vacuum side, there will be also a definite partial pressure pHV of that gas on the high-vacuum side.
The relation of these partial pressures is given by the zero-flow compression ratio K0 of the molecular pump

where n is the rotational speed of the pump (or the vapor jet speed in a diffusion pump), α is a proportionality constant
containing the geometrical data of the pump, and M is the molar mass of the gas involved.

As Eq. (8.14) shows, K0 depends exponentially on the square root of the molecular weight. This means that for a light
gas like helium the compression ratio is orders of magnitude smaller than for nitrogen or water vapor emerging from a
test object. Connecting the test object to the fore-vacuum line of a leak detector therefore results in a measurable helium
signal in the mass spectrometer, whereas the total pressure in the spectrometer is kept sufficiently low. The sensitivity
of the leak detector can be adjusted by the parameter nthat is, the rotational speed of a turbomolecular pump or the
heating power of a diffusion pump, respectively.

The basic vacuum schematic of a simple counterflow leak detector is shown in Fig. 8.7. In such an arrangement, the
partial pressure of helium pMS, He in the mass spectrometer generated by a given leakage rate QHe flowing into the
leak detector is given by

where SFV, He is the fore-vacuum pump's speed for helium.

< previous page page_496 next page >


< previous page page_497 next page >
Page 497

Fig. 8.7
Basic vacuum schematic of a counterflow leak detector
with turbomolecular pump.

The intrinsic sensitivity sCF of a counterflow leak detector, expressed by the relation of the partial pressure of tracer gas
generated in the mass spectrometer by a given leakage rate, may therefore be expressed analogous to Eq. (8.13) as

Comparing this equation with Eq. (8.13) for the direct-flow leak detector, it is obvious that a counterflow leak detector
with the same intrinsic sensitivity as a direct-flow one can be built by making (SFV, He·K0) in the counterflow machine
equal to sHV, He in the direct-flow one. As a typical example a fore-vacuum pump of 0.3 liter·s1 is combined with a
turbomolecular pump with reduced speed to yield a helium compression ratio of 100. In this case we have (SFV, He·K0)
= 30, which is the same as for a typical direct-flow leak detector. If the follow-up detection system is also the same, the
total sensitivity (see Section 8.9) is equal.

When the test object is connected to the fore-vacuum line no "crossover" problems arise. Once the pressure has dropped
below the tolerable limit of the high-vacuum pump, the fore-vacuum valve is opened and measurement is possible. This,
along with no need for an LN2 trap, made possible a completely new generation of helium leak detectors which could
be fully automated because of the ease of the pumping cycle. This makes them simple and rugged enough to be used not
only by scientists but by everybody in industry.

Like all technical systems, the counterflow leak detector is not free of disadvantages. The weakest point is the fore-
vacuum pump now pumping the test object determining the signal stability and the response time for the helium signal.
Because its pumping speed is low, only small test objects can be tested in the reasonable time of some

< previous page page_497 next page >


< previous page page_498 next page >
Page 498

seconds. The detection limit, given by the signal stability, is also limited to some extent (see details in Section 8.9).
Special fore-vacuum pumps with high pumping stability are therefore necessary to build a good counterflow leak
detector.

In the direct-flow leak detector the high-vacuum pump with much higher speed is pumping on the object after the
crossover. Instabilities of the fore-vacuum pump are not contributing to output signal stability because of the high
compression ratio of the high-vacuum pump. That is why for very low detection limits (below some 1011 mbar·liter·s1)
a strongly throttled direct-flow unit with an LN2 trap is still unbeatable.

To combine the simplicity and ruggedness of the counterflow principle with the speed and sensitivity of the direct-flow
principle, advanced counterflow leak detectors were developed. They involve a more complicated pumping and valve
system.

8.8.4
Advanced Counterflow Helium Leak Detectors

One of the first advanced counterflow helium leak detectors was the cabinet model UL500 (Leybold AG, Germany),
whose vacuum diagram is shown in Fig. [8.8]. The heart of this system is a twofold turbomolecular pump [15]. It
incorporates two pumps in one housing pumping in opposite directions and backed by one fore-vacuum pump. The
former high-vacuum side is pumping the leak detector's inlet port, thus boosting the backing pump's speed. This
generates an extraordinary high inlet pumping speed of about 12 liter·s1 for helium. The pumped helium is diffusing
backwards through the other half-pump into the mass spectrometer in a counterflow

Fig. 8.8
Vacuum schematic of advanced cabinet counterflow leak detector for big volume
testing (Leybold model UL500). The turbomolecular pump has two separate sets of
stages in one housing for high inlet pumping speed and counterflow, respectively.
The roughing pump assists the measurement mode via a coupling valve opened to
the fore-vacuum pump when roughing is completed.
< previous page page_498 next page >
< previous page page_499 next page >
Page 499

manner. The roughing pump fulfills an additional purpose after roughing is completed at about 0.1-mbar inlet pressure:
Via the coupling valve some speed is added to the backing pump to enable an easy crossover by keeping the fore-
vacuum pressure below 0.1 mbar. Otherwise, when the turbomolecular pump takes over the inlet pumping, the fore-
vacuum gas flow would suddenly be much too high for the backing pump to keep the pressure below 0.1 mbar. With
this unit, volumes up to 100 liters can be tested down to 2 × 1010 mbar·liter·s1 with a response time of seconds without
the use of LN2.

Another version of a twofold turbomolecular pump was used in an advanced portable counterflow leak detector. In the
HLT150/160 model of Balzers the turbomolecular pump rotor is divided into two stages of pumping in the same
direction. The inlet is connected in between the two stages so that the lower stage works as a highly stable and clean
backing pump for the counterflow procedure, for which the upper pump stage (connected to the mass spectrometer) is
employed. For dirty systems with higher pressure up to 0.5 mbar the complete turbomolecular rotor is used for
counterflow, and the oil-sealed backing pump alone pumps the object. This portable unit can be used for service
purposes where clean components as well as dirty systems with higher pressure have to be tested down to some 1010
mbar·liter·s1.

By replacing the lower half of the turbomolecular pump with a molecular drag type stage (see Chapter 4), it is possible
to create leak detectors with a maximum allowable inlet pressure in the range of millibars or even tens of millibars. This
helps to decrease the pumping time for components and allows the testing of systems in the low vacuum range without
using partial flow and big roughing pumps.

8.8.5
Oil-Free and Dry Helium Leak Detectors

The presence of an oil-sealed rotary pump at the inlet port of a counterflow leak detector always produces a slight oil
contamination of the test object depending on the pressure conditions and the testing time. Advanced counterflow leak
detectors avoid this problem by not roughing down to the low pressure rangethat is, at most only down to 0.1 mbar. In
this pressure region, viscous flow conditions prevent the oil from the fore pump backstreaming into the test object. Leak
detectors of this kind are sometimes called "oil-free" because their inlet port is kept free of oil and thus the test object is
kept reasonably clean for later use under UHV conditions.

There is always the risk of wrong operation producing oil contamination of test objects even with "oil-free" units,
because there are conditions of operation where oil backstreaming from the roughing pump cannot be avoided
completelyfor example, for rather big leaks under low pressure, where the roughing pump is still at the inlet port
operating in partial flow. To keep sensitive test objects completely hydrocarbon-free, especially for the semiconductor
industry, completely "dry" leak detector systems are required. The technical basis for "dry" counterflow leak detectors
are molecular drag pumps with a very high fore-vacuum compatibility up to 10 mbar or more. These enable roughing or
backing with diaphragm or scroll pumps, thus completely avoiding any oil in the system. Dry leak detectors are still
rather expensive and bulky units because they are derived from existing counterflow models by replacing forepumps by
scroll pumps or a combination of diaphragm and small drag pumps.

The most important problem in dry leak detectors is the recovery time after having seen some helium from a leak.
Because the dry fore-vacuum pumps are all operating

< previous page page_499 next page >


< previous page page_500 next page >
Page 500

very close to their ultimate pressure limit, their pumping speed is nearly zero. To overcome this problem, purge gas is
admitted into the fore-vacuum line. Ideally, this gas must be free of helium. Such pure gas is difficult to supply,
especially for a portable unit. In practice, ambient air is used. The natural helium content of atmospheric air then puts a
limit on the detection of small helium leaks.

Since the development of dry detectors has just started, one can expect that by lowering the detection limit and response
time as well as achieving a quick pumpdown, these will be inexpensive and simple solutions in the near future.

8.9
Specifications of Mass Spectrometer Leak Detectors

Helium leak detectors are described by a set of specifications like all analytical instruments. The three most important to
characterize the performance of a leak detector are sensitivity/detection limit, response time, and maximum allowable
inlet pressure.

There is often confusion about the terms "sensitivity" and "detection limit" of a leak detector. The minimum detectable
leakage rate (the detection limit) is determined mainly by the stability of the detection system. It is given by the smallest
detectable electrical signal that the preamplifier and processing unit can distinguish from noise and drift. Overall
sensitivity s0 of a unit is defined as the ratio of the output signal Si to the leakage rate Q. The following relation holds:

This has to be distinguished from the intrinsic partial pressure sensitivity of a leak detector as described in Eqs. (8.13)
and (8.16):

where s is the intrinsic sensitivity sDF or sCF of a direct or counterflow leak detector (s has the unit (liter/s)1), sMS is
the mass spectrometer sensitivity (unit: A/mbar) and samp is the amplifier sensitivity (unit: V/A). The minimum
detectable leakage rate (detection limit Qmin) is given by the overall sensitivity s0 of the unit and the minimum
detectable signal Simin:

Simin has to be defined by means of stability criteria such as signal noise and drift. Such a definition has been given, for
example, in ISO 3530 (and the older AVS 2.1 standard) by stating that the minimum detectable signal is given by peak-
to-peak noise and the amount of drift in a specified short time interval. As the noise can be severely influenced by the
time constant of the signal amplifier or the follow-up averaging software, the time constant has always to be stated
together with the detection limit specification. For modern leak detectors having very low drift, the minimum detectable
signal is merely given by the peak-to-peak noise of the signal. The performance of a specific unit can be assessed by the
time constant it needs to achieve a given detection limit.

< previous page page_500 next page >


< previous page page_501 next page >
Page 501

The shorter the response time of a leak detector, the better it is suitable for industrial testing in short cycles, especially
for leak localization. The required minimum detectable leakage rate normally puts a first limit to quick response
because it requires a minimum time constant. The response time of a leak detector is, however, made up of several
completely independent contributions, nearly all of which also have some influence on the detection limit:

Gas transport time constants are (a) diffusion time in high-pressure connection piping from the test object and (b) the
tracer gas pumping speed at the inlet port of the leak detector (or the outlet port of the test object) together with the test
object's volume (or at least the leak detectors dead volume in its valve system). Whereas diffusion time in the
connection piping is not a property of the leak detector and can be avoided by an appropriate setup, the inlet pumping
speed is most important for the response time in a test. It depends mostly on the leak detector principle. It is smallest in
a simple counterflow, reasonable in a classical direct flow, and maximum in an advanced counterflow leak detector with
a booster pump. Only in the latter can the inlet pumping speed be increased without a reduction in the intrinsic
sensitivity, which would have a negative effect on the detection limit according to Eq. (8.19).

The electrical time constants in a leak detector are (a) the preamplifier's time constant (the longer it is, the more
sensitive it is) and (b) the time constant resulting from any averaging procedure in the software for smoothing the signal
noise. In order to generate a reasonable signal from the very small ion current produced in the mass spectrometer, a
certain gain of the preamplifier at the ion collector is necessary. The higher the gain, the more noise is produced, so
some time constant is required to smooth the signal again. For low noise, one has to achieve high gain in the first stage
of amplification. In order to measure larger signals with the same unit, switchable amplifiers are used. These generate
additional noise, so optimization is necessary.

There have been numerous solutions to the problem of optimizing the time constant, noise, and sensitivity.
Microprocessors combined with advanced analog semiconductor devices enable still new ways of preamplifier design.
Modern leak detector amplifiers are capable of quantitative measurement of 1 × 1015 A with a time constant of less
than a second.

A leak detector is different from an electrical measurement instrument in that not only does it have to measure small
quantities of tracer gas, but the measurement of the tracer gas should be independent of other gases such as air or water
vapor flowing from the test object into the leak detector. For a leak detector to be most useful, the pumpdown or
preparation time should be as short as possible. This means that the test object should be tested at the highest possible
pressure, resulting in a high amount of air and water vapor escaping from the object and flowing through the leak
detection system. On the other hand, a mass spectrometer has specific limits of linearity and sensitivity depending on
total pressure (see Chapter 7, "Partial Pressure Analysis"). Therefore, the spectrometer has to be protected against gases
and vapors different from the tracer gas.

Dealing with mass spectrometer leak detectors, depending on the test specimen, two different concepts to describe gas
handling have to be used. When connected to systems with their own pumps that generate a certain total pressure, the
leak detector has to tolerate this as its inlet pressure. The limit of "maximum tolerable inlet pressure" has to be
specified, and the leak detector with the highest value will be most useful. In testing components, the leak detector itself
pumps down the object. Now the

< previous page page_501 next page >


< previous page page_502 next page >
Page 502

shortest pumpdown time until the detector is ready for a leak measurement will characterize the best unit. The
pumpdown time is given both by the maximum tolerable inlet pressure and by the pumping speed for air at the inlet; in
other words, it is given by the "maximum tolerable total gas flow" that the leak detector vacuum system can bear under
measurement conditions. The actual total gas flow from the test object will decrease exponentially with time above
approximately 0.1 mbar (volume gas flow), but only inversely with time below that pressure (desorption gas flow). A
maximum tolerable inlet pressure of more than 0.1 mbar indicates extremely short pumpdown times because the slow
desorption time regime is completely avoided.

8.10
Quantitative Leakage Rate Measurements

In recent times, the demand for quantitative leakage tests has increased sharply. This is due to the standardization of
quality systems by the ISO 9000 series of standards. Each quantity in a production process has to be measured against
standards traceable to a national standard. For quantitative measurements of leakage rates, such traceable standards are
reference leaks calibrated against transfer standard leaks from the respective national laboratory.

Helium leak detectors are not fundamental measurement instruments because the measurement of leakage rate with such
an instrument is based on a set of parameters that are not precisely known. The most important of these are the pumping
speed of each vacuum pump at the respective total pressure, the sensitivity of the mass spectrometer, and the sensitivity
of the preamplifier. Even when a leak detector is once adjusted to yield the "true" value of leakage rate, its long-term
stability is normally not sufficient to use the leak detector itself as a standard.

Therefore, a helium leak detector needs frequent recalibration with a standard leak to ensure precise measurements. In
practice, such a calibration is often done with only one leak in the 108 mbar·liter·s1 range because of the difficulties to
get a leak calibrated in different leakage rate ranges. Because the measurement range of a modern leak detector typically
reaches from some 1010 mbar·liter·s1 up to some 102 mbar·liter·s1 or even more, the accuracy of measurements in
ranges far away from the calibration point is not really known. In other words, the linearity of the leak detector is not
known. In fact, because no traceable leaks with sufficient accuracy are available in the whole range, the uncertainty of
big or small leakage rates can only be estimated from the uncertainty of the components involved. This is normally
sufficient for acceptance tests of industrial parts, but the operator should be aware of the limited accuracy if measured
values get close to acceptable leakage rates. Safety margins should be used.

Calibration leaks are leak artifacts that produce a definite amount of tracer gas flow under definite pressure conditions
and with a good long-term stability. In principle, both permeation or conductance (capillary)-type leak artifacts may be
used for that purpose. Because of the risk of clogging, however, permeation leaks are used if possible [16].

At present, traceability of the leakage rate is only achieved in the main European countries and the United States, where
the national laboratories have established primary leakage rate standards and a calibration service for helium leaks with
leakage

< previous page page_502 next page >


< previous page page_503 next page >
Page 503

rates in a certain range [17]. Primary standards for helium leakage rates exist in the range of some 106 mbar·liter·s1
down to some 109 mbar·liter·s1. Helium transfer leaks are normally made in the form of a quartz permeation tube
pressurized by several atmospheres of pure helium which is kept in a pressurized reservoir. Such an arrangement is
completely safe against clogging and has a long-term drift (depending on the volume and pressure of the reservoir and
the leakage rate) in the order of 0.5% per year. Such transfer standards are used by national calibration services to
generate calibrated standard leaks which in turn are used to calibrate and adjust helium leak detectors [18].

Having a leak traceable to a primary standard in hand, this may be used to calibrate leaks for everyday work. Two basic
methods are available:

Method A: Calibration by comparison using a mass spectrometer (normally built in a leak detector for measuring helium
leaks) as the transfer device

Method B: Calibration by direct volumetric measurement of gas flow using a calibrated capillary and a timer

Method A is the easiest way to do a calibration, especially when a leak detector can be employed (as for helium leaks).
It is, however, limited to the range of linear measurement of the leak detector and the available standard leaksat present,
leakage rates below 1 × 106 mbar·liter·s1. One can use either one or two standard leaks for comparison. The leakage of
the unknown leak has to be in between the values of the two standard leaks. This allows an estimation of uncertainty
including linearity errors of the leak detector used. Using one leak is only reasonable if its value is at least in the same
decade as the "unknown" leak because otherwise the (normally unknown) linearity error of the leak detector makes it
impossible to state any uncertainty of the measurement. If, however, the leak is very close to the reference leak, the
linearity error can be neglected.

If both the reference and the "unknown" leak are of the same type and hence have the same temperature dependence, no
temperature correction is necessary if both leaks have been stored at this temperature for a sufficiently long time.

Method B is only applicable to leaks greater than 1 × 106 mbar·liter·s1 because of possible temperature effects in the
measurement capillary. It requires a skilled worker capable of handling a calibrated capillary with a water or oil slug
moving under the action of the escaping tracer gas from the leak. Leaks discharging to vacuum and atmosphere can be
calibrated in this way. The uncertainty of measurement is mainly determined by the accuracy of reading the capillary,
how constant the temperature can be kept, and the duration of the measurement.

It is often required to convert leakage rates measured with a given tracer gas into leakage rates for a different gas or
even a liquid present under actual process conditions. Because the different modes of flow depend strongly on geometry
and pressure which may not be constant in the leak channel, an analytical computation is rather difficult when precise
results are required. There are rather precise numerical predictions possible for different gases, if a set of characteristics
for a given leak have been measured with one gas specie [19].

To determine rejection levels for a tracer gas test, first acceptable leakage rates for the process fluid have to be
established from process considerations (lifetime,

< previous page page_503 next page >


< previous page page_504 next page >
Page 504

environmental aspects, cost, etc.). The following formulae can then be used to determine rejection limits for a test with a
specific tracer gas different from the process fluid. Much experience in calculating such limits has been accumulated in
the packaging of radioactive materials [20].

For the conversion of leakage rates when gas species, temperature, or pressure differential are different during test and
operation of a system, the following equation gives the relations in the case of molecular flow:

where T are absolute temperatures, M are molar masses of the flowing gases, and ∆p are the pressure differentials
during test and process, respectively.

For viscous flow conditions, the conversion of leakages has to be done according to

where p1 and p2 are the upstream and downstream pressures for test and process, respectively, and the η are the
dynamic viscosities of the flowing gases. The effect of temperature is accounted for by the effect of temperature on
viscosity, varying as the square root of temperature (see Section 1.10).

To apply the above equations, the conditions of flow have to be stated. Having atmospheric pressure on one side and
vacuum of less than 1 mbar on the other side, it is practical to assume molecular flow (flow dependent on pressure
differential) for leakage rates below 107 mbar·liter·s1. Viscous flow (flow dependent on difference of squared
pressures) can be assumed for leakage rates above 104 mbar·liter·s1. Intermediate leakage rates have to be converted by
application of both flow considerations and a conservative determination of the resulting rejection level.

8.11
Mass Spectrometer Leak Detectors for Other Tracer Gases and Future Developments in Leak Detection

The technique of mass spectrometer leak detection is becoming more and more a standard method of nondestructive
testing in industry. This means that although the operation of the mass spectrometer as the heart of the leak detector is
based on a high-vacuum system, the test objects are increasingly parts of systems such as valves, cans, gas supply lines,
and so onthat is, systems normally operating under atmospheric or overpressure. Some of these systems already contain
a specific gas or liquid that can be used for leak detection purposes. So there is a demand for tracer gas leak detectors
capable of detecting gases different from helium.

A specific application in this field was generated by the introduction of new refrigerant fluids caused by the problem of
ozone layer damage by the old ones. Very sensitive, but at the same time selective, leak detection on closed pressurized
refrigerator systems became necessary. Although this is not a vacuum application, it took only a short time for
manufacturers to modify the existing helium leak detectors and

< previous page page_504 next page >


< previous page page_505 next page >
Page 505

develop powerful tools for the testing of such systems. The sniffing technique (see Section 8.7) and the quadrupole
mass spectrometer well known from residual gas analysis in vacuum vessels were combined to produce a new type of
mass spectrometer tracer gas leak detector. Very compact, transducer-like quadrupole mass spectrometers designed in
recent times made this development possible.

In principle, this new kind of leak detector could easily be modified to operate in a vacuum testing mode. However, as
gases different from helium or hydrogen cannot easily be separated from desorbing water by traps or counterflow
operation, the detection limit can be expected to be very poor or the pumpdown time long. So it seems that light gases
will remain the best choice as tracer fluids in vacuum leak detection for the time being.

For the future of tracer gas leak detection, one can expect that new, less complicated yet well-performing detection
systems will be developed to replace the mass spectrometer and high-vacuum systems of present units.

References

1. R. C. McMaster, ed., Nondestructive Testing Handbook, 2nd ed., Vol. 1. American Society for Nondestructive
Testing and the American Society for Metals, 1982.

2. C. D. Ehrlich, A note on flow rate and leak rate units. J. Vac. Sci. Technol. A 4(5), 2384 (1986).

3. H. Yasuda and V. Stannett, Polymer Handbook. Wiley, New York, 1975.

4. L. Laurenson and N. T. M. Dennis, Permeability of common elastomers for gases over a range of temperatures. J.
Vac. Sci. Technol. A 3(3), 1707 (1985).

5. H. E. Nuss and I. Streuff, Leak rate measurements for large vacuum chambers. Vacuum 46, 845 (1995).

6. L. C. Beavis, Real leaks and real leak detection. Vacuum 20, 233 (1970).

7. M. Moraw and H. Prasol, Leak detection in large vessels. Vacuum 28, 63 (1977).

8. D. A. Howl and C. A. Mann, The back pressurizing method of leak testing. Vacuum 15, 347 (1965).

9. A. Nerken, History of helium leak detection. J. Vac. Sci. Technol. A 9(3), 2036 (1991).

10. A. O. Nier, C. M. Stevens, A. Hustrulid, and T. A. Abbott, Mass spectrometer for leak detection. J. Appl. Phys. 18,
30 (1947).

11. G. Reich, Leak detection with tracer gases: Sensitivity and relevant limiting factors. Vacuum 37, 691 (1987).

12. W. Becker, Erhöhung der Empfindlichkeit des Heliumlecksuchers durch Verwendung einer Turbomolekularpumpe
besonderer Konstruktion. Vak.-Tech. 8, 203 (1968).

13. M. H. Hablanian and W. E. Briggs, New technical developments in helium leak detection. Proc. Int. Vac. Congr.,
7th, Vienna, 1977, p. 199 (1977).

14. W. Becker and W. K. Huber, A novel leak detector with turbomolecular pump. Proc. Int. Vac. Congr., 7th, Vienna,
1977, p. 203 (1977).

15. G. Reich, The principle of He enrichment in a counterflow leak detector with a turbomolecular pump with two
inlets. J. Vac. Sci. Technol. A 5(4), 2641 (1987).

16. W. Jitschin, G. Grosse, and D. Wandrey, Diffusion leak artifacts as a secondary standard for gas flow. Vacuum 38,
883 (1988).
< previous page page_505 next page >
< previous page page_506 next page >
Page 506

17. K. Jousten, G. Messer, and D. Wandrey, A precision gas flowmeter for vacuum metrology. Vacuum 44, 135 (1993).

18. G. Grosse, G. Messer, and U. D. Wandrey, Summary abstract: Calibration and long-term characteristics of helium
reference leaks. J. Vac. Sci. Technol. A 5(4), 2661 (1987).

19. J. L. Chamberlin, The modeling of standard gas leaks. J. Vac. Sci. Technol. A 7(3), 2408 (1989).

20. J. Higson, C. Vallepin, and H. Kowalewsky, A review of information on flow equations for the assessment of leaks
in radioactive transport containers. Proc. The 9th Int. Symp. Packag. Transp. Radioact. Mater. (PATRAM '89), June
1116, 1989, Washington, DC, USA, Vol. I, pp. 195205.

< previous page page_506 next page >


< previous page page_507 next page >
Page 507

9
High-Vacuum System Design

Wolfgang Schwarz

High-vacuum systems cover a wide range of size and complexity from simple bell jar chambers for laboratory tests to
sophisticated industrial production systems and large-size research equipment. Most of these systems operate in a wide
pressure range from atmospheric down to 106 mbar or less.

Designing vacuum systems comprises a number of tasks from different engineering disciplines such as mechanical
engineering, electrical and electronic design, software engineering, and vacuum engineering. The main tasks in the
vacuum engineering step are the selection of appropriate types and sizes of pump sets, sizing of pipework, valves, and
other components influencing pressure or flow distributions in the system according to the process requirements.

In this chapter the design equations for the pumpdown process and for process operations are given. Together with the
effective pumping speed of the pump set and the gas loads in the system, they yield predictions of the system's vacuum
performance. Appropriate analytical and numerical calculation techniques are reviewed.

9.1
Calculations of Vacuum Systems

The pump set on a vacuum system has to perform two tasks. It has to evacuate the system starting from atmospheric
pressure down to a specified pressure within a certain time, and it must be able to maintain a specified pressure during
the vacuum

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_507 next page >


< previous page page_508 next page >
Page 508

Fig. 9.1
Schematic diagram of a
basic vacuum system.

process operation. Since, in general, these process requirements result in completely different requirements on the
pumps, it is advisable to consider both tasks separately in the first step and in a second step to evaluate the results from
both tasks with respect to pump types and sizes.

A schematic diagram of a basic vacuum system is shown in Fig. 9.1.

9.1.1
Basic Pumpdown Equations

In order to derive basic design equations an elementary idealized vacuum system is considered. A chamber of volume V
at a uniform pressure p is connected to a pump set through a component (e.g., a tube) of conductance C. The pump set
has a pumping speed

at its inlet. Since we are interested in the pumping action at the system, we define the effective pumping speed by

using the pressure p in the system instead of the pressure p0 at the pumps inlet. The effective pumping speed can be
calculated from a steady-state flux balance of the flow entering the component and the flow into the pump and the
pressure drop across the component determined from its conductance:
< previous page page_508 next page >
< previous page page_509 next page >
Page 509

Fig. 9.2
Relative effective pumping speed as a function of the relative
conductance between system and pump set.

The effective pumping speed is always smaller than the installed pumping speed S0. The fraction of the speed of the
pump set

which is available at the system is illustrated in Fig. 9.2. In order to obtain an effective pumping speed of 90% of S0 or
more the conductance between the system and the pump set has to be larger than nine times the pumping speed of the
pump set. This emphasizes the importance of high conductances between the pump set and the system.

For all further considerations the effective pumping speed will be used instead of S0, and it is implied that S has been
derived from the known conductance or pressure across the component (e.g., determined as outlined in Chapter 2) and
the intrinsic pumping speed S0 of the pump.

In Fig. 9.1 a gas flow into the system is shown schematically. This flow can be caused by gas entering the system
from the ambient by a purposely fitted gas inlet, by leaks or permeation through seals or chamber walls. It can,
however, also be caused by gas sources physically within the system like outgassing from materials.

From the basic quantities volume of the system, effective pumping speed, and flow into the system the time-dependence
of the pressure in the system can be calculated. If we apply the ideal gas law with the assumption of constant gas
temperature, a flux balance yields the basic differential equation:

< previous page page_509 next page >


< previous page page_510 next page >
Page 510

where is the flow the pump system removes from the chamber:

In order to solve the differential equation the pressure-dependence and time-dependence of the effective pumping speed
and the flow into the system have to be known.

In real systems the pumping speed is markedly pressure-dependent. In some cases, even its time-dependence has to be
considered (e.g., during start-up of a pump set). The flow is generally a function of time and it can be a function of the
pressure as in the case of outgassing. The simplest assumptions, although not necessarily realistic, are pressure- and
time-independent constant pumping speed and flow. This yields a simple expression for the pressure in the system:

Figure 9.3 gives an example of a pumpdown curve of a system with a volume of 200 liters pumped by a pump with S =
10 liter·s1. The flow into the system is 0.75 mbar·liter·s1.

The pressure versus time curve clearly shows two ranges. Starting from atmosphere, the pressure drops exponentially
with time (range I):

During this time the gas flow into the system can be neglected since p·S is large compared to . Hence practically
only gas is pumped which has been contained in
Fig. 9.3
Pumpdown curve for a system with S = const = 10 liter·s1,
a volume of 200 liters, and a flow into the system of
= const = 0.75 mbar liter·s1. In range I the gas flow
from the volume determines p(t); in range II the
pressure is only a function of .

< previous page page_510 next page >


< previous page page_511 next page >
Page 511

the volume. The pressure drop in range I is characterized by the vacuum time constant:

which is only a function of the system volume and the effective pumping speed. It is independent of . In the second
range the pressure is practically constant:

In this range, only the gas flow together with the pumping speed determine the pressure, but not the system volume.
Because has been assumed to be not time-dependent, Eq. (9.10) yields the minimum attainable pressure in the
system under these conditions. The pressure pB is often called base pressure or ultimate pressure.

Solving Eq. (9.7) for t yields the pumpdown time from p0 to p:

From Eq. (9.11) as well as from Fig. 9.2 it is obvious that as p approaches pB the pumpdown time diverges. Small errors
in the (in general predicted) pB can result in huge errors in the pumpdown time.

In the above example, pumpdown started at atmospheric pressure. It should be noted, however, that none of the
equations or assumption makes explicit or implicit use of pressure or flow ranges. Therefore, the solutions derived are
valid for all pressures from UHV to above atmospheric pressure as long as the conditions of constant gas temperature,
constant flow, and constant pumping speed are fulfilled.

9.1.2
Process Pressure

With a few exceptions, vacuum processes are started after the system has been evacuated and the pressure has fallen
below a defined value. Typically the system then operates in range II of Fig. 9.3, that is, in the flow-controlled pressure
range. Therefore in order to calculate the pressure the flow into the system and its time-dependence have to be known.
Since can exhibit arbitrarily complex behavior, no general expression can be derived. Instead the limiting cases of
slowly varying and rapid step-like changing flow are considered.

< previous page page_511 next page >


< previous page page_512 next page >
Page 512

In range II for a slowly varying flow a quasi-steady-state pressure is to be expected. Neglecting the gas storage effect of
the volume compared to the flow removed by the pumping system, we obtain

Equation (9.5) yields a quasi-steady-state pressure:

The condition (9.12) can be rewritten as

or as a condition for the rate of flow change:

For both, the relative change of pressure and the relative change of flow the rate of change has to be small compared to
1/τ to allow the use of the approximation in Eq. (9.13). It is interesting to note that the volume, while not explicitly
needed for the determination of the quasi-steady-state pressure, shows up implicitly in the conditions in form of the
vacuum time constant.

The other limiting case of the time-dependence of the flow is a rapid change from one flow level to an other. For a step
function change of the flow:

the pressure follows an exponential relaxation:

where are the equilibrium pressures before and after the flow change, respectively.

The characteristic time for the relaxation is again the vacuum time constant τ = V/S. It is the same for rising and falling
pressure, and it is independent of the equilibrium pressure as long as the effective pumping speed remains unchanged.
From Eqs. (9.13) and (9.9) the necessary pumping speed S for a defined gas load and a required time constant can be
directly calculated.

< previous page page_512 next page >


< previous page page_513 next page >
Page 513

Applying these equations to practical systems, however, poses a number of problems. Neither the condition for pressure-
independent effective pumping speed nor that for constant gas flow are hardly ever fulfilled in practical systems. Nearly
all pumps exhibit a pronounced pressure-dependence of their pumping speed S0. Furthermore, the conductance of
elements such as tubes, valves, or baffles is also markedly pressure-dependent as shown in Chapter 2. The gas flow in
the system is, in general, not constant but is a function of time, especially when outgassing dominates.

In order to calculate realistic pumpdown curves it is therefore necessary to determine the time-dependence of the flow
(Section 9.2) and to calculate the effective pumping speed S(p) at the system (Section 9.3).

9.2
Gas Loads in High-Vacuum Systems

In every vacuum system a number of gas sources is usually present. Chamber walls and the components within the
chamber release gas which was adsorbed at the surfaces or entrapped in the volume of the material. Air and water vapor
enter the system through leaks or permeate through seals. Finally the process performed in the vacuum system may
release gas from the material being processed or it may require gas from an external source. In order to size the pump
set the gas loads from the different sources have to be known.

9.2.1
Outgassing

One major gas source is the outgassing of the vacuum system itself and the material to be processed. Although much
research has been done in the field of the theory of desorption from surfaces and diffusive outgassing of bulk material, it
is not possible to predict the outgassing behavior from basic material properties. The main reason is that the surface of
real materials is an ill-defined mixture of different microscopic structures with different geometrical and desorption
properties (for details see Chapter 10). Therefore outgassing measurements have to be performed on the materials of
interest.

Figure 9.4 illustrates some examples of the time-dependence of the outgassing rate at constant temperature. The
diagram suggests that at least for a limited time interval the outgassing rate can be fitted to

with the geometrical surface A and the fit parameters a1h and α. The parameter a1h can be identified as specific
outgassing rate after 1 h of pumping. The decay exponent α is the negative slope of the outgassing curve in a loglog plot
(Fig. 9.4).

The specific outgassing rates cover a range over several decades depending not only on material type but also on surface
preparation, cleaning steps, exposure times to atmosphere, and relative humidity. The decay exponent ranges from about
0.2 to 1.2. Its value hints to the type of outgassing mechanism. Desorption from surfaces with

< previous page page_513 next page >


< previous page page_514 next page >
Page 514

Fig. 9.4
Time-dependence of the outgassing flow rate for different materials.

only negligible contribution from the bulk yield values close to 1, and diffusion-controlled outgassing from the bulk of
the material yields values of about 0.5. At room temperature the main contribution to the outgassing flow from materials
which have been exposed to ambient air is water vapor (≥80%) with small amounts of N2 and CO2.

Metals, glasses, and ceramics yield outgassing rates which are well fitted with a single set of a1h and α values for
outgassing times from about 1 s to more than 100 h. The decay exponent for these materials is close to 1. Especially for
metals, details of the preparation and cleaning steps have a larger influence on the specific outgassing rate than do the
differences between different types of materials.

Polymer materials generally show higher outgassing rates compared to metals and decay constants in the range of 0.4 to
0.8, indicating contributions from diffusion from the bulk. In many cases, however, a single-parameter set is not
sufficient for an adequate fit over the time interval of interest. In these cases either two fits of Eq. (9.17) for different
time intervals or other fit functions should be used. Apart from water, some polymer materials contain volatile organic
substances which are released during the outgassing process. In contrast to water vapor, these substances cannot be
readsorbed during exposure to ambient air, resulting in reduced outgassing rates for repeated pumpdowns.

For an entire vacuum system the gas load due to outgassing is obtained from the summation over the contributions of all
surfaces. With Eq. (9.17) the time-dependence of the gas load is

< previous page page_514 next page >


< previous page page_515 next page >
Page 515

Fig. 9.5
Time-dependence of the total outgassing flow rate of a
system consisting of 4.5 m2 steel surface and 0.25 m2
polymer surface.

It should be noted that due to the small decay constants even small surface areas of polymer or rubber material may
cause significant contributions to the total gas load for long pump times as shown in Fig. 9.5. Therefore use of these
materials should be minimized in high-vacuum systems.

Although this analytic approach for the calculation of the gas load for a vacuum system is scientifically correct, it has
some practical limitations. Outgassing measurements are typically performed on small samples of simple-geometric-
shaped plates, which are easy to clean and easy to handle. Due to practical constraints, however, components for
vacuum systems cannot always be treated like outgassing samples. As a consequence, this results in outgassing
characteristics with increased outgassing rates and smaller decay constants compared to samples. Furthermore, for
complex components it may be difficult to access the geometric surface and the outgassing parameters of all kinds of
materials involved.

Some of the problems can be circumvented by measuring the outgassing characteristic of a similar system and then
scaling the results to the system under consideration. In order to measure the outgassing flow rate of a system, it is

evacuated and the time-dependence of the pressure p(t) in regime II is recorded. The flow rate is then calculated
from p(t) and the measured or calculated effective pumping speed S(p):

Assuming negligible leak and permeation rates is the outgassing rate of the system. The flow rate for the system
under consideration is then obtained by scaling with the ratio of the surface areas of the two systems:
Obviously this scaling approach is only valid for systems which are comparable with respect to materials,
manufacturing, and cleaning procedures and similar ambient conditions.

< previous page page_515 next page >


< previous page page_516 next page >
Page 516

As suggested by the curve for the total flow in Fig. 9.5, effective outgassing rates and decay constants can also be fitted
for entire systems using Eq. (9.17), although several fits for different time ranges may be necessary for good accuracy.

9.2.2
Leaks

Leaks are unintended and undesired paths for gases and vapors into a vacuum system. Obviously the flow rate through
leaks, the leak rate, cannot be predicted since no information about the geometry of the leak channel exist a priori.

The smaller the leaks that need to be detected, the more time-consuming and the more expensive the leak checks on
systems and components become (see Chapter 8). It is therefore desirable to have criteria for necessary leak check
sensitivities derived from the systems requirements.

For most systems, leaks are sufficiently small if they do not contribute more than 10% to the base pressure; that is, the
leak rate has to be no higher than 10% of the total flow rate, with the remaining more than 90% being due to outgassing
and permeation. The tolerable leak rate is therefore

Assuming that this leak rate is due to a number of nleak of undetected leaks of equal leak rate, the detection sensitivity
should be

On high-vacuum systems the effective pumping speed ranges from 10 liter·s1 on small load lock chambers to more than
several 10,000 liter·s1 on large systems. The number of undetected leaks should be not more than 10 on small chambers
and less than 100 on large systems. With these rough estimates the necessary leak detection sensitivity for single leaks
as a function of the base pressure are plotted in Fig. 9.6.

It should be kept in mind that these are only crude guidelines. Process requirements may well put more stringent
requirements on leak detection sensitivities.

9.2.3
Permeation

All materials exhibit a certain permeability for gases. This process involves three steps: the adsorption of a molecule on
the high-pressure side of the material, the diffusion through the material, and the desorption on the low-pressure side.
Depending on the details of the kinetics, the permeation flow rate at constant temperature is

< previous page page_516 next page >


< previous page page_517 next page >
Page 517

Fig. 9.6
Estimated leak detection sensitivity for single leaks as a function
of the base pressure of systems in order to keep the leak flow
rate smaller than 10% of the total gas load at base pressure.

Fig. 9.7
Gas permeation through metals.

Here A is the cross-sectional area through which permeation takes place, d is the thickness of the material, and p1 and
p2 are the pressures on both sides of the material. For most of the gases like H2O, N2, and O2 which permeate 'through'
the material as a molecule, Eq. (9.23) applies. Hydrogen, however, is dissociated during adsorption and therefore
according to the law of mass action, Eq. (9.24) has to be used.
The permeation conductivity kperm is a gas-type-dependent and material-dependent parameter which strongly increases
with increasing temperature. Figure 9.7 and

< previous page page_517 next page >


< previous page page_518 next page >
Page 518

Fig. 9.8
Gas permeation through polymer materials.

Fig. 9.8 give examples of the permeation conductivity of some metalgas and polymergas systems. At room temperature
the permeation rates through metals are very small and can be neglected. Permeation through polymers used, for
example, for seals have to be taken into consideration in systems where base pressures of some 107 mbar or less have to
be achieved.

9.2.4
Process Gas

Vacuum processes involve a wide range of physical and chemical mechanisms. In some processes, such as vacuum
annealing to modify crystal structures, vacuum is only needed to avoid contamination of the material being processed.
Gas is neither released nor needed for the process. Other processes, such as degassing of liquid metals, can release huge
amounts of gas. Table 9.1 summarizes examples of typical process gas loads.

Due to this wide range, no general rules exist for the process gas load of vacuum systems. For the design of a dedicated
vacuum system it is therefore necessary in most cases to scale experimental data on process gas loads from similar
systems. The scaling factors have to be chosen carefully, considering the basic reaction mechanisms and the kinetics of
the process. Degassing processes, for instance, can be limited by the transport mechanisms within the material, by the
desorption step at the surface, or by kinetics in the gas phase. Accordingly, the geometrical dimensions of the material,
its surface area, or partial and total pressures should primarily be used for scaling. However, as geometric relations
change, the relative importance of the limiting mechanisms may change. It is therefore advisable to recheck the
mechanisms after scaling.

< previous page page_518 next page >


< previous page page_519 next page >
Page 519

Table 9.1. Typical Ranges of Process Gas Loads in Industrial Vacuum Systems
Process Typical
Process Gas Load
(mbar·liter·s1)
Evaporation of optical layers
0.11
Semiconductor processing
110
Large-area sputter coating of architectural glass
520
Precision casting
0.110
Induction melting 101000
Steel degassing 5000100,000

9.3
Design of High-Vacuum Pump Sets

Pump sets for high-vacuum systems have to operate in the pressure range from atmospheric to the system base pressure which
may be as low as 108 mbar. This pressure range is covered by a combination of high-vacuum pumps which operate typically
below 102 mbar and fore-vacuum pumps which pump from atmospheric to the ''changeover pressure" of the high-vacuum
pumps. In most cases the forepumps also serve as backing pumps for the high-vacuum pumps. Details on the different types of
pumps and their pumping mechanisms can be found in Chapters 35. In this section the combination of different pump types will
be discussed.

9.3.1
Forepump Sets

9.3.1.1
Fore-Vacuum Pumps

In practically all high-vacuum systems, oil-sealed or dry compression pumps are used as forepumps often in combination with
Roots blowers. The pumping speed characteristics of the forepumps are published in the manufacturer's data sheets. For vacuum
design calculations, however, these data need some interpretation. Published pumping speed curves are measured according to
ISO standards under well-defined conditions. These conditions, however, are not necessarily representative of the application
environment. Speed curves and base pressures of oil-sealed pumps, for example, are measured excluding condensable gases. In
vacuum systems, however, water vapor is the dominant gas at low pressures. Therefore the water vapor pumping speed would
be needed. Furthermore, the oil in forepumps is often contaminated with water, which increases the base pressure of the pumps
to values much higher than the data obtained under standard measuring procedures.

For design calculations it is therefore desirable to have a method to generate pumping speed curves of forepumps with
adjustable ultimate pressure (see Fig. 9.9). A simple but surprisingly accurate approximation of the pumping speed is obtained
from the model of an ideal pump with pressure-independent pumping speed Smax and a constant internal backstreaming
At the base pressure pu the backstreaming balances the throughput puSmax; that is, the net pumping speed S(p) vanishes. From

< previous page page_519 next page >


< previous page page_520 next page >
Page 520

Fig. 9.9
Pumping speed of forepumps with different base pressures.

a flux balance the net pumping speed is

At high pressures, Eq. (9.25) yields a pumping speed of Smax which can be taken from the manufacturer's data sheets.
This approximation allows one to set the base pressure pu according to the application conditions of the pump, yielding
realistic speed and throughput close to pu.

9.3.1.2
Roots Combinations

In order to increase the pumping speed of a forepump set or to attain lower ultimate pressures, Roots blowers can be
added in series with the forepumps. This approach is often more economical than using more or larger-size forepumps
or multistage pumps, respectively.

A Roots blower can be considered to consist of a perfect pump with a theoretical displacement Sth plus an internal
backstreaming. The backstreaming is due to clearances between the rotors and gas adsorbed on the rotors at the fore-
vacuum side and subsequently desorbed at the high-vacuum side of the blower. Using this model the net throughput
net = p . S can be determined from the blower's theoretical throughput and the internal backstreaming
:

Assuming that the internal backstreaming is proportional to the pressure difference across the blower, can be
expressed as a function of the maximum compression ratio at zero throughput k0(pF):
< previous page page_520 next page >
< previous page page_521 next page >
Page 521

where pF is the fore-vacuum pressure. From the flux balance equation [Eq. (9.26)] with , the net pumping speed
is obtained:

Since for a given high-vacuum pressure p the corresponding fore-vacuum pressure pF is not known a priori, the
calculation starts with an assumed fore-vacuum pressure for which S is obtained from Eq. (9.28) using k0(pF) and SF
(pF). The corresponding high-vacuum pressure is then determined by the continuity equation for the flow through the
pump set:

assuming the same gas temperatures at the inlet and the outlet of the blower. The efficiency of the blower (i.e., the ratio
of the net speed and its theoretical speed),

is only a function of the staging ratio of the blower speed and fore-vacuum pumping speed Sth/SF and the maximum
compression ratio k0. The maximum compression ratio of a blower mainly depends on the shape of the rotors and the
clearances between the rotors and the housing. Figure 9.10 shows a typical k0 curve for air. Blower efficiency of more
than 70% is achieved with k0≥ 15 and staging ratios of 10 or less (Fig. 9.11). It only weakly depends on the absolute
value of the maximum

Fig. 9.10
Maximum compression ratio of a Roots blower as
a function of the fore-vacuum pressure.
< previous page page_521 next page >
< previous page page_522 next page >
Page 522

Fig. 9.11
Roots blower efficiency as a function of compression
ratio at zero throughput for different staging ratios.

compression ratio but tends to decrease significantly for staging ratios above 10. Therefore in order to achieve high
efficiencies in the entire operating range of the blower, the staging ratio should not exceed 10.

Although relatively high compression ratios are possible at high fore-vacuum pressures, the pressure difference ∆p
which can be safely achieved is limited by mechanical forces on the rotors and the temperature rise of the gas in the
exhaust area of the pump. Therefore pump manufacturers specify a maximum pressure difference ∆pmax for continuous
operation for blowers without integrated bypass valves. This pressure difference also determines the pressure pc, R at
which the blower can be cut in.

In order to calculate pc, R the continuity equation is rearranged, yielding

Equation (9.31) can be iterated with respect to k0. Considering that pc, R only slightly decreases with increasing k0, a
safe value for the cut in pressure can be calculated with k0 = 15. Typically, the cut in pressure ranges from about 10 to
about 100 mbar depending on staging ratio and tolerable pressure difference.

The net pumping speed of Roots blower/fore-vacuum pump combinations, using the compression ratio according to Fig.
9.10, is shown in Fig. 9.12. The pumping speed curve of the combination begins at a considerably lower ultimate
pressure than that of the forepump alone. It then steeply rises to a nearly constant speed plateau. At higher pressures the
blower speed decreases due to the diminishing k0.

As to be expected, the ultimate pressure is approximately that of the forepump divided by the compression ratio at this
pressure. Therefore the combination's ultimate pressure is typically a factor of 2030 lower than that of the forepump.
This would suggest an operating range extended by the same amount. However, the net
< previous page page_522 next page >
< previous page page_523 next page >
Page 523

Fig. 9.12
Pumping speed of Roots blower forepump combinations
with a maximum fore-vacuum pumping speed of 100
m3/h and a theoretical blower speed of 500 m3/h.

speed close to the ultimate pressure is very sensitive to small changes in the ultimate pressure of the forepump as well as
to variations in k0 which might be induced by contamination of the rotors. Therefore the lowest operating pressure of
the combination should be about four to five times the calculated ultimate pressure.

At the high-pressure end of the operating range of the blower, thermal aspects have to be taken into consideration. The
compression performed by the blower is

This power heats the exhaust gas and the blower. For example, a blower with a theoretical speed of 1000 m3/h backed
by a 200-m3/h forepump yields a compression power of about 2 kW when operating at an intake pressure of 25 mbar.
Part of this heat load is transferred to the ambient through the housing of the blower and the pipe work between the
blower and the forepump. A considerable amount, however, is transferred into the forepump by the hot exhaust gas. For
transient operations such as evacuating a system, the heat load to the forepump is usually not a problem, but for steady-
state operations at high intake pressures a gas cooler may be necessary between blower and forepump in order to avoid
overheating the forepump.

For applications where high pumping speeds are needed in the pressure range between 102 and some 10 mbar
multistage blower systems with a comparatively small fore-vacuum pump can be used. The calculation of the net
pumping speed S2 of the entire pump set is analogous to that of S1, the speed of the forepump and blower 1. First S1
(p1) is obtained from Eq. (9.28), then the combination of forepump and blower 1 are considered as forepump for blower
2 and the same algorithm is applied to this blower. Figure 9.13 gives an example of such a pump set. Valve V1 isolates
the pump set from the vacuum system, and V2 is an optional valve in case the forepump has no integrated inlet valve.
The cut in pressures of the blowers are sensed by the pressure switches PS1 and PS2. These can be either two
membrane-type pressure switches or trigger signals derived from a suitable gauge with gauge controller.

< previous page page_523 next page >


< previous page page_524 next page >
Page 524

Fig. 9.13
Schematic diagram of a
forepump set with two Roots
blowers. The pressure switches
PS1 and PS2 start the blowers
at their appropriate cut in
pressures.

In most installations, due to space restrictions, the forepump set has to be placed some distance away from the vacuum
system. Often V1 is then installed at the chamber. With V1 closed and the forepump operating, the blowers and the
tubes between the pump set and the chamber are evacuated and the blowers are started. The entire volume from V1 to
the inlet of the forepump quickly reaches pressures well below 1 mbar. Opening V1 to start evacuation of the system
causes a rapid expansion of the air in the system into the evacuated volume of the pump set. Although the pressure
switches will immediately turn off the blowers because of their inertia, they will generate a pressure peak at the
forepump. The high air speed associated with the expansion may transport dust and particles from the chamber into the
forepump, and the pressure peak may damage exhaust filters in the forepump.

In order to avoid these problems, V1 should be located close to the pump set and the blowers should be switched off and
allowed to spin down for a few seconds before opening V1. An alternative would be a bypass valve with flow restriction
across V1.

9.3.2
High-Vacuum Pump Sets

The majority of high-vacuum pump sets today are equipped with turbomolecular pumps, turbomolecular pumps in
combination with cryo-surfaces, or cryo-pumps. Diffusion pumps are applied mainly in systems, where high pumping
speeds are needed and where considerable amounts of debris from the processes are to be expected.

Although it applies totally different pumping principles, the pumping speed of all these pumps can be described
schematically by three regions as illustrated in Fig. 9.14. The operating range is characterized by a constant-speed
pumping plateau. At low pressures, pumping is limited either by the maximum compression of the transfer
< previous page page_524 next page >
< previous page page_525 next page >
Page 525

Fig. 9.14
Schematic sketch of the three ranges of the
pumping speed of high-vacuum pumps. Detailed
behavior of the speed curve in the overload and
ultimate pressure range vary for different
pump types.

pumps (turbomolecular and diffusion pumps) or by the equilibrium vapor pressure in the entrapment pumps (cryo-
pumps, cryo-surfaces). At high pressures, overload due to high gas load occurs. In turbomolecular pumps, gas friction
between rotor and stator blades causing unacceptable heating of the rotor yields an upper limit for the operating
pressure. In diffusion pumps the jets become unstable when the collision rate between the oil vapor and the gas to be
pumped exceeds a certain value. Cryo-pumps and cryo-surfaces cease stable operation when the heat load on the
condensing surfaces cause their temperatures to rise or when adsorbing surfaces become saturated. The actual limits of
the operating range of a particular pump of course depends on the pumping principle and its technical realization.

For the design of a high-vacuum pump set, all three ranges have to be kept in mind: the ultimate pressure range with
respect to the planned ultimate system pressure, the operating range with respect to process pressure and throughput,
and the overload range with respect to pumpdown and crossover from fore-vacuum to high-vacuum pumps. Regardless
which type of high-vacuum is actually used, a few general design rules apply to the three ranges.

At the planned ultimate pressure of the system, pu, s, a sufficiently large and repeatedly achievable pumping speed is
needed. Therefore the high-vacuum pump and its ultimate pressure should be chosen so that the pumping speed at pu, s
is at least 90% of the plateau pumping speed. In high-vacuum systems equipped with turbomolecular or cryo-pumps,
this criterion is of no concern since their ultimate pressures are smaller than 109 mbar whereas pu, s≥ 108 mbar. For
diffusion pumps and cryo-surfaces, this rule has to be considered in more detail as discussed later in this chapter.

The pumping speed of all high-vacuum pumps as well as the conductances of the vacuum components between the
pumps and the chamber are gas-type-dependent. However, the gas-type dependencies of pumping speed and
conductance often show a tendency to counterbalance: Pumping speeds for light gases such as hydrogen or helium are
in most cases smaller than those for heavier gases, whereas the molecular conductances show the opposite behavior.
Therefore in most cases it is sufficient to calculate the effective speed for nitrogen and use this result for other gas
components too. In systems with cryo-surfaces or cryo-pumps the differences in pumping speed for

< previous page page_525 next page >


< previous page page_526 next page >
Page 526

Fig. 9.15
Selection of the changeover pressure from
fore-vacuum to high-vacuum pumps.

water vapor and all other gases is generally large so that a different approach has to be taken.

During normal operation of a vacuum system the overload range of the high-vacuum pumps should never be reached;
but during pumpdown, short overload periods are hardly avoidable. In order to minimize the overload time the two
criteria for the change over pressure pco have to be met: (i) The effective pumping speed of the high-vacuum pump
must be at least as large as of that of the fore-vacuum pumps (Fig. 9.15), and (ii) the backing pump must be able to keep
the backing pressure below the maximum tolerable foreline pressure when the high-vacuum pump operates at the
changeover pressure.

When criterion (i) is not met, switching from the fore-vacuum to the high-vacuum pumps decreases the effective
pumping speed as illustrated at p2 in Fig. 9.15. As a result, the pressure in the system starts to rise. This may then cause
a change back to the fore-vacuum pump set again, thus starting a cycle of several changeovers.

In order to avoid this situation, a changeover pressure p1 < pco, max should be chosen. Obviously it takes more time to
reach p1; but during this additional fore-vacuum pump time the system continues to outgas, and the outgassing flow
decreases. Switching then from the small fore-vacuum speed to the larger high-vacuum pumping speed causes a rapid
drop in the chamber pressure and shortens the overload time of the high-vacuum pump considerably.

During changeover the high-vacuum pump has to handle the largest throughput. The backing pump must be able to
maintain the foreline pressure below the critical foreline pressure under these circumstances. Otherwise the high-
vacuum pumps will cease pumping; this may result in cycling between fore- and high-vacuum pump, similar to the
situation at too-high changeover pressures.

9.3.2.1
Turbomolecular Pump Sets

Turbomolecular pump (TMP) sets are usually configured according to one of the schematics shown in Fig. 9.16. In
systems where the spin-up time of the turbomolecular pump of several minutes can be tolerated, no high-vacuum valve
between system and pump set is needed, thus reducing cost and avoiding pumping speed losses in the high-vacuum
valve.

The system is pumped down by starting the backing pump set and opening V2. The chamber is evacuated through the
TMP. Usually the internal conductance of the
< previous page page_526 next page >
< previous page page_527 next page >
Page 527

Fig. 9.16
Schematic diagram of turbomolecular pump
sets (a) without and (b) with high-vacuum valve
between system and pump set. The bypass
valve V3 in (a) is optional.

TMP is large enough so that it does not limit the roughing of the chamber. A few minutes before the backing pump set
alone would reach the maximum tolerable backing pressure pB, max of the TMP (typically 0.10.5 mbar for standard
TMPs, 510 mbar for pumps with molecular drag stages), the spin-up of the TMP can be started. When the TMP has
reached its operational speed, the backing pressure is then below pB, max.

In systems where debris from the chamber is expected to move during roughing, it is advisable to bypass the TMP by a
valve (V3 in Fig. 9.16). This avoids particle deposition in the TMP during roughing.

Chambers which require short pumpdown times or which are cycled continuously from atmosphere to high vacuum
have to be equipped with a high-vacuum valve V1 and a bypass valve V3. Before the pumpdown of the chamber starts,
the TMP is evacuated via V2 and ramped up to speed with V3 closed. For roughing, V2 is closed and V3 opened. At
this time the TMP operates without backing pump. Its fore-vacuum pressure is monitored, however. In case it exceeds
pB, max, roughing is interrupted by closing V3. Opening V2 then allows the system to pump below pB, max. Proper
timing of the valves has to make sure that before opening V2 the pressure in the fore-vacuum line between V2 and V3 is
smaller than the maximum backing pressure.

Calculation of the effective high-vacuum pumping speed of a TMP pump set is in principle straightforward. According
to Eq. (9.3), the speed of the TMP as published by its manufacturer and the conductances of all the vacuum components
between the TMP and the chamber are needed for the calculation. For most of these components, molecular
conductances either are published or can be calculated. However, in some of the components such as valves or tubes,
transitional flow may occur. This would suggest that the pressure-dependence of the conductance has to be known
which is rarely the case for complex components. Since the speed of the TMP decreases with increasing pressure, the
conductances play a less important role at higher pressures.

< previous page page_527 next page >


< previous page page_528 next page >
Page 528

Fig. 9.17
Effective speed of a 150 liter · s1 TMP with a 100-mm-
diameter, 300-mm-long tube using the pressure-dependent
conductance and a pressure-independent approximation.

Therefore for most applications, using the molecular conductance in the transitional flow regime yields a sufficiently
accurate approximation of the effective pumping speed. Figure 9.17 gives an example.

Although the pumping speed of TMPs decreases above about 1 × 103 mbar, this is no indication of overload, but the
reduction in pumping speed is due to the pumping mechanism. Most TMPs can continuously operate at inlet pressures
up to 102 mbar and provide very stable pumping speeds.

The backing pump set has to be designed so that the maximum tolerable fore-vacuum pressure pF, max as specified for
the TMP is not exceeded except for a few minutes during pumpdown. From the maximum expected operation pressure
pop, max, the throughput of the pump is obtained using the effective TMP pumping speed S. When the TMP
is equipped with a purge gas system, the purge gas flow has to be added, yielding a backing speed of

It should be noted that SF is the pumping speed required at the fore-vacuum flange of the TMP and that it is a lower
limit for the pumping speed. Larger fore-vacuum pumping speed may be necessary due to, for example, pumpdown
requirements.

9.3.2.2
Diffusion Pump Sets

The schematic diagram of the diffusion pump set is shown in Fig. 9.18. For start-up of the diffusion pump, V1 and V3
are closed and the pump is evacuated by the forepump through V2. When the foreline pressure reaches the maximum
fore-vacuum pressure of the diffusion pump pF, max the heaters are started. After the diffusion pump has reached its
operating temperature pumpdown of the system can be started in the usual way be closing V2 and opening V3. It is
advisable to monitor the backing pressure of the diffusion pump during pumpdown with V1 and V2 closed. The small
backing pump, shown as option in Fig. 9.18, will
< previous page page_528 next page >
< previous page page_529 next page >
Page 529

Fig. 9.18
Schematic diagram of a
diffusion pump set. The
baffle and the additional
backing pump are optional.

maintain sufficiently low backing pressures even in the presence of small leaksfor example, due to particles on the seal
of V1. Without this additional pump, when the backing pressure rises, the roughing has to be interrupted for a short time
by closing V3 and reopening V2 until an acceptable backing pressure is reached again.

Crossover from the fore-vacuum to the diffusion pump should be performed according to the rules outlined in Section
9.3.2. During crossover and high throughput operation the backing pressures of the diffusion pump needs special
attention. Maximum foreline pressures of diffusion pumps as published in the data sheets are usually measured for zero
throughput. In order to run a pump at high throughput, a lower backing pressure is needed for stable operation, typically
one-half to two-thirds of the pressure for zero throughput. Therefore the necessary effective backing pump speed is

The operating pressure range of diffusion pumps extends from about 10 times the ultimate pressure up to the maximum
inlet pressure pmax where the overload range with nearly constant flow begins (Fig. 9.19). pmax ranges from a few 104
mbar for large pumps with inlet diameters of 1 m to a few 103 mbar for small pumps. The effective pumping speed
calculation has to include the conductance of V1 and eventually that of the baffle. Although the conductance of the
valve is pressure-dependent in the upper operating and the overload range, for approximate calculation this can be
neglected. Typically the conductance of angle valves for diffusion pumps as well as the baffle conductance are about as
large as the nominal pumping speed. Therefore as a rule of thumb the effective pumping speed of a diffusion pump with

< previous page page_529 next page >


< previous page page_530 next page >
Page 530

Fig. 9.19
Operating pressure range for diffusion pumps.

Fig. 9.20
Inlet pressure changes caused by flow changes in
the operation and overload range in a diffusion pump.

valve is about half of the nominal speed. With valve and baffle it is about one-third of S0.
Without special measures, stable operation at pressures above the maximum inlet pressure is hardly possible. Small
relative changes in the flow which are reflected in proportional pressure changes at low pressure yield large pressure
excursions above pmax due to the flat curve (Fig. 9.20). To enable stable operation at pressures in the system
above pmax, the pump is throttled by a suitable throttle valve. Figure 9.21 illustrates the operating principle. For a given
flow and no conductance losses

< previous page page_530 next page >


< previous page page_531 next page >
Page 531

Fig. 9.21
Extension of the operating range of a diffusion pump
by throttling.

between the system and the pump, the pressure pP at the inlet of the pump and the pressure in the system pS are the
same. Throttling the pumpthat is, purposely introducing conductance losses between chamber and pump yields a lower
effective pumping speed in the system. For the same flow a higher chamber pressure results while at the same time
the pressure at the pump remains at pP, thus ensuring stable operation.

9.3.2.3
Pump Sets with Cryosurfaces

In high-vacuum systems most of the outgassing flow of the chamber consists of water vapor. This suggest the use of
pump sets with high water vapor pumping speed. Cryo-cooled surfaces are able to provide large water vapor pumping
speeds at reasonable cost. Cryosurfaces are used in many different forms such as cooled tubes or panels within the
system or cryo-cooled baffles in front of pumps. Cooling is provided either by boiling liquid nitrogen or by a
refrigerator recirculating a cooling mixture.

Pumping Speed of Cryosurfaces. The pumping speed for a condensable gas of a cold surface of surface area A is
determined by the surface temperature Ts, the saturation pressure ps at Ts, and vapor pressure pg and temperature Tg:

For water vapor at room temperature with Mg = 18 g/mol and a sticking coefficient σc = 1, Eq. (9.35) yields a pumping
speed of

The water vapor saturation pressure can be approximated by


< previous page page_531 next page >
< previous page page_532 next page >
Page 532

Fig. 9.22
Water vapor pumping speed of cryosurfaces at constant
surface temperature.

in the temperature range from 150°C to 0°C with sufficient accuracy for pumping speed calculations. With Eq. (9.36)
and Eq. (9.37) the pumping speed can be calculated explicitly from pH2O in front of the condensing surface and the
surface temperature Ts. Figure 9.22 illustrates the strong influence of the cryo-surface temperature on the pumping
speed at low pressures.

The heat load on the cryo-surface consists of the heat of condensation and the thermal radiation from the ambient:

The contribution of thermal radiation is nearly independent of the surface temperature, but strongly dependent on the
ambient temperature. At room temperature the radiative heat load is approximately 400 W/m2, in an environment with
an ambient temperature of 80°C the load more than doubles to about 850 W/m2. This underlines the importance of
shielding cryo-surfaces and cryo-pumps against high temperature radiation.

The condensation rate is proportional to pumping speed and gas pressure:

For water vapor at room temperature and a pressure of 103 mbar the maximum possible condensation rate is about 0.11
g/(m2·s).

Due to the thermal load by thermal radiation and the heat of condensation, the surface temperature of the condensing
surface Ts is higher than the temperature Tliq of the coolant because of the temperature jump between coolant and
cryopanel wall ∆Twliq, the temperature drops across the wall ∆Tw, and the temperature drops

< previous page page_532 next page >


< previous page page_533 next page >
Page 533

Fig. 9.23
Temperatures on a liquid-cooled cryopanel.

across the frost ∆Tf on the surface (Fig. 9.23):

Even when water vapor is pumped at pressures above 103 mbar and heat loads of a few kilowatts per square meter of
cryosurface have to be removed, the total temperature drop normally does not exceed 20°C. With liquid nitrogen as
coolant the surface temperature stay below 180°C, so a constant pumping speed of the cryosurface from UHV to
medium vacuum is obtained.

Refrigerators do not provide the low temperatures or the power handling capacity achievable with liquid nitrogen. Most
refrigerators have a cooling power characteristic starting at a temperature TR,0, the minimum coolant temperature
without heat load and increasing nearly linear with increasing thermal load:

The parameter AR characterizes the power handling capability of the refrigerator. Just as TR,0, the minimum coolant
temperature, AR, depends on the refrigerator type and its coolant mixture.

From the combination of Eqs. (9.36) to (9.41) the pumping speed of a refrigerator-cooled cryosurface can be calculated
using appropriate refrigerator parameters and estimates for the temperature drops between condensing surface and
coolant. Figure 9.24 gives an example of a 0.5-m2 cryosurface operated with a refrigerator with AR = 50 W/K and
different minimum coolant temperatures. For the design of a pumping system with a cryosurface, Fig. 9.24 suggests
choosing the minimum coolant temperature so that the ultimate pressure of the cryosurface is at least one decade lower
than the planned ultimate pressure of the system.

The overload range for this refrigerator cryosurface combination starts at about 102 mbar water vapor pressure. The
operating range can be extended to higher pressures either by reducing the cryosurface with the same refrigerator or by
using a more powerful refrigerator with the size of the cryosurface unchanged. It should be kept in mind that the coolant
heats up as it travels through the cryocoil. In order to avoid redistribution effects from the hotter to the cooler parts of
the cryosurface, the

< previous page page_533 next page >


< previous page page_534 next page >
Page 534

Fig. 9.24
Water vapor pumping speed of a 0.5-m2 cryosurface for a
refrigerator with AR = 50 W/K and different minimum
coolant temperatures.

design of the cryosurface should be based on its maximum temperaturethat is, on the outlet temperature of the coolant.

Pump Combinations with Cryo-surfaces. In high-vacuum systems, cryosurfaces with surface temperatures above about
80 K (193°C) can only be used to pump water vapor. For the noncondensable ''permanent" gases, other pumps are
required. In order to size the water vapor pumping speed SH2O relative to pumping speed for permanent gas Sp, we
consider the total and partial pressures:

Obviously, as soon as the water vapor partial pressure is small compared to the permanent gas partial pressure, a further
increase in water vapor pumping speed will not significantly reduce the total pressure. For the calculation of the total
pressure it is convenient to define a total pumping speed,

and express the water vapor flow rate as a fraction fH2O of the total flow rate,

yielding the following for the total pumping speed:


< previous page page_534 next page >
< previous page page_535 next page >
Page 535

Fig. 9.25
Gain of total pumping speed as a function of water
vapor pumping speed for different water vapor fractions.

For the fraction of water vapor in the gas of a vacuum system, no general rules exist. Mass spectrometer measurements
in high-vacuum systems taken during pumpdown and close to the base pressure resulted in fH2O valves for between
80% and 95%. Figure 9.25 illustrates the gain in total pumping speed with respect to the permanent gas pumping speed
as a function of the relative water vapor pumping speed. As suggested by the diagram, water vapor pumping speeds
which are 10 to 15 times the permanent gas pumping speed are probably an optimum for water vapor fractions of 95%
or less.

9.3.2.4
Cryopump Sets

Today, high-vacuum cryopump sets nearly exclusively make use of refrigerator-type pumps. Therefore the discussion of
cryopump set design will focus on this pump type. As an entrapment-type pump, cryopumps show design characteristics
which are not found in transfer pumps. In contrast to, for example, turbomolecular or diffusion pumps, cryopumps have
a limited storage capacity for pumped gas and therefore require regeneration at periodic intervals.

Like all other high-vacuum pumps, cryopumps exhibitin principlepumping speed characteristics as illustrated
schematically in Fig. 9.14. Published pumping speed data, however, only show the operating range with pressure-
independent speed.

For the overload range a pumping speed curve similar to that of cryosurfaces (Fig. 9.24) would be expected. With
cryopumps, however, this range is not accessible. In contrast to cryosurfaces, overload in a cryopump usually causes a
complete breakdown of the pumping action. As soon as the temperature on the low-temperature stage rises markedly
above 20 K, pumped gas is released from the adsorbing surfaces, resulting in a pressure rise within the cryopump. This
in turn by conduction through the gas increases the heat load from the pump housing at room temperature to the
cryopanels. Further temperature rise of the adsorber and gas release then usually require a shutdown and regeneration of
the pump. Therefore overload of cryopumps, even for short times, has to be avoided.

The operating range of most cryopumps begins at about 103 mbar and extends to the UHV range. In vacuum systems,
where operating pressures higher than the

< previous page page_535 next page >


< previous page page_536 next page >
Page 536

maximum inlet pressure of the cryopump are required, a throttle valve has to be added between the system and the
pump. The function is the same as discussed for throttled diffusion pumps (see Section 9.3.2.2).

The water vapor pumping speed of cryopumps is typically 3 to 4 times larger than the pumping speed for nitrogen and
argon. In order to calculate the pressure achievable with a cryopump, either the total pumping speed approach [Eq.
(9.44)] or explicit partial pressure calculations analogous to Eq. (9.42) should be taken.

The maximum operating time tmax before regeneration is necessary is limited by the amount of gas which can be stored
in the pump without noticeably degrading its performance. tmax can be calculated from the pump capacity Qmax as
published by the manufacturer and the average flow

For hydrogen the capacity is typically between 10 and 40 bar·liter, whereas for all heavier gases the capacity is on the
order of 1000 bar·liter. Hydrogen is therefore the critical gas in many cases. It should be noted that in some processes
such as plasma processesalthough not purposely introducedhydrogen may be produced by dissociation of water vapor.

The ultimate pressure attainable with a cryopump is not only a parameter of the design of the pump. Instead, it depends
on the type and amount of gas which has been already stored in the pump. For operation in high vacuum systems,
however, the ultimate pressure is usually low enough as long as the pump has not yet reached its capacity. In order to
attain ultimate pressures below 108 mbar cryopumps have to be well-regenerated and operated not too close to their
capacity limits.

Changeover from the fore-vacuum pump to the cryo-pump imposes a peak heat load on the cryo-surface since the gas
from the systems volume has to be pumped in a short time. In order to maintain the cryopanels at sufficient low
temperatures the changeover pressure pco has to be chosen according to

The changeover value Qco depends on the cooling capacity of the refrigerator and is specified by the pump
manufacturers. For a cryo-pump with an inlet flange of 200 mm, typically Qco ranges from 50 to 200 mbar·liter.

For regeneration of cryo-pumps, different procedures are applied for warming up and removal of the pumped gas (see
Chapter 5).

At the end of the regeneration cycle the cryopump has to be evacuated below 5 × 102 mbar before the refrigerator is
started. Since the adsorber panels are very sensitive to oil contamination, oil backstreaming from the pump set has to be
avoided. As illustrated in Fig. 9.26, this can be achieved either by an adsorber trap or alternatively by purge gas
introduced in the pump line between cryopump and forepump. The adsorber should be isolated by V4 during
pumpdown in order to avoid loading the adsorber material with water vapor. For the purge gas to be efficient in
avoiding oil backstreaming, the flow should be adjusted so that the pressure at the

< previous page page_536 next page >


< previous page page_537 next page >
Page 537

Fig. 9.26
Schematic diagram of a
cryopump set V4 is only
needed in combination
with the adsorber trap.
Alternatively to the
adsorber purge, gas can
be introduced via V5.

inlet of the forepump is 0.1 to 0.2 mbar and the gas inlet should be located near the cryopump fore-vacuum port.

9.4
Calculation Methods for Vacuum Systems

In order to derive solutions to the basic design equations in Section 9.1, highly simplifying assumptions have been
made. Conditions such as constant pumping speed or time-independent flow are hardly ever met in real systems.
Therefore these solutions can provide approximations for real systems; but for sufficiently accurate predictions of the
vacuum performance, calculations have to be based on realistic data of pumps, conductances, and gas flows.

In principle there are two ways to include real data in the design calculations: (i) Model all relevant data such as
pumping speeds and flows in an analytical form and solve the basic equations analytically or (ii) use the data and solve
the equations by numerical methods. Both approaches have their advantages and their limitations. Analytical methods
yield closed-form solutions which provide insight into the dependencies of different parameters. However, analytical
solutions are only available for simple approximations of, for example, pumping speed curves. Numerical methods on
the other hand are very powerful in including the most complex dependencies and system structures, but they require
appropriate software tools and they do not show dependencies directly.
In this section, analytical approximations and numerical methods are discussed.

< previous page page_537 next page >


< previous page page_538 next page >
Page 538

9.4.1
Analytical Approximations

Solutions for the basic differential equation [Eq. (9.5)] have been derived assuming constant pumping speed and
constant gas flow. In real systems these conditions are certainly not fulfilled for the entire pressure range. For limited
ranges, however, these assumptions are often well-approximated. Combining analytical solutions with appropriate
approximations can therefore provide useful practical solutions.

For the approximation of the pumping speed as discussed in Section 9.3.1, we obtained

and constant gas flow the basic differential equation has the solution

for the pumpdown of a system. During pumpdown the major gas load is due to outgassing. Since the outgassing flow
has a slow time-dependence compared to typical vacuum time constants, Eq. (9.47) also yields a reasonable
approximation for the pumpdown of a system with outgassing. Using Eq. (9.17) for the outgassing flow rate the
pressure during pumpdown is

To check the accuracy of this approximation, it has been compared with numerical solutions using the same pumping
speed and gas flow formulae. The largest differences occur in the transition range between volume-gas-controlled and
flow-controlled pumpingthat is, between range I and range II in Fig. 9.3. Here the deviations in the calculated pressures
are less than 20%. In range I as well as in range II the deviations between the approximate and the numerical solutions
are typically less than 5%.

Equation (9.48) provides an explicit pressure versus time curve. Although pumpdown times to specified pressures can
be extracted from calculated p(t) data, often it is desirable to have an explicit expression for the pumpdown time. Since
Eq. (9.48) cannot be solved for t explicitly, an alternate approach is needed. An estimate for the pumpdown time is
obtained considering the evacuation of the volume (range I) and pumping of outgassing flows (range II) independently.
For both steps a separate pumping time is calculated and the results are added to obtain the total pumpdown time:

< previous page page_538 next page >


< previous page page_539 next page >
Page 539

For a pumping speed according to Eq. (9.25) and an outgassing flow according to Eq. (9.17) the times are

This approach yields an estimate for the pump time which is larger than the actual pump time. As to be expected, the
largest errors occur in the transition region between range I and II. Here the estimated times can be too large by a factor
of 1.7. In range I the deviations are a few percent at most. In range II the time is less than 30% too large. Although not
very precise, this estimate makes it immediately obvious whether the pumpdown process is dominated by range I or
range II pumping by comparing tI and tII.

Up till now, only constant pumping speeds and ultimate pressure-limited pumping speeds have been considered. For
pump sets with Roots blowers or with high-vacuum pumps, at least step changes in the pumping speed have to be taken
into account. The calculation of the pumpdown curve using a step function speed is straightforward in principle: The
pressure range is subdivided into constant pumping speed ranges as illustrated in Fig. 9.27. Then p(t) is calculated for
each range separately from Eq. (9.48) using appropriate matching conditions.

From atmosphere to the cut in pressure pc,R the pressure is

As the pressure reaches the cut in pressure at t = tc,R the pumping speed increases to S2:

Finally after the changeover to the high-vacuum pump we obtain

It is important to note that the time argument in the exponential function and the associated starting pressure have to be
adopted to the different pumpdown ranges

< previous page page_539 next page >


< previous page page_540 next page >
Page 540

Fig. 9.27
(a) Step function approximation of the pumping speed.
(b) Calculated pumpdown curve. Note the change in the
slope of p(t) caused by the change of the pumping speed.

while the time argument for the outgassing flow remains unchanged. Although Eqs. (9.50) yield p(t) explicitly, the time-
matching conditions for switching from one pumping speed to the next have to be iterated numerically. The accuracy of
this approximation is mainly determined by the choice of the constant pumping speed representing the real speed data.
During the changeover from one pumping speed to the other, transient errors are to be expected. In reality the
changeover is not an instantaneous process, but it takes some time to activate valves and to run blowers up to speed.
Furthermore, during crossover the high-vacuum pump has to cross its overload range. Depending on the overload
handling capability, this can take some time. Fortunately these transient errors do not noticeably influence the pressure
in the long run.
Analogous to the pressuretime curve the estimate of pumpdown time, Eq. (9.49), can be applied to step function
pumping speeds. The volume gas pump times in the

< previous page page_540 next page >


< previous page page_541 next page >
Page 541

different speed ranges are added to the outgassing pump time. For the pumpdown to a pressure below the crossover
pressure to the high-vacuum pump the estimated pump time is

For this estimate, about the same errors as for Eq. (9.49) are to be expected. On top of these intrinsic errors the
uncertainties in S1 to S3 have to be considered.

For pumpdown calculations Eqs. (9.50) and (9.51) constitute about the maximum complexity of analytical
approximation which can be handled with reasonable effort. Better approximations of pumping speeds or gas flows
should be analyzed using numerical methods.

9.4.2
Numerical Methods

The analytical methods outlined in Section 9.4.1although powerful and easy to usereach their limitations when precise
predictions of vacuum system performance are required or when multiple interacting vacuum chambers have to be
considered. In these cases, numerical methods have to be used which can handle the complexities involved in the
calculations.

Two different approaches have been taken for the numerical solution of vacuum calculation problems: Dedicated
vacuum system design software has been developed and general simulation software has been adapted to vacuum
problems.

9.4.2.1
Dedicated Software

Facing the challenge of precise pumping speed and pumpdown predictions the vacuum industry developed numerical
methods and dedicated programs for vacuum system calculations.

For the calculation of the pumping speed, approximate approaches are usually not precise enough when accuracies of
10% or better are required. Instead, measured pumping speed data have to be used whenever possible. From a numerical
point of view, pumps can be classified into two categories: (1) pumps whose pumping speed is
< previous page page_541 next page >
< previous page page_542 next page >
Page 542

directly available from measured data and (2) pumps where the pumping speed has to be calculated from characteristic
data of the pump and from properties of the remaining pump set. Typical examples for the first group are fore pumps,
diffusion pumps, and turbomolecular pumps, while Roots blowers and cryo-surfaces belong to the second category.

For pumps for which measured pumping speed data are directly available the construction of the S(p) curve is
straightforward. The pumping speed is interpolated on a logSlogp basis as required by the pumpdown calculation
algorithm.

The basic calculation steps for pumps of the second category have been outlined in Section 9.3. In order to predict the
vacuum performance of Roots blowers and cryo-surfaces accurately, their individual characteristic data have to be
known. For instance, the maximum compression ratio of Roots blowers is influenced by the detailed internal geometry
of the pump, resulting in different k0 curves for blowers of different size. Furthermore, not only Sth but also k0 for a
given machine depends on its rotational speed.

In order to obtain accurate effective pumping speed data the pressure-dependent conductance of all flow restrictions
between the pump set and the system has to be considered. In some pump sets there are noticeable flow resistances
within the pump set which have to be taken into account: tubes of considerable length or filters installed between the
different pump stages. Furthermore, purge gas may be needed at different points within the pump set, which also
influences the effective pumping speed.

With the so determined effective pumping speed and appropriate gas load data the basic differential equation [Eq. (9.5)]
has to be solved. Considering the large pressure range to be coveredin most cases more than nine decadesand the step
changes in pumping speed or gas flow, stable numerical methods are needed. One proven approach is the use of an
ansatz function similar to Eq. (9.47) for small time steps. For the limiting cases of constant speed and gas flow, this is
the exact solution independent of the size of the time steps. For all other caseswhich are by far the majorityin addition to
the ansatz function a suitable predictorcorrector scheme has been implemented. This method results in high numerical
stability. It handles sudden changes of speed or flow without any problem, and it also handles the complete pump-down
from atmospheric pressure to the system base pressure.

Such programs, which include all the pump and conductance aspects, measured gas flow versus time data, and suitable
solution strategies, are able to predict the pumpdown with accuracies of better than 10% over the entire pressure range
from atmospheric to high-vacuum. Besides providing reliable predictions for new systems, the accuracies allow the use
of the results of the simulation as an analytical tool for existing vacuum equipment. By comparing measured and
calculated pressure versus time data, pump problems or hidden flow restrictions can be identified.

9.4.2.2
Network Approach

All systems considered so far are based on the schematic diagram in Fig. 9.1that is, a chamber with one pump set and
different gas sources assumed to be at a uniform pressure. This model requires that the pressure drop within the system
due to the conductance Csystem between the gas sources and the pump port be small compared to the absolute pressure;
that is, Csystem must be much larger than the effective pumping speed.

While adequate for typical single-chamber systems the model fails in most multichamber systems and in extended
systems, where distributed pumps and gas sources

< previous page page_542 next page >


< previous page page_543 next page >
Page 543

Fig. 9.28
Schematic diagram of a
vacuum system with two
chambers coupled by a
conductance C12.

have to be taken into account. Figure 9.28 gives an example for a two-chamber system. Through the conductance
between the chambers, gas flow between the chambers is possible. The straightforward but tedious approach would be
to set up a system of coupled differential equations for both chambers and solve it for p1 and p2; but for networks of
elements, other methods already exist.

It has long been known that an analogy exists between vacuum systems and electrical networks. For electrical network
analysis, sophisticated software tools are available. This suggests that we should apply these tools to vacuum system
calculations.

The relation between vacuum metrics and electrical metrics are found by comparing terms in similar equations. Ohm's
law relates an electrical current I with the voltage drop V across an electrical conductance G by

The corresponding equation for the gas flow is

where C is the conductance of the element and p is the pressure drop across C. Charging and discharging an electrical
capacitor of capacitance C is described by

whereas the change of the gas content in a volume V follows:


It is important to note that the symbols V and V and C and C have completely different meaning in electrical and
vacuum context and they should not be confused.

< previous page page_543 next page >


< previous page page_544 next page >
Page 544

Table 9.2. Correspondence Between Electrical and Vacuum Metrics


Electrical Metric Electrical Symbol Electrical Vacuum Metric Vacuum Symbol Vacuum Units
Unit
Voltage V V Pressure p mbar
Pa

Current I A Gas flow mbar·liter·s1


Pa·m3·s1
Conductance G 1/Ω Conductance C liter·s1
m3·s1
Capacitance C F Volume V liter
m3

Matching terms in these four equations yields a correspondence table between vacuum metrics and electrical metrics (Table 9.2). Although for
electrical units a single standard is established for engineering purposes in vacuum technology, different systems are in use. Any of these
systems can be used with circuit simulation software as long as self consistency is maintained within the units. Table 9.2 gives two examples for
sets of units. With the correspondence table, vacuum components can be translated into electrical circuit elements (Figure 9.29). A volume is
represented by a capacitor with one side connected to ground potential. In vacuum metrics this ground potential corresponds to zero absolute
pressure, the reference level relative to which the pressures are measured.

Flow resistances without internal volume, such as orifices, are modeled by a simple resistor. Tubes, where not only the conductance but also
the volume has to be considered, are represented by resistorcapacitorresistor networks. The resistors have to be chosen so that their series
connection yields the total conductance of the tube, whereas the capacitor represents the entire volume. For long tubes and in cases where
pressure distributions along tubes are to be investigated, several of these networks can be connected in series.

Pumps with constant pumping speed can be represented by a conductance to the zero-pressure reference nodethat is, to ground potential. In
cases where the ultimate pressure of the pump shall be included according to Eq. (9.25), the equivalent circuit consists of a conductance G
corresponding to Smax in series with a constant voltage source, which is to be set to voltage corresponding to pu.

Most gas loads can be approximated by current sources injecting currents irrespective of the pressure.

For the vacuum system shown in Fig. 9.28, the equivalent electrical circuit is illustrated in Fig. 9.30. With this network, many aspects of the
vacuum performance of the system can be studied. Starting the simulation with the volumes ''charged" to atmospheric pressure and imposing
appropriate time-dependent outgassing flows 1 and 2 yields the pumpdown curve. Similarly, process operation can be simulated by adding
process gas flow where necessary. Propagation of pressure changes from one chamber to the other are easily investigated and even opening and
closing of a valve between the two chambers can be simulated by varying the conductance C12 from zero (i.e., valve closed) to the conductance
of the open valve. This great flexibility has turned circuit simulation into an indispensable tool for calculations in complex vacuum systems.

There are, however, some requirements on the circuit simulation software in order to be useable for vacuum system simulation. The pressure
range to be covered in

< previous page page_544 next page >


< previous page page_545 next page >
Page 545

Fig. 9.29
Translation of vacuum elements to equivalent circuits.
Fig. 9.30
Electrical equivalent circuit of the two-
chamber vacuum system in Fig. 9.28.

a vacuum simulation often extends over more than nine decadesfor example, from atmospheric pressure to less than 106
mbar. The software must be able to handle this large range with sufficient accuracy. Vacuum elements such as
conductances and pumps usually exhibit considerable nonlinear behavior. Therefore for an adequate

< previous page page_545 next page >


< previous page page_546 next page >
Page 546

simulation it must be possible to use nonlinear elements defined by tabulated data or formulae. Finally, arbitrarily
definable time functions for the current sources are needed to simulate gas loads.

General References

D. Degras, Le Vide 64, 155 (1956).

S. Dushman, in Scientific Foundations of Vacuum Technique (J. M. Lafferty, ed.), 2nd ed. Wiley, New York, 1962.

J. Elsley, Vacuum 25, 229 (1975).

J. Elsley, Vacuum 25, 347 (1975).

Th. Gebele and W. Buschbeck, LEYBOLD SYSTEMS, private communication.

A. Haefer, Kryo-Vakuumtechnik, Grundlagen und Anwendungen. Springer, Berlin, 1981.

G. Horikoshi, J. Vac. Sci. Technol. A 5, 2501 (1987).

K. Kanazawa, J. Vac. Sci. Technol. A 7, 3361 (1989).

J. F. O'Hanlon, A User's Guide to Vacuum Technology, 2nd ed. Wiley, New York, 1989.

J. Santeler, J. Vac. Sci. Technol. A 5, 2472 (1987).

W. Schwarz, J. Vac. Sci. Technol. A 5, 2568 (1987).

S. R. Wilson, J. Vac. Sci. Technol. A 5, 2479 (1987).

M. Wutz, H. Adam, and W. Walcher, eds, Theorie und Praxis der Vakuumtechnik, 3rd ed. Vieweg Verlagsges.,
Braunschweig, 1986.

< previous page page_546 next page >


< previous page page_547 next page >
Page 547

10
Gas-Surface Interactions and Diffusion

John B. Hudson

At the pressures encountered in typical vacuum systems, the mean free path of gas molecules is very long compared to
the dimensions of the apparatus. Consequently, the behavior of the gas is dominated by gassurface collisions. In many
cases, the energetic interactions between the gas molecules and surfaces within the system are sufficiently strong that
the gas molecules will be trapped at the surface, or sorbed, for a period ranging from nanoseconds to essentially forever.
The magnitude of this surface lifetime is a critical factor in a number of the processes involved in vacuum system
operation. Pumping processes such as sorption or cryopumping depend on the trapping of gases on cold surfaces by
physical adsorption. Getter pumping depends on the uptake of active gases by a combination of chemisorption and
absorption into the bulk of the getter material. The gas load that must be handled by the system pumps consists, except
for a brief period at the beginning of the pumpdown cycle, of gases leaving the surface by a combination of (a)
permeation through the surface from the bulk of the material exposed to the vacuum and (b) desorption from adsorbed
layers present on the surfaces themselves.

In the material that follows, these processes of sorption and desorption will be discussed in detail, with emphasis on the
effect that these phenomena have on the performance of vacuum systems and on the operations carried out within these
systems. In this discussion the interplay of four processes will be seen to control the net rate of uptake or emission of
gas from the surfaces of materials exposed to the vacuum environment.

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_547 next page >


< previous page page_548 next page >
Page 548

Fig. 10.1
Processes involved in the overall
rate of sorption or desorption of
a gas at a surface.

The sorption of gases on surfaces, either as adsorption on the free surface or as solution into the bulk below the surface,
can be treated by either a thermodynamic or a kinetic approach. In most processes of interest in vacuum technology, the
rates of adsorption and desorption phenomena will be of greater importance than the equilibrium amounts of gas
adsorbed on, or dissolved in, the surfaces of the system. Moreover, it will be shown that the equilibrium conditions can
always be obtained by equating the rates of all sorption processes to the rates of the corresponding desorption processes
for any set of system conditions. Consequently, the kinetic approach to these processes will be emphasized here.

The rates of the surface-related processes that may occur, in the most general case, may be represented as shown in Fig.
10.1. This figure shows, schematically, the rates of the competing surface processes such as adsorption into the adlayer,
Ra, and dissolution into the bulk, Rs, which remove material from the gas phase, and processes such as permeation from
the solid, Rp, and desorption from the adlayer, Rd, which liberate material into the gas phase. For any set of system
operating conditions, the net rate of pumping or outgassing associated with the surface will be set by the balance of
these rates.

10.1
Adsorption

10.1.1
Basic Equations

In many cases, the rates of one or more of the processes described above will be either negligible or extremely fast, and
the behavior of the system can be characterized completely in terms of one or two of these rates. Adsorptiondesorption
phenomena can be described, in the absence of effects due to bulk processes, in terms of the parameters that control the
rates of adsorption and desorption. These are the molecular impingement rate, I, which was defined in Section 1.4 as

which sets an upper limit on the rate at which material can accumulate on the surface; the desorption frequency, vd,
which is the probability of desorption per adsorbed species per unit time and which will depend upon (1) the nature of
the gassurface interaction in any specific system and (2) the temperature and the amount of the gas adsorbed; the
instantaneous surface coverage, na; and the probability that a molecule striking the surface will be accommodated into
the adsorbed layer, S, which may be a function of the amount adsorbed.

< previous page page_548 next page >


< previous page page_549 next page >
Page 549

The amount adsorbed at equilibrium, neq (molecules/unit area), is related to these parameters through the equation

where n represents the order of the desorption process and is usually a small integer, and the subscript eq indicates that
S and vd are the values appropriate to equilibrium coverage. It must be borne in mind that because n may differ from
unity and because vd and S may depend on the adlayer coverage, a linear relation between neq and pressure will be
observed only in the simplest of adsorption systems.

The temperature dependence of the equilibrium coverage is contained in vd. The physical significance of vd can be
described in terms of the energetic interaction between the adsorbed species and the surface. In the simplest case, this
interaction can be represented in terms of a one-dimensional potential energy well, as shown in Fig. 10.2. This figure
shows the change in system potential energy that takes place as a molecule from the gas phase approaches an adsorption
site on the surface. Once the molecule is trapped in this potential well, the probability of escape is given by absolute
reaction rate theory as
Fig. 10.2
One-dimensional potential well for a gas molecule approaching a solid
surface.

< previous page page_549 next page >


< previous page page_550 next page >
Page 550

where ν'0 is an attempt frequency, on the order of the vibrational frequency of the adsorbed species (~ 1013 s1), and
is the molar free energy of activation for the process. For a potential well of the form shown in Fig. 10.2, this
latter term may be approximated as

where f and f* are the molecular partition functions of the system in the equilibrium and activated states, respectively.
Thus

where ν0 is ν'0/(f*/f).

Alternatively, rather than looking at the desorption frequency, it is often more useful to use the reciprocal of this term,
which is known as the mean stay time for adsorption or mean surface lifetime. This is given by

Note that this implies that the deeper the potential wellthat is, the greater the magnitude of ∆Hdthe longer will be the
mean stay time. The values of τ0 measured experimentally range from about 1016 to 109 s, implying that the ratio f/f*
ranges between about 103 and 104. Equation (10.3) may be rewritten in terms of the mean stay time to yield

The existence of a finite value for τa means that a given molecule will spend a proportionally larger fraction of its time
near the surface than anywhere else in the system. Thus, the time-averaged concentration is going to be higher near the
surface than in bulk of the gas phase. This is just the condition that is described classically as adsorption.

< previous page page_550 next page >


< previous page page_551 next page >
Page 551

The value of τa in any situation is a strong function of both ∆Hd and T. The value of ∆Hd is in turn a function of the type of attractive
forces present in the system. Systems in which the only attractive forces are of the van der Waals, or dispersion, type show values in the
range from 100 cal/mol to 5000 cal/mol. Customarily, adsorption processes involving forces of this magnitude are called physical
adsorption or physisorption processes. Systems in which hydrogen bonding, covalent chemical bonding, or metallic bonding can take
place will show values of ∆Hd ranging from 5 kcal/mol to as high as 150 kcal/mol. Adsorption processes involving forces of this type are
referred to as chemisorption processes. The effect of this wide range of observed ∆Hd values on the stay time is summarized in Table
10.1 [1], based on the assumption that T = 300 K and τ0 = 1013 s. As can be seen from the table, the values of τa cover a range from
times essentially equal to τ0 at the low end, to inconceivably long times at the high end. Note, however, that there is a fairly wide range of
∆Hd values for which τa is within a few orders of magnitude of 1 s, and that this range can be greatly extended by changing the
temperature. For example, if the temperature were 600 K instead of 300 K, the value of τa associated with ∆Hd = 40,000 cal/mol would
drop from 1017 s, which is approximately the age of the earth, to 1 s, a readily conceivable and experimentally measurable value.

10.1.2
Adsorption Isotherms

Consider next a number of possible assumptions concerning the value of n, the relation between S and na, and the form of the parameter
τa, and also consider the effect of these assumptions on the form of the relation between neq and p at constant temperaturethat is, the
adsorption isotherm. The simplest set of assumptions that one can make are that n = 1 and that τa and S are independent of na. The form of

Table 10.1. Mean Stay Time for Adsorbed Molecules at 300 K for Various Values of the Adsorption Energy, Assuming τ0 =
1013 sa
∆Hd Typical Cases τa (s)
100 cal/mol Helium
1.2 × 1013
1.5 kcal/mol H2 physisorbed
1.3 × 1012
3.54 kcal/mol Ar, CO, N2, CO2 (physisorbed)
1 × 1011
1015 kcal/mol Weak chemisorption
Organics physisorbed 3 × 106
2 × 102
20 kcal/mol H2 chemisorbed
100
25 kcal/mol
6 × 105
(1 week)
30 kcal/mol CO chemisorbed on Ni
4 × 109
(> 100 yr)
40 kcal/mol
1 × 1017
(≈ age of the earth)
150 kcal/mol O chemisorbed on W
101100
(≈ 101090 centuries)
a Adapted from DeBoer [1].

< previous page page_551 next page >


< previous page page_552 next page >
Page 552

Eq. (10.3) in this case will be

which at constant temperature leads to

This behavior, which is known as Henry's law, predicts a linear increase of neq with p, as shown in Fig. 10.3. The
obvious difficulty with this model is that it predicts an unlimited adlayer coverage as p increases, which is not observed
in practice.

In order to account for the fact that many adsorption systems appear to show adsorption on specific surface sites and
that adsorption is limited to one adsorbed species per site, we may introduce what is known as the Langmuir model of
the adsorption process [2]. In this model it is assumed that the adspecies are bound to a fixed number, n0, of adsorption
sites per unit area, with no more than one adspecies per site, that the value of ∆Hd is independent of coverage and is the
same for all sites, and that the adsorption probability is finite for impingement on empty sites, but zero for impingement
on occupied sites. This leads to

Fig. 10.3
Henry's law isotherms showing equilibrium adlayer
coverage versus pressure for a range of temperatures.
< previous page page_552 next page >
< previous page page_553 next page >
Page 553

and to

Defining the fractional coverage, θ, as

and dividing Eq. (10.13) by n0 yields

This may be solved to yield

This equation is known as the Langmuir adsorption isotherm and is plotted in Fig. 10.4. Note that at low pressures,
where this relation reduces to the Henry's law relation

At high pressures, where the equation reduces to

and all adsorption sites are filled.


The Langmuir model provides a good description of the adsorption process in many systems in which strong
chemisorption occurs. It does not provide a very good description in systems involving relatively weak adsorption
forces (physical adsorption), because it neglects lateral interactions among adspecies, surface mobility, surface
heterogeneity, and the possibility of adlayers thicker than one monolayer.

< previous page page_553 next page >


< previous page page_554 next page >
Page 554

Fig. 10.4
The Langmuir adsorption isotherm.

One may also consider another set of assumptions, similar to those of the Langmuir model, for the fairly common case
of the adsorption of a diatomic gas, in which the gas molecule dissociates in the adsorption process to yield two
adsorbed atoms, one per adsorption site, and in which the two adsorbed atoms must recombine in order to desorb. In this
case, the following equation is based on the assumption that in order for the molecule to adsorb dissociatively, two
adjacent unoccupied sites are necessary:

The desorption rate for this case will be

leading to

Substituting θ = neq/n0 and solving for θ yields


< previous page page_554 next page >
< previous page page_555 next page >
Page 555

This equation, which is sometimes referred to as the dissociative Langmuir isotherm, is shown in Fig. 10.5.

In physical adsorption, in which strong chemical binding forces are not involved, the potential well associated with the
adsorption process arises from the relatively weak van der Waals interactions between the adsorbate and the atoms in
the solid surface. This potential well can be approximated by summing the pairwise interactions between the adsorbate
and the atoms in the surface as shown in Fig. 10.6, summing over as many atoms as necessary to include all atoms that
make a significant contribution to the system potential energy. These individual pairwise interactions

Fig. 10.5
The dissociative Langmuir isotherm.
Fig. 10.6
Summation of pairwise interatomic interactions to
determine the energy of adsorption for physical
adsorption.

< previous page page_555 next page >


< previous page page_556 next page >
Page 556

can generally be represented by an expression of the form

where n and m are integers. For the case in which the attractive forces involved are the induced dipole forces (van der
Waals or dispersion forces), the appropriate values of these parameters are n = 6, m = 12,
the so-called LennardJones potential. Here ε is the maximum negative value of the
pair potential and re is the internuclear separation associated with this energy minimum. In the event that the surface is
that of a crystalline solid, with regular atomic spacing, the system energy, u(xyz), will depend on the x, y, z coordinates
of the adsorbing species relative to some reference point on the crystal surface. We may represent the z dependence of
this energy by plotting u(z) versus z, as shown schematically in Fig. 10.7. Here the zero of potential energy is taken as
the gas-phase species at infinite distance from the surface. As z decreases, the energy of the system becomes negative
(implying attractive forces between the gas-phase species and the surface), reaches a maximum negative value at some
value ze, and then rises and becomes positive (repulsive) at very small values of z. The variation of u(x, y, z) as a
function of x or y is most conveniently expressed in terms of the variation of the minimum of the u(z) curve, u0, with x
or y. A schematic plot of u0 versus x is shown in Fig. 10.8. The regular variation of u0 with x reflects the regular
interatomic spacing on the crystal surface. A plot of u0 versus y would be similar in appearance. This plot indicates that
there are favored values of x (and y) at which u0 has a maximum negative value. These sites, characterized by u00 on
the figure, represent the equilibrium binding sites for the species adsorbed on

Fig. 10.7
One-dimensional potential energy curve showing potential energy
as a function of distance to the surface for the Lennard-Jones
612 interatomic potential.

< previous page page_556 next page >


< previous page page_557 next page >
Page 557

Fig. 10.8
Variation of adsorption potential well depth with lateral
position along a crystalline solid surface.

Fig. 10.9
A two-dimensional potential energy surface showing
the variation of potential well depth with impact
parameter for an atom approaching a solid surface.
Dotted contours indicate areas of strongest
attractive interaction; solid contours indicate
regions of weakest interaction.
the surface. The intervening minima in the absolute value of u0, designated by v0, represent less stable configurations
that the absorbed species must pass through in order to migrate from one favored site to another. This variation of
adsorption energy as a function of position on the surface may also be shown by plotting contours of equal u0 as a
function of x and y, as shown in Fig. 10.9. If the value of kT is small compared to v0, negligible migration will occur,
and the adsorbed species can be considered as bound at the equilibrium sites, a situation known as immobile adsorption.
In the alternative case, kT comparable to or greater than v0, migration will be possible, and mobile adsorption results.

< previous page page_557 next page >


< previous page page_558 next page >
Page 558

In many cases of physical adsorption, it has been observed that the adsorption process continues well beyond the
monolayer coverage limit imposed by the Langmuir model. Several treatments have been developed to describe the
isotherm equation for this case. The most commonly used is the so-called BrunauerEmmettTeller [3], or BET, model.
These workers concluded that many of the experimentally observed adsorption isotherms reported in the literature for
physical adsorption showed evidence of multimolecular layer adsorption. These isotherms could be characterized as
belonging to one of five types, shown in Fig. 10.10. The isotherms of type I are, of course, Langmuir isotherms and are
characteristic of adsorption of only a single molecular layer. The other four types are generally considered to represent
multilayer adsorption.

The derivation of the simple form of the BET equation is based on the same assumptions as those involved in the
derivation of the Langmuir equation, with the added hypothesis that the formation of successive layers can occur by the
adsorption of molecules on top of an already formed previous layer. The model assumes that when adsorptive
equilibrium has been attained, the total surface area can be broken down into patches, each of which is covered with
zero, one, two, or more monolayers of adsorbed material, as shown in Fig. 10.11. Over a period of time, the fraction of
the surface having a given coverage will remain constant. That is, if θi is that fraction of the surface covered by i
monolayers, then at equilibrium we obtain dθi/dt = 0 for all i, and the total amount of material adsorbed is

Fig. 10.10
The five isotherm types observed in physical adsorption, according to Brunauer et al. [3].

< previous page page_558 next page >


< previous page page_559 next page >
Page 559

Fig. 10.11
Model used to develop the BET adsorption equation.
Patches on the surface are classified in terms of
the number of monolayers in that patch.

or

Considering the dynamic nature of the adsorptive equilibrium, it is evident that this implies that the net rate of processes
contributing to an increase of a given θi must be the same as the net rate of those processes contributing to a decrease in
that same θi. Processes contributing to the increase in a given θi are (a) the impingement of molecules on patches of θ(i
1) to convert them to patches of θi and (b) desorption of molecules from patches of θ(i + 1). Processes contributing to a
decrease in θi include the adsorption of molecules on patches of coverage θi, to convert them to patches of θ(i + 1), and
desorption of molecules from patches of θi to convert them to patches of θ(i 1). Thus, for each θi one may write an
expression stating that
< previous page page_559 next page >
< previous page page_560 next page >
Page 560

Subtracting the expression for the change in θ0 from that for the change in θ1 leaves

By a similar process for the other θi, subtracting the expression for changes in θ(i 1) from the expression for changes in
θi leaves a set of simultaneous equations of the form

When this set of equations is solved, subject to the assumption that τa2 = τa3 = τai one obtains an expression for na of
the form

Substituting I = p/(2πmkT)1/2 in the expression for x and making the further substitution that β = 1/(2πmkT)1/2, one
obtains

Note that x is dimensionless, because βτa2/n0 has units of (pressure)1. It is thus convenient to define another parameter,
q, having units of pressure, as

The final expression for adlayer coverage is thus


< previous page page_560 next page >
< previous page page_561 next page >
Page 561

Equation (10.41) has the property that as p → q, the denominator in the expression approaches zero. Thus, as p → q, θ→
∞.

It is observed in many systems in which physical adsorption occurs that as p approaches p0 (the equilibrium vapor
pressure of the bulk condensed phase of the adsorbed species at the temperature of the experiment), the amount
adsorbed rises sharply and appears headed for infinity. It has thus been customary to associate q with p0 and,
consequently, to associate τa2 with ∆Hv, the enthalpy of vaporization of the condensed adsorbate. No fundamental
significance is to be attached to this choiceit is merely a convenient choice based on empirical observation. The form of
the final expression for the adsorption isotherm in the BET model leads to the adsorption isotherm shapes shown in Fig.
10.12. In this figure, the curve for q = p0 corresponds to the type III isotherm, while the curves for q < p0 and q > p0
correspond to types V and I respectively.

The isotherm equation may be rearranged into a form convenient for determining the values of the parameters n0 and kB
by presenting the experimental results as

where the substitution q = p0 has been made, and the adlayer coverage has been expressed in terms of the volume of gas
at STP adsorbed at equilibrium, V, and the volume of adsorbed gas associated with monolayer coverage, Vm. This is
equivalent to stating

Thus, for an adsorption isotherm which obeys this relation, a plot of the expression pV1 (p0 p)1 versus p/p0 should yield
a straight line for which the intercept is 1/(VmkB) and the slope is (kB 1)/(VmkB).

Fig. 10.12
The BET adsorption equation for various values of the relation of q to p0.
< previous page page_561 next page >
< previous page page_562 next page >
Page 562

If the adsorption is limited to n layers, Eq. (10.42) must be replaced by the relation

where x = p/p0.

It should be noted that this equation has two limiting cases: When n = infinity, it reduces to Eq. (10.42) (since x must be
less than 1), whereas if n = 1 (that is, if only one layer can be adsorbed) it assumes the form

that is, the Langmuir equation with Langmuir's constant χ having the value kB/p0 and θ = V/Vm.

If it is assumed that the adsorption process is influenced by the presence of capillaries and that bulk liquid condenses in
the capillaries as the pressure of the adsorbate gas approaches saturation, a rather more complex equation can be
derived. This was done by Brunauer et al. [4]. Their extended equation, although not commonly used, is of interest
because it represented the first equation to be derived which could describe isotherms of all of the five types shown in
Fig. 10.10. The subject of capillary condensation is treated in more detail later in this chapter.

The BET equation has found widespread use for the determination of surface areas. It has been found that the
adsorption isotherms for nitrogen and the rare gases argon and krypton at low temperatures (commonly liquid nitrogen
temperature, 77.4 K, is used) on a large variety of adsorbents fit the BET equation over a considerable range. A large
number of investigations have yielded reasonable values for Vm on various adsorbents when these isotherms are
determined and plotted according to the BET equation. The measurement of ''BET surface areas" is a standard technique
in many laboratories. The determination of a surface-area value involves the measurement of adsorption over a range of
pressures up to p/p0 about 0.3. These data are then plotted in accord with Eq. (10.42) as p/V (p0 p) versus p/p0. The
slope and intercept of the resulting linear plot then give the values of Vm and kB. In the model used to derive the
equation, Vm is the volume adsorbed in the first monolayer; kB is a constant which is proportional to exp {(∆H1∆Hv)/
RT}, where ∆H1 is the average heat of adsorption in the first layer and ∆Hv is the heat of liquefaction of the bulk
adsorbate.

A second class of models of the physical adsorption process treats the adsorbed layer as a two-dimensional gas. This
treatment will be justified in those cases where the barrier to motion between sites is small compared to kT, and the
adsorbed species can be considered to be free to move over the surface. The equation of state for the gas is stated in
terms of a two-dimensional pressure, Π, and an area per molecule, σ, analogously to the three-dimensional pressure, p,
and volume, V. Models of this type are useful, in that they permit one to take into account the interactions among
adsorbed species.

One may treat this behavior in the simplest case by assuming that adatoms interact only with the edges of the two-
dimensional surface and make no allowance for the fact that each adatom occupies a finite amount of surface area (that
is, assuming that the

< previous page page_562 next page >


< previous page page_563 next page >
Page 563

whole surface is available to each adatom). Furthermore, one may assume that adatoms collide elastically on the surface
with no energetic interaction, and that this two-dimensional gas can be described by the two-dimensional analog of the
three-dimensional ideal gas law.

Consider the application of these assumptions to a clean surface bounded by some kind of barrierfor example, the
surface of a liquid in a vessel. There will be a force, γ, exerted on the barrier because of the surface tension of the liquid.
When adsorption takes place on this surface, γ will be reduced. The tension on the barrier is consequently reduced as γ
is reduced. One can equally well look on this reduction in tension as a pressure exerted on the barrier by the two-
dimensional gas.

One can develop this relation mathematically by beginning with the Gibbs absorption isotherm,

where µi is the chemical potential per molecule for species i. For the case of adsorption from a one-component gas
phase onto a surface, using

one can determine that

If, now, a two-dimensional pressure, Π, is defined as

this leads to

Integration of this expression yields

where na may be a function of p.

Consider the case of Henry's law adsorption, as discussed previously. In this case

at any given temperature. Thus


< previous page page_563 next page >
< previous page page_564 next page >
Page 564

or

Substituting these results into Eq. (10.51) yields

where σ = 1/na is the surface area per adatom. This expression has the same form as the three-dimensional ideal gas law,

where NAv is Avogadro's number. The behavior of the ideal two-dimensional gas is thus seen to be exactly analogous
to that of the ideal three-dimensional gas.

Alternatively, one may make more realistic assumptions concerning the behavior of the two-dimensional gas. One may
assume that each molecule occupies some finite area on the surface, b2, and that there is an interaction energy between
any two adsorbed molecules, leading to a lateral interaction force, a2. These are essentially the same assumptions that
are made in developing the van der Waals equation of state for a three-dimensional gas and lead in this case to the two-
dimensional equation of state:

Making the additional assumption that b2 is 1/n0 and using

Eq. (10.59) may be rearranged to yield


< previous page page_564 next page >
< previous page page_565 next page >
Page 565

Using this relation and the previously stated relation between Π and p [Eq. (10.51)], one can determine that

where K is a constant related to the adsorption energy, ∆Ea. This relation, known as the HillDeBoer equation, is the
adsorption isotherm equation for a two-dimensional gas with lateral interactions on a homogeneous surface. The effect
of the lateral interaction energy appears in the last term in the brackets on the right-hand side of Eq. (10.62).

The form of the HillDeBoer equation is shown in Fig. 10.13. At high temperature and low coverage, where the effects
of excluded area and lateral interactions will be minimal, the exponential term approaches unity, (1 θ) approaches unity,
and the isotherm equation reduces to

which is the Henry's law isotherm. At lower temperatures, or for larger values of a2, and at higher coverages, increasing
deviations from Henry's law are observed. Below a certain temperature, Tc2 (the two-dimensional critical temperature),
the curve described by Eq. (10.62) becomes double-valued, and a first-order phase change is observed, just as in the
case of a three-dimensional van der Waals gas. Within the region bounded by the dashed curve, two phases exist at
equilibrium: a two-dimensional gas and a two-dimensional solid or liquid. At higher pressure, only the condensed phase
will be present.

Fig. 10.13
Adsorption isotherms according to the HillDeBoer
equation showing two-dimensional condensation.

< previous page page_565 next page >


< previous page page_566 next page >
Page 566

An additional complication in the characterization of physical adsorption behavior is that of surface heterogeneity. To
this point, it has been tacitly assumed that the heat of adsorption, or desorption energy, is the same for all potential
adsorption sites on the surface. In practice, this is seldom the case. The surfaces of real crystalline materials contain a
number of surface defects. Even on single-crystal surfaces, defects such as surface steps, vacant surface sites, adsorbed
atoms, and sites of dislocation emergence from the bulk are present. The configurations associated with a number of
these defects are shown schematically in Fig. 10.14. The surfaces of polycrystalline materials are even more complex,
because surfaces made up of different crystal planes will have different atomic arrangements and consequently different
adsorption energies.

The net effect of these surface defects is a quasicontinuous range of adsorption site energies. This distribution manifests
itself in adsorption isotherms on nearly perfect surfaces as an increase in the amount adsorbed at low adsorbate
pressures relative to that expected for the perfect surface, because the adsorption energies associated with the defect
sites are in most cases larger than those associated with perfectly ordered surface regions. At any given temperature, the
surface lifetimes on these high-energy sites will be longer than for the perfect surface sites, resulting in a higher
equilibrium coverage. There have been no direct correlations between heterogeneous adsorption behavior and specific
types of surface defects, but empirical models based on assumed distributions of adsorption site energies, such as a
Gaussian distribution, have had considerable success in fitting adsorption isotherms on heterogeneous surfaces. This
approach has been treated in detail by Ross and Olivier [5].

For completeness, it is necessary to mention two empirical adsorption isotherms that have been used extensively in the
past to describe adsorption data, especially physical adsorption data. These are the Freundlich isotherm,

and the Temkin isotherm,

Fig. 10.14
Typical surface sites and defects on a simple cubic (100) surface.

< previous page page_566 next page >


< previous page page_567 next page >
Page 567

Both of these isotherms have the property that the adlayer coverage increases more slowly than linearly with
temperature, as is commonly observed. In the case of the Freundlich isotherm, the parameter n is temperature-
dependent, and n tends to be large at low temperature. The parameter K is also temperature-dependent, and it decreases
with increasing T. Comparing the forms of these isotherm equations with the other isotherm equations developed above,
the parameter K in both cases must contain the surface lifetime, explaining its decrease with increasing temperature. The
parameter n in the Freundlich isotherm corresponds in part to the exponent of p in the Langmuir and dissociative
Langmiur isotherms. The fact that it is temperature-dependent probably is related to surface heterogeneity, because this
isotherm has been most frequently used to characterize adsorption on nonuniform surfaces. Similarly, the logarithmic
dependence of coverage on pressure in the Temkin isotherm would arise from a combination of site blocking and
surface heterogeneity effects, both of which would lead to a less rapid than linear rise of coverage with pressure.

10.1.3
Heat of Adsorption

In many cases, measurement of the heat liberated in the course of the adsorption process may be used to characterize the
variation of the adsorption energy with coverage due either to surface heterogeneity or to adsorbateadsorbate
interactions. For systems in which the adsorption process is reversible at the temperature of measurement, one may
determine the adsorption energy as a function of coverage using a form of the ClausiusClapeyron equation relating the
change in equilibrium pressure required to attain a given adlayer coverage to the temperature of adsorption, namely

where qst is the so-called isosteric heat of adsorption and will be numerically equal to the adsorption energy at the
coverage of interest. If isotherms are available at two or more slightly different temperatures, qst may be calculated from

where p2 and p1 are the equilibrium pressures at the temperatures T2 and T1, respectively, for a constant coverage.
Measurements of qst for a range of coverages will show the effects of surface heterogeneity as a decrease in qst with
increasing coverage. Adsorbateadsorbate interactions can lead to either an increase or a decrease of qst with coverage,
depending on whether the interaction is attractive or repulsive. In most cases of physical adsorption, the interaction will
be attractive (e.g., van der Waals). In the case of chemisorption, the nearest-neighbor adsorbateadsorbate interaction
may be repulsive due to polarization of the adsorbed species in its interaction with the surface.

< previous page page_567 next page >


< previous page page_568 next page >
Page 568

10.1.4
Observed Behavior

Over a period of many years, adsorption isotherms have been obtained for a wide variety of systems over a wide range
of temperatures. Representative examples, chosen to show the correspondence between the theoretical equations
developed above and experimental measurements, are given below.

Consider first physical adsorption isotherms. At low pressures, on energetically uniform surfaces, the simple Henry's
law behavior may be observed, as in the case of the adsorption of CO2 on wood charcoal shown in Fig. 10.15 [6]. Note
that even at the relatively low coverages involved here, the isotherm data show deviations from Henry's law at the lower
temperatures measured. More typical physical adsorption isotherms are shown in Fig. 10.16, taken from the work of
Ross and Olivier [7]. These figures show isotherms for argon on two different samples of carbon black, having different
degrees of surface heterogeneity. P33 is a carbon black graphitized by heating to the temperature shown in parentheses.
The isotherms show effects due to both adsorbateadsorbate interactions and surface heterogeneity. Isosteric heat of
adsorption curves derived from these isotherms are shown in Fig. 10.17. The initial decrease in the heat of adsorption
with increasing coverage is due to surface heterogeneity (the 2700°C sample being much more homogeneous), while
the increase in heat at higher coverages reflects adsorbateadsorbate interactions. Similar effects are seen in Fig. 10.18
for argon adsorption on Linde molecular sieve 13X, a material commonly used in sorption roughing pumps [8].

Physical adsorption isotherms taken at low enough temperatures on very energetically uniform surfaces often show the
first order phase changes in the adsorbed layer predicted by the two-dimensional gas treatment described previously. A
typical example showing this behavior is shown in Fig. 10.19, for the case of CFCl3 adsorbed

Fig. 10.15
Isotherms for the sorption at low pressures of
carbon dioxide by wood charcoal [6].

< previous page page_568 next page >


< previous page page_569 next page >
Page 569

Fig. 10.16
Comparison of experimental adsorption isotherms (individual points) with the
theoretical description of a mobile adsorbed film (solid line). (a) Argon
adsorbed by P33 (2700°C) at 77.5 K; (b) argon adsorbed by P33 (1000°C)
at 77.5 K [7].

on a very energetically uniform graphite surface [9]. Note that the extent of the discontinuous rise in coverage with
pressure decreases in extent with increasing temperature and would disappear at Tc2, the two-dimensional critical
temperature.
Currently, the bulk of the physical adsorption isotherm measurements being made are in connection with the
determination of surface area using the BET technique discussed above.

The literature on chemisorption isotherms is much more sparse than that for physical adsorption. The reason for this is
that in order to attain equilibrium between adsorbed and gas phases, the molecular impingement rate must be in balance
with the

< previous page page_569 next page >


< previous page page_570 next page >
Page 570

Fig. 10.17
Isosteric heat curves as a function of coverage for argon
adsorbed by two thermally conditioned carbon blacks [7].

Fig. 10.18
Comparison of experimental adsorption isotherm (individual points) with
the theoretical description of a mobile adsorbed film (solid line) for argon
on Linde molecular sieve 13X at 77.5 K [8].
desorption rate, as set by the surface lifetime and equilibrium adlayer coverage. In most chemisorption systems, the
surface lifetime, except at very high temperatures, is so long that the equilibrium pressure in the submonolayer range
would be below the capability of ultrahigh-vacuum systems. As a result, most of the available data on chemisorption
systems are in the form of uptake curves, showing the amount adsorbed as a function of gas exposure, or in the form of
desorption kinetic measurements. Both of these types of measurement will be discussed in Section 10.1.5, "Adsorption
kinetics."

In a few cases, however, in systems showing relatively weak chemisorption, isotherm data have been obtained. One
such example is the adsorption of oxygen on

< previous page page_570 next page >


< previous page page_571 next page >
Page 571

Fig. 10.19
Comparison of experimental adsorption isotherms (individual points)
with the theoretical description of a mobile adsorbed layer
(solid lines) for CFCI3 adsorbed on P33 (2700°C) [9].

silver. This system has been studied by Buttner et al. [10], using the Gibbs adsorption isotherm,

where γ is the surface tension and k is Boltzmann's constant. Results of this measurement are shown in Fig. 10.20.

Chemisorption isotherms have also been measured by conventional techniques in a few systems, primarily for diatomic
gases on refractory metal filaments. Data obtained by Hickmott [11] for H2 on tungsten are shown in Fig. 10.21,
covering the temperature range from 77 K to 373 K. These results were shown to fit the Temkin isotherm, characterized
by a coverage-dependent heat of adsorption. This apparent temperature dependence may be related to the fact that the
measurements were made on a surface showing a range of crystalline planes, each of which may have had a different
interaction energy with hydrogen. Measurements of N2 adsorption on tungsten, in the temperature range from 1330 K to
1540 K by Kisliuk [12], are shown in Fig. 10.22. These data are well fit by the dissociative Langmuir isotherm equation.

Because of the difficulty in measuring adsorption isotherms for systems in which strong chemisorption occurs, much of
the available equilibrium data on chemisorption systems is in the form of qualitative measurements of which gas surface
combinations lead to the formation of chemisorbed monolayers and which don't. Because the chemisorption process
leads to the formation of strong chemical bonds, the process is quite specific, permitting classification of systems such
as those shown in

< previous page page_571 next page >


< previous page page_572 next page >
Page 572

Fig. 10.20
Variation of the surface tension of silver with
oxygen partial pressure.
Reprinted with permission from J. Phys. Chem. 56, 657
(1952) [Ref. 10]. Copyright 1952 American
Chemical Society.

Table 10.2 [13]. This table presents a classification of a large number of metals in terms of their tendency to chemisorb
a group of common gases. Note that the transition metals, as a group, tend to be very active in chemisorption, while the
noble metals are somewhat less so, with the non-transition metals and the semiconductors being much less active. The
selectivity in chemisorptive properties has been used to selectively purify less active gases of more active impurities.

10.1.5
Adsorption Kinetics

For the case of physical adsorption, the energy associated with the minimum in the potential well will be comparable to
the heat of condensation of the adsorbate, typically between 1.5 and 2.0 times ∆Hv. For this case, the desorption
frequency will be given by

Because there is no barrier to adsorption, the accommodation coefficient into the physically adsorbed state is generally
unity, except for light adsorbates or very low surface temperatures.
The overall from of the expressions for Ra and Rd for this case will depend on the effect that the presence of one
adsorbed molecule has on the adsorption or desorption of another molecule and on the energetic uniformity of the
surface. A number of models have been developed to describe the consequences of various assumptions concerning
these parameters. The simplest assumption that one can make is that molecules adsorb and desorb independently of one
another, on an energetically

< previous page page_572 next page >


< previous page page_573 next page >
Page 573

Fig. 10.21
Hydrogen isotherms on tungsten over a range
of temperatures [11].
Fig. 10.22
Nitrogen isotherms on tungsten. Here P′ is the pressure multiplied
by (Ts/Tg)1/2, where Ts is the temperature of the solid and Tg is the
temperature of the gas. The lines are drawn with a slope of 0.5 [12].

< previous page page_573 next page >


< previous page page_574 next page >
Page 574

Table 10.2. Classification of Metals and Semimetals Based on Adsorption Propertiesa


Gasesb
Group Metals O2 C2H2 C2H4 CO H2 CO2 N2
A Ca, Sr, Ba, Ti Zr, Hf, V, Nb Ta, Cr, Mo, W Fec, Red A A A A A A A
B1 Ni, Cod A A A A A A NA
B2 Rh, Pd, Pt, Ird A A A A A NA NA
C Al, Mn, Cu, Aue A A A A NA NA NA
D K A A NA NA NA NA NA
E Mg, Agc, Zn, Cd In, Si, Ge, Sn A NA NA NA NA NA NA
Pb, As, Sb, Bi
F Se, Te NA NA NA NA NA NA NA
a From Bond [13].
b A, adsorption; NA, no adsorption.
c The adsorption of N2 on Fe is activated, as is the adsorption of O2 on Ag films sintered at 0°C.
d These metals probably belong to the group in which they are included, but the behavior of films is not
known.
e Au does not adsorb O2.

uniform surface, with no limit on the amount adsorbed. These assumptions lead to the Henry's law model for the
adsorption process described above. This model is an idealization; but it should, in principle, be approached by all
physically adsorbed systems at low pressure and high temperature (i.e., as na→ 0).

The net rate of adsorption or desorption in this case is given by

Integration of this relation leads to

for the case of na < neq, and to

for the case of na > neq. Equation (10.71) would describe the behavior expected when a clean surface is exposed to the
adsorbate, as would be the case in a sorption or getter pump. Equation (10.72) describes the expected behavior when a
system is pumped

< previous page page_574 next page >


< previous page page_575 next page >
Page 575

from high pressure to a lower pressure. If the partial pressures of adsorbable gases in the system are rapidly reduced to
low values, as is often the case in pumping a vacuum system, then the adlayer population will decrease exponentially
with time, with the rate being proportional to vd.

The behavior of real systems is more complicated than the Henry's law model suggests. In general, even in physical
adsorption systems, the adsorption rate will depend on whether the incoming molecule strikes the bare surface or
whether it strikes an area already covered by adsorbate. The desorption rate will be sensitive to any factor that changes
∆Hd, such as energetic interactions between adsorbed species, the position of the incoming atom relative to the surface
(i.e., first layer, second layer, etc.), and the presence of sites of different ∆Hd on the surface. Complications of this sort
are difficult to treat analytically.

10.1.6
Chemisorption Kinetics

The kinetics of the chemisorption process, in which strong chemical bonds are formed between the adsorbate and the
substrate and which may involve dissociation of the adsorbed molecules, are considerably more complicated than the
physical adsorption process considered above. In such cases, the shape of the potential well will be much more
complicated than that shown in Fig. 10.2, leading to more complicated expressions for the desorption frequency. These
expressions in general show a dependence on na and involve multiple adsorbed states, some of which may represent
dissociation of the adsorbed species. This process may again be considered in terms of potential energy diagrams for the
adsorbate substrate system.

The form of the potential energy curves in this case will depend on whether the incoming molecule is adsorbed intact,
without the breaking of any intramolecular bonds, or whether the molecule must dissociate in order for its component
atoms to reach the chemisorbed state. The simpler case of nondissociative adsorption is shown in Fig. 10.23. For the
particular case of a homonuclear diatomic molecule, the case of

Fig. 10.23
One-dimensional potential energy curve for the
nondissociative chemisorption of a diatomic
molecule showing both a weakly physisorbed state
and a more strongly adsorbed chemisorbed state.
< previous page page_575 next page >
< previous page page_576 next page >
Page 576

dissociative adsorption is shown in Fig. 10.24. In the first of these cases, the observed energy of desorption will be
approximately equal to the well depth relative to the zero of potential energy, ∆Ea in the diagram. In the case of
dissociative adsorption, one must consider both the possibility of desorption as individual atoms, for which the
desorption energy is approximately (∆E)a per atom, and the possibility of recombination to reform the diatomic
molecule and subsequent desorption of this species, for which the desorption energy is approximately 2(∆E)m per
molecule. These two desorption energies are related by

where ∆ED is the dissociation energy of the diatomic molecule. (Recall that, as they have been defined, adsorption
energies are inherently negative; dissociation energies are inherently positive.)

The case of heteronuclear diatomic molecules that adsorb dissociatively is even more complex. Figure 10.25 is the
diagram appropriate to CO adsorption on a metal surface, in the case where the adsorption is nonactivated. This case
again shows a potential well for the molecularly adsorbed species; but here one must show a single potential well for the
combination (Ca + Oa), because the potential wells for the adsorbed carbon and oxygen atoms will not necessarily
coincide. Moreover, one must use as a reference for the desorption of (C + O) atoms the dissociation energy of CO.
Additional possible outcomes for the desorption process also arise in this case, because it is possible that O2 or C2 or
CO2 species as well as CO could desorb from a layer containing adsorbed carbon and oxygen atoms. This case thus
takes one over the boundary between simple adsorption processes and surface chemical reactions, which will be covered
later in this chapter.

A final complication that may arise in the case of dissociative chemisorption is one in which the crossover between the
curves for the physisorbed and chemisorbed states

Fig. 10.24
One-dimensional potential energy curves for the
dissociative adsorption of a homonuclear diatomic
molecule, showing both a weakly physisorbed
state and a dissociatively chemisorbed state.
< previous page page_576 next page >
< previous page page_577 next page >
Page 577

Fig. 10.25
One-dimensional potential energy curves for the
dissociative chemisorption of a heteronuclear diatomic
molecule, showing both a weakly physisorbed
state and a dissociatively chemisorbed state.

Fig. 10.26
One-dimensional potential energy curves for the
dissociative adsorption of a homonuclear diatomic
molecule for the case of a finite activation
energy barrier to chemisorption.
occurs at an energy greater than that of the zero of energy, as shown in Fig. 10.26. In this case, the chemisorption
process will be activated, with a finite activation energy being required to surmount the energy barrier leading to the
chemisorbed state. The kinetics of the adsorption process may show a complicated temperature dependence in this case,
and the adsorption rate may depend on the kinetic energy of the incident gas molecules. The rate equations for this
process will be considered in the section on activated adsorption (Section 10.2.3).

These more complicated adsorption energetics lead to more complicated adsorptiondesorption kinetic behavior as well.
The simplest case to consider is that of the

< previous page page_577 next page >


< previous page page_578 next page >
Page 578

nondissociative adsorption of a molecule into a strongly bonded configuration followed by the desorption of the intact
molecule. Typical systems showing this behavior include CO on nickel or platinum. The potential energy diagram for
this case was shown in Fig. 10.23. The shallow potential well for physical adsorption of the molecule discussed above is
still present; but in this case, there is an additional well, with a much larger ∆Ea (typically in the 25-kcal/mol range) and
a considerably shorter re. In the simplest case, there is no significant barrier to the capture of the adsorbing species into
the shallow potential well, so that a constant fraction, α, will be accommodated into this state at a rate governed by the
impingement rate, I. In the deeper potential well state, however, since strong chemical bonds are formed between the
adsorbate and one or more surface atoms, molecules can only be accommodated onto vacant sites.

The fact that we have two possible adsorbed states in this case, the physically adsorbed state and the chemisorbed state,
leads to a potential complication in the adsorption and desorption rates. Referring to Fig. 10.23, if the energy difference
between the bottom of the physical adsorption well and the crossover point between the two states is ≥ kT, then it is
possible that the incoming molecule may be trapped in the physical adsorption state. Once in this state it may either
desorb, with a rate constant kp, or pass into the chemisorbed state with a rate constant ka. Once in the chemisorbed state,
it may return to the physical state with a rate constant kd. The overall adsorption and desorption rates in this case may
be obtained by considering the system in terms of a generalized chemical reaction scheme

Formulation of the overall rate equations for the adsorption and desorption process using this scheme is complex and
generally requires that some assumptions be made in order to obtain a specific result. For example, in the present case,
if one assumes that chemisorption occurs on a fixed number of sites, that there are no lateral interactions among
adsorbed species, that there is a maximum of one chemisorbed species per site, that the weakly adsorbed molecular or
''precursor" species behaves the same over both filled and empty chemisorption sites, and that the surface lifetime in the
molecular state is short (so that the coverage in this state is always small and the population in this state will always be
at its steady-state value), then one can write for the overall rate of the desorption reaction that

where θ is the fractional coverage in the chemisorbed state. The rate of adsorption can be written

< previous page page_578 next page >


< previous page page_579 next page >
Page 579

where n0 is the number of chemisorption sites per unit area as before and α is the accommodation coefficient for
trapping in the weakly bound precursor state. When adsorptive equilibrium has been reached, Ra and Rd will be equal,
so that

This may be solved to yield the equilibrium coverage in the chemisorbed state as

which is simply the Langmuir isotherm developed earlier.

The net rates of uptake or emission in this case will depend on the relative values of the rate constants, ki. These in turn
will depend on the various activation energy terms, if we assume that the form of the ki is

For the case where is small compared to kT, the presence of the weakly bound precursor state will not affect the
adsorption and desorption processes, and we will have

for the case of desorption into a vacuum and

for the case of adsorption into a state from which the desorption rate is negligible. In this case, will be
numerically equal to ∆Ea. In either of these cases, the rate of change of surface coverage with time will be exponential.

The case of the dissociative chemisorption of a homonuclear diatomic molecule may be treated similarly, using the
chemical reaction sequence

using the same notation as in the previous case of nondissociative adsorption. In this case, making the same assumptions
as before, the expressions for Rd and Ra become
< previous page page_579 next page >
< previous page page_580 next page >
Page 580

In this case, equating Ra and Rd to find the equilibrium coverage yields

which was previously shown to be the "dissociative Langmuir isotherm."

The rates of adsorption and desorption are considerably more complicated in this case than in previous cases. However,
for cases in which is small compared to kT the rate expressions reduce to

for the case in which the adsorption rate is negligible and

for the case in which desorption may be neglected.

Alternatively, the effects of the precursor state may be considered explicitly for both the adsorption and desorption
processes. This is of interest for a number of reasons. First, for many chemisorption systems, the binding energy in the
chemisorbed state is so high that θeq = 1 at ambient temperature, even for very low gas-phase pressures. Consequently,
measurement of the equilibrium coverage does not provide any information on the details of the adsorption process. In
this case, kinetic measurements of the rate of approach to equilibrium offer the only means of determining the actual
reaction sequence and reaction order and, in turn, the values of the various rate constants and their associated energy
terms. As a practical matter, the major importance of chemisorption is in relation to the rate of surface processes such as
crystal growth, corrosion, catalysis, and the operation of getter pumps such as the titanium sublimation pump or bulk
getters. The rate of chemisorption is in many cases the rate-controlling step for the overall process.

Consider as an example the adsorption rate for the case of the dissociative adsorption of a homonuclear diatomic
molecule, as discussed above. In order to describe the rate of adsorption in such a process, one generally talks in terms
of the sticking coefficient for chemisorption, defined as

that is, the unit collision adsorption probability. Clearly, the upper limit on S is unity. In practice, S is often much less
than unity and in general will change as adlayer coverage changes. For the case of the dissociative adsorption of a
diatomic molecule,

< previous page page_580 next page >


< previous page page_581 next page >
Page 581

it has already been shown that

One may consider the form of this equation in terms of the relative values of ka and kp. If as shown in Fig.
10.24, the surface lifetime in the molecular state will be long compared to the time required for the dissociation process,
leading to

and the sticking coefficient will be independent of coverage right up to the point where θ→ 1.

It is also useful to look at the effect of various ratios of the rate constants on the overall rate of the desorption process.
Consider the case in which For this case

and the desorption process will follow second-order kinetics. Alternatively, if we obtain

which would show a significant departure from second-order kinetics, especially at large values of θ. Further examples
of this type will be found later in this chapter when surface reaction kinetics are considered in detail.

Consider finally the kinetic behavior observed in systems in which the crossover between the molecularly physisorbed
and dissociatively chemisorbed states occurs at an energy greater than zero, as shown in Fig. 10.26. In such cases, a
finite barrier is present for both the adsorption and desorption processes. This not only influences the temperature
dependence of the adsorption rate but also gives rise to two new phenomena, namely, the desorption of molecules
having excess kinetic energy and a dependence of the adsorption rate on the kinetic energy of the incident molecules.

Consider first the dependence of the adsorption rate on substrate temperature. In terms of the precursor mechanism
discussed previously, since the ratio of the rate constant kp to ka, which is

will decrease with increasing temperature, leading to an increasing chemisorption probability. The alternative case,
leads to a decreasing chemisorption probability with increasing temperature.
< previous page page_581 next page >
< previous page page_582 next page >
Page 582

In the case of activated adsorption, it is also possible to have direct desorption from the chemisorbed state, without a
finite residence time for the desorbing molecule in the precursor state. This leads to the desorption of molecules having
an excess energy equivalent to the barrier height shown in Fig. 10.26. This energy may be carried off either as
translational kinetic energy or as excess energy in the internal modes of the molecule. Excess translational kinetic
energy can be observed either by a time of flight measurement of the desorbing molecules or as a departure from the
expected cosine spatial distribution of the desorbed species.

Direct adsorption from the gas phase is also possible in the case under discussion, and measurement of the adsorption
probability as a function of the kinetic energy of the incident molecules provides a means of measuring the height of the
barrier to chemisorption. If one considers the behavior of an impinging molecule in terms of the concept of a classical
turning point, defined by the distance from the surface at which the kinetic energy of the impinging molecule is just
balanced by the repulsive part of the adsorptive potential well, one sees that an impinging molecule having an incident
kinetic energy greater than the barrier height will have a classical turning point beyond the location of the barrier and
can reach the chemisorbed state directly, in a single collision, without trapping in the precursor state. This behavior has
been observed in a number of systems by using seeded free jet expansions to produce fluxes of adsorbate molecules of
known, high kinetic energies and measuring the chemisorption probability as a function of this energy.

10.1.7
Kinetic Measurements

The kinetics of adsorption and desorption processes have been studied in detail for many systems, using a wide variety
of techniques. The parameter τa has been measured directly in a number of systems using molecular beam scattering
techniques, using a system similar to that shown schematically in Fig. 10.27 [14]. In these measurements, the adsorbate
surface, usually a single crystal, is mounted on the axis of the system, and exposed to pulses of the adsorbing species
generated by the modulated molecular beam source. A mass spectrometer, mounted in line of sight to the sample
surface, measures the flux of gas desorbed from the surface. The surface lifetime may be determined either from a plot
of desorbed flux versus time, using Eq. (10.72), or by measuring the phase and amplitude of the desorbed flux relative
to the modulation wave form the incident molecular beam [15]. This technique was first reported by Scheer and Fine
[16] and has been used extensively by Hudson et al. [14, 1721], Wharton et al. [22], and Campbell et al. [23], among
others. Data obtained in a range of systems are summarized in Table 10.3, reported as τ0 and ∆Ed for the systems
studied. Additional values of τa, determined from capillary transit time measurements, are summarized later in this
chapter in the section on capillarity effects (Section 10.1.8).

The sticking coefficient for chemisorption has been measured in many systems at temperatures low enough that the
desorption rate in negligiblethat is, under conditions in which the adsorption rate should be described by equations such
as Eq. (10.89) or Eq. (10.90). A summary of sticking coefficient measurements for nitrogen chemisorption on various
tungsten surfaces is shown in Fig. 10.28, taken from the work of King [24]. It can be seen that both the initial value of
the sticking coefficient and the variation of the sticking coefficient with coverage differ greatly from one

< previous page page_582 next page >


< previous page page_583 next page >
Page 583

Fig. 10.27
Schematic top view of an ultrahigh-vacuum molecular beam scattering
system used for direct measurement of the surface lifetime, τa. Reprinted
from D. A. Hoffman and J. B. Hudson, "The Adsorption and Decomposition
of N2O on Nickel (100)," Surf. Sci. 180, [Ref. 14] with kind permission of
Elsevier Science NL, Sara Burgerhartstraat 25, 1055 KV,
Amsterdam, the Netherlands.

Table 10.3. Experimental Measurements of Surface Lifetime by Mass Spectrometric Molecular Beam Techniques
Adsorbate Substrate τ0 (s) ∆Ed Temperatures (K) References
(kcal/mol)
Cd W(110)
9 × 1011 41.1 7501000 17
Ag W(110)
3 × 1013 66 9501300 18
C2H4 C on Ni(110)
4 × 1010 11.9 300380 19
N2O Ni(110)
1012 6.2 120160 14, 20
N2O O on Ni(110)
1011 4.7 120160 14
CO C+O on Fe(110)
3 × 1010 13.7 > 700 21
CO Pt(111)
4 × 1014 31.1 419505 22
O2 Pt(111)
42a 51 686780 23
a Second-order recombinative desorption. Units of τ0 are s/cm2.
crystal face to another. Note that all faces have sticking coefficient values less than unity even at zero coverage, indicating either that α is
less than unity or that there is a finite probability of desorption from a precursor state rather than passage into the chemisorbed state. Some
of the faces show a range of nearly constant sticking coefficient, suggesting the importance of precursor states in the adsorption process.

< previous page page_583 next page >


< previous page page_584 next page >
Page 584

Fig. 10.28
Sticking coefficient for the chemisorption of nitrogen on various
tungsten single crystal planes. Reprinted from CRC Crit. Rev. Solid
State Mater. Sci. 7, 167 (1977) [Ref. 24]. Copyright 1977
CRC Press, Boca Raton, Florida.

More detailed measurements, over a range of temperatures, would be necessary in order to fully characterize the
adsorption process in this system.

An example of desorption over a barrier, to produce desorbed species having excess kinetic energy, is shown in Fig.
10.29 for the case of D2 desorption from Fe (110) [25]. The experimental results show an approximate cos2θ
distribution, indicative of excess kinetic energy normal to the surface. Desorption of molecules in thermal equilibrium
with the surface would be expected to show a cos θ dependence. Results of adsorption measurements in a system
showing activated adsorption are presented in Fig. 10.30, for the case of O2 chemisorption on W (110) [26]. In this
study, the sticking probability, S0, increased rapidly as the translational kinetic energy of the impinging O2 molecules
was increased. The rate of increase implies a barrier height on the order of 0.20.3 eV (47 kcal/mol). The fact that the
sticking coefficient scales as the normal component of the incident kinetic energy implies that the barrier is essentially
one-dimensional.

10.1.8
Capillarity Effects

Up to this point, it has been implicitly assumed that the surfaces on which adsorption was taking place were flat, or
nearly flat. The structures of many practical adsorbent materials contain networks of small-diameter pores. In some
cases, such as the so-called "activated carbons," this porosity arises from the method of manufacture. In others, such as
the synthetic zeolites, the porosity is an inherent feature of the crystal structure of the material. In either case, the
presence of small-diameter pores can have
< previous page page_584 next page >
< previous page page_585 next page >
Page 585

Fig. 10.29
Desorption flux as a function of detector angle relative to the
surface normal for angle resolved thermal desorption of
deuterium from Fe(110). Reprinted from E. A. Kurz and J. B.
Hudson, "The Adsorption of H2 and D2 on Fe(110) II," Surf.
Sci. 195, [Ref. 25] 34 (1988), with kind permission of Elsevier
Science NL, Sara Burgerhartstraat 25, 1055 KV, Amsterdam,
the Netherlands.

appreciable effects on both the equilibrium adsorption behavior and the rate of uptake into the adsorbed phase.

Consider first the effect of pores on adsorptive equilibrium. This can be explained by a purely thermodynamic argument
based on the effect of surface curvature on the equilibrium vapor pressure of the adsorbing species. The physical
situation in this case is shown in Fig. 10.31, which shows a pore of circular cross section containing a liquid which wets
the surface of the pore. The relation between the equilibrium vapor pressure and the geometry of the system for a case
like this is given by the Kelvin equation,
where pβ is the vapor pressure over the liquid in the pore, p0 is the vapor pressure over the flat surface of the liquid, γ is
the surface tension of the liquid, and Vα is the molar volume of the liquid. As an example of the magnitude of this
effect, consider the case of water adsorbed into a 10-nm-diameter pore at 300 K. A calculation based on Eq. (10.94)
leads to a value of pβ/p0 of 0.313, implying that the pores would be filled by

< previous page page_585 next page >


< previous page page_586 next page >
Page 586

Fig. 10.30
Sticking coefficient for chemisorption of oxygen on
tungsten, as a function of incident oxygen molecule
kinetic energy [26].

Fig. 10.31
Schematic representation of the condensation of
an adsorbed gas in a small-diameter pore.

condensed water if they were exposed to the atmosphere at a relative humidity exceeding 32%. This illustrates the
source of the large amounts of water that are released during pumpdown of systems containing porous materials
following exposure to the atmosphere.
The kinetics of the adsorption process will also be significantly changed from the values determined previously if the
adsorbate must penetrate into a porous material to achieve adsorbed equilibrium. This kinetic effect can be explained in
terms of the diagram shown in Fig. 10.32, which shows the path of a gas molecule through a tube under conditions such
that the mean free path in the gas, λ, is large compared to the tube diameter, d. This process can be one of the significant
factors controlling the net

< previous page page_586 next page >


< previous page page_587 next page >
Page 587

Fig. 10.32
Schematic representation of the path of a gas molecule
through a capillary having a diameter that is short compared
to the mean free path in the gas phase.

rate of flow of gases through small-diameter tubes at pressures low enough that few gasgas collisions occur. This is
relevant both to the time required to saturate a sorption roughing pump and to the time required to pump down a
chamber through a capillary tube.

Consider the average time required for a molecule to traverse a capillary of length l and diameter d as shown in Fig.
10.32. For the case of λ > d, will be the product of the number of collisions with the wall required to traverse the
capillary, multiplied by the sum of the average time of flight per jump and the mean surface lifetime per collision. It can
be shown from kinetic theory considerations that

where the term outside the parentheses is the number of collisions required to traverse the capillary, the first term in the
parentheses is the average time of flight, and the second term in the parentheses is the surface lifetime. This expression
may be rearranged to separate the effects of the two processes, yielding

Note that the first term depends on T1/2, due to the presence of , and that the second is exponentially dependent on T,
due to the temperature dependence of τa, as shown in Eq. (10.7). If the molecule moves over the surface due to surface
diffusion while it is adsorbed, this will contribute a third term to Eq. (10.96). For the present we will ignore this effect.

The values of calculated for various gases for various-sized tubes are summarized in Table 10.4, taken from DeBoer
[27]. Each entry in this table shows separately the effects of the two terms in Eq. (10.96). The cases shown represent
values for the surface lifetime that would be representative of H2 (1012), N2 (1010), and an organic molecule (104)
adsorbed at ambient temperature.

The general trend shown here is that for heavier, and consequently more strongly adsorbed, molecules and for large
values of 1/d, the adsorption term becomes more important. For the organic molecule it is the dominant term for all of
the cases shown. Note in passing that this is how a gas chromatograph works: The more strongly adsorbed species in a
mixture take longer to pass through the tube of the chromatograph, resulting in a separation of the components of a
mixture. Note too that in
< previous page page_587 next page >
< previous page page_588 next page >
Page 588

vacuum systems, where pumping speed and the conductance of tubes are the factors of importance, the effects shown in Table 10.4 are
transient. At steady state the surface concentration will become larger at the high-pressure end of the tube relative to the low-pressure end,
with the net effect of canceling out the adsorption-dependent term in the flow rate.

This technique has been used in practice to determine surface lifetimes in a number of systems. A summary of available data, taken from
Redhead et al. [28], is shown in Table 10.5 [2931].

The implications of these results for the operation of sorption roughing pumps are obvious. Saturation of the external surfaces of
adsorbents such as the zeolites typically used in sorption roughing pumps will occur rapidly at liquid nitrogen temperature. The bulk of
the time required to reach adsorptive equilibrium will be the time required for adsorbing gas to penetrate the pore structure of the zeolite.
Based on the figures shown in Table 10.4, the required time for a typical adsorbent particle will be on the order of 5000 s.

Table 10.4. Delay Times for transmission of pressure pulses through a capillary of length / and diameter d. Mean stay
time for adsorption is τa and mean molecular velocity is
τa = 1012 s τa = 1010 s τa = 104 s

= 15 × 104 cm·s1 = 5 × 104 cm·s1 = 104 cm·s1


l = 10 cm
d = 101 cm 3 × 103 + 5 × 109 102 + 5 × 107 5 × 102 + 5 × 101
l = 10 cm
d = 104 cm 3 × 104 + 5 × 107 103 + 5 × 105 5 × 103 + 50
l = 102 cm
d = 106 cm 3 × 104 + 5 × 105 103 + 5 × 103 5 × 103 + 5000
l = 103 cm
d = 107 cm 3 × 105 + 2 × 105 104 + 5 × 103 5 × 104 + 5000
a Adapted from DeBoer [27].

Table 10.5. Mean Stay Time for Adsorption of Molecules on Surfaces Measured by Time Delay Methoda
Gas Surface Range of Measurement τa (s) References
Ar Glass 7890 K 29
τa = 1.7 × 1014 exp(3800/RT)
He Glass 13.820.4 K 30
τa1 = 109 exp(299/RT)
τa2 = 109 exp(530/RT)
(two states)
NH3 Analcite 278308 K 31
τa = 3.1 × 1011 exp(4300/RT)
SO2 Analcite 278308 K 31
τa = 6.6 × 1011 exp(3010/RT)
a Adapted from Redhead et al. [28].

< previous page page_588 next page >


< previous page page_589 next page >
Page 589

10.2
Absorption

Consider next the complications that arise when a species is not only present in the gas phase and as an adsorbed layer,
but is also dissolved in the bulk of a material in the system or in the system walls. In this case, one must consider not
only the equilibrium and fluxes of gas molecules between gas and surface, but also equilibrium and fluxes associated
with flow in an out of the bulk of the solid. In most cases, the gases that will dissolve readily in the materials found in
vacuum systems are simple gas molecules, and in most cases they dissolve as single atoms. Thus the typical process that
we deal with here, say for the flow of a gas into and out of a solid, can be represented as

where A(sol) represents A atoms dissolved in the bulk of a solid and we will have n = 1 for gases that dissolve
nondissociatively and n = 2 for dissociative dissolution. The potential energy diagram for this process is shown in Fig.
10.33. The new features that we see here are, first, that the energy does not rise indefinitely as the molecule approaches
the surface, but goes through a maximum, then decreases to a minimum representing a stable site for the dissolved
species, and, second, that there will be a succession of such sites extending into the bulk, separated by maxima in the
potential energy. Note that the energy associated with the minimum in the potential well representing the absorbed
species may be higher or lower than the minimum associated with the adsorbed species and may be higher or lower than
the energy of the species in the gas phase. The energy difference between the minimum in the absorbed species
potential well and the gas-phase species is the heat of solution, ∆Hs. In the event that the minimum for the absorbed
species is lower than the energy for the gas-phase species, solution will be an exothermic process, and the solubility will
decrease with increasing temperature. In the alternative case of a positive heat of solution, as shown in Fig. 10.33,
solubility will increase with increasing temperature.

Fig. 10.33
One-dimensional potential energy diagram for the transport of
gas through a solid by adsorption followed by dissolution and
bulk diffusion.

< previous page page_589 next page >


< previous page page_590 next page >
Page 590

10.2.1
Equilibrium Solubility

Consider first the case in which the solid and gas phases are in equilibrium. Based on purely thermodynamic
considerations, one can represent the process shown in Eq. (10.97) as a generalized chemical reaction and write, for the
equilibrium constant

where aA(sol) is the activity of A in solution and aA(gas) is the activity in the gas phase. If one makes the substitutions

the equilibrium concentration in the solidthat is, the solubilitymay be written as

where C0 is a constant for any particular gassolid combination and ∆Hs is the heat of solution defined by the potential
energy diagram of Fig. 10.33 above (that is, the heat released in the sorption process, per gram mole of gas absorbed.) S
is the so-called solubility constant. Solubilities may be expressed in terms of the concentration in the solid, as mole
fraction of solute, as grams per unit volume, or as the volume of gas (at STP) taken up by a given mass of solid, or it
may be expressed in terms of the mass of gas taken up by a given mass of solid.

A similar result may be obtained by a kinetic argument that takes specific account of the role of an adsorbed layer in the
sorption process. This argument will be presented later, in a discussion of the combined effects of permeation and
desorption on the rate of outgassing.

10.2.2
Diffusion Rates

The rate of transport in the bulk can be characterized by a jump frequency, the rate at which absorbed atoms move from
one equilibrium site to another. Using absolute reaction rate theory, this jump frequency can be represented as

The one-dimensional flux arising from this jumping process is given, assuming that the jump direction is random, by

< previous page page_590 next page >


< previous page page_591 next page >
Page 591

which is known as Fick's first law [32]. In this equation, D is the diffusivity and is related to the jump frequency by

Fick's first law is adequate to define the diffusive flux in the case of steady-state transport, as would occur in the steady-
state permeation of gas through a solid slab, such as the wall of a vacuum system, with constant gas pressures being
maintained on both sides of the slab. In order to describe the transport process in the more common case of the diffusion
of a gas into or out of a solid, in which the concentration of the gas is changing with time at any point in the solid, one
must use a form of Fick's law which takes account of the change with time in the amount of material in any small
volume of the solid. For one-dimensional diffusion, the form of this equation is

which is known as Fick's second law. Solution of this equation requires specification of the boundary conditions
appropriate to any specific situation.

10.2.3
Kinetics of Absorption and Permeation

Absorption involves successive passage from the gas phase to the adsorbed phase and then into a near-surface bulk site,
followed by penetration into the bulk by a series of diffusive jumps. Desorption involves the reverse sequence. The
overall process of removal of gas from the bulk of a solid can thus be treated as a process of bulk diffusion, followed by
desorption. The desorption steps can be treated as described earlier in this chapter and, in many cases, will be rapid
compared to the bulk diffusion process, especially when material must be removed from deep within the bulk. For
simplicity, the absorptiondiffusion and the adsorptiondesorption processes will first be treated separately.

Consider first the kinetics of the uptake of gas into the solid by absorption followed by diffusion processes, for the case
where the details of the adsorptiondesorption sequence do not affect the overall transport rate. For this case, there will
be an equilibrium between the gas phase and the near-surface bulk phase set by

where J is the net diffusion flux from the bulk at the surface, and S is the sticking coefficient for adsorption. At
equilibrium, this flux will be proportional to the concentration at the surface of the solid, Cx=0. The impingement rate is
set by the gas-phase pressure, and the value of S will depend on the nature of the adsorption process. For the case of
nondissociative adsorption, at adsorptive equilibrium, S will be a constant; for dissociative adsorption, S will be
proportional to p1/2.

< previous page page_591 next page >


< previous page page_592 next page >
Page 592

10.2.4
Steady-State Permeation

If it is assumed that the adsorption and desorption processes are fast relative to absorption and permeation, then the rate
of gas uptake or release from a solid will be controlled by the rate of diffusion of the gas through the solid. Consider
first the case in which a region at high pressure is separated from a region of low pressure by a solid material that
permits passage of the gas from the high-pressure region to the low-pressure region by bulk diffusion. In this case, after
an initial transient, the rate of permeation will be constant. The resulting concentration gradient through the solid will be
linear, as shown in Fig. 10.34, and the permeation process can be described by Fick's first law. The steady-state
permeation rate will be given by

where d is the thickness of the solid and the concentrations on the high- and low-pressure sides will be given (assuming
that the adsorption and dissolution processes involved are rapid relative to the diffusion process) by the equilibrium
solubilities at the two pressures. For the case of gases that dissolve without dissociation, this leads to

or, taking account of the temperature dependence of the diffusivity,

The temperature dependence of the permeation process will thus depend on the algebraic sum of the activation energy
for diffusion and the heat of solution.

Fig. 10.34
Schematic representation of steady-
state permeation through
a solid of thickness d.
< previous page page_592 next page >
< previous page page_593 next page >
Page 593

For the case of diatomic gases that dissociate on solution, the corresponding expression will be

The steady-state permeation of gases through solids has been the subject of numerous studies. Those cases relevant to
contemporary vacuum technology include the permeation of gases from the atmosphere through materials used as
vacuum walls or as gasketing materials, as well as the use of materials that have a high permeability for specific gases
relative to all other gases, and may thus be used as a method of admitting pure gases to vacuum systems. In the former
category, the only case of current interest is that of the permeation of helium, and possibly other atmospheric gases,
through glass, especially quartz glass. This process has been extensively studied over a long period of time. In most of
the measurements made, Eq. (10.108) has been rewritten as

where K is the permeability, K0 and Ep are given by

and it is assumed that

Results have commonly been reported as the volume of gas at STP per second per square centimeter of surface per
millimeter of thickness, or as Qµl = micron-liters at 0°C per square centimeter of surface per millimeter thickness per
unit pressure. The permeation rate can thus be characterized by an equation of the form

The most comprehensive studies of the permeation of gases through various glass compositions are those of Norton
[33], who has reported permeation measurements in which not only the amount of gas but also its composition was
studied with the aid of a mass spectrometer. His observations on the heliumsilica system covered the range 78°C to 700°
C, and they agree fairly well with those of previous investigators.

< previous page page_593 next page >


< previous page page_594 next page >
Page 594

In addition, he studied helium permeation through a number of glasses, as well as the transport of several other gases
through silica. The results of his helium permeability measurements are shown in Fig. 10.35, and the data reported for
other gases are given in Tables 10.6 and 10.7. Data for the permeation of helium, neon, hydrogen, and nitrogen through
Vycor glass have been given by Lieby and Chen [34]. It should be noted that Alpert [35] has shown that permeation of
atmospheric helium is one of the limiting factors in the attainment of ultrahigh vacua.

Urry reported that Pyrex is permeable to hydrogen; and this result was confirmed by Taylor and Rast [36], who reported
a permeability of 1.86 × 1013 g/cm2/s/mm/atm. They also emphasized the importance of the history of the glass sample
in affecting the permeability observed. In this connection, it has been found, for example, that the permeation rates are
altered if the glass is under stress.

The other situation in which steady-state permeation is encountered is the use of selectably permeable membranes for
the purification of gases, or for the admission of pure gases into a vacuum system. This technique has been used by
Young and Whetten [37] for the purification of helium using a quartz tube assembly for the admission of the gas into a
vacuum system. With this device, they found that the only impurity remaining above one part per million was hydrogen.
Similar techniques have

Fig. 10.35
Permeability of various glasses to helium [33]. K is the permeation velocity of helium
through glass in units of cm3 gas (N.T.P.) per second per cm2 area per mm thickness
per cm Hg gas pressure difference.
< previous page page_594 next page >
< previous page page_595 next page >
Page 595

Table 10.6. Permeability Constant K through Silicaa


Gas 700°C 600°C
Helium 2.1 × 108
Hydrogen 2.1 × 109 1.25 × 109
Deuterium 1.7 × 109
Neon 4.2 × 1010 2.8 × 1010
Argon Under 1015
Oxygen Under 1015
Nitrogen Under 1015
a From Norton [33].

Table 10.7. Hydrogen Permeation Constant, Ka


Glass 665°C
No. 1720 (Corning) 4.5 × 1012
Mullite 1.8 × 1011
a From Norton [33].

been used for the admission of hydrogen into a vacuum system using a palladium or palladium alloy tube, as well as for
the admission of oxygen using a silver tube. In these two cases, the appropriate expression for the permeation rate is Eq.
(10.109), because in both of these cases the absorption process involves the dissociation of the diatomic molecule.

10.2.5
Transient Permeation

There are many situations in which one must deal with transient permeation. One such case is that in which the material
separating two regions of different pressure is initially clean and is abruptly exposed to a change in conditions, such as
an increase in pressure on one side or an increase in temperature that allows gas uptake and diffusion to begin. To treat
this transient case one must use Fick's second law, as stated in Eq. (10.104). For the case described above, which is
shown diagrammatically in Fig. 10.36, we have the boundary conditions

for 0 ≤ x ≤ d at t = 0,
C=0
for x = d at t > 0,
C = Chi
for x = 0 at t > 0.
C=0

(This last condition is satisfied if we assume that plo is maintained essentially at zero.) The solution to Fick's second
law for these boundary conditions is
< previous page page_595 next page >
< previous page page_596 next page >
Page 596

Fig. 10.36
Transient permeation curves showing
the approach of the concentration profile
in the solid to the steady-state
value of a constant dC/dx.

This yields an instantaneous outgassing rate on the low pressure side of

At very small values of Dt, dQ/dt is approximately zero. That is, it takes a finite time for the first gas to diffuse
completely across the slab. At large values of Dt, dQ/dt → DChi/d, the value determined previously for the steady-state
case. At intermediate times, dQ/dt will lie between these two extremes.

Alternatively, one may be concerned with the process of gas initially present in a solid diffusing out into the vacuum
space over a period of time, or the opposite case, the diffusion of a species present in the gas phase into the bulk of a
material exposed to that gas phase. The first case is representative of outgassing from the walls of a vacuum chamber or
from objects within the chamber. The second case represents the process of pumping by a bulk getter.

As a practical matter, the first of these processes is the principal source of gas in well-prepared ultrahigh-vacuum
systems, and it can be a significant fraction of the total gas load in a vacuum system in many other cases. Two situations
in which these processes can take place may be considered, both of which involve desorption of gas from the walls of
the vacuum system, or from objects completely within the system, into the gas phase.

The first case, which involves desorption from the vacuum wall, or desorption from a thick piece of material within the
vacuum system, can be treated as a process of diffusion from a semi-infinite slab. The second case, which involves
diffusion from a thin source, such as a heated filament, may be treated as diffusion from a finite slab.

Consider first the case of the semi-infinite slab, shown schematically in Fig. 10.37. For this case, one must again use
Fick's second law, but this time the appropriate
< previous page page_596 next page >
< previous page page_597 next page >
Page 597

Fig. 10.37
Concentration profiles for the diffusion of a dissolved gas
out of a semi-infinite slab.

boundary conditions are

for x ≥ at t = 0,
C = C0
for x = 0 at t > 0,
C=0

assuming again that the pressure in the vacuum space is essentially zero. The solution in this case is

The right-hand side of Eq. (10.118), except for the term C0, is what is known as the error function. Values of this
function have been tabulated for a wide range of values of the argument y. The instantaneous rate of outgassing for this
case may be obtained from Eq. (10.118) as
< previous page page_597 next page >
< previous page page_598 next page >
Page 598

The total amount of gas removed per unit area in time t will thus be

Note that dQ/dt increases with increasing values of D, which implies an exponential increase of Q with temperature, and
decreases as t increases. Consequently, if one wishes to thoroughly outgas the system walls it is advisable to heat the
system to the highest possible temperature to maximize D. After such a process, when the system is brought back to
room temperature, C0, and consequently dQ/dt, will be much lower than they were initially.

The case of the finite slabfor example, a thin piece of material within a vacuum system or a hot filament initially
containing a dissolved impuritymay be treated similarly. Here again, one must solve Fick's second law, this time for the
boundary conditions

for 0 ≤ x ≤ d at t = 0,
C = C0
for x = 0, x = d at t > 0.
C=0

This case is shown schematically in Fig. 10.38. The solution in this case is

This leads to an instantaneous outgassing rate of

The complication in this case relative to the semi-infinite slab is that C0 is not constant with time at some distance from
x = 0. However, at short timesthat is, times for
Fig. 10.38
Concentration profiles for the diffusion of a dissolved gas
out of a finite slab.

< previous page page_598 next page >


< previous page page_599 next page >
Page 599

which (Dt)1/2 < dEq. (10.124) reduces to the same form as that for the semi-infinite slab, because under these
conditions the concentration deep within the slab is still C0. Consequently, at short times we obtain

At longer times, dQ/dt will decrease more rapidly than this equation predicts.

One may also consider the case of uptake of gas from the vacuum system into an infinite or semi-infinite slab, as would
be the case in the operation of a bulk getter pump. In this case, the appropriate boundary conditions are

for x = 0, t ≥ 0,
C = C0
for x > 0, t = 0.
C=0

This case is shown in Fig. 10.39. Solution of Fick's second law in this case leads to

where the second term in the parantheses is again the error function defined in Eq. (10.118).

The rate of uptake of material per unit area per unit time is, in this case,
Fig. 10.39
Concentration profiles for the uptake of a gas into an initially
clean semi-infinite slab. Note that the depth of penetration for
a given concentration ratio, such as C', increases as

< previous page page_599 next page >


< previous page page_600 next page >
Page 600

leading to a total uptake in time t of

This behavior is commonly observed in the oxidation rate of metal surfaces and is described as parabolic kinetics.
Similar equations may be derived for other geometries, and additional examples are treated in Chapter 5, Part I, ''Getters
and Getter Pumps."

10.2.6
Effect of Desorption Kinetics on Permeation

To this point, it has been assumed that the kinetics of the adsorption and desorption processes that must take place at the
surfaces of materials undergoing permeation processes are rapid relative to the permeation rates involved, and thus do
not influence the overall permeation rate. In many cases, this assumption is not valid. For example, if the surface of a
material is covered with an adsorbed layer that inhibits the adsorption process that must take place prior to absorption,
then the near-surface concentration of the permeating species may not be the equilibrium solubility that was tacitly
assumed in the development of the permeation rate equations developed above. This is an especially important problem
in cases in which a gas must adsorb dissociatively prior to solution and permeation. The inherent kinetics of the
adsorption process may also effect permeation rates even on an atomically clean surface. This case has been treated by
Arbab and Hudson [38] for the case of hydrogen permeation through pure iron. These authors measured steady-state
permeation rates for hydrogen through a thin iron membrane under conditions in which surface cleanliness could be
ensured, and they found that the permeation rate showed an anomalous decrease at low temperatures. This decrease was
explained in terms of the potential energy diagram shown in Fig. 10.40, which is based on the observed energetics of the
hydrogeniron system. In this figure, the potential energy of a hydrogen atom is shown as a function of distance across
the permeation wall. The zero of energy is taken as a hydrogen molecule in the gas phase. The potential wells at each
surface represent the sites for stable chemisorption of hydrogen as hydrogen atoms. The series of minima between these
chemisorption wells represent stable sites in the bulk. The intervening maxima represent the barriers surmounted in bulk
diffusion.

As has already been shown, under steady-state conditions, Fick's first law for steady-state one-dimensional transport
through the bulk of a solid is normally stated as [39]

where J is the steady-state flux, D is the diffusion coefficient, Chi and Clo are the volume concentrations of the
diffusing species at the planes just below the entrance and exit interfaces of the permeation wall, respectively, and L is
the wall thickness.

Assuming that the concentration in the near-surface side of the membrane is in equilibrium with the gas phase, one may
write the following equation using Eq. (10.100):

< previous page page_600 next page >


< previous page page_601 next page >
Page 601

Fig. 10.40
Potential energy diagram for the permeation of hydrogen through iron.
Reprinted from M. Arbab and J. B. Hudson, "The influence of Desorption
Kinetics on Hydrogen Permeation in Iron," Appl. Surf. Sci. 29, 8 (1987) [Ref. 38],
with kind permission of Elsevier Science NL, Sara Burgerhartstraat
25, 1055 KV, Amsterdam, the Netherlands.

where the solubility, S, is a function of temperature as previously shown. The near-surface concentration on the high-
vacuum side of the sample, Clo may be assumed to be zero, leading to

(It will be shown later that surface effects can raise Clo to a finite value, approaching Chi at low enough temperatures.)

For convenience in the kinetic analysis that follows, the hydrogen concentrations may be stated in terms of ni (atoms
per cm2), on any plane parallel to the interface. That is,
where d is the interplanar spacing. The factor of ½ is necessary to account for the fact that permeation data are usually
expressed on a per H2 molecule basis.

< previous page page_601 next page >


< previous page page_602 next page >
Page 602

Mass balance expressions for the hydrogen atom fluxes onto and away from planes 2 and 3 (as defined in Fig. 10.40), at
steady state (i.e., dn2/dt = dn3/dt = 0), may be written as

where n1 = 2Sp1/2, n0 is the number of surface sites per unit area available for hydrogen adsorption, kij is the
probability for the transition of a hydrogen atom from a site on plane i to a site on plane j, and kd is the desorption
probability. The parameter (1 n3/n0) in the above equations expresses the assumption that a site on plane 3 already
occupied by an adsorbed hydrogen atom is unavailable to other hydrogen atoms attempting to arrive on that plane. A
similar term for plane 2 can be neglected if we assume

Simultaneous solution of Eqs. (10.133) and (10.134) yields

The steady-state permeation flux, J, can be expressed in terms of n3 as

with n3 obtained from the solution of Eq. (10.135). These equations contain five temperature-dependent terms, namely
S, D, k23, k32, and kd, whose activation energies are shown in Fig. 10.40.

The application of this treatment to the authors' data for hydrogen in iron is shown in Fig. 10.41. In evaluating the
theoretical equation, literature values were used for the crystallographic parameters for iron, the equilibrium solubility
[40], and the absorptiondesorption behavior of hydrogen on iron [41]. The activation energy for permeation obtained in
their study and the value for the activation energy for desorption were adjusted to obtain a best fit to the data. The
solution of Eqs. (10.135) and (10.137), using ∆Hd = 25 kcal/mol up to a saturation coverage of one monolayer, was
found to agree well with the experimental results, as shown by the solid curve in Fig. 10.41. This value is in good
agreement with the results of previous studies. Application of the analysis developed above to the data reported by
Nelson and Stein [42],

< previous page page_602 next page >


< previous page page_603 next page >
Page 603

Fig. 10.41
Arrhenius plot of hydrogen flux as a function of temperature
for hydrogen permeation through iron. The straight line
shows the result of a linear regression analysis of data for
T > 200°C. The solid line is based on the model developed
in the text. Reprinted from M. Arbab and J. B. Hudson,
"The influence of Desorption Kinetics on Hydrogen
Permeation in lron." Appl. Surf. Sci. 29, 6 (1987) [Ref. 38],
with kind permission of Elsevier Science NL, Sara
Burgerhartstraat 25, 1055 KV, Amsterdam, the Netherlands.

using the heat of adsorption assumed above, showed that the apparent activation energy for permeation at high
temperatures increased only by 0.1% over that experimentally measured. Therefore, unless much stronger surface
reactions are shown to be operative, the scatter in previous results at high temperatures may be attributed to bulk
processes or to experimental errors.

An interesting feature of the results shown in Fig. 10.41 is that the deviation from Arrhenius behaviour occurs quite
abruptly. Therefore, the value of the true activation energy for permeation may be reliably found by linear regression
analysis of the high-temperature data. At temperatures even slightly below the point of deviation, the hydrogen flux is
controlled almost entirely by the desorption kinetics; an evaluation of the activation energy for permeation at low
temperatures shows it to be equivalent to that for hydrogen desorption. Hence, if Eq. (10.131) is applied to low-
temperature permeation data, anomalously high values of the apparent activation energy for diffusion, equivalent to
∆Hd∆Hs, will be obtained. This deviation from Arrhenius behavior would occur at higher temperatures for planes
which absorb the hydrogen atom more strongly.
< previous page page_603 next page >
< previous page page_604 next page >
Page 604

Fig. 10.42
Effect of membrane thickness on the ratio of calculated apparent
diffusion, Dapp, to that extrapolated from the high-temperature data, Dtrue.
Reprinted from M. Arbab and J. B. Hudson, "The Influence of Desorption
Kinetics on Hydrogen Permeation in Iron." Appl. Surf. Sci. 29, 15
(1987) [Ref. 38], with kind permission of Elsevier Science NL, Sara
Burgerhartstraat 25, 1055 KV, Amsterdam, the Netherlands.

Figure 10.42 shows the effect of the permeation wall thickness on the value of the apparent diffusivity, as predicted by
the analysis of Arbab and Hudson. The calculated value of diffusivity approaches the value expected from extrapolation
of the high-temperature measurements as the wall thickness increases, causing the bulk transport process to become
dominant in the determination of the overall rate of permeation.

The trend in apparent diffusivity with wall thickness is similar to that reported by Wach and Miodownik [43] in their
electrochemical permeation study at 25°C. Similarly, Palczewska and Ratajczyk [44] considered the effect of wall
thickness on gas permeation between 30°C and 40°C. They reported that diffusivity was independent of membrane
thickness above 0.78 mm, while for thicknesses below 0.42 mm, decreased values of apparent diffusivity are observed.
Nelson and Stein [42] also studied the influence of wall thickness on permeation for temperatures of 240°C and higher.
By geometrical arguments they concluded that, based on that part of their study, permeation was not influenced by
surface processes. However, since the data used in their analysis describe the permeation behavior above the deviation
temperature observed in their experiments (100°C), they are outside the range in which the permeation rate is affected
by surface processes and thus do not provide a critical test of the effect of these surface processes on permeation.

Surface contaminants may also lead to the observation of incorrect values of the apparent diffusivity if their presence
results in the retardation or enhancement of the desorption process. For example, preadsorbed submonolayers of sulfur,
oxygen, and

< previous page page_604 next page >


< previous page page_605 next page >
Page 605

carbon were observed to decrease both the energy and the frequency factor for desorption of hydrogen from Fe(100)
surfaces, while preadsorbed potassium seemed to increase the value of the former [45]. Direct evidence for this effect
was observed for the permeation of hydrogen in palladium in an electrochemical permeation study [46]. In that
investigation it was noticed that sulfur deposition on the membrane surface resulted in a reduced permeation flux and
consequently low values for the apparent diffusivity.

The treatment developed above may also be applied to other hydrogenmetal systems. Among the transition metals, the
interaction of iron, nickel, and palladium with hydrogen has received considerably more attention than others.
Therefore, a comparison of the results for iron may be most appropriate with those for the latter two metals. The
diffusivity values obtained by various permeation methods for these metals do not indicate the anomalies observed for
iron, even at room temperature or below [47].

Hydrogen permeation in nickel was examined using the model developed above. Because of the consistency in the
values of diffusivity reported by various authors [47], it is sufficient to consider the experimental results of only one
studyfor example, the work of Ebisuzaki et al. [48] for the permeation parameters. The required parameters relevant to
the desorption of hydrogen from various nickel surfaces were taken from the work by Christmann et al. [49]. The
analysis showed excellent agreement with the experimental data of Ebisuzaki et al. [48]. A deviation from Arrhenius
behavior was predicted only for temperatures lower than 20°C.

Similar to nickel, there is good consistency among various experimental values of hydrogen diffusivity in palladium
[47]. There is, however, substantial evidence that hydrogen adsorption on this metal is more complex than that shown
by the potential energy diagram of Fig. 10.40. Adsorbed hydrogen atoms were observed to readily occupy subsurface
sites on Pd(111) surfaces [50, 51], while desorption from Pd(100) surfaces was observed to follow "quasi-first-order"
kinetics beyond very low coverages [52], in contrast with iron [41, 53]. Therefore, the model described here cannot give
an adequate description of the effect of the desorption kinetics on hydrogen permeation in palladium. In spite of the
above evidence, Engel and Kuipers [54] found that a potential energy diagram similar to that shown herewith
parameters chosen to yield a best fit to the experimental dataadequately described their results on the interaction
between hydrogen and deuterium on a Pd(111) surface for temperatures above 75°C. Their study indicated that
transport between the surface and the bulk is an important factor for hydrogen adsorption on the Pd(111) surface.

The kinetic model developed above was applied to the case of hydrogen in palladium by curve fitting, using the values
of k32 and kd as the fitting parameters. The values of all other parameters involved in Eq. (10.135) were taken from the
literature [55, 56]. The equation developed in this fitting process predicts an anomalous decrease in the value of
diffusivity, similar to that found in iron, for temperatures below 75°C for vicinal surfaces. However, no experimental
studies report such a deviation. Only contaminated surfaces have shown values of the apparent diffusivity below that
expected from the extrapolation of the high temperature data, as mentioned above [42]. The reason for this lack of
agreement is not obvious, but may be related to the existence of a strongly bound subsurface hydrogen site.

< previous page page_605 next page >


< previous page page_606 next page >
Page 606

10.3
Surface Chemical Reactions

Many of the surfaces present in vacuum systems can, especially at high temperatures, serve as catalysts for surface
chemical reactions. Those of importance in vacuum technology fall into three classes:

1. Reactions in which a gas is decomposed to form reactive radical speciesfor example, H2 on tungsten filaments.

2. Reactions in which a gas reacts to deposit material in a layer on the surfacefor example, C2H4→ C + 4H2.

3. Reactions in which a gas reacts with an adsorbed species, or with the surface itself, to form new volatile speciesfor
example, O2 + C → CO + CO2, H2O + C → CO + CH4, O2 + Mo → MoO3.

All classes of surface reactions can be thought of in terms of a sequence of steps. In its most general form, the sequence
involves adsorption from the gas phase into a molecular precursor state, chemisorption, surface migration to the reaction
site, the actual reaction step, surface diffusion away from the reaction site, return to a physically adsorbed state, and
finally desorption into the gas phase. Figure 10.43 diagrams this process. It is seldom that all of these steps will be
involved in any given reaction, and in many cases only one of these steps will control the overall reaction rate. In
principle, however, all of these steps must be considered in formulating the detailed reaction mechanism.

A number of these processes have already been discussed in detail previously in this chapter, such as accomodation into
an adsorbed state, physisorption and chemisorption, and surface migration. The only new step in the sequence is the
actual reaction step itself. In discussing this step, it is useful to introduce some new terms that describe this step in
various cases.

Reactions are often referred to as being either structure-sensitive or structure-insensitive, depending on the way in
which the reaction rate changes as a function of

Fig. 10.43
Schematic view of the possible steps involved in a surface
chemical reaction between two homonuclear diatomic molecules.

< previous page page_606 next page >


< previous page page_607 next page >
Page 607

either (a) surface perfection or (b) the particle size of the catalytic material. Reactions whose rate depends only on the
amount of surface present, and not on its structure, are said to be structure-insensitive. Those that show a rate dependent
on surface defect structure are said to be structure-sensitive.

Depending on the importance of chemisorption before reaction, reactions are classified as following either a
LangmuirHinshelwood or an EleyRideal mechanism. A bimolecular reaction that proceeds by chemisorption of both
reactants prior to the reaction step is said to follow a LangmuirHinshelwood mechanism. A bimolecular reaction
between one chemisorbed species and a second species that impinges directly from the gas phase is said to follow an
EleyRideal mechanism.

In the catalysis literature there is also frequent mention of active sites for reaction. This is a term that developed in the
early days of the study of catalytic reactions, as an explanation of the observed structure sensitivity of many reactions.
Much current research in the field is aimed at elucidating the nature of surface configurations that provide especially
favorable sites for the surface reaction step. To date, positive correlations have been found between surface reactivity
and the presence of ledge or kink sites on a crystal surface, but a detailed understanding of their effect on reactivity is
lacking.

A wide variety of reactions can take place between the residual gases in a vacuum system and heated surfaces within the
system. The most obvious place for such reactions to take place is at the surfaces of heated filaments used as electron
emitters. Possible reactions of this type have been discussed by Redhead [57] and by Alpert [58]. Possible reactions for
the case of hydrogen include

H2(g) → 2H(a),
H(a) + O(a) → H2O(g),
H(a) + C(a) → CxHy(g).

The dissociation reaction was observed on tungsten as long ago as 1915 by Langmuir [59]. Hickmott [60] observed the
formation of atomic hydrogen at a tungsten surface at an appreciable rate for surface temperatures above 1000 K, and he
measured an activation energy for the production of atomic hydrogen of 67 kcal/mole. The rate of formation as a
function of filament temperature and residual hydrogen pressure is shown in Fig. 10.44, taken from the paper by
Hickmott.

Atomic hydrogen produced at hot filaments can also react with other adsorbed species both at the heated filament and at
vacuum chamber walls. This reaction has been used as a practical source of atomic hydrogen in many studies. These
include qualitative studies of adsorption of hydrogen on semiconductors, which show a vanishingly small sticking
coefficient for H2(g). It has also been used by Mesters et al. [61] in a study of the removal of adsorbed oxygen from a
Cu(111) surface. Results of this study indicated that the rate of removal by gas phase H2 was limited by inhibition of
dissociative adsorption of hydrogen by the adsorbed oxygen, whereas the adsorption of atomic hydrogen and
subsequent removal of the absorbed oxygen suffered no such limitation. The use of atomic hydrogen produced at a
heated filament as a reactant has also been applied quantitatively, as in studies of hydrogen adsorption on silicon by
Smentkowski et al. [62].

< previous page page_607 next page >


< previous page page_608 next page >
Page 608

Fig. 10.44
Dependence of the rate of formation of atomic hydrogen
(VA) on tungsten filament temperature and hydrogen
pressures (p2). Gas temperature = 298 K [60].

An example of the second type of reaction listed above is the decomposition of ethylene, C2H4, on the Ni(110) surface.
This reaction has been studied by Zuhr and Hudson [63]. Results indicated the rapid, complete dehydrogenation of
adsorbed ethylene, with desorption of the hydrogen liberated as H2, at temperatures above 150°C. At temperatures
below 350°C, the decomposition process led to an adsorbed carbon phase at monolayer coverage. Decomposition at
higher temperatures led to the formation of a graphitic monolayer, which was stable up to 550°C, and dissolved into the
bulk of the nickel crystal at higher temperatures. The resulting surface carbon concentration as a function of surface
temperature during the reaction, as measured by Auger electron spectroscopy (AES), is shown in Fig. 10.45.
Comparison with other studies of ethylene adsorption of nickel surfaces, using surface spectroscopic techniques [64,
65], indicate dissociative adsorption at temperatures as low as 25°C in this system. This is consistent with the
observations of Zuhr and Hudson [63], because hydrogen evolution would have been too slow to detect by the mass
spectrometric techniques used in their study.

An example of the third type of reaction listed abovethat is, one in which a gas-phase species reacts with an adsorbed
species to form a new gas-phase productis the reaction of adsorbed carbon with gaseous O2. In studies of this reaction
by Sau and Hudson [66, 67], adsorbed carbon layers were formed by the decomposition of ethylene on a Ni(110)
surface as described above, then removed by reaction with
< previous page page_608 next page >
< previous page page_609 next page >
Page 609

Fig. 10.45
Steady-state surface carbon coverage on Ni(110), arising
from decomposition of ethylene, as determined by AES at
various substrate temperatures. Reprinted from R. A. Zuhr
and J. B. Hudson, "The Adsorption and Decomposition of
Ethylene on Ni(110)," Surf. Sci. 66, 414 (1977) [Ref. 63], with
kind permission of Elsevier Science NL, Sara Burgerhartstraat
25, 1055 KV, Amsterdam, the Netherlands.

O2 to form CO. This is an example of a so-called cleanoff reaction and is similar to volatilization reactions. This
reaction was studied using molecular beam relaxation spectroscopy, with an O2 molecular beam used as reactant and
the gas-phase O2 and CO fluxes from the surface being measured mass spectrometrically. In addition, the surface
carbon and oxygen coverages were measured by AES. These measurements were made for both (a) the monolayer
carbon structure formed by ethylene adsorption at low temperatures and (b) the graphitic monolayer formed at high
temperatures.

The results of measurements on the graphitic monolayer, for a surface temperature of 873 K, are summarized in Fig.
10.46, which contains much kinetic information. If one looks first at the rate at which carbon disappears from the
surface, on the basis of the carbon AES signal, one sees that the rate is at first slow, accelerates to a maximum, then
drops back to zero as the carbon adlayer is depleted. This same behavior is mirrored in the curve of CO production rate
obtained mass spectrometrically. The phase lag of this CO signal initially increases with increasing extent of reaction,
then saturates. The surface oxygen coverage remains low until very late in the reaction sequence.
< previous page page_609 next page >
< previous page page_610 next page >
Page 610

Fig. 10.46
Summary of kinetic measurements of the oxidation of a graphitic carbon layer on Ni(110).
Reprinted from R. Sau and J. B. Hudson, ''The Oxidation of Graphitic Monolayers on
Ni(110)," Surf. Sci. 95, 468 (1980) [Ref. 66], with kind permission of Elsevier Science NL,
Sara Burgerhartstraat 25, 1055 KV, Amsterdam, the Netherlands.

The reaction sequence deduced from these data is shown in Fig. 10.47. It involves initiation of the reaction at defect
sites in the carbon layer, followed by the growth of "holes" in the adlayer, at which oxygen chemisorption occurs
readily. The adsorbed oxygen, which is mobile on the surface, diffuses to the edge of the hole and reacts to form CO,
which is readily desorbed. This reaction takes place quite efficiently initially, until the hole size grows to the point
where there is competition for adsorbed oxygen between the surface reaction and dissolution of oxygen into the bulk of
the crystal. It is at this point that the phase lag of the product signal saturates. A theoretical rate equation has been
developed embodying this reaction sequence. The fit of the theoretical curves to the experimentally measured carbon
coverage curve is shown in Fig. 10.48.

Results of the reaction of oxygen with the low-temperature carbon layer were more complex. The significant finding as
far as vacuum technology is concerned was that after heating to 525 K, the adsorbed layer was reactive with gaseous O2
at temperatures as low as 333 K, but was unreactive with previously chemisorbed oxygen or oxygen present as NiO.
This suggests the involvement of a weakly chemisorbed oxygen species on the carbon-covered surface, and points to the
possibility of CO production from adsorbed carbon on the walls of metal vacuum chambers, even in cases where the
walls are covered by an oxide layer.

< previous page page_610 next page >


< previous page page_611 next page >
Page 611

Fig. 10.47
Schematic view of the proposed reaction mechanism for the oxidation of a graphitic
carbon monolayer on Ni (100).
Fig. 10.48
Comparison of rate of removal of surface carbon from Ni(110) by
O2 gas with theoretical equations based on the mechanism shown in
Fig. 10.47. Reprinted from R. Sau and J. B. Hudson, "The
Oxidation of Graphitic Monolayers on Ni(110)," Surf. Sci. 95,
471 (1980) [Ref. 66], with kind permission of Elsevier Science NL,
Sara Burgerhartstraat 25, 1055 KV, Amsterdam, the Netherlands.

< previous page page_611 next page >


< previous page page_612 next page >
Page 612

A second reaction in which an adsorbed species is removed by reaction with a gas-phase species is the oxidation of
adsorbed CO by the overall reaction

This reaction has been extensively studied by many investigators using a variety of techniques on a wide range of
surfaces [68]. The major questions that these studies have tried to answer are whether the reaction proceeds by a
LangmuirHinshelwood or an EleyRideal mechanism; the effects of varying either the gas-phase pressure or the surface
adlayer populations of the two reactants; and the effect of surface structure on the reaction rate and mechanism. The
overall reaction scheme deduced is

The overall mechanism is thus the LangmuirHinshelwood mechanism, with both reactants being chemisorbed (and the
O2 molecule dissociated) prior to CO2 formation. Because of the differences in the heats of adsorption of CO and O2,
the reaction rate as a function of temperature and the relative partial pressures of the two reactants is complex. The rate
also depends on whether CO is adsorbed and the surface then exposed to O2 or vice versa. All of the results obtained,
however, are consistent with the reaction scheme presented in Eq. (10.139).

Measurements have also been made in a few cases of reactions between adsorbed species and gas reaching the surface
by permeation through the bulk. Arbab and Hudson [69] have studied the reaction between permeating hydrogen and
oxygen chemisorbed of the surface of a thin, polycrystalline iron membrane to form water. The hydrogen was supplied
to the surface as atomic hydrogen by maintaining an atmosphere of hydrogen on the back side of the iron sample while
the front side was maintained at ultrahigh vacuum in a conventional surface science research system. Reaction rates
were measured both by mass spectrometric detection of the desorbed product water and by AES measurement of the
reduction in the surface oxygen concentration as the reaction proceeded. Results indicated a rapid removal of oxygen as
water, as long as the surface oxygen coverage exceeded one monolayer. Submonolayer coverages were inert to reaction.
A typical mass spectrometer plot of water evolution rate versus time is given in Fig. 10.49, showing an initial rapid rise
as hydrogen is admitted to the back side of the sample, followed by a slower decrease as the surface oxygen layer is
consumed. Measurements of water formation during oxygen exposure to the surface concurrent with steady-state
hydrogen permeation showed the reaction to be half-order in O2 pressure. Kinetic analysis of the results indicated that
the reaction proceeded by the formation of surface hydroxyl groups, followed by their disproportionation to form water
and adsorbed oxygen. The overall activation energy for the process was 16 kcal/mol, of which 8.5 kcal/mol is associated
with the temperature dependence of the hydrogen permeation.

Another reaction sequence of importance in vacuum technology is the so-called water cycle in tungsten filament lamps.
This cycle involves two sets of reactions. At the heated surface of a tungsten filament, water vapor in the system may
react

< previous page page_612 next page >


< previous page page_613 next page >
Page 613

Fig. 10.49
Evolution of water during the titration of adsorbed oxygen by hydrogen
permeating through a thin iron membrane, as measured mass
spectrometrically. Hydrogen permeation was initiated at t = 305.
Reprinted from M. Arbab and J. B. Hudson, "The Titration of Oxygen
Adsorbed on a Polycrystal-line Iron Surface by Hydrogen Permeation,"
Surf. Sci. 209, 191 (1989) [Ref. 69], with kind permission of Elsevier
Science NL, Sara Burgerhartstraat 25, 1055 KV,
Amsterdam, the Netherlands.

according to

H2O(g) → H2(g) + O(a),


O(a) + W(s) → WO3(s) → WO3(g).

At cold surfaces adjacent to the filament, the reverse reaction will be favored, leading to
WO3(a) + H2(g) → WO3(a) + 2H(a),

WO3(a) + 2H(a) → WOx(s) + H2O(a) → H2O(g).

This reaction sequence is responsible for the blackening observed on the walls of ionization gauge tubes operated in
high partial pressures of water vapor.

< previous page page_613 next page >


< previous page page_614 next page >
Page 614

In electric illumination technology, the development of halogen lamps was a reaction to this problem and the related
problem of an excessively high tungsten evaporation rate at high filament temperatures. In these lamps, a low pressure
of a halogen gas such as iodine is used to effect the return of tungsten to the heated filament by the series of reactions

W(s) + I2(g) → W(s) + 2I(a)

W(s) + 3I(a) → WI3(g)

at the relatively cold bulb surface, and

WI3→ W(s) + I2(g)

at the filament surface, thus reversing the reaction that led to tungsten loss initially. This cycle permits operation of the
tungsten filament at higher temperature, and it consequently provides a brighter more efficient light source.

10.4
Outgassing Behavior

Except during the early stages of a pumpdown cycle, when the air or other gas in a system is being removed, the bulk of
the gas load to the pumping system consists of gas evolved from surfaces in the system, including both the system walls
and any objects within the vacuum system. This gas load arises from a combination of desorption from adsorbed layers
and permeation of gases dissolved in the system walls or other surfaces. Since the system ultimate pressure is directly
related to the magnitude of this gas load, measures taken to reduce outgassing are critical to the production of high and
ultrahigh vacuum. Measures include both treatments applied to the materials of system construction, prior to or in the
course of system fabrication, and treatment of the system as a whole in the course of system operation. The first
category includes both (a) heat treatments to reduce the amount of dissolved gases in the materials of construction and
(b) surface treatments intended either to reduce the total surface area or to develop a surface layer that will be
impermeable to gases or unreactive to chemisorption or surface reactions. The second category includes various
"bakeout" techniques intended to desorb adsorbed layers present prior to pumpdown and to deplete the near surface
region of materials of dissolved gases.

The question of outgassing will be considered below, in terms of both (a) the development of pressure versus time
curves based on various assumptions concerning the adsorption, desorption, and diffusion behavior of the gases
involved and (b) measures that may be used to decrease the contribution of adsorption, desorption and diffusion
processes to system residual pressure.

10.4.1
Desorption of Adsorbed Gases

Earlier in this chapter, equations were developed for the rate of desorption of adsorbed gases based on a range of
assumptions concerning desorption energetics and mechanism. In the simplest cases, the desorption rate decreased
exponentially with time as the adlayer population was depleted by desorption. The time constant for this

< previous page page_614 next page >


< previous page page_615 next page >
Page 615

exponential decrease is the mean stay time for adsorption, which in turn depends on the activation energy for the
desorption process. Because of this, the effects of various adsorbed species on system pumpdown rate and ultimate
pressure fall into three classes. Adsorbates with a very short surface lifetime, such as weakly physisorbed gases
including the rare gases, oxygen, and nitrogen, will be desorbed and removed by the pumping system rapidly and will
not appreciably affect the overall pumpdown rate. At the other extreme, gases that are very strongly adsorbed and have
essentially infinite surface lifetimes at ambient temperature, such as chemisorbed oxygen, nitrogen, or halogens, will
have a vanishingly small desorption rate and will likewise not significantly affect pumpdown rates. It is the gases
having surface lifetimes in the intermediate range of seconds to hours that provide the principal impediment to rapid
system pumpdown. Included in this category are most organic species, water, and weakly chemisorbed species such as
CO and CO2. This effect is shown graphically in Fig. 10.50 [71], which shows the pressure versus time for a 1-liter
system pumped by a 1-liter/s pump and having a surface area of 100 cm2, initially covered with a monolayer of gas
having the desorption energy shown in the figure. Note that for desorption energies less than 15 kcal/mol, the desorption
is rapid; also note that for desorption energies of 25 kcal/mol and higher, the outgassing rate contributes a partial
pressure below 1011 Torr. Also shown in Fig. 10.50 is the rate of decrease in system pressure in the course of a bakeout
at 300°C for adsorbates having desorption energies in the troublesome 15- to 25-kcal/mol range. At these temperatures,
the contribution to system pressure from these gases effectively disappears in a matter of seconds. Note that this
calculation includes only gases adsorbed at the surface of the system and does not account for replenishment of the
adsorbed layer by permeation from the bulk, which is a much more serious problem in the bakeout of practical systems.

Fig. 10.50
Pressure vs. time curves for the pumpdown of vacuum systems having adsorbed layers of
gas with the adsorption energies shown. System parameters are V=1 liter, S=1 liter/s,
A=100 cm2. Initial adlayer coverage is one monolayer at t=0 [71].

< previous page page_615 next page >


< previous page page_616 next page >
Page 616

10.4.2
Dissolved Gases

The release of species by permeation through the bulk of materials exposed to the vacuum is usually the final factor
limiting system ultimate pressure. Here the major contributors are usually hydrogen released by permeation of dissolved
hydrogen atoms followed by recombination and desorption as H2 molecules, and CO or CO2 formed by permeation of
carbon atoms and surface reaction with water followed by desorption. In glass systems, CO and CO2 remaining in
glasses made by calcining carbonates can also permeate as molecules and be desorbed into the vacuum.

10.4.3
Overall Pumpdown Curves

The pressure vs. time behavior of a typical vacuum system will depend on both the inherent adsorption-desorption and
diffusion behavior of the gases present in the system when pumpdown begins, and on the past history of the system. It
will also depend on system configuration, in particular the extent of the surface area of the system relative to the system
pumping speed or the area of any orifice that limits that pumping speed. In typical systems, the surface area will be
large compared to the size of the orifice through which gas must be pumped. As a result, a gas molecule desorbed from
the surface will be much more likely to readsorb on another site on the surface than to be removed by the pumping
system. Consequently, treatments of the pumpdown process must take account of both the desorption and readsorption
processes. Moreover, in cases in which there is a finite rate of replenishment of the adsorbed layer by diffusion of gas
from the bulk of the system walls or objects within the system, the effect of this process and the extent to which
dissolved gases are replenished when the system is exposed to the atmosphere must also be considered.

There have been numerous treatments of the rate of gas removal and its effect on system pressure. These generally fall
into one of two categories: (1) treatments of the adsorption desorption process, which are most applicable to the early
stages of pumpdown and have been primarily concerned with the removal of adsorbed water, and (2) treatments of the
diffusion and desorption of dissolved gases, important primarily in later stages of the pumping process and which
usually assume that desorption is not rate-limiting and that no readsorptionabsorption processes are important.

Consider first treatments of the effect of adsorptiondesorption processes on pumpdown behavior. This case has been
treated in detail by Redhead [72, 73] for both monolayer and multilayer adsorption. The basic conservation of mass
equation relating system pressure and pumping time in this case is given by Redhead as

where K is a constant to convert from pV units to molecules (3.27 × 1019 molecules/Torr-liter or 2.45 × 1019 molecules/
mbar-liter at 295 K), S is the pumping speed and N is the total number of adsorbed molecules. This equation has been
solved by several authors, subject to various assumptions concerning the form of the desorption rate equation and the
readsorption probability. Hobson [74] and Bills [75] have solved this equation subject to the assumption of adsorption
with a constant sticking

< previous page page_616 next page >


< previous page page_617 next page >
Page 617

coefficient and a first-order desorption rate with desorption energy independent of coverage. This leads to an equation
of the form

The problem with this approach is that it assumes that both the adsorption probability and the desorption energy are
coverage-independent. A more realistic approach is to assume a quasistatic system, in which the adlayer is in
equilibrium with the gas phase. Use of this assumption is based on the fact that readsorption of a desorbed molecule is
much more probable than removal by the pumping system, as mentioned above. This assumption, along with the
assumptions of perfectly reversible adsorption and a form for the adsorption isotherm, permits solution of Eq. (10.140).
Substitution of a generalized adsorption isotherm equation of the form

where n is the number of adsorbed molecules per unit area and the subscript m refers to monolayer coverage (assumed
to be the maximum coverage), into Eq. (10.140) yields

which may be integrated to yield

where p0 is the pressure at t = 0. The solution is completed by using the expression for dθ/dp found by differentiating
the isotherm equation chosen.

This approach has been taken by Venema [76], using the Langmuir isotherm, by Mizuno and Horikoshi [77, 78], using
the Freundlich isotherm, by Weiss [79], using the DubininRadushkevich isotherm, and by Redhead [72], using an
equation based on the Temkin isotherm.

Redhead [73] has extended the treatment based on the Temkin isotherm to cover multilayer adsorption, using an
empirical multilayer adsorption isotherm of the form

in which Ψ = p/ps and ps is the saturated vapor pressure of water at the temperature of interest. Application of this
isotherm equation to Eq. (10.140) showed that the system pressure decreases exponentially with a time constant V/S as
the adlayer coverage

< previous page page_617 next page >


< previous page page_618 next page >
Page 618

drops through the multilayer region. This is the same behavior as one would expect in the absence of adsorption.

Treatments that consider the effect of gas diffusing out of the bulk of the system walls are based on the diffusion
equations presented in Section 10.2. In the simplest case, in which the gas being removed is present in the system walls
at the time of construction and is not replenished by exposure of the system to the atmosphere, the mass balance
equation is again

where in this case dN/dt = dQ/dt for the appropriate solution to Fick's second law. For example, for the case of diffusion
from a semi-infinite slab, we have, from Eq. (10.121),

Integration of this equation is complicated by the fact that C0 is not an explicit function of p. However, as a practical
matter, the system pressure will be dominated by permeationdesorption processes only after the gas initially in the
system volume and the bulk of the adsorbed gases have been removed. The pressure in the system can thus be
represented by the quasi-steady-state expression

for the case of the semi-infinite slab.

The more complicated case of a dissolved or adsorbed gas source that may be replenished by exposure to the
atmosphere has been considered by Dayton [80]. The analysis is complicated in this case by the fact that the behavior
observed in any situation depends on the extent to which the dissolved or adsorbed gas has been replenished, and
consequently on the time that the system has been exposed to the dissolving gas since the previous pumpdown and on
the extent to which the dissolved gas was desorbed during previous pumpdown cycles. Dylla et al. [81] have observed
empirically that the initial pumpdown of such a system follows an outgassing equation of the form

where n is a constant of order unity, whose exact value depends on the details of the system involved.

< previous page page_618 next page >


< previous page page_619 next page >
Page 619

10.4.4
Mitigation of Outgassing

Finally, consider techniques that may be used to reduce the rate of outgassing of surfaces in vacuum. These generally
fall into three categories; (1) surface treatments carried out prior to or during system fabrication aimed at developing
surfaces that are relatively inert to chemisorption or provide low diffusivity barriers to the release of absorbed gases, (2)
surface treatments carried out in situ following system construction, and (3) bakeout procedures aimed at reducing the
amount of adsorbed or absorbed gases present in the system subsequent to pumpdown.

10.4.5
Surface Treatments during Construction

First consider treatments carried out during the construction of the system. For the case of stainless steel systems, the
use of vacuum-melted alloys, or the vacuum degassing of construction materials prior to or during system construction,
has been used to reduce the amount of dissolved gases (primarily hydrogen) and to refine the structure and composition
of the surface oxide. This technique has been found to significantly reduce hydrogen outgassing in baked UHV systems
[82], but has yielded only slight improvements in water outgassing rates in unbaked systems [81, 83].

A number of techniques have been developed aimed at reducing the ratio of real surface area to the geometrical area of
the surface (the surface roughness factor) and at producing a thin, dense surface oxide layer that will be relatively inert
to chemisorption. For the case of stainless steel systems, these have been primarily electropolishing techniques. In this
process, the surface to be treated is made the anode in an electrochemical cell. As the cell operates, protrusions on the
surface are dissolved preferentially, and the surface is progressively flattened as the process proceeds. Yoshimura et al.
[84] found that an electropolished system, after a 150°C, 20-hour bakeout, yielded an outgassing rate of 1.1 × 1013
mbar-liter/s-cm2. Okamura et al. [85] found that electropolishing or electrolytic abrasive polishing produced surfaces
that had an outgassing rate below 1012 mbar-liter/s-cm2 following a 250°C, 80-hour bakeout. In contrast to these
results, two studies of water outgassing in unbaked electropolished stainless steel systems [81, 86] found that the
electropolishing process had little effect on water outgassing. For the case of aluminum systems, Ohi and Konno [87]
found that electropolishing produced a highly hydrated oxide and had little effect in reducing water outgassing in an
unbaked system.

Surface machining processes such as diamond machining [87], or mirror polishing by mechanical polishing techniques
[8891], have been found to reduce water outgassing in aluminum systems by as much as a factor of 10 relative to
untreated surfaces. Suemitsu et al. [88] also found that the outgassing rate was proportional to the surface roughness
factor, with the surface roughness of the mirror polished surface being roughly a factor of three smaller than that
produced by other machining or electropolishing procedures.

Extensive use has also been made of "alcohol lathing" techniques in the construction of aluminum systems [9294]. In
this process, a thin layer of material, roughly 500 µm, is machined from the surface of the material, using an alcohol as
the cutting fluid. This process produces a smooth surface, covered with a thin, dense oxide layer. A study by Suemitsu
et al. [92] indicated that isopropanol gave the best results of

< previous page page_619 next page >


< previous page page_620 next page >
Page 620

a series of alcohols and reported an outgassing rate of 7 × 1012 mbar-liter/s-cm2 after a 100°C, 24-hour bakeout.

The oxide layers on stainless steel may also be modified by various passivating techniques involving exposure to
oxygen or fluorine at high temperatures [95, 96]. Treatments of this sort leave the surface covered with a thin, dense
oxide or fluoride coating, which is inert to chemisorption of most gases and has a very low specific surface area for
water adsorption.

10.4.6
In Situ Surface Treatments

In situ treatments in the course of system commissioning, primarily glow discharge or plasma processes, have been used
to reduce the amount of adsorbed water and also to remove surface carbon-containing compounds by oxidation. This
process involves either (a) applying a high voltage to some element in the system in order to initiate a plasma discharge
or (b) attaching a separate plasma source to the piece being cleaned. The desorption rate of adsorbed gases is enhanced
by bombardment of the interior surfaces of the system by energetic ions, electrons, or excited neutral species, which
cause desorption by a number of energetic particleadsorbate interactions, and, in the case of oxygen plasmas, react with
surface carbon-containing compounds to produce volatile species such as CO. This process is, of course, limited to
systems in which production of the required plasma is possible, and it is not deleterious to fixtures within the chamber.
Dylla et al. [81] found small but significant improvements in water outgassing in both aluminum and stainless steel
systems when using helium glow discharge cleaning. Similarly, Chou [83] found that oxygen glow discharge cleaning
of an unbaked aluminum system reduced outgassing due to photodesorption in an ultraviolet synchrotron beam line. Ota
et al. [97], using an oxygen plasma source attached to a high-purity aluminum synchrotron housing, found that the dc
oxygen plasma could reduce outgassing due to photon bombardment by a factor of 10 when the housing was placed in
service. Auger electron spectrometric analysis of surfaces treated by this process showed virtually complete removal of
carbon from the surface oxide layer.

10.4.7
Bakeout Processes

The fact that both desorption and permeation rates depend exponentially on temperature is used extensively in the
reduction of outgassing through various bakeout procedures. Typically, the system is pumped into the range where the
residual pressure is dominated by desorption or permeation processes. The temperature is then raised, either by external
heaters or by a high-intensity light source within the vacuum chamber. After a period of hours, the system is returned to
room temperature.

Adsorbed materials will be removed from the system rapidly by this process. The analyses of system pumpdown rate
mentioned earlier [72, 73, 7679] also apply to conditions during the bakeout process, with the difference that the
desorption rate and equilibrium adlayer coverage are those appropriate to the bakeout temperature. This will result in a
much higher system pressure at the beginning of the bakeout process, and consequently to a much more rapid rate of gas
removal for a given system pumping speed. As an example, consider the case of water adsorbed on

< previous page page_620 next page >


< previous page page_621 next page >
Page 621

stainless steel. If a heat of adsorption of 15 kcal/mol and a bakeout temperature of 150°C are assumed, the rate of gas
removal will be increased by a factor of roughly 5000 relative to the value at room temperature. This will result in
essentially complete removal of the adsorbed water in a matter of seconds.

Removal of adsorbed gases by permeation through the bulk followed by desorption is a more lengthy process. Here one
may consider the case of a thick-walled system, so that the diffusion step can be treated as diffusion out of a semi-
infinite slab. For this case, the equation developed previously for the rate of outgassing for diffusion out of a semi-
infinite slab, namely

may be used to calculate the rate of outgassing both before and after the bakeout process. Prior to bakeout, the
outgassing rate will be given by

where Dlo is the diffusivity of the gas involved at room temperature and tlo is the time that the system has been pumped
at room temperature. Following bakeout for a time, thi, the outgassing rate will be given by

Using the literature values for the diffusion of hydrogen in iron [98] (namely, QD = 1600 cal/mol and D0 = 1 × 103
cm2/s), and assuming a 10-hour bakeout at 150°C shows that the rate of outgassing after the bakeout is a factor of five
smaller than that prior to bakeout. This relatively modest reduction is due to the very low activation energy associated
with the permeation of hydrogen through iron. A large increase in permeation rate can be attained only at a much higher
temperature, as was used in the vacuum degassing process described earlier [82, 83]. Species with higher activation
energies for permeation will be more efficiently removed by bakeout at temperatures in the 150°C range, and higher
bakeout temperatures would result in an even greater reduction in outgassing. As a practical matter, the limit on bakeout
temperature is usually set by the presence of temperature-sensitive components in the system, or by the desire to avoid
oxidation of the external surfaces of the system. Pyrex ultrahigh-vacuum systems were typically baked out at 450°C, the
highest temperature that could be used without exceeding the softening point of the glass. Stainless steel systems are
typically baked out at 150°C to avoid surface oxidation.

High-intensity ultraviolet light sources located within the vacuum chamber are also finding increasing use in bakeout
processes. In this case, the mechanism of removal of adsorbed gases involves both heating of surfaces in the vacuum
and, possibly, direct photon-excited desorption of adsorbed species. This approach has the advantage of concentrating
the heating on the surfaces where the adsorbate resides, rather than heating the whole system to the required bakeout
temperature.

< previous page page_621 next page >


< previous page page_622 next page >
Page 622

References

1. Adapted from J. H. DeBoer, The Dynamical Character of Adsorption, p. 35. Oxford University Press (Clarendon),
Oxford, 1953.

2. I. Langmuir, J. Am. Chem. Soc. 40, 1361 (1918).

3. S. Brunauer, P. H. Emmett, and E. Teller, J. Am. Chem. Soc. 60, 309 (1938).

4. S. Brunauer, L. S. Deming, W. E. Deming, and E. Teller, J. Am. Chem. Soc. 62, 1723 (1940).

5. S. Ross and J. P. Olivier, On Physical Adsorption. Wiley, New York, 1964.

6. A. Magnus and H. Kratz, Z. Anorg. Chem. 184, 241 (1929).

7. S. Ross and J. P. Olivier, On Physical Adsorption, pp. 188189. Wiley, New York, 1964.

8. A. D. Crowell, J. Chem. Phys. 26, 1407 (1957).

9. W. Machin and S. Ross, Proc. Soc. London. Ser. A 265A, 455 (1962).

10. F. H. Buttner, E. R. Funk, and H. Udin, J. Phys. Chem. 56, 657 (1952).

11. T. W. Hickmott, J. Chem. Phys. 32, 810 (1960).

12. P. Kisliuk, J. Chem. Phys. 31, 1605 (1959).

13. G. C. Bond, Catalysis by Metals. Academic Press, New York, 1962.

14. D. A. Hoffman and J. B. Hudson, Surf. Sci. 180, 77 (1987).

15. R. H. Jones, D. R. Olander, W. J. Siekhaus, and J. A. Schwarz, J. Vac. Sci. Technol. 9, 1429 (1972).

16. M. D. Scheer and J. Fine, in Condensation and Evaporation of Solids (E. Ruttner, P. Goldfinger, and J. P. Hirth,
eds.), p. 327. Gordon & Breach, New York, 1964.

17. J. B. Hudson and J. S. Sandejas, J. Vac. Sci. Technol. 4, 230 (1967).

18. C. M. Lo and J. B. Hudson, Thin Solid Films 12, 261 (1972).

19. R. Zuhr and J. B. Hudson, Surf. Sci. 66, 405 (1977).

20. R. Sau and J. B. Hudson, J. Vac. Sci. Technol. 18, 607 (1981).

21. E. A. Kurz, S. Lassig, and J. B. Hudson, J. Vac. Sci. Technol. A 1, 1266 (1983).

22. D. H. Winicur, J. Hurst, C. A. Becker, and L. Wharton, Surf. Sci. 109, 263 (1981).

23. C. T. Campbell, G. Ertl, H. Kuipers, and J. Segner, Surf. Sci. 107, 220 (1981).

24. D. A. King, CRC Crit. Rev. Solid State Mater. Sci. 7, 167 (1977).

25. E. A. Kurz and J. B. Hudson, Surf. Sci. 195, 31 (1988).


26. C. T. Rettner, L. A. DeLouise, and D. A. Auerbach, J. Chem. Phys. 85, 1131 (1986).

27. Adapted from J. H. DeBoer, The Dynamical Character of Adsorption, p. 37. Oxford University Press (Clarendon),
Oxford, 1953.

28. P. A. Redhead, J. P. Hobson, and E. V. Kornelsen, The Physical Basis of Ultrahigh Vacuum, p. 76. Chapman &
Hall, London, 1968.

29. P. Clausing, Ann. Phys. (Leipzig) [5] 7, 521 (1930).

30. D. Müller, Z. Phys. 188, 326 (1965).

31. R. M. Barrer and D. M. Grove, Trans. Faraday Soc. 47, 826 (1951).

32. J. Crank, The Mathematics of Diffusion, 2nd ed. Oxford University Press (Clarendon), Oxford 1975.

33. F. J. Norton, J. Am. Ceram. Soc. 36, 90 (1953).

34. C. C. Lieby and C. L. Chen, J. Appl. Phys. 31, 268 (1960).

35. D. Alpert and R. S. Buritz, J. Appl. Phys. 25, 202 (1954).

36. N. W. Taylor and W. Rast, J. Chem. Phys. 6, 612 (1938).

37. J. R. Young and N. R. Whetten, Rev. Sci. Instrum. 32, 453 (1961).

< previous page page_622 next page >


< previous page page_623 next page >
Page 623

38. M. Arbab and J. B. Hudson, Appl. Surf. Sci. 29, 1 (1987).

39. J. Crank, The Mathematics of Diffusion, 2nd ed., Chapter 4. Oxford University Press (Clarendon), Oxford, 1975.

40. O. D. Gonzalez, Proc. Conf. Fundam. Aspects Stress Corros. Cracking, Ohio State University, 1967, p. 43 (1969).

41. E. A. Kurz and J. B. Hudson, Surf. Sci. 195, 15 (1988).

42. H. G. Nelson and J. E. Stein, NASA Tech. Note NASA TN D-7265 (1973).

43. S. Wach and A. P. Miodownik, Corros. Sci. 8, 271 (1968).

44. W. Palczewska and I. Ratajczyk, Bull. Acad. Pol. Sci. Ser. Chim. 9, 267 (1961).

45. J. Benziger and R. J. Madix, Surf. Sci. 94, 119 (1980).

46. R. V. Bucur, Int. J. Hydrogen Energy 10, 399 (1985).

47. J. Völkl and G. Alefeld, in Hydrogen in Metals I (J. Völkl and G. Alefeld, eds.), Chapter 12. Springer, Berlin, 1978.

48. Y. Ebisuzaki, W. J. Kass, and M. O'Keeffe, J. Chem. Phys. 46, 1378 (1967).

49. K. Christmann, O. Schober and G. Ertl, Surf. Sci. 99, 320 (1980).

50. T. E. Felter, S. M. Foiles, M. S. Daw, and R. H. Stulen, Surf. Sci. 171, L379 (1986).

51. G. D. Kubiak and R. H. Stulen, J. Vac. Sci. Technol. A 4, 1427 (1986).

52. R. J. Behm, K. Christmann, and G. Ertl, Surf. Sci. 99, 320 (1980).

53. F. Boszo, G. Ertl, M. Grunze, and M. Weiss, Appl. Surf. Sci. 1, 103 (1977).

54. T. Engel and H. Kuipers, Surf. Sci. 90, 162 (1979).

55. S. A. Koffler, J. B. Hudson, and G. S. Ansell, Trans. Metall. Soc. AIME 245, 1735 (1974).

56. H. Conrad, G. Ertl, and E. E. Latta, Surf. Sci. 41, 435 (1974).

57. P. A. Redhead, AVS Natl. Vac. Symp., Trans. 6, 12 (1960).

58. D. Alpert, Vide 17, 19 (1962).

59. I. Langmuir, J. Am. Chem. Soc. 37, 417 (1915).

60. T. W. Hickmott, J. Chem. Phys. 32, 810 (1960).

61. C. M. A. M. Mesters, T. J. Vink, O. L. J. Gijzeman, and J. W. Geus, Appl. Surf. Sci. 26, 367 (1986).

62. V. A. Smentkowski, H. Jänsch, M. A. Henderson, and J. T. Yates, Jr., Surf. Sci. 330, 207 (1995).

63. R. A. Zhur and J. B. Hudson, Surf. Sci. 66, 405 (1977).

64. J. E. Demuth and D. Eastman, Phys. Rev. Lett. 32, 1123 (1974).
65. G. Dalmai-Imelik and J. C. Bertolini, C. R. Hebd. Seances Acad. Sci., Ser. C. 270, 1079 (1970).

66. R. Sau and J. B. Hudson; Surf. Sci. 95, 465 (1980).

67. R. Sau and J. B. Hudson, Surf. Sci. 102, 239 (1981).

68. Results of these measurements are summarized in G. Ertl, Surf. Sci. 299/300, 742 (1994).

69. M. Arbab and J. B. Hudson, Surf. Sci. 209, 183 (1989).

70. P. A. Redhead, J. P. Hobson, and E. V. Kornelsen, The Physical Basis of Ultrahigh Vacuum, p. 370. Chapman &
Hall, London, 1968.

71. J. P. Hobson, Trans. 8th Natl. Vac. Symp., 26 (1961).

72. P. A. Redhead, J. Vac. Sci. Technol. A 13, 467 (1995); 14, 2680 (1996).

73. P. A. Redhead, J. Vac. Sci. Technol. A 13, 2791 (1995).

74. J. P. Hobson, Proc. Int. Vac. Congr. 2nd, 1961, p. 26 (1962).

75. D. G. Bills, J. Vac. Sci. Technol. 6, 166 (1969).

< previous page page_623 next page >


< previous page page_624 next page >
Page 624

76. A. Venema, Proc. Int. Vac. Congr. 2nd, 1961, p. 1 (1962).

77. H. Mizuno and G. Horikoshi, KEK Rep. 84-11 (1984).

78. G. Horikoshi, J. Vac. Sci. Technol. A 5, 2501 (1987).

79. D. R. Weiss, Notes of Workshop on Control and Measurement of Water in Vacuum Systems. NIST, Gaithersburg,
MD, 1994.

80. B. B. Dayton, J. Vac. Sci. Technol., A 13, 451 (1995).

81. H. F. Dylla, D. M. Manos, and P. H. LaMarche, J. Vac. Sci. Technol. A 11, 2623 (1993).

82. N. Yoshimura, H. Hirano, T. Sato, I. Ando, and S. Adachi, J. Vac. Sci. Technol. A 9, 2326 (1991).

83. T. S. Chou, J. Vac. Sci. Technol. A 9, 2014 (1991).

84. N. Yoshimura, H. Hirano, K. Ohara, and I. Ando, J. Vac. Sci. Technol. A 9, 2315 (1991).

85. S. Okamura, E. Miyauchi, and T. Hisatsugu, J. Vac. Sci. Technol. A 9, 2405 (1991).

86. K. L. Siefering and W. H. Whitbeck, J. Vac. Sci. Technol., A 12, 2685 (1994).

87. T. Ohi and O. Konno, J. Vac. Sci. Technol. A 12, 3186 (1994).

88. M. Suemitsu, H. Shinoyamada, N. Miyamoto, T. Tokai, Y. Moriya, H. Ikeda, and H. Yokohama, J. Vac. Sci.
Technol. A 10, 570 (1992).

89. K. Okada, Y. Ishikawa, and M. Furuse, J. Vac. Sci. Technol., A 5, 2902 (1987).

90. N. Yoshimura, T. Sato, S. Adachi, and T. Kunazawa, J. Vac. Sci. Technol. A 8, 924 (1990).

91. S. Kato, M. Aono, K. Sato, and Y. Baba, J. Vac. Sci. Technol. A 8, 2860 (1990).

92. M. Suemitsu, H. Shinoyamada, N. Matsuzaki, and N. Miyamoto, J. Vac. Sci. Technol. A 10, 188 (1992).

93. M. Suemitsu, T. Kaneko, and N. Miyamoto, J. Vac. Sci. Technol. A 5, 37 (1987).

94. M. Suemitsu, T. Kaneko, and N. Miyamoto, J. Vac. Sci. Technol. A 7, 2658 (1989).

95. T. Okumura, Submicron ULSI Process Technology II, p. 101. Tokyo, 1989.

96. M. Matagoro, Submicron ULSI Process Technology II, p. 119. Tokyo, 1989.

97. N. Ota, M. Saitoh, K. Kanazawa, T. Momose, and H. Ishimaru, J. Vac. Sci. Technol. A 12, 826 (1994).

98. R. F. Miller, J. B. Hudson, and G. S. Ansell, Metall. Trans. A 6A, 117 (1975).

< previous page page_624 next page >


< previous page page_625 next page >
Page 625

11
Ultrahigh and Extreme High Vacuum

Paul A. Redhead

Ultrahigh vacuum (UHV) is defined by the American Vacuum Society (AVS) [1] as the range of pressure between 107
and 1010 Pa (7.5 × 1010 to 7.5 × 1013 Torr), and extreme high vacuum (XHV) is defined as the range of pressure
below 1010 Pa (7.5 × 1013 Torr). British [2] and German [3] standard definitions differ from those of the AVS; both
these standards define ultrahigh vacuum as the range of pressure below 106 Pa (108 Torr).

The development of vacuum techniques adequate to reach ultrahigh vacuum started with the investigations of Langmuir
and his associates in 19121913. At that time a considerable number of physicists believed that electron emission from
an incandescent metal surface at low pressures was due to the presence of residual gas and that consequently the
thermionic emission would disappear in a good vacuum. In order to demonstrate the existence of pure thermionic
emission, as well as the validity of the space-charge relation, Langmuir [4] used a bulb containing two hairpin filaments,
either of which could serve as an electron emitter and the other as anode. During exhaust (by a Gaede rotary mercury
pump) the glass bulb was heated for 1 hour at 360°C, which was the highest practical temperature without deformation
of the bulb by atmospheric pressure. A liquid-air trap was placed between bulb and pump. The filaments were heated to
27002800°C in order to degas them, and the bulb was sealed off. In order to improve the vacuum, the filaments were
aged at 2400°C for about 24 hours, which served to clean up the residual gases oxygen and

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_625 next page >


< previous page page_626 next page >
Page 626

nitrogen. ''To still further improve the vacuum in some cases," Langmuir states that "the entire bulb was immersed in
liquid air and the filaments heated to a high temperature for a short time." As shown subsequently by Dushman [5],
measurements with the molecular-drag gauge indicated that the lowest pressure obtained in Langmuir's experiments was
well below 5 × 105 Pa.

Langmuir's techniques became general practice in the 1920s and 1930s to obtain very low pressures (with the diffusion
pump replacing the older types of pump and pressure measured by a hot-cathode ionization gauge). Even with the
rigorous processing techniques discussed above, the lowest pressure indicated by the contemporary ionization gauge (i.
e., a hot-cathode gauge with a cyclindrical ion-collector of large area) was about 106 Pa. However, there was
considerable evidence from measurements of the rate of change of surface properties (work function, thermionic
emission, etc.) in the 1930s and 1940s that much lower pressures were being obtained than were indicated by the
ionization gauges. In the late 1930s Anderson [6] and Nottingham [7] showed, from measurements of the rate of change
of work function, that the pressure of adsorbable gases in their systems was much less than that indicated by the
ionization gauge. It can be estimated from Anderson measurements of the rate of change of the work function of a
tungsten surface that the pressure of adsorbable gases in his system was below 109 Pa. For a review of these early
attempts to approach UHV in the period before 1950, see Redhead [8].

In 1947 Nottingham [9] suggested that the limit to the lowest measurable pressure was caused not by the pumps but by
an x-ray effect in the hot-cathode ionization gauge. Nottingham proposed that soft x-rays, produced when electrons
impinged on the anode with energies of about 150 volts, released photoelectrons from the ion collector; this
photocurrent was indistinguishable in the measuring circuit from the current due to positive ions arriving at the ion
collector. This hypothesis was soon confirmed, and in 1950 the BayardAlpert ionization gauge [10] was announced
which reduced the lowest measurable pressure by a factor of about 100 (i.e. to about 108 Pa) by reducing the size of the
ion collector, from a large cylinder surrounding the other electrodes in the conventional gauge, to a fine wire on the axis
of the grid. The ion collector in the form of a fine wire intercepted only about 1% of the soft x-rays that would strike the
cylindrical ion collector of a conventional gauge. This simple and elegant gauge is still the most widely used form of
hot-cathode ionization gauge. Since the invention of the BayardAlpert gauge (BAG), several other types of total
pressure gauges, both hot- and cold-cathode, suitable for use in UHV and XHV have been developed and mass
spectrometers have been modified to measure partial pressures in UHV and XHV.

Equipment to measure and produce UHV is now commercially available in great variety, and UHV technology is
widely used in industry and the laboratory whenever it is essential to keep surfaces free from contamination, gases pure,
plasmas undefiled, and charged particles unscattered. UHV is widely used in industry in the development and
manufacture of semiconductor devices to keep surfaces uncontaminated during processing and testing. In space
research, UHV and XHV technologies are needed to simulate space conditions for testing components and for the
measurement of gas density in space (e.g., UHV gauges have been used to measure the atmospheric pressure on the
surface of the moon). Particle accelerators and storage rings require UHV conditions to minimize the loss of charged
particles from the beam by collisions

< previous page page_626 next page >


< previous page page_627 next page >
Page 627

Table 11.1. Gas-Phase Parameters at UHV/XHV


Pressure Molecular Molecular Fluxb Molecular Mean Time for 1/10 Monolayerc
(Pa) Densitya (molecules· cm2·s1) Free Pathb
104 1.9 s
2.5 × 1010 cm3 2.9 × 1014 34 m
107 32 min
2.5 × 107 cm3 2.9 × 1011 34 km
1010 22 days
2.5 × 104 cm3 2.9 × 108 3.4 × 104 km
1013 60 years
25 cm3 2.9 × 105 3.4 × 107 km
1017 600 millenia
0.25 m3 29 0.03 light years
a At 298 K.
b For N2 at 295 K.
c Assuming a constant sticking probability of 0.1 and a density of adsorption sites of 1015 cm2.

with gas molecules. Magnetic confinement devices for research in thermonuclear fusion require UHV conditions to ensure the purity of the
confined plasma and to minimize plasma loss. UHV technology is essential to experimental surface science in order to maintain surfaces
uncontaminated for the duration of experimental procedures. It was the development of UHV technology in the 1950s that permitted the start of
modern surface science where the initial state of the surface could be rigorously controlled. The commercial availability of UHV hardware in the
1960s led to the almost exponential expansion in surface science to the point where it is now one of the most active sectors of research in
physics and chemistry.

XHV technology is still in the development stage; it is in use in some accelerators and storage rings and has potential application in the processing
of advanced semiconductor devices.

Table 11.1 indicates in a general way the properties of the UHV and XHV environment which make it essential for studies requiring clean
surfaces, pure gases, or unscattered particles. UHV technology has been discussed in detail in several texts [11, 12].

The problems in achieving UHV or XHV can be seen in a general way by examining the conservation of mass equation for a vacuum system of
volume V and surface area A, pumped by a pump of speed S with a leak rate of L and an outgassing rate per unit area of Q; then if L, Q, and S
are constant and a single gas is predominant, we obtain

When the ultimate pressure p∞ is reached, then dp/dt = 0 and assuming that L → 0 (which is essential for UHV/XHV) we obtain

< previous page page_627 next page >


< previous page page_628 next page >
Page 628

Thus a value of p∞ in the UHV or XHV range may be achieved by decreasing Q or increasing S. Since the maximum
value of S is limited both by available space and by budgets, it is usually necessary to take vigorous action to reduce the
outgassing rate Q from the walls and internal parts of the system, and in particular from vacuum gauges, residual gas
analyzers, and other components with heated cathodes.

The pumping speed of a surface at which all impinging molecules are removed from the gas phase is

where T is the temperature of the gas and M is its molecular weight. For hydrogen (the most prevalent gas at UHV/
XHV) at 300 K we obtain

thus the maximum pumping speed of any type of pump for H2 is 44 liter·s1 for each square centimeter of effective
pumping area. The above equations are a considerable simplification since they are only strictly correct if the gas
density is uniform throughout the vacuum chamber. At very low pressures in the presence of adsorbing surfaces or
localized pumps (e.g., gauges acting as pumps with limiting conductance to the chamber), this is not necessarily so; in
practice, most extended UHV/XHV systems tend to exhibit nonuniform pressure distributions, particularly of the
chemically active gases which are readily adsorbed on surfaces in a system at room temperature (hydrogen is a problem
in this respect). The problems of pressure and temperature nonuniformity at low pressures has been considered to some
extent by Da [13], Grigorev [14] and Haefer [14a].

In general, the maximum pumping speed that can be applied to a vacuum system of arbitrary size is approximately
proportional to the surface area; this is true for all types of pump from cryopumps to diffusion pumps. Thus we see from
Eq. (11.2) that when the maximum pumping speed is applied to a system, the ultimate pressure is proportional to the
outgassing rate Q and is independent of the system volume or surface area.

In the following sections the limitations to the measurement and attainment of UHV/XHV are discussed.

11.1
Limits to the Measurement of UHV/XHV

Only the various types of ionization gauge or residual gas analyzer (RGA) are capable of making pressure
measurements in the UHV/XHV region. Several physical and chemical processes limit the measurement of very low
pressures by both hot-cathode and cold-cathode ionization gauges and by RGAs; these processes are examined in the
following sections. Detailed description of total pressure gauges may be found in Chapter 6, and partial pressure
measurements by residual gas analyzers are described in Chapter 7.

< previous page page_628 next page >


< previous page page_629 next page >
Page 629

11.1.1
Residual Currents

The residual current in an ionization gauge is defined as the current to the ion collector if the pressure were suddenly
reduced to zero; in other words, it is that part of the collector current which is independent of gas pressure. The residual
current consisted of two components: (1) the x-ray photocurrent described above and (2) the positive ion current to the
collector caused by electron stimulated desorption (ESD) of gas adsorbed on the anode (grid). Both of these effects are
proportional to the electron current in a hot-cathode gauge.

The electron current to the anode of a cold-cathode gauge is proportional to pressure, so that the x-ray and ESD effects
do not cause a limit to the lowest measurable pressure in these gauges.

The x-ray effect as described above causes a limit to the lowest pressure measurable by all types of hot-cathode gauges.
The design of hot-cathode gauges may be modified to reduce the x-ray effect by:

1. Reducing the surface area of the ion collector to minimize the fraction of x-rays intercepted. This was first introduced
by Bayard and Alpert [10] and leads to an x-ray limit (pressure at which the ion current equals the x-ray photocurrent)
of about 5 × 109 Pa in modern BA gauges. Watanabe [15, 16] has taken this approach to the limit by placing the
collector within a spherical grid and reducing its length to 0.05 mm; an x-ray limit of 2.7 × 1011 Pa was achieved.

2. Placing the collector outside the grid and shadowing it from the source of x-rays. This method, though very effective,
is limited by the reflection of soft x-rays from metal surfaces with a reflection coefficient of some tens of percent. The
extractor gauge [17] and the bent-beam gauge [18] (also known as the Helmer gauge) are examples of this arrangement,
the former having an x-ray limit of 2 × 1010 Pa (8 × 1012 Pa in a modified form [19]) and the latter having an x-ray
limit of about 1012 Pa when modified [20].

3. Suppressing the x-ray photocurrent by placing a negatively biased grid in front of the ion collector to reflect
photoelectrons back to the collector. This method is limited by photoelectrons being generated at the grid and then
attracted to the positive collector. This method was first used by Metson [21] in an early rival to the BA gauge design
and later in the bent-beam gauge and other designs.

4. Compensating for the residual current either by subtracting it electronically from the collector current or by
balancing the photocurrent from the collector with the photocurrent to the collector from a close spaced metal envelope
[22] or other electrode at about the same potential as the collector. By careful choice of the potential and/or spacing of
the other electrode, it is possible to reduce the net photocurrent at the collector to near zero. The principle of
compensation of a "forward" photocurrent by balancing it with a "reverse" photocurrent from another electrode was first
described [23] for a BA gauge in 1963 and has since been used to reduce the x-ray limit in BA gauges. The
counterbalancing of two photocurrents in opposite directions has an advantage over the electronic subtraction method
since it is not affected (to first order) by changes in electron current. There are no published data on the long-term
stability of this arrangement; exposure to chemically active gases may cause a shift in contact potential between the
collector and the other electrode involved and thus change the balance conditions. Unless a method of frequently
checking the residual current is

< previous page page_629 next page >


< previous page page_630 next page >
Page 630

available, such as the modulation method (see page 635), the compensation method of reducing the x-ray limit may be
unreliable when chemically active gases are present.

Positive ions may be desorbed by the impact of electrons on adsorbed layers on the anode (grid) surface of ionization
gauges or the electron collector of the ion source of mass spectrometers; this process is known as electron stimulated
desorption (ESD). The current of ESD ions constitutes part of the residual current because it is not proportional to
pressure. The ESD effect is most troublesome after exposure of the gauge or RGA to chemically active gases (typically
O2, H2O, CO, CO2, CxHy, and halogens or halides), and the ESD current can be reduced by heating the grid or by
electron bombardment (thus operating the gauge or ion source at a high electron current tends to minimize the ESD
current). The ESD effect is the limitation to UHV/XHV measurements which is most difficult to eliminate.

Several methods have been developed to reduce the limit due to the ESD effect; they all depend on distinguishing the
ESD ions from gas-phase ions by the difference in their kinetic energy. This differences arises from the following two
effects:

1. The mean, initial energy of most ESD ions is several electron-volts (e.g., O+ from O2, 6 eV; O+ from CO, 2 eV),
whereas the initial energy of gas-phase ions is usually below 1 eV.

2. The electron space charge in an initially field-free space (such as the grid of an extractor or bent-beam gauge or the
ionizing region of the ion source of a mass spectrometer) can cause a depression of potential of several tens of volts at
electron currents of a few milliamps. Thus the ions from the grid surface have a considerably higher energy than the gas-
phase ions after extraction from the ionizing region.

Three types of hot-cathode gauge are insensitive to the ESD effect. The first is the modulated BayardAlpert gauge
(MBAG), where the ESD ions are separated from the gas-phase ions by the modulation process; this has been confirmed
experimentally and reasonably accurate measurements of oxygen pressure can be made with an MBAG [2426]. The
reason for the lack of modulation of the ESD ions is the low collection efficiency in a BAG for ions with significant
initial kinetic energy (see Redhead [24] and Comsa [27] for a more detailed explanation). The H+ ions produced by
ESD at the grid of an MBAG are modulated like gas-phase ions [28] because of the low kinetic energy of the H+ ions.

The second type of gauge which is insensitive to ESD effects is the extractor gauge, in which the ion collector is
surrounded by a hemispherical reflector electrode at grid potential [29, 30]. This gauge is insensitive to ESD effects
because ions released from the grid surface by electron bombardment have sufficient energy to reach the reflector
(which is at grid potential) while the collection efficiency of these ions on the fine collector wire is small. Gas-phase
ions cannot reach the reflector because of the depression of potential within the grid caused by electron space charge.
Thus to improve the separation of ESD and gas-phase ions the gauge should be operated at high electron current.

The third group of hot-cathode gauges that can separate ESD from gas-phase ions have electrostatic energy analyzers.
This group includes (a) several types of bent-beam

< previous page page_630 next page >


< previous page page_631 next page >
Page 631

Fig. 11.1
Ion collector current versus deflector voltage
for the 180° bent-beam gauge (ion
spectroscopy gauge) for various electron
currents [33], after exposure of the Mo grid
to O2 at about 107 Pa.

gauge [18, 20, 31, 32] with a 90° analyzer and (b) Watanabe's bent-beam gauge [33] with a 180° hemispherical analyzer
(also known as the ion spectroscopy gauge). The gauges with 90° analyzers do not provide complete separation of the
two types of ions. The gauge with the 180° analyzer has sufficient resolution to achieve complete separation; this can be
seen in Fig. 11.1, which shows the collector current as a function of deflection voltage for various values of electron
current after the gauge has been exposed to about 107 Pa of oxygen. It can be seen that complete separation of ESD and
gas-phase ions is achieved while the position of the peak of the gas-phase ions shifts with electron current because of
the space charge effect. The Bessel box gauge uses a "Bessel box" energy analyzer to separate ESD from gas-phase ions
[34a]; this gauge is shown schematically in Fig. 11.2. This type of analyzer is capable of complete separation of the two
types of ions, as shown in Fig. 11.3.

Methods for reducing the errors caused by ESD in the measurement of total pressure in the UHV/XHV range are
reviewed in Watanabe [35].

In an RGA there is no x-ray limit since any x-ray induced photocurrent only adds to the mass independent background
current. However, ESD ions are produced by electron bombardment of material adsorbed on the electron collector of the
ion source; the ESD ions most frequently seen are mass 1 (H+), 10 and 11 (B+ when LaB6 cathodes are used), 16 (O+),
19 (F+), 23 (Na+), 28 (CO+), 35 and 37 (Cl+), and 39 (K+).
< previous page page_631 next page >
< previous page page_632 next page >
Page 632

Fig. 11.2
Schematic diagram of Bessel box gauge showing dimensions (mm) and applied
potentials [34a].
Fig. 11.3
Ion energy spectrum of the Bessel box gauge [34a] at a pressure of
2.3 × 1010 Pa (H2 eq.). The background signal, 63 counts s1,
is the x-ray background equivalent to 3.5 × 1011 Pa.

< previous page page_632 next page >


< previous page page_633 next page >
Page 633

Fig. 11.4
Mass spectrum with a 90° magnetic sector instrument [37] in an
aluminosilicate glass and stainless steel system. The peaks at mass
2 (H2), 4 (He), and 28 (CO) are true gas-phase peaks; the peaks at
mass 1 (H+), result from ESD ions
from the surfaces in the ion source.

With a double-focusing mass spectrometer (i.e., one that focuses ions with both spatial and energy dispersion), such as a
cycloidal EXB mass spectrometer, it is impossible to differentiate between ESD and gas-phase ions. With a single-
focusing instrument (i.e., one that focuses the ions for spatial dispersion only), such as magnetic sector instruments and
quadrupoles, the two types of ions can be readily separated on the basis of their initial energy. With a single focusing
90° magnetic sector instrument the ESD ion peaks are displaced upward on the mass scale [36]; the shift in the position
of the ESD peaks is larger when low accelerating voltages are used. Figure 11.4 shows a mass spectrum taken with a
90° magnetic sector mass spectrometer [37] in an aluminosilicate glass and stainless steel system. The peaks at mass 2
(H2), 4 (He), and 28 (CO) are true gas-phase peaks; and the peaks at mass 1 (H+), 16 1/3 (O+), and 19 1/4 (F+) result
from ESD ions released from the grid of the ion source.

Figure 11.5 shows a mass spectrum with a quadrupole mass spectrometer [38] at a total pressure of 3 × 1010 Pa; the
peaks at mass 16 (O+), 19 (F+), and 35 (Cl+) are ESD ions. The mass-filter action of a quadrupole MS is very sensitive
to the kinetic energy with which the ions are injected into the quadrupole structure. A complete separation of the two
types of ions is possible either by modulating the ion accelerating voltage (typically at 125 Hz) and adjusting the mean
value of the ion accelerating voltage while detecting the ion current with a lock-in amplifier [39] or by placing an
electrostatic energy analyzer (e.g., a Bessel box analyser as used in the Bessel box gauge [34]) between the ion source
and the quadrupole structure [40].

< previous page page_633 next page >


< previous page page_634 next page >
Page 634

Fig. 11.5
Mass spectrum with a quadrupole mass spectrometer [38] at a total pressure of
3 × 1010 Pa. The peaks at mass 16 (O+), 19 (F+), and 35 (Cl+) are ESD ions from
surfaces in the ion source.

To measure very low pressures with a hot-cathode gauge, when the residual current is a non-negligible fraction of the
total collector current, it is necessary to measure the residual current; this may be done in four general ways [41]:

1. By reducing the pressure in the gauge to near zero. This is not often possible but has been achieved for several types
of gauge [41] (BA, suppressor, bent-beam, and extractor).

2. By plotting the collector current as a function of the electron accelerating voltage (the Alpert method [10]). In this
method it is assumed that the collector current versus electron voltage characteristic is the superposition of a "gas
ionization" curve and a "residual current" curve. It is also assumed that the residual current results from x-rays only and
that the x-ray photocurrent is given by

where Ve is the electron accelerating voltage (the grid-filament voltage in a BA gauge) and m is 1.2 to 1.8. At
sufficiently low pressure, when the gas ionization part of the characteristic plotted on a loglog scale can be subtracted, a
linear extrapolation through the operating voltage is taken to be the residual current. This method has several
difficulties: (1) It is a lengthy procedure and the grid is bombarded with electrons up to 1000 volts, which may result in
outgassing or a change in surface conditions; (2) only the x-ray effect is measured, and the presence of an ESD effect
makes the measurement more unreliable; and (3) the photocurrent versus electron energy characteristic on a loglog scale
is not usually a straight line but exhibits changes of slope [41]. Figure 11.6 demonstrates the separation of gas-phase ion

< previous page page_634 next page >


< previous page page_635 next page >
Page 635

Fig. 11.6
Determination of residual current (ir) by the Alpert method
for a modulated extractor gauge [42]. The measured
collector current is ic, and the estimated gas phase ion
current gas is i+.

current from the x-ray photocurrent [42], which is the basis of the Alpert method of determining ir.

3. By measuring the collector current as a function of the pressure measured with a reference gauge having a much
lower residual current limit.

4. By modulation of the ion current. The modulation method was first applied to the BA gauge [43] and has since been
applied to most types of hot-cathode gauge. In general, it consists of modulating the potential of a suitable electrode so
that the current from gas-phase ions is modulated while the modulation of the residual current is insignificant. With the
potential of the modulating electrode at Vm1, the ion collector current is

where i+ is the current of gas-phase ions and ir is the residual current. When the potential of the modulating electrode is
at Vm2, the collector current becomes
< previous page page_635 next page >
< previous page page_636 next page >
Page 636

If the modulation of the residual current is negligible (ε = 0), then i+ and ir can be readily determined:

where ∆i = ic1 ic2. The modulation factor k may be determined by making modulation measurements at pressures
sufficiently high that or by making modulation measurements at two different pressures (a and b) when

ε is not always negligible when making pressure measurements well below the residual current limit. Hobson [44] made
measurements on a modulated BA gauge in a cryopumped system capable of reaching a pressure of 9 × 1013 Pa. After
estimating the quantity εir by modulation measurements at three different pressures, measurements could be made down
to 4 × 1012 Pa even though the residual current was equivalent to a pressure of about 7 × 109 Pa. The value of εir was
so small that Eqs. (11.7) were applicable to measurements above 109 Pa. There have been many investigations of
different modes of modulating a BA gauge, as well as methods of modulating other types of hot-cathode gauges; these
are summarized in Redhead [45]. Figure 11.6 shows the result of determining i+ and ir by the modulation method as a
function of grid-filament voltage in a modulated extractor gauge [42].

The modulation method of measuring ir is very convenient since it gives a rapid method of observing changes in the
current due to ESD ions during the operation of a system; this gives a clear indication of whether the gauge has become
contaminated by a chemically active gas. Modulation also allows accurate pressure measurements of gases which cause
ESD such as oxygen. Singleton [25] has shown that the pressure indicated by a modulated BA gauge was in good
agreement with the oxygen pressure indicated by a mass spectrometer while the unmodulated BA gauge readings were
considerably too high.

11.1.2
Effects at Hot Cathodes

The hot cathode in an ionization gauge or RGA can cause problems in UHV/XHV measurements, which may be
divided into four categories:

1. Increased outgassing resulting from the heating of electrodes and the envelope by radiation from the hot cathode.

2. Evaporation of neutrals and positive ions from the cathode.

3. Chemical reactions of gas molecules at the hot surface resulting in changes in gas composition.

4. Light from an incandescent filament (e.g., a pure tungsten filament) causing photoelectron emission, which can be a
problem if an electron multiplier is used.

These effects all increase rapidly with temperature, so that the most effective remedy is to use a cathode with a low
work function which can provide the required emission

< previous page page_636 next page >


< previous page page_637 next page >
Page 637

Table 11.2. Comparison of Cathodes Used in UHV/XHV


Cathode φ Te R η References
(V) (K) (g·cm2·s1) (A·W1)
W
4.5 2180 6.4 × 1012 1.3 × 104 46
LaB6/C
2.7 1370 ~1015 ~103 47
LaB6/Rh
2.7 1370 1010 ~103 48
ThO2/W, Ta, Mo
2.6 1400 ~2 × 1016 ~103 49, 50
Impregnated
1.51.6 ~1000 <1015 51
Y2O3/W
2.0 Less than 50
ThO2
Y2O3/Re, Nb
2.9
[BaSr]O/Ni
1.01.2 750 ~1017 ~102 52
φ average work function.
Te temperature for emission density of 102 A·cm2.
R rate of evaporation at T = Te.
η emission efficiency at T = Te, amps of emission per watt of heating power.

at a low temperature. The most widely used low work function cathode materials are thoriated tungsten, thin coatings of thorium or yttrium
oxides on tungsten, rhenium, or iridium, and lanthanum hexaboride on a carbon substrate (LaB6 coated on a refractory metal such as tungsten
results in the boron diffusing into the metal and producing an unacceptably high rate of evaporation of lanthanum; a carbon layer prevents the
diffusion of boron). Barium/strontium oxide cathodes are not normally used in UHV/XHV instruments (although they have the lowest work
function of any cathode material) because their emission is easily poisoned by chemically active gases and they emit negative ions in
substantial amounts (Cl, O, and H mainly). Table 11.2 [4652] compares the properties of thermionic cathodes used in UHV/XHV.
Commercially available impregnated cathodes have been used in an ionization gauge [53] at pressure down to 7 × 1011 Pa; electron currents
of about 7 mA could be obtained at a cathode temperature of 1080 K.

Field emission arrays of the Spindt type [54] have been tested as electron sources in a quadrupole RGA [55] and in a 256.4° bent-beam gauge
[56], in the latter case, stable emission of 1 mA was obtained and the outgassing was sufficiently low to achieve a pressure 1011 Pa. This type
of cathode operating at room temperature is potentially very useful in UHV/XHV, provided that problems with outgassing and stability after
exposure to chemically active gases can be minimized in routine use.

The temperature increase caused by radiation from the hot cathode can be reduced by using a metal of high conductivity and low emissivity
[57] for the envelope of the gauge or ion source; an example is aluminum, which has an emissivity of only one-tenth that of stainless steel. An
ion source using a thoria-coated rhenium filament and an aluminum alloy envelope could be turned on and off at a pressure of 3 × 1010 Pa
without any observable change in the measured total pressure [57].

The evaporation of material from a hot cathode can cause a significant contribution to the residual current at sufficiently high cathode
temperatures. Alpert and Buritz first measured this effect in a BA gauge with a pure tungsten filament [58] and showed

< previous page page_637 next page >


< previous page page_638 next page >
Page 638

that the apparent pressure (pev, Pa) was related to the rate of evaporation of tungsten (R, molecules·m2·s1) by the
expression

where A is the surface area of the cathode (m2). The measured limit due to tungsten evaporation is about 1010 Pa in the
BA gauge and about 7 × 1011 Pa in the bent-beam gauge at electron currents of 10 mA. The pure tungsten filament in a
bent-beam gauge has been replaced by a thoria-coated filament resulting in a reduction of a factor of 10 in the pressure
limit due to evaporation [20]. Measurements of the modulation of the ion current due to evaporated material from
tungsten filaments show that it is modulated in the same way as gas-phase ions [59]; thus the modulation method is
incapable of separating this component of the residual current.

The emission of positive ions of impurities from a hot tungsten filament (mainly sodium) can cause false pressure
readings; to prevent this effect the filament should be completely immersed in a positive electric field. The evaporation
of neutrals or positive ions by various types of cathodes is reviewed in Redhead et al. [12, pp. 299303].

A hot cathode can behave like a chemical factory causing substantial changes in gas composition in the UHV/XHV
region; some of the common reactions of molecules with heated tungsten are outlined below:

1. H2 + W → H + H

2. O2 + W + C → O, CO, WxOy

3. H2O + W → 2H + WxOy

4. H2O + W + C → O, CO, WxOy, H, H2

5. CH4 + W → WC(s), H, H2

The atomic species produced in the above reactions are very active, and the following interactions with other unheated
surfaces in the system are possible (where M is a metal):

6. H + H + M → H2

7. H + C + M → CxHy

8. H + C + MxOy→ CO

9. H + MxOy→ H2O

10. O + M → MxOy

The atomic hydrogen formed at a hot cathode (reaction 1 above) is adsorbed on the other surfaces of the system,
resulting in an anomalously high pumping speed for hydrogen when the cathode temperature exceeds 1000 K.
Contaminant speciesin particular, CO (reaction 8 above)are produced by the interaction of atomic hydrogen with the
chamber walls and electrodes. The complex interactions of hydrogen in ion sources with hot filaments at pressures near
109 Pa have been studied in detail [60]. A fuller account of the interaction of gases with hot electrodes may be found in
Redhead et al. [12, pp. 275280].

< previous page page_638 next page >


< previous page page_639 next page >
Page 639

11.1.3
Gauges with Long Electron Paths

Magnetic ionization gauges suitable for UHV/XHV employ electron trapping to obtain very long electron paths in a
small volume and hence increase the sensitivity of the gauge; these include the magnetic cold-cathode gauges based on
a Townsend discharge (magnetron and inverted-magnetron gauges) and the hot-cathode magnetron. The space charge
in the magnetic gauges is electronic at low pressures and behaves like a pure electron plasma which can have many
different modes of oscillatory behavior and anomalous electron transport mechanisms [61]. Although not fully
understood, these effects may lead to anomalous behavior of the discharge such as mode jumping, resulting in abrupt
changes of sensitivity with pressure, low-frequency fluctuations, and nonlinear relations between ion current and
pressure.

To minimize instabilities and nonlinearities in magnetic cold-cathode gauges, it has been suggested that (a) the magnetic
and electric fields should be orthogonal and cylindrically symmetric, (b) all electrodes surrounding the plasma should be
of high conductivity material, and (c) the cathode end-plates should be electrically separated so that only the ion current
to the cylindrical portion of the cathode is measured, thereby permitting the application of a small axial electric field to
help stabilize the plasma. Since high voltages (up to 6 kV) are used in these gauges, it is necessary to have good
insulation of the high-voltage electrode to ensure that leakage currents do not affect low-pressure measurements. These
problems are offset by the advantages of cold-cathode gauges in the UHV/XHV region, which include:

1. Absence of a hot filament, thus avoiding the effects outlined in Section 11.1.2 above.

2. Absence of an x-ray limit; the low pressure limit is usually set by the noise level of the discharge. Pressures as low as
1011 Pa have been measured.

3. Low outgassing (cold-cathode gauges usually are significant pumps).

4. Rugged construction.

At very low pressures the time to establish the discharge may be quite long; this delay is composed of two parts, the
statistical lag (which is the time before a chance event, such as the arrival of a cosmic ray, initiates the discharge) and
the formative lag (which is the time for the discharge to build to its stable condition). The statistical lag can be reduced
to a very short time at all pressures by the addition of a weakly radioactive source to the ionizing volume. Ni63 (a beta
emitter) was first used in the 1960s to trigger cold-cathode gauges and sputter-ion pumps. In 1996 Welch et al. showed
[62] that the use of a 1-µCi Am241 source (a 5.6-MeV alpha emitter used in smoke detectors) reduced the starting time
at 5 × 109 Pa for inverted-magnetron gauges to about 10 minutes; gauges without the Am241 source required as much
as 12 hours for ignition to occur at this pressure. Kendall and Drubetsky [63] used the same type of Am241 source in a
double inverted magnetron gauge (i.e., a single set of electrodes immersed in two opposed magnetic fields giving the
effect of two inverted magnetrons in parallel) and showed that the product of starting time and pressure was
approximately constant; at 7 × 109 Pa the starting time was 22 ± 10% seconds; it was also found that a 0.5-µCi Ni63
source was much

< previous page page_639 next page >


< previous page page_640 next page >
Page 640

less effective than the Am241 source in reducing the starting time. The use of a pulse of photons has also been shown to
be effective in starting cold-cathode gauges at low pressures, either from an external photographic flash unit for a glass
envelope gauge or an internal source for metal-envelope gauges.

Most experimenters have found that both the magnetron gauges (MGs) and inverted-magnetron gauges (IMGs) show a
nonlinear response to pressure in the UHV/XHV range. The ion current is given by i+ = kpn, where n is in the range 1 to
1.6; the value of n is dependent on the anode voltage (Va), the magnetic field (B), and the dimensions of the gauge. An
abrupt change in the value of n as the pressure is changed has been observed by many experimenters for both the
magnetron and inverted-magnetron gauge; the cause of this change of slope is not fully understood.

With the IMG, Feakes and Torney [64] found for p < 105 Pa that n = 1.2 when Va = 6 kV and B = 0.22T. Nichiporovich
[65] has measured an IMG with the anode formed by a tungsten wire spiral which could be heated either to trigger the
gauge or to outgas it by electron bombardmentin this case, n = 1 from 4 × 1010 to 104 Pa at Va = 6 kV and B = 0.2 T.
Mennega and Schaedler [66] found that n = 1.65 for p < 107 Pa with Va = 3.3 kV and B ≈ 0.1 T, a 7-µCi source of Ni63
was used to trigger the gauge at low pressures. Peacock and Peacock [67] measured the response of an IMG with Va = 4
kV: below 107 Pa, n = 1.35 ± 0.03 for B = 0.16 to 0.22 T; for B < 0.1 T, n increased to 1.8 at 0.062 T.

Feakes and Torney [64] have calibrated the MG to 5 × 1011 Pa with Va = 4.8 kV and B = 0.105 T: below 107 Pa, n =
1.35. Nichiporovich and Khanina [68] have calibrated the MG down to 1010 Pa for Va from 3 to 8.5 kV and for B =
0.118 and 0.099 T; at Va = 34 kV and B = 0.099 T, n = 1.04 below 107 Pa. This MG had a spiral wire cathode that
could be heated both to trigger the gauge and to degas it by electron bombardment. Grishin and Grishina [69] have
immersed an MG in a liquid He bath whose temperature was varied by pumping on the liquid He; hydrogen was
introduced into the gauge and the H2 pressure calculated from the known variation of H2 vapor pressure with
temperature; it is claimed that pressures down to 1013 Pa were obtained. With Va = 2.5 kV and B = 0.12 T, n = 1 from 4
× 1013 to 104 Pa; the reported sensitivity of 4 × 103 A Pa1 for H2 is about 10 times higher than that observed by any
other experimenter. In general, the sensitivity of the MG is found to be about 2.5 × 102 A·Torr1, about 10 times that of
the IMG. More details on cold-cathode gauges can be found in Redhead et al. [12, pp. 329 et seq.].

The hot-cathode magnetron gauge (also known as the Lafferty gauge [70]) is operated at a very low electron emission to
avoid the instabilities and nonlinearities observed in cold-cathode magnetic gauges which operate at the maximum
electronic space charge. The sensitivity of the Lafferty gauge with an operating anode current of 2.5 × 109 A (107 A
emission at zero magnetic field) is 6 × 104 A Pa1, which is comparable to a typical BA gauge sensitivity. However, the
x-ray limit of the hot-cathode magnetron gauge is only 3 × 1012 Pa. Modifications to the original Lafferty gauge design,
including the addition of an electron multiplier, have been made by several experimenters; the addition of a suppressor
electrode [71] to reduce the current of photoelectrons leaving the ion-collector is predicted to reduce the x-ray limit to
less than 1014 Pa. The hot-cathode magnetron gauge is potentially capable of measuring lower pressures in the XHV
range than any other design.

< previous page page_640 next page >


< previous page page_641 next page >
Page 641

11.1.4
Comparison of UHV/XHV Gauges

It should be pointed out that many of the hot-cathode gauge designs capable of making total pressure measurements in the XHV range are
complex and expensive; in most cases a quadrupole RGA is more suitable since it provides information on the gas species present and, as
indicated above, with suitable care to reduce ion source outgassing, can measure to about 1014 Pa.

Tables 11.3A and 11.3B [7277] compare the lower limit of pressure and the sensitivity of some selected hot-cathode total pressure gauges
suitable for the UHV/XHV range. It can be seen that the immersed collector gauges (i.e., gauges which have the ion collector immersed in
the ionizing volume such as the BA gauge)

Table 11.3A. Comparison of Some Hot-Cathode Gauges for UHV/XHV: Immersed Collector Gauges
Gauge Collector Diameter Grid-Filament Voltage px Sensitivity (A·Pa1)b References
Type (µm) (V) (Pa)a
BAG
175 105 4 × 109c 7.5 × 104 44
BAG
150 170 5 × 109 4 × 104 72
BAG
50 100 2.4 × 1010 1.3 × 103 73
BAG
4 75 9 × 1011 4 × 104 74
Point
30 110 3 × 1011 1.5 × 103 15, 16
collector (50 long)
a px is the pressure at which the ion current equals the residual current (the x-ray limit). The minimum measurable pressure is
given approximately by pm = 10px when px is not measured and by pm = px/10 when px is measured.
b Sensitivity to nitrogen at electron current of 4 mA.
c For modulated BA gauge pm = 4 × 1011 Pa when ε is measured.

Table 11.3B. Comparison of Some Hot-Cathode Gauges for UHV/XHV: External Collector Gauges
Gauge Electron px Sensitivity References
Emission (Pa)a (A·Pa1)
(A)
Extractor 1.3 × 103 75
(Leybold IE511) 2 × 1010 1 × 104 (N2)
Bent-beam 90° 3 × 103 20
(Helmer gauge) < 2 × 1012 1 × 103 (N2)
Bent-beam 180° 5 × 103 76
(Ion spectroscopy gauge) < 3 × 1013 4 × 104 (N2)
Bent-beam 256.4° 1 × 104 77
< 6 × 1012b 1.8 × 106 (H2)
Bessel box 3 × 105 34
4 × 1011c 2.5 × 107 (Ar)
a px is the pressure at which the ion current equals the residual current (the x-ray limit).
b The lowest measurable pressure is about 4 × 1013 Pa for 1 min measurement.
c The lowest measurable pressure is estimated to be 3 × 1012 Pa or lower.

< previous page page_641 next page >


< previous page page_642 next page >
Page 642

are capable of measurements in the XHV range; for example, the point collector gauge with modulation is capable of
measuring to about 3 × 1012 Pa. All the external collector gauges, except the extractor gauge, have electrostatic energy
analyzers which permit the separation of ESD ions from gas-phase ions on the basis of their initial energy; the 90°
analyzer gives only partial separation. Of this group of gauges, the simplest in construction are the extractor gauge
(which is commercially available) and the Bessel box gauge.

There are only three gauge types using magnetic fields that have found use for UHV/XHV measurements. When
equipped with an electron multipler [78], the hot-cathode magnetron gauge has an estimated lower limit of 4 × 1016 Pa
in an ion counting mode; when provided with a suppressor electrode in front of the collector [31] but no multiplier, the
estimated lower limit is 1014 Pa with a sensitivity of about 7.5 × 104 A Pa1. The cold-cathode magnetron and inverted-
magnetron gauges appear to have a lower limit of measurable pressure of about 1011 Pa which is not set by any x-ray or
ESD effect but rather by the decrease in sensitivity with pressure and the noise level. Both gauges tend to be nonlinear
in their response in the UHV/XHV region, following a power law relation i+ = kpn, where n = 1.0 to 1.5 depending on
the gauge design and the operating conditions.

11.2
Limits to Pumps at UHV/XHV

We next examine the limits to the lowest pressure attainable with those types of vacuum pumps which are suitable for
UHV and XHV. These may be considered in two categories: kinetic pumps, which impart momentum to the gas
molecules and remove them from the vacuum system; and capture pumps, which trap gas molecules by ionic
entrapment, condensation, adsorption, or chemical reaction at a surface within the vacuum system. A more complete
discussion of vacuum pumps can be found in Chapters 4 and 5.

11.2.1
Kinetic Pumps

Two types of kinetic pump are suitable for UHV/XHV: They are the diffusion pump and the turbomolecular pump (or
turbopump). The advantage of kinetic pumps is that they can maintain a low pressure indefinitely and remove large
quantities of gas permanently from the system, whereas most capture pumps require some form of regeneration after
they have pumped a specific quantity of gas and do not remove the gas from the system.

Many of the early experiments to achieve UHV in the 1950s used oil [79] or mercury [80] diffusion pumps with liquid
nitrogen or other types of trap to prevent the backstreaming of pump fluid into the UHV chamber. Diffusion pumps
retain their pumping speed indefinitely as pressure is reduced, and they are available with H2 pumping speeds from 0.1
to over 3.5 m3·s1, but the forepressure must be maintained as low as possible because hydrogen in the fore-vacuum can
backdiffuse to some extent through the vapor jets, thus reducing the effective hydrogen pumping speed. In modern oil
diffusion pumps, about half the molecules crossing the input plane are pumped away, while the other half return to the
UHV chamber. Uncontaminated pressure in the UHV/XHV range cannot be reached with a diffusion pump without

< previous page page_642 next page >


< previous page page_643 next page >
Page 643

extremely careful trapping to prevent any backstreaming of the diffusion pump fluid or the rotary pump oil. Alpert was
the first to demonstrate [79] that an oil diffusion pump with a copper foil trap at room temperature maintained its
pumping speed to below 109 Pa. It has been claimed that with properly trapped oil diffusion pumps it is possible to
achieve 1012 Pa [81]. Diffusion pumps are not now widely used for UHV or XHV, partly as a result of these difficulties
in ensuring dependable, long-term trapping.

Turbomolecular pumps (TMPs) have considerable advantages for UHV/XHV since they can provide a completely oil-
free system, and, unlike diffusion pumps, most designs can be operated in any position. This is possible because the
TMP can be designed to operate against a high backing pressure which can be provided by an oil-free diaphragm pump,
and the rotor can be magnetically suspended to avoid lubricated bearings. The predominant gas at UHV/XHV is
hydrogen; thus special steps must be taken to increase the low compression ratio for hydrogen (about 104) in the TMP,
and the conventional TMP is only capable of achieving about 108 Pa. In 1990 a tandem TMP structure (with magnetic
suspension) consisting of two TMPs on the same shaft was developed [82]; the upper section was designed for a high
compression ratio and the lower portion for high throughput. This pump had a H2 compression ratio of 5 × 108 and a
maximum backing pressure of 0.5 Torr, and it achieved a pressure of 109 Pa. It was shown [83] in 1995 that the eddy-
current losses in the magnetic suspension can cause a temperature increase in the rotor of 60°C, resulting in outgassing.
By coating the lower half of rotor and stator with a high emissivity layer (SiO2 with emissivity of 0.9) the rotor
temperature was reduced to 25°C, allowing a pressure of < 1010 Pa to be obtained. Cho et al. [84] used a TMP with
magnetic suspension backed by a small molecular drag pump followed by a dry diaphragm pump. The stainless steel
system was first baked at 450°C and then opened to dry nitrogen and pumped again with a 200°C bake for 140 hours;
the final pressure was 1010 Pa. Thus it has been clearly demonstrated that pressures in the XHV range can be achieved
with a turbopump alone, provided that steps are taken to achieve a high compression ratio for H2.

11.2.2
Capture Pumps

Capture pumps suitable for the UHV/XHV range include ion, getter, and cryopumps (and combinations thereof), and
they have been reviewed in some detail by Welch [85]. No traps are required and no fluids are introduced into the
system. The main disadvantage of these pumps is that the pumped gas is stored within the vacuum system and is
potentially available for later release into the vacuum.

Ion pumps may be divided into two categories: (a) sputter-ion pumps (SIP), which pump chemically active gases by
sputtering fresh adsorbing surfaces and pump rare gases by ionic entrapment, and (b) getter-ion pumps (GIP), which
contain an evaporable getter (usually titanium). Ion pumps have the advantage at UHV/XHV that the ion current is a
rough measure of pressure. The main problem with an SIP at UHV/XHV is that the pumping speed tends to decrease
with pressure [86] as the discharge intensity (current per unit pressure) decreases. Re-emission of previously pumped
rare gases is also a problem with SIPs. The main purpose of an ion pump in

< previous page page_643 next page >


< previous page page_644 next page >
Page 644

Fig. 11.7
Discharge intensity (A·Pa1) versus pressure for sputter-ion pumps
[89] at three different magnetic fields. Anode voltage 5 kV, anode
diameters 19 and 25 mm. Curves 1,2,3: 19 cells of 33-mm diam.; B=0.1,
0.15, 0.2 T, respectively. Curves 4,5,6: 25 cells of 19-mm
diameter; B=0.1, 0.15, 0.2 T, respectively.

the UHV/XHV range is to provide some pumping speed for the gases which are not pumped by getter pumps (methane
and the rare gases).

It was shown in the early 1960s that some speed was retained by an SIP at very low pressures: In 1961 Klopfer [87]
achieved a pressure of 8 × 1010 Pa with a small SIP, and in 1962 Davis [88] achieved pressures in the 1011 Pa range
with a commercial 5 liter·s1 SIP; in both cases the pressures were measured with mass spectrometers. The rate of
decrease of the pumping speed of an SIP at pressures in the UHV/XHV range is critically dependent on magnetic field,
anode voltage, and dimensions of the pumping cell. Pumping speed is proportional to discharge intensity (i+/p); Fig.
11.7 shows i+/p as a function of pressure in a diode sputter-ion pump for various magnetic fields and anode diameters
[89]. It can be seen that i+/p increases with the larger anode diameter and that increasing the magnetic field tends to
raise i+/p at low pressures. The long time delays in starting a SIP discharge at very low pressures have been reduced by
the addition of a radioactive source (e.g., Ni63, which is a beta emitter) to trigger the discharge [90].

Getter-ion pumps have the advantage that re-emission of previously pumped gas can be prevented by evaporating a
layer of getter material (usually titanium) over the surfaces where gas has been pumped. Getter-ion pumps have not
been widely used in the UHV/XHV range in spite of their advantage in preventing re-emission. Kornelsen [91] has
described a small cold-cathode magnetron pump containing a titanium evaporator which maintains some speed in the
108 Pa range and is capable of pumping 67 Pa·liter of helium and 13 Pa·liter of argon without re-emission. An inverted-
magnetron SIP containing a titanium evaporator has been described by Komiya [92, 93] which maintains some pumping
speed to 1011 Pa; both of the

< previous page page_644 next page >


< previous page page_645 next page >
Page 645

above-mentioned pumps are in effect combined sputter-ion and getter-ion pumps. Hot-cathode GIPs, such as the
orbitron pump which has achieved pressures in the 1010 Pa range [94], have not been widely used at UHV/XHV in
spite of their potential advantage that i+/p does not decrease at low pressures.

Getter pumps, both evaporable and nonevaporable, are widely used in UHV/XHV; however, they do not pump the rare
gases and their speeds for hydrocarbons are negligible. Evaporated titanium films at room or liquid nitrogen
temperatures have been used, in combination with other pumps needed, to remove the rare gases and methane down to
pressures of 1011 Pa and below. For chemically active gases (e.g., H2, H2O, CO, CO2, O2, and N2) a sticking
probability of about 0.5 is achieved on freshly evaporated metal films; thus pumping speeds of about 5 liter·s1·cm2 can
be achieved. Renewal of the film is necessary after about a monolayer of gas has been adsorbed. For a titanium a
capacity of about one pumped molecule to one evaporated titanium atom is possible. Freshly evaporated titanium films
tend to outgas methane unless very pure titanium is used or the film is baked at 100°C for a few hours. Nonevaporable
getters (NEGs) are very effective in the UHV/XHV range because of their high pumping speed to hydrogen. NEGs are
particularly useful in accelerators and storage rings since they can be placed very close to beam lines. For example,
Benvenuti [94a] has achieved pressures as low as 5 × 1012 Pa in a 3 m long section of an accelerator ring using a Zr-V-
Fe NEG strip (43.5 m long), a sputter-ion pump (400 liter·s1), and a titanium sublimation pump cooled with liquid
nitrogen.

Cryopumps used in the UHV/XHV range fall into two categories: (a) cryo-condensation pumps, which physisorb more
than 23 monolayers of gas on a smooth surface of relatively small area, and (b) cryo-sorption pumps, which have a
porous surface of very large effective area with less than 23 monolayer of physisorbed gas.

For the cryo-condensation pump the limiting pressure is the vapor pressure of the adsorbed gas at the temperature of the
surface, and the capacity of the cryo-condensation pump is essentially infinite, at least until the thickness of the layer
presents problems. Condensation coefficients of about 0.5 are typical, and hence maximum pumping speeds of about 5
liter·s1·cm2 are possible; however, maximum speeds cannot be obtained in most situations because the cryosurface
must be protected from room temperature radiation by suitable baffles at an intermediate temperature; the exceptions
are when the complete wall of the vacuum system is a cryosurface (e.g., in the cold bore of an accelerator or storage
ring with superconducting magnets). The low-pressure limit can in principle be made infinitely low since re-emission
can be reduced to near zero by lowering the temperature. Table 11.4 shows the temperatures at which the vapor pressure
of some common gases is equal to 1.3 × 108 and 1.3 × 1011 Pa, respectively [95]; it can be seen that only hydrogen,
neon, and helium may limit the pressure in a cryo-condensation pump at 10 K to more than 1011 Pa.

For the cryo-sorption pump the limiting pressure is established by an appropriate isotherm relating the equilibrium
pressure to the surface coverage; for the UHV/XHV region the DubininRadushkevich isotherm has been found
appropriate for most gases [96]. Theory indicates that at very low pressures and surface coverage the equilibrium
pressure should be proportional to the coverage (Henry's law); so far this has never been observed. A cryo-surface with
less than a monolayer coverage at 4.2 K will pump all gases to extremely low pressures; for example, a system of 1- to
2-liter volume pumped to UHV, sealed off, and then totally immersed in liquid helium, will drop to an estimated
pressure of about 1028 Pa based on an extrapolation of the

< previous page page_645 next page >


< previous page page_646 next page >
Page 646

Table 11.4. Temperatures (K) for Two Selected Vapor Pressures of Common Gases
[95]
Gas Vapor Pressure Vapor Pressure
1.3 × 108 Pa 1.3 × 1011
(1010 Torr) Pa (1013 Torr)
H2
3.21 2.67
He
0.303 0.250
CH4
28.2 24.0
H2O
130.0 113.0
Ne
6.47 5.5
CO
23.7 20.5
N2
21.1 18.1
O2
25.2 21.8
Ar
23.7 20.3
CO2
68.4 59.5
Kr
32.7 27.9
Xe
45.1 38.5
Data mainly from R. E. Honing and H. O. Hook, RCA Rev. 21, 360 (1960).

DubininRadushkevich isotherm [97]. After pumping to UHV, Thompson and Hanrachan [98] immersed a complex
surface analysis system containing a mass spectrometer in liquid helium at 4.2 K; the pressure dropped to 1012 Pa at 30
K as the system cooled and then became immeasurable.

In the preparation of highly porous surfaces for cryo-sorption pumps at UHV/XHV it is necessary to obtain good
thermal contact between the porous material and the underlying cooled surface to achieve rapid thermal equilibrium.
Metal or metal oxide sponges have good thermal contact to the substrate. Hobson [96] has used porous silver bonded to
a metal substrate and achieved an effective area about 1000 times that of a flat surface. Figure 11.8 shows adsorption
isotherms for hydrogen on several different surfaces [99] at 4.3 K; the sudden drops at pressures in the 1013 Torr range
are probably due to the x-ray limit of the extractor gauge used to measure pressure. It can be seen that the porous
aluminum oxide surface (anodized aluminum with an oxide layer having small pores and a thickness of about 40 µm)
has a hydrogen adsorption capacity more than 103 times as great as a smooth stainless steel surface.

11.2.3
Comparisons of Pumps for UHV/XHV
The choice of pump(s) for UHV/XHV depends on the application and the resources available. For relatively small
systems requiring the absence of hydrocarbon contamination a good choice is the magnetically suspended
turbomolecular pump backed with a diaphragm pump. This arrangement is completely oil-free, does not require
periodic regeneration, and can obtain pressures below 1010 Pa. When pumps requiring starting pressures lower than
those of turbopumps are used (e.g., ion pumps), then to obtain hydrocarbon free conditions it is necessary either to use
oil-free backing pumps or sorption backing pumps or to trap the oil vapor from the backing pump

< previous page page_646 next page >


< previous page page_647 next page >
Page 647

Fig. 11.8
Hydrogen adsorption isotherms at 4.3 K on various surfaces [99].

with great care. For extremely low pressures (< 1010 Pa) the use of liquid helium cryopumps, including the possibility
of complete immersion, is a practical solution.

For large systems the choice of pumps is dominated by costs and frequently by the difficulty of access to the vacuum
chamber (as in accelerators and storage rings). Nonevaporable getters combined with sputter-ion or cryopumps are a
frequent choice. Where superconducting magnets are used, it is possible to use the cold bore of the beam line as a cryo-
pump; this has been done in some storage rings and accelerators.

11.3
Leak Detection at UHV/XHV

As was pointed out in connection with Eq. (11.1) above, it is necessary to reduce the leak rate into an UHV or XHV
system to as close to zero as possible; this raises the problem of the measurement of very small leak rates. The general
problems of leak detection have been described in Chapter 8, and we only discuss here the measurement of extremely
small leak rates. The sensitivity of a conventional helium leak detector is about 1012 Pa·m3·s1 (1011 atm·cm3·s1); a
higher sensitivity is desirable to test UHV and especially XHV systems. To achieve this increased sensitivity it is
necessary to accumulate the helium that enters the system through the leak, either in the gas phase or in the adsorbed
phase.

The sensitivity of a helium leak detector can be increased by reducing the helium pumping speed to zero in a closed
system and allowing the helium to accumulate in the gas phase; the pumping speed to active gases can be kept high by
the use of nonevaporable getters or a cryopump operating at about 18 K, neither of which will

< previous page page_647 next page >


< previous page page_648 next page >
Page 648

pump helium. The leak rate is determined, after calibration, by the rate of rise of the helium ion current in the mass
spectrometer (usually a quadrupole). Bergquist and Sasaki [100] have described a helium leak detector system using
nonevaporable getters and the rate of rise method; they claim that this system can routinely detect leaks as small as 1016
Pa·m3·s1 (1015 atm·cm3·s1).

Another leak detection method has been developed [101] in which the inleaking helium is adsorbed on a smooth metal
plate maintained at 9 K or lower in a closed system; the active gases are adsorbed on surfaces at about 20 K which may
have porous carbon or nonevaporable getter material on them. The temperature of the helium adsorbing plate is raised to
1020 K by an internal heater, and the pressure of the desorbed helium is monitored with an RGA. This cycle is repeated
with the plate cooled to 9 K or lower for 1 min and then heated to 1020 K for 1 min. With a 1-min accumulation time at
< 9 K, a sensitivity of 1016 Pa·m3·s1 (1015 atm·cm3·s1) is claimed; with longer accumulation times, a sensitivity of
1019 Pa·m3·s1 (1018 atm·cm3·s1) is claimed.

11.4
Outgassing

To achieve pressures in the UHV, and more particularly the XHV, range it is essential to minimize the outgassing from
the walls of the vacuum chamber and the internal parts, as is evident from Eqs. (11.1) and (11.2). In the early days of
UHV, most systems were of borosilicate glass which could be baked at 450°C to minimize outgassing of water. The
remaining outgassing resulted from the permeation of atmospheric helium through the glass walls; this could be
minimized by the use of aluminosilicate glass, which has a much lower He permeation rate at room temperature.
Pressures as low as 1012 Pa were obtained by Hobson [44] in 1964 with an aluminosilicate glass system and a
cryosorption pump. Modern systems are almost entirely constructed of metal, most commonly stainless steel or an
aluminum alloy, where after processing and bakeout the principal component of the outgassing is usually hydrogen.
This section is concerned with methods to minimize hydrogen outgassing from stainless steel and aluminum alloys and
with the measured values of outgassing rates that result from the best practices to reduce outgassing.

Reduction of outgassing from the internal parts of an UHV/XHV system is normally achieved by degreasing, chemical
cleaning, and vacuum firing of the parts before assembly followed by electron bombardment, I2R heating, or radio-
frequency (rf) heating in the vacuum system. The in situ degassing process is particularly important for any parts that
operate at elevated temperaturesfor example, electrodes of hot-cathode devices.

The SI unit for outgassing rate per unit area is Pa·m·s1 (i.e., Pa·m3·s1·m2); another widely used unit is
Torr·liter·s1·cm2, which equals 1.33 × 103 Pa·m·s1.

11.4.1
Reduction of Outgassing Rates

Reduction of the outgassing rates of the metals used in the construction of vacuum chambers is essential if pressures in
the UHV/XHV range are to be achieved efficiently. Outgassing rates from metals can be reduced by:

< previous page page_648 next page >


< previous page page_649 next page >
Page 649

1. High-temperature vacuum firing to reduce the amount of dissolved hydrogen (as high as 1000°C for stainless steel).

2. Baking of the vacuum system to remove water (150450°C). It is important to achieve as uniform a temperature
distribution as possible while baking.

3. Degreasing and chemical cleaning.

4. Surface treatment to reduce surface roughness and remove porous oxides. This includes electropolishing, surface
machining under special conditions, and glow-discharge cleaning.

5. Surface treatments to create oxide or other films on the surface that act as a barrier to diffusion of hydrogen from the
bulk.

6. Deposition of films of low hydrogen permeability on a metal substrate (e.g., titanium nitride or boron nitride).

7. Choice of metal with a low solubility for hydrogen.

Examples of the effects of these treatments are given below.

Outgassing rates in the first few hours of pumpdown of an unbaked system are dominated by water. Barton and Govier
[102] have compared the outgassing rate of an unbaked stainless steel specimen (a) after vapor degreasing and (b) after
vapor degreasing, vacuum baking, and exposure to atmosphere. The short-term outgassing rates achieved by the two
methods were within a few percent of one another. Mathewson [103] has compared three methods of cleaning
aluminum alloy vacuum chambers: vapor degreasing only, weak alkaline etch, and strong alkaline etch. The strong
alkaline etch was found to be best. Suemitsu [104] has prepared mirror-polished surfaces of aluminum alloys by
electrochemical buffing techniques and has shown that measured outgassing rates at 10 h are proportional to the surface
roughness factor. Dylla [105] has compared the effects of several different surface treatments on the measured
outgassing rates of unbaked stainless steel (type 304) and aluminum alloys (type 6061/63) at relatively short pumping
time. The outgassing rates at 100 min did not differ by more than a factor of 9, lying between 8 × 107 and 7 × 106
Pa·m·s1. The short-term outgassing rate was not much affected by surface roughness. Li and Dylla [106] have measured
the outgassing rate of an electropolished stainless steel surface after it was exposed to a glow discharge following
venting for 1 h to the atmosphere. The most effective treatment was a helium discharge for 2 h at a dose of more than
0.2 C·cm2·h which reduced the outgassing rate by a factor 13.

The ability to reach UHV without having to bake has considerable advantage for many applications. Kato [107] has
achieved 1010 Pa without baking in a stainless steel system after electrochemical buffing. Pressures of 3 × 107 Pa were
reached in 4 days, and the turbopumps and the titanium sublimation pump (TSP) only were then baked at 120350°C for
72 h; the pressure then dropped on day 15 to 8 × 1010 Pa with the TSP cooled in liquid nitrogen. A pressure of 3 × 109
Pa could be maintained without cooling or operating the TSP.

Systems capable of very fast pumpdown to the UHV range without bakeout have great potential for semiconductor
processing. Miki [108] has developed an aluminum system machined in a controlled atmosphere of oxygen and argon
(the EX process), which was first pumped to 109 Pa after a bakeout and then vented to dry nitrogen. The system was
then repumped by a turbopump and a cryopump reaching 106 Pa in

< previous page page_649 next page >


< previous page page_650 next page >
Page 650

2.5 min and 108 Pa in 170 min. The system had been vented to dry nitrogen with only a few ppb of water or other
contaminant gases; the nitrogen was introduced over a moisture trap and a NEG trap.

Baking the vacuum system rapidly removes the adsorbed water leaving hydrogen as the main constituent of the
outgassing in metal systems, which then limits the lowest pressure attainable in the UHV/XHV range. Thus when
system bakeout is possible, the main effort should then be directed to the reduction of the hydrogen outgassing rate at
room temperature.

In a classic paper, Calder and Lewin [109] studied both theoretically and experimentally the reduction of outgassing
from stainless steel. After baking for about 25 h at 300°C the measured outgassing rates were about 4 × 109 Pa·m·s1
(H2). After vacuum furnacing for 3 h at 1000°C followed by baking in the vacuum system for 25 h at 360°C, the
outgassing rate had dropped to well below 1011 Pa·m·s1 (H2), which was the lower limit of measurement. Strausser
[110] has compared the outgassing rates of stainless steel after (a) vapor degreasing only and (b) vapor degreasing,
chemical cleaning, and glass-bead blasting. Minimum outgassing rates for H2 of about 5 × 109 Pa·m·s1 were observed,
with process (b) above being the better of the two by less than a factor of 5.

The effects of oxidation of a stainless steel surface (200°C for 3 h in atmospheric air) on outgassing rates has been
studied [111]; the reduction in outgassing rate of the oxidized surface compared to the well-outgassed unoxidized
surface was a factor of 4. Ohmi [112] has developed a method to produce a continuous film of chromium oxide (Cr2O3)
on stainless steel which passivates the surface and prevents corrosion by HCl gas; this process may prove effective in
reducing outgassing in UHV/XHV chambers.

A high-current Penning discharge in oxygen has been used to clean the surface of aluminum (99.99%) vacuum
chambers used in a storage ring [113, 114]. Hydrocarbon contaminants were removed by sputtering, and the
photodesorption yield was reduced by a factor of 10.

Coating the interior surface of a vacuum chamber with a thin film of material having a low permeation rate for
hydrogen is a promising way to reduce outgassing rates. Titanium nitride has been found to be a suitable material [115],
when deposited on stainless steel by ion plating in a film 12 µm thick; the outgassing rates with and without the coating
were 1.7 × 1011 and 2.2 × 109 Pa·m·s1 (H2), respectively, a reduction of about two orders of magnitude.

Ishimaru [116] has described an aluminum alloy system where the chamber was machined in an atmosphere of dry
oxygen and argon (the EX process). After bakeout at 150°C for 24 h the outgassing rate was about 1010 Pa·m·s1; an
ultimate pressure of 4 × 1011 Pa was claimed.

Copper has been used in the manufacture of vacuum tubes since the 1930s but has not been widely used for UHV/XHV
systems. Watanabe [117] has measured outgassing by the throughput method from an electropolished, oxygen-free-high-
conductivity copper chamber; the lowest outgassing rate observed after bakeout at 300°C was 5 × 1012 Pa·m·s1 (H2).

Titanium coated with titanium nitride, deposited by the hollow cathode discharge ion plating method, has also been
examined [118] as a low outgassing material; preliminary results indicate that this combination has an outgassing rate
very similar to well degassed stainless steel.

< previous page page_650 next page >


< previous page page_651 next page >
Page 651

Table 11.5. Some of the Lowest Measured Outgassing Rates (H2)


Material Measurement Methoda Surface Treatment Hydrogen Outgassing Rate Gauges in References
(Pa·m·s1) Test Chamberb
St. St.c (321Ti) GA 1 None 119
2 × 1010
St. St. (316L) T 2 RGA + IG 120
1.3 × 1010
St. St. (316LN) GA 3 RGA + IG 121
2 × 1010d
St. St. (316L) GA 4 SRG 122
2.3 × 1011

Al (99.99%) T 5 IG 116
~1011
Al alloy (A6063-T6) T 5 IG 116
~1010

Cu (OFHC) T 1 None 119


2 × 1013
Cu (OFHC) T 6 IG 117
5.4 × 1012

TiN on St. St.e St. St.e T 7 None 115


1.7 × 1011
a GA, gas accumulation method; T, throughput method.
b Type of pressure sensor(s) used in test chamber: IG, ionization gauge; RGA, residual gas analyzer; SRG, spinning rotor gauge.
c St. St., stainless steel.
d Outgassing rates for other gases; ≤ 1 × 1015 (CH4), ≤ 3 × 1014 (H2O), ≤ 4 × 1013 (CO), ≤ 2 × 1014 (CO2).
e TiN film 12 µm thick.
Note: All surface treatments included an initial chemical cleaning, alkali detergent wash, and/or vapor degrease.
1. Glass bead blast + 250°C bake for 3 days + 400°C in UHV furnace for 3 days + 250°C bake for 1 day.
2. 950°C in vacuum furnace for 2 h + 200°C bake for 2 days.
3. 150°C bake for 10 days.
4. Vacuum fired + 250°C bake for 3 days.
5. Machined in O2 + Ar atmosphere (EX process), or machined under ethanol (EL process) + 150°C bake for 24 h.
6. 300°C bake for 3 days.
7. Electropolish + TiN ion-plated + 150°C bake for 2 days.

Table 11.5 [119122] lists some examples of the lowest measured outgassing rates for hydrogen from stainless steel, aluminum and its
alloys, and copper surfaces. Two methods of measurement have been used: (1) the gas accumulation method (also called the pressure rise
method) where the test chamber is sealed off after pumping to the ultimate pressure and the pressure rise with time is measured and (2) the
throughput method where the test chamber is pumped through an orifice of known conductance and the pressure drop across the orifice is
measured. The presence of a hot-cathode RGA or ion gauge in the test chamber can cause extra outgassing; thus the measurements made
with a spinning rotor gauge or with no gauge in the test chamber should be more reliable. The low outgassing rates observed with copper
[117] and with stainless steel after rigorous vacuum furnacing [119] are noteworthy.

< previous page page_651 next page >


< previous page page_652 next page >
Page 652

11.5
UHV/XHV Hardware

The expanding use of UHV technology in the 1960s depended on the development of demountable metal systems which
required reliable all-metal valves, motion transmitters, demountable flanges, and other demountable components which
were free from any leakage. Previous versions of these devices used at higher pressures, which relied on greases or
elastomers to achieve a vacuum-tight seal, were not applicable at UHV where high bakeout temperatures were used and
the vapors from the greases and elastomers were not tolerable. The various methods that had been developed prior to the
start of UHV technology for making permanent joints between metal parts and between metals and ceramics or glasses
were in general suitable for use at UHV/XHV, provided that sufficient care was taken to produce clean, leak-free joints.
Tungsten-inert gas welding and electron beam welding have proved very effective for UHV/XHV systems. The
historical development of vacuum hardware has been described by Singleton [123], and the various types of sealing
methods and the theory of metal gasket seals have been described by Roth [124].

The early types of metal gasket seals using aluminum or gold wire and flat flanges were compatible with UHV but were
not very reliable. More consistency was achieved in 1954 with copper gaskets compressed between knife edges of a
harder metal [125]. The development of the captured copper gasket seal by Wheeler [126] in 1962 led to a thoroughly
reliable seal for UHV/XHV and is now almost universally used in a wide variety of commercially available sizes.

Alpert described the first all-metal valve in 1951 that was both bakeable and suitable for UHV use [127]; this valve had
a closed conductance of 1013 m3·s1. Bills and Allen [128] developed in 1955 an improved valve with a constrained
silver insert as the sealing element which achieved a closed conductance of 1017 m3·s1. Another type of seal for larger
UHV valves, using a hard knife-edge driven into a copper gasket by pneumatic pressure, was developed [129] in 1962.

A wide variety of demountably flanged components compatible with UHV/XHV are now available such as windows,
electrical and liquid feed-throughs, and motion transmitters (using both metal bellows and magnetic coupling) from
which complex UHV/XHV systems can be built.

References

1. Dictionary of Terms for Vacuum Science and Technology. (M. S. Kaminsky and J. M. Lafferty, eds.), AIP Press,
Woodbury, NY, 1980.

2. Glossary of Terms Used in Vacuum Technology. Part I. Terms of General Application, BS 2951. British Standards
Institution, London, 1969.

3. Vakuumtechnik Bennenungen und Definitionen, DIN 28400, (1972).

4. I. Langmuir, Phys. Rev. 2, 450 (1913).

5. S. Dushman, Phys. Rev. 5, 212 (1915).

6. P. A. Anderson, Phys. Rev. 57, 122 (1940).

7. W. B. Nottingham, Phys. Rev. 55, 203 (1939).

8. P. A. Redhead, in Vacuum Science and Technology: Pioneers of the 20th Century (P. A. Redhead, ed.), p. 133. AIP
Press, Woodbury, NY, 1993.

< previous page page_652 next page >


< previous page page_653 next page >
Page 653

9. W. B. Nottingham, Conference on Physical Electronics. M.I.T., Cambridge, MA, 1947.

10. R. T. Bayard and D. Alpert, Rev. Sci. Instrum. 32, 571 (1950).

11. G. F. Weston, Ultrahigh Vacuum Practice. Butterworth, London, 1985.

12. P. A. Redhead, J. P. Hobson, and E. V. Kornelsen, The Physical Basis of Ultrahigh Vacuum. AIP Press, Woodbury,
NY, 1993.

13. D. Da and X. Da, J. Vac. Sci. Technol. A 5, 2484 (1987).

14. G. I. Grigorev, K. K. Tzatzov, and I. N. Martev, Proc. Int. Vac. Congr., 8th, Cannes, 1980, Vol. 2, p. 291.

14a. R. A. Haefer, Vak. Tech. 16, 149 and 185 (1967).

15. F. Watanabe, J. Vac. Sci. Technol. A. 5, 242 (1987).

16. F. Watanabe, J. Vac. Soc. Jpn. 34, 17 (1991).

17. P. A. Redhead, J. Vac. Sci. Technol. 3, 173 (1966).

18. J. C. Helmer and W. H. Hayward, Rev. Sci. Instrum. 37, 1652 (1966).

19. Z. Q. Tang, H. Y. Chen, and Z. H. Lu, J. Vac. Sci. Technol. A 5, 2384 (1987).

20. C. Benvenuti and M. Hauer, Proc. Int. Vac. Congr., 8th, Cannes, 1980, Vol. 2, p. 199.

21. G. H. Metson, Br. J. Appl. Phys. 2, 672 (1951).

22. F. Watanabe, J. Vac. Soc. Jpn. 34, 25 (1991).

23. W. H. Hayward, R. L. Jepsen, and P. A. Redhead, Trans. Natl. Vac. Symp. 10, 228 (1963).

24. P. A. Redhead, Vacuum 13, 253 (1963).

25. J. H. Singleton, J. Chem. Phys. 45, 2819 (1966).

26. J. H. Singleton, J. Vac. Sci. Technol. 4, 103 (1967).

27. G. Comsa, Vide 147, 137 (1970).

28. J. P. Hobson and J. Earnshaw, Proc. Int. Vac. Congr., 4th, London, 1968, p. 619.

29. P. A. Redhead, J. Vac. Sci. Technol. 3, 173 (1966).

30. U. Beeck and G. Reich, J. Vac. Sci. Technol. 9, 126 (1972).

31. C. D. Suen, J. Z. Chen, and Y. H. Kuo, J. Vac. Sci. Technol. A 8, 3888 (1990).

32. W. Li and Z. Zhang, J. Vac. Sci. Technol. 5, 2447 (1987).

33. F. Watanabe, J. Vac. Sci. Technol, A 10, 3333 (1992).

34. H. Akimichi, T. Tanaka, K. Takeuchi, Y. Tuzi, and I. Arakawa, Vacuum 46, 749 (1995).
34a. H. Akimichi, T. Arai, K. Takeuchi, Y. Tuzi, and I. Arakawa, J. Vac. Sci. Technol. A 15, 753 (1997).

35. F. Watanabe, J. Vac. Sci. Technol. A 11, 1620 (1993).

36. J. L. Robins, Can. J. Phys. 42, 886 (1964).

37. J. P. Hobson and J. W. Earnshaw, Can. J. Phys. 46, 2517 (1968).

38. W. K. Huber, N. Müller, and G. Rettinghaus, Vacuum 41, 2103 (1990).

39. F. Watanabe and H. Ishimaru, J. Vac. Sci. Technol. A 3, 2192 (1985).

40. S. Watanabe, M. Aono, and S. Kato, Vacuum 47, 587 (1996).

41. P. A. Redhead, J. Vac. Sci. Technol. A 10, 2665 (1992).

42. L. G. Pittaway, Philips Res. Rep. 29, 283 (1974).

43. P. A. Redhead, Rev. Sci. Instrum. 31, 343 (1960).

44. J. P. Hobson, J. Vac. Sci. Technol. 1, 1 (1964).

45. P. A. Redhead, J. Vac. Sci. Technol. 7, 182 (1970).

46. H. A. Jones and I. Langmuir, Gen. Electr. Rev. 30, 310 (1927).

< previous page page_653 next page >


< previous page page_654 next page >
Page 654

47. J. M. Lafferty, J. Appl. Phys. 22, 299 (1951).

48. J. D. Buckingham, Br. J. Appl. Phys. 16, 1821 (1965).

49. E. Shapiro, J. Am. Chem. Soc. 74, 5233 (1952).

50. G. A. Haas, Methods of Experimental Physics, vol. 4A, p. 1. Academic Press, New York, 1967.

51. S. Yamamoto, I. Watanabe, S. Sasaki and T. Yagushi, Vacuum, 41, 1759 (1990).

52. S. Wagener, The Oxide-Coated Cathode. Chapman & Hall, London, 1951.

53. T. Satoh and C. Oshima, J. Vac. Soc. Jpn. 37, 77 (1994).

54. I. Brodie and C. A. Spindt, Adv. Electron. Electron Phys. 83, 1 (1992).

55. N. Ogiwara, Y. Miyo, and Y. Ueda, Vacuum 47, 575 (1996).

56. C. Oshima, T. Satoh, and A. Otuka, Vacuum 44, 595 (1993).

57. F. Watanabe, J. Vac. Sci. Technol. A 8, 3890 (1990).

58. D. Alpert and R. S. Buritz, J. Appl. Phys. 25, 202 (1954).

59. P. Repa, Vacuum 36, 559 (1986).

60. J. P. Hobson and J. W. Earnshaw, Can. J. Phys. 46, 2517 (1968).

61. P. A. Redhead, Vacuum 38, 901 (1988).

62. K. M. Welch, L. A. Smart, and R. J. Todd, J. Vac. Sci. Technol. A 14, 1288 (1996).

63. B. R. F. Kendall and E. Drubetsky, J. Vac. Sci. Technol. A 14, 1292 (1996).

64. F. Feakes and F. L. Torney, Trans. Natl. Vac. Symp. 10, 257 (1963).

65. G. A. Nichiporovich, Instrum. Exp. Tech. (Engl. Transl.) 1440 (1966).

66. H. Mennega and W. Schaedler, Proc. Int. Vac. Congr., 4th, London, 1968, p. 656.

67. R. N. Peacock and N. T. Peacock, J. Vac. Sci. Technol. A 8, 2806 (1990).

68. G. A. Nichiporovich and I. F. Khanina, Proc. Int. Vac. Congr., 4th, London, 1968, p. 666.

69. S. F. Grishin and E. Y. Grishina, Sov. Phys.Tech. Phys. (Engl. Transl.) 16, 1199 (1972).

70. J. M. Lafferty, J. Appl. Phys. 32, 414 (1961).

71. J. Z. Chen, C. D. Suen, and Y. H. Kuo, J. Vac. Sci. Technol. A 5, 2373 (1987).

72. B. Angerth, Vacuum 22, 7 (1972).

73. J.-M. Laurent, C. Benvenuti, and C. Scalambri, Proc. Int. Vac. Congr., 7th, Vienna, 1977, Vol. 1, p. 133.
74. A. van Oostrom, J. Sci. Instrum. 44, 927 (1967).

75. F. Watanabe, J. Vac. Sci. Technol. A 9, 2744 (1991).

76. F. Watanabe, J. Vac. Sci. Technol. A 11, 1620 (1993).

77. C. Oshima and A. Otuka, J. Vac. Sci. Technol. A 12, 3233 (1994).

78. J. M. Lafferty, Proc. Int. Vac. Congr. 4th, London, 1968, p. 647.

79. D. Alpert, Vac. Symp. Trans., Committee on Vacuum Technique, Boston, 1956, p. 69.

80. A. Venema, Vacuum 9, 54 (1959).

81. D. J. Santeler, J. Vac. Sci. Technol. 8, 299 (1971).

82. H. Enosawa, C. Urano, T. Kawashima, and M. Yamamoto, J. Vac. Sci. Technol. A 8, 2768 (1990).

83. H. Ishimaru and H. Hisamatsu, J. Vac. Sci. Technol. A 12, 1695 (1995).

84. B. Cho, S. Lee, and S. Chung, J. Vac. Sci. Technol. A 13, 2228 (1995).

85. K. M. Welch, Capture Pumping Technology. Pergamon, Oxford, 1991.

86. T. S. Chou and D. McCafferty, J. Vac. Sci. Technol. 18, 1148 (1981).

< previous page page_654 next page >


< previous page page_655 next page >
Page 655

87. A. Klopfer, Vak.-Tech. 10, 113 (1961).

88. W. D. Davis, Trans. Nat Vac. Symp. 9, 363 (1962).

89. T. Koizumi, Y. Kawasaki, T. Kurita, M. Kondou, and Y. Hayashi, J. Vac. Soc. Jpn. 34, 505 (1991).

90. C. Hayashi, J. Vac. Sci. Technol. 3, 286 (1966).

91. E. V. Kornelsen, Trans. Natl. Vac. Symp. 7, 29 (1960).

92. S. Komiya, N. Takahashi, and M. Xu, J. Vac. Soc. Jpn. 36, 125 (1995).

93. M. Xu, N. Takahashi, and S. Komiya, J. Vac. Soc. Jpn. 36, 271 (1995).

94. D. G. Bills, J. Vac. Sci. Technol. 10, 65 (1973).

94a. C. Benuenuti and P. Chiggiato, Vacuum, 44, 511 (1993).

95. J. P. Hobson, Proc. Int. Vac. Congr. 9, 35 (1983).

96. J. P. Hobson, J. Phys. Chem. 73, 2720 (1969).

97. J. P. Hobson, J. Vac. Sci. Technol. 3, 281 (1966).

98. W. Thompson and S. Hanrachan, J. Vac. Sci. Technol. 14, 643 (1977).

99. M. G. Rao, P. Kneisel, and J. Susta, Cryogenics 34, 377 (1994).

100. L. E. Bergquist and Y. T. Sasaki, J. Vac. Sci. Technol. A 10, 2650 (1992).

101. G. R. Mynemi, U.S. Pat. 5,343,740 (1994).

102. B. S. Barton and B. P. Govier, Vacuum 20, 1 (1970).

103. A. G. Mathewson, J.-P. Bacher, K. Rooth, R. S. Calder, G. Dominichini, A. Grillot, N. Hilleret, D. Latorre, F.
LeNormand, and W. Unterlerchner, J. Vac. Sci. Technol. A 7, 77 (1989).

104. M. Suemitsu, H. Shimoyamada, N. Miyamoto, T. Tokai, Y. Moriya, H. Ikeda, and H. Yokoyama, J. Vac. Sci.
Technol. A 10, 570 (1992).

105. H. F. Dylla, D. M. Manos, and P. H. LaMarche, J. Vac. Sci. Technol. A 11, 2623 (1993).

106. M. Li and H. F. Dylla, J. Vac. Sci. Technol. A 13, 571 (1995).

107. S. Kato, M. Aono, K. Sato, and Y. Baba, J. Vac. Sci. Technol. A 8, 2860 (1990).

108. M. Miki, K. Itoh, N. Enomoto, and H. Ishimaru, J. Vac. Sci. Technol. A 12, 1760 (1994).

109. R. Calder and G. Lewin, Br. J. Appl. Phys. 18, 1459 (1967).

110. Y. E. Strausser, Proc. Int. Vac. Congr., 4th, London 1968, p. 469.

111. K. Odaka and S. Ueda, J. Vac. Sci. Technol. A 13, 520 (1995).
112. T. Ohmi, A. Ohki, M. Nakamura, K. Kawada, T. Watanabe, Y. Nakagawa, S. Miyoshi, S. Takahashi, and M. S. K.
Chen, J. Electrochem. Soc. 140, 1691 (1993).

113. M. Saitoh, K. Kanazawa, T. Momose, and H. Ishimaru, J. Vac. Sci. Technol. A 11, 2518 (1993).

114. N. Ota, M. Saitoh, K. Kanazawa, T. Momose, and H. Ishimaru, J. Vac. Sci. Technol. A 12, 826 (1994).

115. K. Saito, S. Inayoshi, Y. Ikeda, Y. Yang, and S. Tsukahara, J. Vac. Sci. Technol. A 13, 556 (1995).

116. H. Ishimaru, J. Vac. Sci. Technol. A 7, 2439 (1989).

117. F. Watanabe, Y. Koyatsu, and H. Miki, J. Vac. Sci. Technol. A 13, 2587 (1995).

118. M. Minato and Y. Itoh, J. Vac. Sci. Technol. A 13, 540 (1995).

119. G. Messer and N. Treitz, Proc. Int. Vac. Congr., 7th, Vienna, 1977, p. 223.

120. H. C. Hseuh and X. Cui, J. Vac. Sci. Technol. A 7, 2418 (1989).

121. J. R. J. Bennett and R. Elsey, Vacuum 44, 647 (1993).

< previous page page_655 next page >


< previous page page_656 next page >
Page 656

122. K. Jousten, J. Vac. Soc. Jpn. 37, 678 (1994).

123. J. H. Singleton, J. Vac. Sci. Technol. A 2, 126 (1984).

124. A. Roth, Vacuum Sealing Methods. Pergamon, Oxford, 1966, and AIP Press, Woodbury, NY, 1994; Vacuum
Technology, 2nd ed. North-Holland Publ., Amsterdam, 1982.

125. H. H. Pattee, Rev. Sci. Instrum. 25, 1132 (1954).

126. W. R. Wheeler, Trans. Natl. Vac. Symp. 9, 159 (1962).

127. D. Alpert, Rev. Sci. Instrum. 22, 536 (1951).

128. D. G. Bills and F. G. Allen, Rev. Sci. Instrum. 26, 654 (1955).

129. R. J. Connor, R. S. Buritz, and T. von Zweek, Trans. Natl. Vac. Symp. 8, 151 (1962).

< previous page page_656 next page >


< previous page page_657 next page >
Page 657

12
Calibration and Standards

Karl Jousten

The pressure p in an enclosed gaseous system is basically defined as the force dF per area dA exerted by the gas in the
chamber:

The SI unit of pressure is the pascal (Pa),

which will be used throughout this chapter. The pressures in vacuum are divided into various ranges by the scale shown
in Table 12.1 [1].

Since in the vacuum range <10 Pa for practical areas of a few square centimeters the force becomes very small and
difficult to measure, in most vacuum gauges other pressure-dependent physical properties of the gas such as viscosity,
thermal conductivity, or particle density are being used to indicate pressure. For these gauges it is generally not possible
to state an equation by which the pressure can be calculated from well-known parameters of the gauge. In these cases a
calibration of the gauges is needed, by which the gauge's indication is related to the true pressure applied to it. This is
also true for gauges which still use the mechanical force of the gas, but the magnitude of the force or the area cannot be
quantified absolutely. Practically, two calibration methods exist:

Foundations of Vacuum Science and Technology, Edited by James M. Lafferty.


ISBN 0-471-17593-5  1998John Wiley & Sons, Inc.

< previous page page_657 next page >


< previous page page_658 next page >
Page 658

Table 12.1. Classification of Vacuum Ranges from AVS [1]


Vacuum Range Pressure Range (in pascals)
(Degree of Rarefication)
Low (rough)
3.3 × 103 to 105
Medium
101 to 3.3 × 103
High
104 to 101
Veryhigh
107 to 104
Ultrahigh
1010 to 107
Extreme ultrahigh
< 1010

1. The more accurate calibration method is to subject the gauge head to a pressure generated by a primary standard. A
primary standard is defined [2] as standard that is designated or widely acknowledged as having the highest
metrological qualities and whose value is accepted without reference to other standards of the same quantity. So
vacuum primary standards provide the most accurately known pressures in vacuum, and they deduce pressure directly
from other units as mass, length, and time according to Definition 12.1 and basic laws of physics. Not all vacuum
ranges can be covered by a single primary standard, so that various methods of generating pressures and also various
realizations of these methods exist (Section 12.1). The most accurate and well-known vacuum primary standards have
been established by the vacuum sections of some National Metrological Institutes, among them the NIST (National
Institute of Standards and Technology) in the United States, PTB (Physikalisch-Technische Bundesanstalt) in Germany,
NPL (National Physical Laboratory) in England, IMGC (Istituto di Metrologia ''G. Colonnetti") in Italy, and the NPL of
India.

2. With less accuracy the reading of a gauge may be compared to the reading of a stable gauge, called secondary
standard or reference standard, which was calibrated on a primary standard. This method is known as the comparison
method (Section 12.2).

In both calibration methods, it is a crucial point that the generated pressure in the primary standard or the pressure
applied to the reference gauge is identical to the pressure applied to the gauge to be calibrated. This means that in all
calibration chambers the pressure has to be as homogeneous as possible, and this is one of the main criteria for a good
design of a calibration system for pressures in vacuum.

According to the kinetic theory of gases the momentum of gas molecules depends on temperature and so does the
pressure. Although this is well known, it is an often forgotten fact that, when pressure values are being compared, also
the temperature of the gas of which the pressure was measured has to be named. It is mostly tacitly assumed that
pressure was measured at room temperature whatever this exactly was. In this chapter, pressure will always be referred
to a temperature of 23°C, if not specified otherwise.

12.1
Primary Standards

Above about 10 Pa, it is possible to measure directly the force exerted by the gas, for example in a liquid manometer.

< previous page page_658 next page >


< previous page page_659 next page >
Page 659

Below 10 Pa, the idea of all vacuum primary standards used today is the following: The gas is measured at a pressure as
high as possible and then expanded in a calculable manner to the desired pressure. The initial pressure of the gas should
be as high as compatible with the expansion method, because the accuracy of pressure measurement increases with
pressure, but low enough that the gas can still be considered as ideal (practically less than 300 kPa; see Section 12.1.2).
The expansion is carried out mainly in two ways: statically or continuously.

In the static expansion, the BoyleMariotte law

is used and the gas is expanded from a small volume into a large volume. Static means that no pump is used for the
expansion itself. Under isothermal conditions, the ratio of the small volume to the arithmetic sum of the small and large
volume gives the factor by which the initial pressure is being reduced (see Section 12.1.2).

In the continuous expansion, the gas at the high pressure is continuously pumped through two orifices, the first being of
very small conductance, the second of much larger conductance. The ratio of the two conductances gives the factor by
which the initial pressure is being reduced (Section 12.1.3).

There is another expansion method, which is rarely used: the expansion by molecular beam. Herein, a molecular beam
is created by the exit orifice of an effusion cell and by the entrance orifice of a calibration chamber. From the solid
angle of the entrance orifice of the calibration chamber with respect to the exit orifice of the effusion cell and the
Maxwellian velocity distribution, the ratio can be calculated by which the initial pressure in the effusion cell is being
reduced (Section 12.1.4).

In this sense, realizing the vacuum pressure scale is like stepping down a ladder. Each step refers to an expansion from a
higher pressure to a lower one, and this lower one may serve again as a starting point for a further expansion, and so on.

12.1.1
Liquid Manometers and Piston Gauges

The initial step of this ladder and the starting point for all vacuum primary standards are the instruments which directly
measure the pressure as force per unit area at a specified temperature. These are the liquid manometers (described in
Section 6.2) and the piston gauges (described in Section 6.4).

Figure 12.1 shows the principle of a gas-operated piston gauge (also called pressure balance). A cylindrical piston
rotates in a closely fitted circular cylinder. The pressure at the base of the piston is defined as the ratio of the total
downward force on the piston to the effective area of the gauge when floating at its operation level.

As with many vacuum primary standards, a piston gauge is basically a pressure generator, not a vacuum meter. And to
be even more specific, it generates a pressure difference between the piston's bottom and the volume above the
pistoncylinder assembly. If this volume is enclosed by a bell jar and evacuated to <0.1 Pa, the piston gauge can be used
as an absolute pressure generator. Typical pressure ranges of piston gauges used for vacuum gauge calibrations are from
2 kPa to 300 kPa.

The gap between piston and cylinder is typically a few tenths of a micron at a cross section of 10 cm2 of the piston. It is
clear from this that manufacturing a

< previous page page_659 next page >


< previous page page_660 next page >
Page 660

Fig. 12.1
The principle of piston gauges. The
weight force acting on the base of a
rotating piston in a closely fitted cylinder
defines the pressure underneath.

pistoncylinder assembly requires elaborated precision techniques and materials (often tungsten carbide and in recent
times also ceramics).

The accuracy with which the pressure can be generated depends on the accuracy with which measurements of both
force and effective area can be performed. If magnetic forces between piston and cylinder can be excluded and gas
friction effects are considered [3], the effective area can be determined by dimensional measurements which agree well
with values received from pressure comparisons with liquid manometers [37]. Comparisons with liquid manometers are
more convenient and more precise in determining effective piston areas, because disturbing effects such as inclination
of the piston axis against the cylinder axis, unroundness of the parts, frictions effects, and so on, are calibrated into the
effective area. Also, possible dependencies of the effective area on gas species, operation height, and pressure [3, 8] can
be easily determined by comparisons with liquid manometers.

The force can be calculated from the product of the sum of all masses mi acting gravimetrically on the piston's bottom
and the local acceleration constant glocal due to gravity, reduced by the buoyancy of the masses in the surrounding gas.
If there is vacuum in the bell jar surrounding the masses, the latter influence can be neglected, and the generated
pressure is calculated from

where A0 is the effective area of the piston, (αpist + αcyl) is the sum of the linear thermal expansion coefficients of
piston and cylinder, T0 is the temperature at which A0 was determined, the actual temperature and pres the residual
pressure in the bell jar.
< previous page page_660 next page >
< previous page page_661 next page >
Page 661

The uncertainties (approximately 95% confidence internal) of the pressure generated by a piston gauge are given by a
pressure-independent term and a pressure-dependent one with typical values of

The piston gauges are used to calibrate vacuum gauges from 2 kPa up to 100 kPa or for generating accurate initial
pressures for static expansion systems.

12.1.2
Static Expansion

The principle of the static expansion is as follows (Fig. 12.2): Gas contained in a small volume, whose pressure p1 was
accurately measured (or generated with a piston gauge), is expanded into a large evacuated volume V2 by opening a
valve in between. To apply the BoyleMariotte law [Eq. (12.3)] it must be assumed that the expansion is isothermal,
meaning that both vessels and the intermediate tubing are at equal temperatures and the gas temperature does not
change during the expansion. Since these assumptions are usually not realized [9], we will apply the more general gas
law pV/T = const. If V1 is at temperature T1 and V2 at T2, we receive the following for the pressure p2 in V2 after
expansion:
Fig. 12.2
The principle of static expansion used for generating low pressures in the vacuum regime.
Gas contained in a small, enclosed volume V1 is expanded into another enclosed volume
V2 which is much larger Under isothermal conditions the pressure in V1 is
reduced by V1/(V1+V2).

< previous page page_661 next page >


< previous page page_662 next page >
Page 662

The ratio V1/V1 + V2 is called the expansion ratio and is a constant for each expansion system. Hence, to determine p2,
the expansion ratio has to be known, and the initial pressure p1 and the temperatures T1 and T2 of the two vessels have
to be measured. This method was first used by Knudsen in 1910 [10].

The initial pressure, which ranges typically from 1 kPa to 200 kPa, is measured either with a liquid manometer or a
piston gauge, or a suitable secondary standard like a quartz Bourdon spiral manometer (QBS; see Section 6.5). It is a
good idea to design the expansion ratio such that an expansion from the highest measurable pressure p1 gives a p2
larger than the lowest pressure measurable with the gauge for initial pressure measurement. By this overlap it is possible
to check if pressures generated by the expansion are consistent with the pressures directly measured with the initial
pressure measurement device.

The gases which can be used for the static expansion method have to meet two requirements: Their virial coefficients
should be small (no significant deviations from ideal gas law) and they should not be adsorbed on the walls of the
vacuum vessel. In practice, this limits the application of the static expansion method to the rare gases, N2 and CH4
[1114]. H2 can be used for p2 > 102 Pa, while oxygen or oxygen containing diatomic gases are hardly manageable in
static expansion systems [14].

Three methods are applied to determine the expansion ratio: the gravimetric technique and the constant pressure
technique, with which the absolute volumes are determined, and expansion techniques, with which the expansion ratio
is determined.

1. Gravimetric Technique. For volumes >0.1 liter, this is the most accurate method. Highly distilled water (4×) or other
liquids like alcohol or mercury [15] is filled into the volume and its weight is measured. To remove all air from the
water and air bubbles sticking to the wall, the volume has to be evacuated. With fiber optics it may be checked that all
bubbles have been removed. When the water temperature including eventual temperature gradients in larger vessels are
accurately measured and buoyancy corrections are applied, volumes can be determined with relative uncertainties in the
low 104 range.

2. Constant Pressure Technique (Fig. 12.3). Volumes <0.11 are best measured with the constant pressure technique
[16]. A known variable volume (for example, a precision piston driven into a vacuum cylinder) and a stable vacuum
gauge are connected to the small volume to be measured. After an expansion into the small evacuated volume to be
measured, the variable volume is adjusted such that the gauge shows the same reading as before the expansion. For that
the valve to the small volume has to be carefully (slowly) closed again in order to remove the effect of the moving
volume in the valve. The size of the change of the known variable volume is identical to the size of the volume to be
measured.

3. Expansion Technique. In this technique the pressures before and after the expansion are used for the determination of
the expansion ratio. This has the advantage that the measurement is carried out in situ. The point is that the pressure
ratio has to be measured with high accuracy. For this it is possible to use two calibrated gauges (or primary standards
like liquid manometers) or to use a single uncalibrated gauge with a strictly linear pressure response.

< previous page page_662 next page >


< previous page page_663 next page >
Page 663

Fig. 12.3
The measurement of small volumes enclosed between two valves: A certain
pressure is set in the volume of the cylinder and the differential gauge with the
bypass valve open. After closing the bypass valve, a piston of known volume
is driven into the cylinder, until the pressure on the gauge shows the same
reading both before and after expansion of the gas into the volume to be measured
(L1L2) times the piston cross section area gives the unknown volume.

Elliott and Clapham [17] used a piston gauge to generate the initial pressure and a calibrated QBS for the pressure
measurement of the expanded gas. The idea of their experiment was to repeat the expansion until a sufficiently high gas
pressure is built up in the large vessel (no pumping between the expansions) which could accurately be measured by the
QBS as secondary standard. When taking corrections due to temperature gradients between the vessels, temperature
drift, and departures from the ideal gas law into account, it is possible to determine expansion ratios of 1 : 100 with a
relative (one standard) uncertainty of close to 1 × 103. The determination of even lower expansion ratios down to 1 :
3000 with this method has been reported [18].

Berman and Fremerey [19] used a single expansion where both the initial pressure p1 and the final pressure p2 was
measured with the same spinning rotor gauge. Since the pressure response of the spinning rotor gauge is not strictly
linear above 102 Pa, they applied a linearization procedure; and by repeating the expansion over a range of initial
pressures and for three gases, they were able to measure an expansion ratio of 1 : 250. This is about the upper limit
which can be determined with sufficient accuracy with this expansion technique.

The disadvantage of the expansion techniques is that the volumes are not determined to their absolute values, which is
inconvenient when volumes have to be added
< previous page page_663 next page >
< previous page page_664 next page >
Page 664

Fig. 12.4
A multiple-stage static expansion standard in use at the National Physical Laboratory
(NPL), United Kingdom. (By courtesy of NPL, UK.)

to the system (for example, the volumes of gauges to be calibrated), and the expansion ratio must be corrected for these.
Therefore the small volume is often determined with one of the first two methods, or the expansion ratio is deliberately
changed by inserting an object with known volume into the system [19] to determine the absolute values.

Although it is possible to determine expansion ratios up to the order of 1 : 10,000 with the gravimetric and constant
pressure technique, in practice mostly expansion ratios up to 1 : 250 are realized, since the expansion technique is much
easier to use than the absolute techniques. Therefore, it is necessary to use multiple expansions to generate lower
pressures. This can be accomplished by designing a system either with a series of small and large volumes, called stages
[11, 20] (Fig. 12.4), or with repeated expansions with the large volume being evacuated after each expansion [21], or
with a combination of both. It is advisable that there be an overlap in the calibration pressures generated by a higher-
valued expansion ratio with a small initial pressure p1 and the ones generated by the next lower-valued expansion ratio
with a high initial pressure p1 in order to check whether the determined expansion ratios are consistent with each other.

In one of the NPL primary standards [11, 20] a five-stage expansion system (Fig. 12.4) is used, while in one of the two
PTB static expansion standards [2224] only a two-stage system is used (Fig. 12.5), so that for lower pressures repeated
expansions have to be applied.

The lower limit of calibration pressures, which can be generated in a static expansion system, is determined by the
lowest residual pressures in the chambers, the outgassing rate of the inner walls of the system, and adsorption effects,
which become significant below 104 Pa even for the gases mentioned above, as investigated

< previous page page_664 next page >


< previous page page_665 next page >
Page 665

Fig. 12.5
A static expansion system used at the Physikalisch-Technische
Bundesanstalt (PTB), Germany.
Reprinted from [22], Copyright 1990, with kind permission
from Elsevier Science Ltd., Kithington, UK.

by Messer [14]. Typically, a lowest calibration pressure, 106 Pa, can be achieved in a well-baked-out stainless steel
system with a specific outgassing rate of <1011 Pa · liter/s · cm2.

The uncertainties of the generated pressures in static expansion systems are mainly determined by the uncertainties of
the expansion ratios and by the uncertainties in the determination of the temperatures. The uncertainties of the system at
PTB are shown in Fig. 12.6.

No national or international written standards have so far been implemented for the realization of this sort of expansion
technique for vacuum primary standards.

12.1.3
Continuous Expansion

In the continuous expansion method, also called "orifice flow method" or "dynamic method" [11,25,26], the calibration
gas is continuously expanded from a volume at high pressure through at least two conductances into the vacuum pump.
In between the two conductances is the calibration chamber (Fig. 12.7). When there are no sinks or sources of gas
between the two conductances, the equation of continuity holds and the net flow through the two orifices must be equal
(isothermal conditions):
< previous page page_665 next page >
< previous page page_666 next page >
Page 666

Fig. 12.6
Uncertainties of the generated pressures in the static
expansion system shown in Fig. 12.5. The main contributions
to the total uncertainty (2σ) are also shown: 1, uncertainty
of expansion ratio; 2, uncertainty of temperature due to
gradients; 3, uncertainty of gas temperature due to expansion
itself; 4, uncertainty of initial pressure generated by a piston
gauge; 5, uncertainty of temperature measurement due to
drift; 6, uncertainty of pressure due to outgassing.

Fig. 12.7
The principle of continuous expansion for generating pressures in the
high and very high vacuum regime. The gas which is initially on
pressure p1 is continously expanded through two conductances of
largely different size and pumped away. The calibration
pressure p2 is given by p2=p1·C1/C2.
where (p1p2) is the pressure drop across conductance C1, and (p2p3) is the pressure drop across C2. If p2 is much
smaller than p1, and p3 is much smaller than p2, we can approximate

where p2 is the calibration pressure. This is the basic equation of the continuous expansion method. Instead of two
volumes as in the static expansion, two orifices of

< previous page page_666 next page >


< previous page page_667 next page >
Page 667

Fig. 12.8
One of the flowmeters used at the Physikalisch-Technische Bundesanstalt.
Flowmeters generate precisely known low flow rates for vacuum metrology.
Reprinted from [31], Copyright 1993, with kind permission from Elsevier
Science Ltd., Kithington, UK.

largely different size are used to reduce the initial pressure p1. The typical range of values for C1 is 106105 liter·s1, and
that for C2 is 10100 liter·s1.

The gas flow with the throughput

at a specified temperature is usually generated in a separate device, called a flowmeter, which is actually also a flow
generator. Herein, p1 (10 Pa to 100 kPa) is measured by a suitable secondary standard such as a QBS or a capacitance
diaphragm gauge (CDG). The conductance is mostly determined by measuring the volumetric speed ∆V/∆t to keep p1
constant (constant-pressure flowmeter). A review of several types of flowmeters has been given by Peggs [27].

A few written standards exist for calibration systems based on this method [2830]. As an example of a flowmeter, the
scheme of the constant pressure flowmeter used at PTB [31] is shown in Fig. 12.8. The gas reservoir consists of two
sections, the reference volume and the working volume, which are separated by valve V3 and the diaphragm of a
sensitive CDG (A). The working volume is enclosed additionally by the valves V4 (constant conductance C1) and V5
(variable conductance C1), and it includes a displacement bellows as a variable volume. The reference volume ends at
valve V2.

In the beginning of the operation, both volumes are equally pressurized with V3 open. After closing V3, gas will leak
out of the working volume through the open valve V4 (or V5) and a pressure difference across CDG (A) will develop.
This signal is used to drive the bellows such that the pressure remains constant inside the working volume. In principle,
the bellows can be driven continuously with constant speed or from time to time so that the pressure will slightly vary
(±5×104) in a sawtooth-like manner. The measured volume speed of the calibrated bellows is identical to the
conductance C1 of valve V4 (or V5) at this pressure.
< previous page page_667 next page >
< previous page page_668 next page >
Page 668

Fig. 12.9
One of the calibration systems based on the
continuous expansion method used at the National
Institute of Standards and Technology, USA. The
chamber above the orifice plate is the calibration
chamber, Reprinted from [40], Copyright 1988, with
kind permission from AIP, Woodbury, New York.

A similar flowmeter used at NIST has been described in McCulloh et al. [32].

From the flowmeter, the gas is injected into the calibration system. This consists of (1) the inlet system, (2) the
calibration chamber, (3) the pump orifice with conductance C2, and (4) the pumping system (Fig. 12.9).

1. Inlet System. The inlet system has to be designed such that any beam effect of the molecular flow through the tubing
from the flowmeter is transformed to a Maxwellian distribution inside the calibration system. This can be accomplished
by building a separate chamber with a small orifice to the calibration chamber. A simpler solution is to form the inlet
tubing such that the outcoming particle will hit a portion of the wall of the calibration chamber far away from the pump
orifice, or hit a baffle plate.

2. Calibration Chamber. The Maxwellian distribution in the calibration chamber is disturbed by both the gas inlet and
the pump orifice. To minimize these disturbances, the following precautions have to be taken concerning (a) the shape
and the size of the chamber and (b) the position and orientation of the flanges.
• The ideal shape of the chamber is the sphere, but a more practical approach is the cylinder with equal length and
diameter. These shapes minimize pressure gradients inside the vessel and also minimize the ratio of inner surface
to volume.

< previous page page_668 next page >


< previous page page_669 next page >
Page 669

According to the German Industry Standard DIN 28416 [29], the surface of the largest inscribed sphere in the
chamber should be at least a factor of 1000 larger than the orifice area. This is to ensure that the pressure within the
chamber is homogeneous to within about 1×103. However, by means of Monte Carlo simulations of molecular
scattering in the chamber, it is possible to estimate the particle density distribution in the chamber and to determine
corrections for the pressure at the position of the flanges used for the gauges under calibration. If this is done, or
measurements of the density distribution inside the chamber are carried out [33, 34], orifice areas larger than
1/1000 of the largest inscribed sphere may be used. The volume of the chamber should be much larger (a factor of
50) than the sum of all volumes added by the gauges to be calibrated.

• The position and orientation of the flanges where the gauges to be calibrated are mounted on (called test flanges),
have to meet the following requirements:

No direct interaction between attached gauges should be possible.

Only a small fraction, if any, of incoming particles should be allowed to hit a test flange directly.

No direct path of molecules from the test gauge to the exit orifice should be possible.

3. Pump Orifice. Since only the molecular flow of particles through orifices can be mathematically described with high
accuracy, the pump orifice and the input flow rate should be sized such that only pressures in the molecular regime will
develop in the calibration chamber. The conductance C2 of a circular pump orifice with open area A and a thin edge is
given by

where is the mean velocity of the gas particles. Particle scattering on the edge is considered by the correction factor
K1, which is [35]

where d is the diameter of the circular orifice and t is its much smaller thickness (t<<d). It is advantageous to install the
orifice on the surface of the largest inscribed sphere of both the calibration vessel and the pump chamber to ensure free
molecular flow through the orifice in both directions. If this is not the case, a further correction has to be applied for the
determination of C2.

Intermolecular scattering is accounted for by the Knudsen correction factor K2:

where a is a constant and the mean free path length of the molecules. The value of a lies between 0.07 and 0.125
according to older literature [26,36,37]; experimental data recently obtained at NIST, however, suggest a = 0.08 [38].

4. Pumping System. One of the assumptions in deducing Eq. (12.8) is that the pressure p3 downstream the pump orifice
in the pumping system can be neglected

< previous page page_669 next page >


< previous page page_670 next page >
Page 670

against the pressure p2 in the calibration chamber, or in practice should be p3<104 p2. This is true if a cryocondensation
pump directly behind the orifice is used [34]. If turbomolecular pumps or diffusion pumps are being used, the
conductance C2 has to be corrected to

and p3/p2 is in the range from 103 for heavier gases up to 0.2 for light gases. To measure this ratio [3942], which is
sometimes called the backstreaming factor, it is necessary to have built another chamber behind the pump orifice of
about the same size as the calibration chamber in order to receive a more uniform pressure distribution in this pump
chamber [43].

Since isothermal conditions rarely exist in practical systems, corrections for the temperature differences between the gas
in the flowmeter and in the calibration chamber are necessary. DIN 28416 recommends to refer both temperatures to a
common temperature of 23°C and calculates different correction terms for gauges which respond directly to gas density
and gauges which respond directly to pressure.

In continuous expansion systems a stationary equilibrium is established: Depending on the gas species, after a certain
equilibrium time the number of particles leaving the calibration chamber will balance the number entering it. Therefore,
ad- and desorption effects are only important until this equilibrium is reached, and more gases are applicable with this
method than with the static expansion. Even water vapor has been used with at least some success [44].

The lower limit of calibration pressure in continuous expansion systems is determined by the residual pressure in the
calibration chamber and by the lowest flow rate which can be generated in the flowmeter with sufficient accuracy. The
typical value of the lower limit is 1×107 Pa. It can be further reduced if flow divider or multiple stages techniques are
used [45,46].

The upper limit of calibration pressure in the continuous expansion is determined by the transition from the molecular
flow to the viscous flow regime. In principle, this can be shifted to quite high values as recently demonstrated at NIST
by using small capillaries of a few-micron diameters of a channel plate structure for the realization of C2; the typical
value for the upper limit, however, is 102 Pa.

The uncertainties which can be achieved in a primary standard of this type are illustrated in Fig. 12.10.

12.1.4
Molecular Beam Expansion

For the last step of the pressure realization toward XHV, it is no longer possible to deduce the pressure directly from the
SI units. Instead, an ionization gauge, calibrated with one of the two methods described above, is used for the
measurement of the initial pressure before expansion.

The method of attenuating the pressure is the following (Fig. 12.11): A stable, but adjustable, pressure of calibration gas
is established in an effusion cell measured with the calibrated ionization guage. Gas effused out of the circular orifice
(open area A1) of this cell is mostly pumped away by a cryo-panel. Only a small fraction of it is formed into a molecular
beam by another circular orifice (A2), which is the entrance to the calibration chamber. The gauge head to be calibrated
is mounted in a position so that

< previous page page_670 next page >


< previous page page_671 next page >
Page 671

Fig. 12.10
Total uncertainty (2σ) of nitrogen pressures generated in one
of the continuous expansion systems in use at the Physikalisch-
echnische Bundesanstalt. Also shown is the uncertainty of the
generated flow rate of the flowmeter, which dominates the
uncertainty below about 105 Pa. The major contributions
above 105 Pa are uncertainties due to pressure inhomogeneities
in the calibration vessel and due to temperature measurement.

Fig. 12.11
Principle of molecular beam expansion. The pressure is reduced by forming a
molecular beam defined by the exit orifice of an effusion cell and the entrance orifice of
a calibration chamber. The remainder of the molecules effused is pumped by a cryo-panel.
it is not hit by the molecular beam directly. Both orifices have to be small enough to ensure a homogeneous Maxwellian
distribution inside both chambers.

In equilibrium, the same number of particles will enter and leave the calibration chamber per unit time:

< previous page page_671 next page >


< previous page page_672 next page >
Page 672

The number of particles per unit time leaving the effusion cell orifice in solid angle ∆Ω at angle θ normal to the orifice
plane is

where n1 is the particle density and 1 the mean velocity in the effusion cell. Molecular beam systems are designed for
θ = 0°. The solid angle of the entrance orifice of the calibration chamber is

where l is the distance between A1 and A2 and we obtain for the gas flow rate into the calibration chamber

while the flow out is

where n2 is the particle density and the mean velocity in the calibration chamber. Equalizing these two equations
according to Eq. (12.14) and assuming results in

For typical values l = 100 mm and A1 = π·1 mm2, it is n1/n2 = 10,000 and so p1/p2, since we have assumed equal
temperatures in the effusion and calibration cell.

In practice, the ratio p1/p2 is determined experimentally by using two calibrated ionization gauges. This is possible,
because in the molecular regime the ratio is pressure independent, so that for the lower p2 a value > 107 Pa can be
chosen, for which the gauge can be calibrated. This experimental approach in determining the pressure attenuation has
the advantage that a number of idealizations in the theory, such as thin orifices, well-defined temperatures, and pressure
homogeneity, do not have to be realized in the apparatus.

Since practically the conductance of A2 has to be of the order of 1 liter·s1 (nitrogen), outgassing of the ionization gauge
becomes a problem for the residual pressure in the calibration chamber. Typical outgassing rates of ionization gauges
are in the range 109106 Pa·liter·s1. So even for ionization gauges with very low outgassing rates, the residual pressure
is limited to about 109 Pa nitrogen equivalent. This problem can be overcome if the calibration chamber is layered by
titanium, which effectively pumps the outgassing gas species but does not pump the rare gases,

< previous page page_672 next page >


< previous page page_673 next page >
Page 673

Fig. 12.12
The calibration system based on molecular beam expansion used at the Physikalisch-
Technische Bundesanstalt. The effusion cell is called the Knudsen cell herein. (From [48].)

which are used as calibration gas. It has also been shown [47], however, that, when the residual signal is carefully
determined and subtracted from the signal at calibration, the gauge can be calibrated for partial pressures even lower
than the residual pressure.

The calibration system designed by Grosse and Messer [48] is illustrated in Fig. 12.12. The recently estimated relative
uncertainty (2σ) of the calibration pressure between 1010 Pa and 107 Pa is somewhat below 7% [24].

12.2
Calibration by the Comparison Method

If the vacuum gauge to be calibrated is not exposed to a calculable pressure generated in a primary standard, but its
reading is compared to the pressure indicated by a so-called reference gauge (secondary standard; see next section), it is
called calibration by comparison.

In the viscous flow regime (> 100 Pa) the setup for the calibration by comparison is relatively simple, since disturbances
such as residual pressure, outgassing, adsorption, desorption, temperature gradients, and so on, are usually of no
significance. The
< previous page page_673 next page >
< previous page page_674 next page >
Page 674

reference gauge and the device to be calibrated will be connected to the same vacuum pipe, evacuated by a simple
pumping system (in many cases a roughing pump will do the job); and via a gas inlet system a constant pressure will be
applied to both gauges. Usually, it is sufficient to adjust a static equilibrium in the calibration system. The inner
diameters of the connecting tubes just have to be large enough that no pressure gradients in the tubes with eventually
large relaxation times will occur. If the calibration system has to be pumped continuously for a stationary equilibrium,
care has to be taken that there are no pressure differences between the gauges. This can be accomplished by a
symmetrical arrangement of the test gauge and the secondary standard to gas sinks and sources.

In the molecular and transition flow regime, where the pressure at specific points in the calibration system may be
dependent on the local temperature, a much greater effort has to be made. In the molecular flow regime (< 0.1 Pa) the
calibration pressure is usually established by a continuous, stationary flow. The DIN standard 28418 [49] and a draft of
ISO [50] exist as guideline for the realization of this method. According to them, the calibration system must fulfill the
following requirements, which are very similar to the requirements for a continuous expansion primary standard.

• The volume of the calibration chamber should be large compared to the sum of all volumes added by the devices to be
calibrated. DIN 28418 requires a factor of 50.

• To minimize the effects of outgassing and sorption, the ratio of inner surface to volume should be as small as possible.
Ideally, the shape of the calibration chamber would be a sphere. A reasonable choice is also a cylinder with equal length
and diameter. DIN 28418 requires that the ratio not exceed the one of a straight cylinder whose length is four times its
diameter, which may be a length value too high.

• The gas inlet system must be designed such that a significant fraction of the incoming gas particles will neither hit
directly any gauge nor hit the orifice of its tubing.

• Similarly the pump system should be designed such that there is no direct path of molecules from the vacuum gauge
into the pump system possible (or only with a very small probability).

• The ports for the secondary standards and the test gauges should be arranged symmetrically to the pump outlet and gas
inlet. Also, they should be arranged such that no direct interaction between each two gauges can occur. Interactions may
be thermal radiation, direct gas or charged particle flow from one gauge to the other, electromagnetic disturbances, and
so on.

• The tubing to the gauges must be of a conductance that will not build up significant pressure gradients due to
outgassing or pumping of the gauge.

• DIN 28418 requires that the base pressure in the chamber not exceed 2% of the lowest calibration pressure. Therefore,
considering the expected outgassing from the walls (after a bake-out, if necessary) and vacuum gauges, the effective
pumping speed to the calibration chamber must be sized accordingly. Besides this, it is always a good idea to subtract
any residual signal from the measured signal to minimize disturbances from that.

< previous page page_674 next page >


< previous page page_675 next page >
Page 675

Fig. 12.13
A commercially available calibration system (Balzers PSK 110)
based on the comparison method. This system also has the
option to apply the continuous expansion method according to
Eq. (12.7) with the use of the conductances C1 and C2. The
CDG and the SRG on the left-hand side are used to measure
the pressure before C1.
(By courtesy of Balzers Instruments, 1996.)

In commercial calibration systems the calibration procedure is preferably automized. The pressure is steadily increased
in the system, and the reference and the test gauge readings are taken at several set points per decade. Since in these
cases the pressure is changing in time, it is important that, when taking the readings at the set points, the pressure is
constant to approximately within ± 0.5% for 510 min to ensure that a stationary equilibrium is established in both the
vessel and the gauge heads, and no errors occur due to a dynamic measurement.

The temperature of the calibration chamber should be measured on several points to get a reasonable estimate of the
mean temperature, because temperature gradients, especially in stainless steel chambers, are always present.

A gas impurity of 0.1% at the inlet into the calibration chamber (impurities of the gas inlet system as well as the
impurities in the reservoir have to be considered) is acceptable.
A typical comparison calibration system as commercially available is shown in Fig. 12.13.

We should note at the end of the sections describing the systems for calibration that also other methods or significant
variations of it existfor example, so-called

< previous page page_675 next page >


< previous page page_676 next page >
Page 676

pressure-time dependence methods as reviewed by Kuz'min [51], which, however, are not used by the widely known
National Metrological Institutes and which usually can be considered of less accuracy than the described ones.

12.3
Calibration of Vacuum Gauges and Mass Spectrometers

In this section we will describe specific points to consider, when vacuum gauges which are suitable as secondary
standards and mass spectrometers are calibrated. A secondary standard is defined as a standard whose value is assigned
by comparison with a primary standard. After calibration they are often used as reference standards in comparison
calibration systems.

The most accurate and stable gauges in the specific vacuum range are used as secondary standards. Nowadays these are
the Quartz Bourdon spiral manometer (QBS) from 1 kPa up to 100 kPa, the capacitance diaphragm gauges (CDGs)
from 0.1 Pa up to 1 kPa (sometimes also up to 100 kPa), the spinning rotor gauges (SRGs) from about 103 Pa to 1 Pa,
and below 103 Pa the ionization gauges (IGs).

In Chapter 5 the physical basis and technical details of these instruments have been described, and we will therefore
restrict ourselves to the points important for the calibration and stability of them.

The quartz Bourdon spiral manometer is quite straightforward to calibrate in the pressure range 1 kPa up to 100 kPa
(see previous section), so that we will start our notes with the CDGs.

12.3.1
Capacitance Diaphragm Gauges

A CDG sensor [52] is a two-sided cavity, separated by a nonporous diaphragm. The side which is exposed to the test
gas is called the test side, the other the reference side, of the CDG. The diaphragm forms the movable side of a variable
capacitor. The change of capacitance will give an indication of the pressure exerted on the diaphragm. The capacitance
is measured by means of an oscillator whose frequency is converted to a voltage, amplified, and sometimes corrected
for predictable system errors (linearization).

The zero stability and the accuracy of the instrument is improved when the diaphragm serves as electrode for two
capacitors, which change their capacitance differently under pressure. Since the dielectric constant is dependent on gas
species and pressure, today's gauges operate in a ''bull's-eye" configuration; that is, both capacitors plates are located on
the reference side of the CDG, one electrode being opposite to the center of the diaphragm, the other being ring-shaped
and concentric to it. This also has the advantage that corrosive or dirty gases do not deteriorate the sensing capacitors
and the lifetime of the sensor head is increased.

Two configurations are commonly used (Fig. 12.14):

1. Absolute-Type: The sensor is permanently vacuum-sealed on the reference side (usually with a getter material inside),
and the inlet port (test port) on the other side allows for absolute pressure measurement. The vacuum on the sealed side
has to be smaller than the lowest detectable pressure, by a factor of 102 or less.

< previous page page_676 next page >


< previous page page_677 next page >
Page 677

Fig. 12.14
The two important types of capacitance diaphragm
gauges. In the absolute type the reference chamber is
evacuated and only absolute pressure can be measured;
with the differential type, also differential pressure
can be measured.

2. Differential-Type: The sensor has two inlet ports, a test port and a reference port, to allow for differential pressure
measurements between the ports. If the pressure on the reference port (called line pressure) is low enough, typically 105
Pa, the test port may be used for absolute pressure measurement also.
Since the modulus of elasticity is temperature-dependent and also the geometry is changing with temperature (it should
be noted here that CDGs sense diaphragm deflections as low as 0.5 nm), CDGs respond to temperature changes in the
head. To minimize the effect on pressure indication, CDGs are available with temperature control units, which stabilize
the sensor head temperature to typically 45°C to within 0.1°C or, for high-accuracy models, even to within 0.02°C. As a
consequence, as soon as the pressure drops below the viscous flow regime at about 100 Pa, the

< previous page page_677 next page >


< previous page page_678 next page >
Page 678

higher temperature (T2) volume inside the head will be at higher pressure than the lower temperature (T1) volume of the
chamber which is usually at room temperature. In the molecular flow regime the ratio p(T2)/p(T1) will be a constant:

In the transition range between viscous and molecular flow the ratio has to be calibrated, and it depends not only on
pressure and on the geometry and the surface of the tubing but also on the gas species (See Section 1.10. on thermal
transpiration).

In the other regimes the indication of a CDG is independent of the gas species if the dielectric constant of the
capacitance is not changed, which occurs when the capacitance is measured on the reference side only. Several authors
have published equations to describe p(T2)/p(T1) in the transition regime [5355]. The theory behind these equations,
however, is not complete and based on assumptions, so that experimentally determined parameters have to be used.
Therefore, if CDGs are used as reference standards, they should be calibrated also in the transition regime instead of
relying on equations with relatively uncertain parameters. Figure 12.15 shows a typical calibration curve of a 133-Pa
full-scale CDG for two gases. The molecular regime of helium is larger than that for nitrogen, because the mean free
path length of helium is larger at the same pressure and temperature.

A valuable guide on the calibration and use of CDGs has been published by a subcommittee of the "Recommended
Practices" committee of the American Vacuum Society [56]. According to these recommendations, before a CDG is
calibrated, the following precautions should be honored (Table 12.2):

Fig. 12.15
A typical calibration curve (carried out at the
Physikalisch-Technische Bundesanstalt, PTB) of a 133-Pa
full-scale CDG (MKS-Baratron) at elevated temperature
for helium and nitrogen. Three regimes can be
distinguished: the molecular regime below 1 Pa, the
viscous flow regime above 100 Pa, and the transition
regime in between.
< previous page page_678 next page >
< previous page page_679 next page >
Page 679

Table 12.2. Points to Consider when Calibrating Capacitance Diaphragm Gauges (CDGs)
CDG Calibrations
For installation consider:
Vibrations
Local air currents
Orientation of sensor head
Connections to differential gauge heads
Isolation valves for absolute-type CDG with full scale <100 kPa
Leak test

Before calibration consider:


Warm-up period ~ 24 h
Zeroing
Precycling

During calibration consider:


Pressures from low to high
Hysteresis
Zero checks

After calibration consider:


Close isolation valve (if existent) before airing
Switch off heater one hour before airing
Disconnect measurement port of differential-type CDG first

The CDG sensor head should be installed such that it is not subject to vibrations, and it should be protected from local
air currents. Since also the gravitational field is acting upon the diaphragm, the orientation of the sensor should be the
same during calibration and use. It is advisable to use the test port for the connection to the calibration system, since
most CDGs are linearized only in this direction and since the reference side should be kept as clean as possible. The
calibration system should be leak tested.

If temperature-controlled CDGs are being used, a warm-up period of 24 h is recommended. After this period the
instrument, including the signal conditioner, has to be zeroed. With single-sided CDGs, one has to apply a base pressure
lower than the smallest detectable pressure. The base pressure should be measured with an independent high- or
ultrahigh-vacuum gauge. For zeroing a differential CDG, the two ports should be "short-circuited" (Fig. 12.16). We note
that the zero indication is a function of line pressure, and for absolute pressure measurement the base pressure of the
reference side should be applied also to the test side for this reason.

To minimize hysteresis effects, it is recommended that CDGs be precycled from zero to full scale and back before
calibrations.

The calibration should proceed from the lowest to the highest pressure and then return, if any hysteresis effects are
expected. The calibration consists of applying pressure, waiting an appropriate time for thermal equilibrium to be
achieved, and recording the pressure indications of the standard and device under calibration as

< previous page page_679 next page >


< previous page page_680 next page >
Page 680

Fig. 12.16
The experimental setup for calibrating a differential CDG for absolute
pressures. The isolation valve is closed and the bypass valve open for
zeroing the instrument. Another gauge (not shown) should be used to
check the pressure, pR, on the reference side.

close to simultaneously as possible. When the CDG is calibrated for pressures below the viscous flow regime, also the
temperature of the gas in the primary or secondary standard has to be recorded. This is important, since when different
gas temperatures T1 will exist when the calibrated gauge is used, corrections have to be applied [see Eq. (12.20)]. For
best accuracy, zero- and full-scale indications should be checked after each calibration point, if possible.

Although most commercial absolute-type CDGs are safe against atmospheric overpressure, one should be aware of the
fact that overpressurizing CDGs with full scales < 133 kPa may seriously change the calibration curve. Therefore, it is
recommended that an isolation valve is used to prevent the sensor from being exposed to pressures higher than its full
scale.

The accuracy of a CDG calibration is mainly determined by the uncertainty of the generated or measured pressure in the
primary standard. It can be as low as 0.01% at 100 kPa, but as high as 0.3% at 0.1 Pa ((2 σ) uncertainties).

The long-term stability of all types of vacuum gauges is within a certain range an individual value of each specific
gauge. It is best estimated by frequently (at least no more than a year apart) recalibrating a gauge over a long period of
time to determine a standard deviation and/or drift of the calibration factors. For instruments with a long metrological
history (and these instruments are quite rare and very precious) the future stability can then be predicted within some
sort of confidence interval (there will never be a "guarantee"!). Typical for all kinds of vacuum secondary standards in
this chapter are discontinuous shifts in calibration factors rather than a steady drift.

CDG stabilities depend on their full scale, and the stability is much better in the viscous flow regime than in the others.
Typical values of long-term instabilities for good instruments over a one-year period in the viscous flow regime near
their full scale are: 0.1% for 100 kPa and 10 kPa full scale and 0.30.4% for 1 kPa and 100 Pa full scale [57, 58].

12.3.2
Spinning Rotor Gauges

Spinning rotor gauges (SRGs) [5961] are the preferred secondary standards from 103 Pa to 1 Pa. They can be used as
secondary standards even down to 104 Pa, if the residual drag and the single scattering of the data is low.

< previous page page_680 next page >


< previous page page_681 next page >
Page 681

In the SRG a steel (often stainless steel) sphere of 4.5 mm or 4.76 mm diameter is magnetically suspended in a small
tube, called a thimble, and put into rotation of about 400 Hz. Due to gas friction the rotor will slow down and the
relative deceleration rate is proportional to the pressure in high vacuum.

The SRG can be used for pressure measurement > 1 Pa [62, 63], but it is less suitable as secondary standard compared
to CDGs in this pressure range.

The pressure in the high-vacuum regime measured by an SRG is calculated from [60, 61]

where the square root is the mean thermal velocity of the gas species with molecular mass m and temperature T; d and ρ
are the diameter and density of the rotating sphere; σ is the dimensionless accommodation coefficient of tangential
momentum of the gas particles on the rotor surface; is the deceleration rate of the rotor due to pressure; and
RD (residual drag, also called offset) is the deacceleration (in s1) caused by induced eddy currents and possible thermal
drifts. RD is measured at p ≤ 106 Pa.

The accommodation coefficient σ is the parameter to be calibrated in an SRG. It is basically an effective


accommodation coefficient, because not only the surface itself but also the surface roughness contributes to its value.
For this reason, σ can be larger than 1. σ may vary from slightly below 1.0 to up to 1.27 for very rough rotor surfaces
[60]. It depends slightly on gas species, where the lightest gases hydrogen or helium give usually either the minimum or
maximum value of σ, depending on the rotor.

After calibrating, σ can be stored into the SRG control unit and the pressure indication will be correct according to the
calibration.

The deceleration rate versus pressure characteristic is, according to our recent knowledge, linear for pressures < 0.1 Pa.
Above 0.1 Pa it becomes increasingly non-linear due to the nonisotropy of incident molecules which have been whirled
up by collisions with molecules coming from the rotor. This can be considered by introducing a (slightly) pressure-
dependent σeff(p) or (as is done in the commercial control units) by a linearization procedure which includes gas
viscosity.

From zero up to 2 Pa, σeff is a strictly linear function of p, as was shown in Messer and Röhl [64] (Figure 12.17).
Therefore, two possibilities exist for determining σ:

•σeff vs. p is determined for 0.1 Pa < p < 2 Pa and extrapolated to p = 0: σ = σeff(p = 0).

• A pressure p significantly below 0.1 Pa is applied to the SRG and the reading pind of the SRG compared to the true
pressure p will give σ = pind/p.

In both cases, σ = 1 has to be entered into the control unit before calibration; in the first case also the viscosity value has
to be set to zero so that no linearization procedure is performed for the pressure indication. We note that for the radius
and the density of the rotor, usually nominal values are entered into the controller, and also for this reason the calibrated
value of σ will be only "effective."

< previous page page_681 next page >


< previous page page_682 next page >
Page 682

Fig. 12.17
The pressure dependence of the (effective) accommodation coefficient for a
special sphere. It is generally linear up to 2 Pa. From [64].

Table 12.3. Points to Consider when Calibrating Spinning Rotor Gauges (SRG)
SRG Calibrations
Consider:
Vibrations
Baking (may change accommodation coefficient)
Stable temperature around gauge head
Warm-up period ~ 6 h
Offset measurement
Frequency-dependence of offset
Transport: Rotor fixed and save from corrosion

Everything that changes the surface of the rotor may change its (effective) accomodation coefficient σ and may
therefore invalidate the calibration. The rotor surface may be subject to corrosion, so that keeping the rotor under
vacuum all the time is certainly a good idea. Mechanical friction may change the surface roughness so that the rotor
should be fixed during transporation. Both requirements can be accomplished by a special transport device as developed
by Röhl and Jitschin [65]. Also, adsorbed molecules on the rotor surface may change the accommodation coefficient. σ
may be different by 2% for a rotor after baking compared to the same rotor before baking. Some rotors, however, do not
change σ after baking [66].

During the calibration procedure of an SRG, the following should be considered (Table 12.3):
< previous page page_682 next page >
< previous page page_683 next page >
Page 683

• The gauge head must not be subject to significant vibrations.

• Before calibration, the rotor should run for at least 6 h. This is because after start-up of the SRG, the temperature drift
of the rotor caused by the eddy currents will give a systematic drift of the residual drag in the first hours.

• Since temperature drifts will falsify the RD measurement, temperature drifts of the thimble (and rotor) should be
avoided.

• The residual drag may be frequency-dependent due to frequency-dependent changes in the axis of rotation and due to
the frequency-dependence of the eddy currents [67]. The size of this frequency-dependence (020% for |∆f| = 10 Hz) is
different from rotor to rotor, and also it is dependent on the orientation of the rotor in the magnetic field. The frequency-
dependence of the residual drag should be carefully determined when high accuracy of the calibration is required or
signals close to the offset value are taken.

• It is recommended that the residual drag be measured over a longer period of time (12 h) before calibration to
determine its value with a small uncertainty due to random effects. If the temperature of the thimble does not change
significantly and the rotor is continuously kept in the same orientation in the same magnetic field, the offset value RD
will be stable [68]. If, however, the temperature is not stable, the offset is better measured immediately before and after
the calibration.

The relative uncertainty of a primary SRG calibration is typically 0.30.5%, which includes the uncertainty due to the
fact that σ also depends slightly on the orientation of the rotor in the magnetic field, typically by 0.00.2%.

The long-term instability over a one-year period of the σ of a well-handled SRG is typically 0.30.5% [69], but can also
be much better. Shifts of 1%, however, after a one-year period have been observed, and this value should be used for a
first-time calibrated rotor.

12.3.3
Ionization Gauges

Although, in general, ionization gauges (IGs) have only a modest stability, they have to be used as secondary standards
below 103 Pa, since for now there is no alternative in this pressure range.

The most common type of ionization gauges nowadays is the BayardAlpert gauge [70], where a hot cathode outside a
cylindrical anode grid provides the electrons for ionization and the thin ion collector wire is centered in the cylindrical
anode. They are mainly available as nontubulated systems ("nude"-type) or as systems within a glass enclosure. In the
German Calibration Service (DKD), some nude ionization gauges have been successfully used as reference standards.

Extractor gauges, which have the ion collector outside the anode grid to reduce the x-ray limit due to high-energy
photons generated at the anode, may also be used as secondary standards.

Cold-cathode gauges, on the other hand, have been found less suitable as secondary standards [71, 72].

Common to all types of ionization gauges is the measurement of an ion current, which ideally should be linearly
proportional to the molecule density in the gauge.

< previous page page_683 next page >


< previous page page_684 next page >
Page 684

Table 12.4. Points to Consider when Calibrating Ionization Gauges


Ionization Gauge Calibrations
Consider:
Tubulation of nude gauges
Orientation of gauge head
Warm-up period ~ 12 h
Conditioning procedure of gauge heads
Cathode heating and electron emission regulation
Residual current measurement
Pressures from low to high

Therefore, in a calibration of an IG, the sensitivity S (also called the vacuum gauge constant) is determined:

where p0 is the pressure due to residual gases in the calibration chamber, p is the pressure due to the residual gases plus
the calibration pressure pcal generated by the standard, Ic is the collector current at pressure p, Ic0 is the residual
collector current at residual pressure p0, and Ie is the emission current measured on the anode.

Since physically in an ionization gauge the gas density is being measured, while the sensitivity is determined by the
pressure, it is always necessary to state the temperature at which the calibration was performed. For clarity, S should
always be corrected for and stated for a temperature of T0 = 23°C. If S was determined for a temperature T1, it is
expressed as

When ionization gauges are being calibrated as secondary standards, the following should be considered (Table 12.4):

• The potential distribution inside the gauge head depends on its surrounding. If a nude gauge is calibrated, it is
therefore advisable to calibrate the gauge in the same configuration as in use or to surround the gauge head at its full
length with a tube which will not be removed. Recently, Bayard-Alpert gauges completely immersed in a stainless-steel
tube have become commercially available.

• The orientation of the gauge head should be identical during calibration and use, since geometrical deformations due
to different orientations may affect the potential distribution and the electron trajectories in the gauge head.

• Due to the hot cathode, the temperature of the gauge head (Tga) is higher than the temperature of the vessel (Tch), and
therefore gas pressure and particle density will be different in the gauge head and the vessel (see Section 1.10 on

< previous page page_684 next page >


< previous page page_685 next page >
Page 685

thermal transpiration). This difference is dependent on the special geometry, the flow conditions, temperature gradients,
accommodation coefficients, and so on, and cannot be calculated. At most the pressures will differ by the factor

assuming molecular flow and complete accommodation of the gas particles to either Tga or Tch. If the
difference is smaller, it will be proportional to the square root factor with a proportionality constant < 1. Hence, to make
the calibration as valuable as possible, the ratio Tga/Tch should be as close as possible during calibration and further
use. This can be accomplished when the ionization gauge is always enclosed in the same tube, so that the heat
conductivity to the immediate environment will be the same, and when the vessel temperatures are very similar during
calibration and use. Additionally, the emission current (heating power) should not be changed significantly from
calibration to use.

• Since effects like the secondary electron and photon production in IGs depend on the surface state, the surfaces should
be as clean as possible to obtain reproducible results from these effects [73, 74]. Therefore the calibration system
including the IG should be baked out to obtain a residual pressure of at least a factor of 10 lower than the lowest
calibration pressure. The cathode should be cleaned by applying a higher temperature than during operation, and the
anode should be cleaned by electron bombardment. Both procedures are usually done in the "degas" mode of
commercial control units. Longer degas operating periods than stated in the manual for the specific gauge, however,
must be avoided, since excessive heating will irreversibly damage the electrodes. The ion collector, whose surface is
critical due to the secondary electron production on it, can be cleaned by ion bombardmentthat is, operating the
ionization gauge at a high pressure (102 Pa or more, if possible) for about 1 h [75, 76]. After these conditioning
procedures, the gauge has to be operated in normal mode (regular emission current) for 12 h before calibration.

• Whenever an ionization gauge is being calibrated, it should be operated either with its control unit or at least with the
same cathode heating power supply and electron emission regulations. It has been found that different cathode heaters,
though stated with the same nominal data, may result in different sensitivities, possibly because the electron emission
distribution on the cathode is changed.

• Instabilities in determined sensitivities of IGs may be due to either changes in the gauge head or changes in the
controller. To distinguish between the two, the controller has to be calibrated separately (emission and ion current
meters, voltages).

Although in principle it should be possible to calibrate a gauge for one gas and use this calibration for another gas with
the ionization probability ratio for the two gases as scaling factor [77], investigations have shown [78, 79] that this is
not the case if high accuracy (uncertainty < 10%) is required. Even for isotopes of the same gas speciesfor example,
hydrogen and deuteriumsignificant differences in sensitivities can be found [80].

If the sensitivity is determined according to Eq. (12.22), the residual current in the IGcaused by outgassing, electron- or
photon-stimulated desorption, or x-ray-induced photoelectronsis subtracted from the signal. A user should be aware of

< previous page page_685 next page >


< previous page page_686 next page >
Page 686

this, because, when the gauge is in use, its residual current reading (or the equivalent pressure reading) has to be
subtracted from the signal as well.

For high accuracy, IGs should be calibrated with three points per decade over their operating range, because in many
cases the sensitivity is at least slightly pressure-dependent (with about a few percent over several decades), even below
103 Pa [81].

The relative (2 σ) uncertainty of sensitivity determination, which can be reached at the time of calibration of an IG,
increases with decreasing pressure and has typical values of 0.51% at 102 Pa, 23% at 106 Pa, and up to 40% at 1010 Pa.

The long-term instability of any IG as for other gauges is a very individual feature [82]. For an IG that is of high quality
and is carefully treated, a value of 36% over a one-year period may serve as reasonable estimate for its instability for
this period [83].

12.3.4
Mass Spectrometers

Mass spectrometers are used extensively for the qualitative and quantitative analysis of gas mixtures in many
applications in industry and research. As an example, in the microelectronic industry the purity of gases has been
continuously improved for the higher integrity of the devices, and a check of impurities in the process gases has to be
done on a routine basis.

Today most of the mass spectrometers for general purposes are of the quadrupole-type, dominating with about 95% of
the market; the magnetic sector type has a significant fraction of the remaining 5% [84]. Therefore, in the following we
will mainly refer to quadrupole mass spectrometers (the calibration methods, however, are independent of the
spectrometer type).

Quantitative interpretation of the mass spectrum can be obtained if the mass spectrometer has been calibrated before. A
useful factor to define for calibration purposes is a sensitivity for each gas component x measured:

I(x) is the ion current at partial pressure p(x) when the instrument is tuned to the molecular peak of component x, and I0
(x) is the corresponding value at some reference pressure p0(x). A number of investigations, however, have shown that
mass spectrometers cannot be calibrated absolutely in a manner, which would allow quantitative interpretation of any
possible mass spectrum [8588]. For example, if the sensitivity for a gas component x was determined by a calibration, in
which x was the only component in the system, this sensitivity for x may be very different if another gas component
with a much higher partial pressure (sometimes called ''matrix gas") is present. This is because at higher pressures,
space charge in the ion source can alter ion extraction efficiencies and ion-molecule collisions can alter ion species
ratios by charge transfer [85]. In this case, for calibration the specific gas mixture very similar to the one when the mass
spectrometer is in use has to be applied. For calibrations of relatively low total pressures (< 104 Pa), however, it is
usually sufficient to calibrate for each gas species separately.

< previous page page_686 next page >


< previous page page_687 next page >
Page 687

Table 12.5. Some Specific Points to Consider when Calibrating Quadrupole Mass Spectrometers
Quadrupole Mass Spectrometer Calibrations
Before calibration consider:
Cleanliness of gauge head and calibration system
Uniform bake-out (no localized sources of outgassing)
Background partial pressure < 10 times calibration partial pressure
Grounding
Warm-up of instrumentation (6 h)
Regular peak shape of mass peaks
Stable ion currents (repeatibility of mass peaks)
Tuning adjustments (set resolution, scan speed, etc.)

During calibration consider:


Protocol of all settings
Background scan
Instability of background signal
Outgassing of mass spectrometer
Fragmentation
Linearity
Dependence of sensitivity on presence of other gas species

The settings of voltages and resolution have a significant influence on the sensitivity and must be clearly stated at the time of
calibration so that they can be reset when using the calibration data.

Additionally the stability has been found to be very poor in an investigation of a group of mass spectrometers, where the
sensitivity for a single-component gas varied up to a factor of two over a 220-day period [85]. The use of electron multipliers is
particularly critical for the stability of mass spectrometers due to aging, especially of new multipliers units, and sensitivity
changes after bake-outs [89]. For these reasons, the use of quadrupole mass spectrometers as reference standards is problematic.

Another problem is that the outgassing of mass spectrometers is quite high and many components measured in a residual gas
spectrum are being produced by the mass spectrometer itself. It was shown by several authors [9092] that the production of
some gases such as methane, water vapor, carbon monoxide, and carbon dioxide is enhanced when hydrogen is introduced into
the mass spectrometer.

A subcommittee of the "Recommended Practices" committee of the American Vacuum Society has published recommendations
for the calibration of mass spectrometers for partial pressure analysis [93], which we will partly refer to in the following (see
also Table 12.5).

The most accurate calibration method of partial pressure analyzers is the use of two or more flowmeters in a continuous flow
primary standard (Section 12.1.3). Each flowmeter injects a well-known gas flow of different species into the calibration
chamber at which the mass spectrometer is installed. Each partial pressure is then given by Equation (12.8), when each variable
is evaluated for the specific gas component. Due to the use of a primary standard and at least two

< previous page page_687 next page >


< previous page page_688 next page >
Page 688

flowmeters, this calibration method is probably restricted to the National Metrological Laboratories.

Useful calibrations can also be obtained using ionization gauges as reference standards, since ionization gauges are
much more linear, stable, and predictable than partial pressure analyzers [85]. It is even possible to calibrate a mass
spectrometer with an IG for gas mixtures with components in moderate ratios from about 1:1 to 1:10 if each component
can be separately injected in the gas chamber, since for p < 103 Pa IGs react quite linearly on the addition of another gas
component. Only the specific sensitivity value of each gas species of the IG has to be considered in the calculation of
partial pressure. For the calibration of mass spectrometers with ionization gauges it is necessary to obey the rules as
stated for general comparison calibration systems (Section 12.2).

In principle, it is also possible to calibrate a mass spectrometer with a spinning rotor gauge. The useful overlap from
104 Pa up to 102 Pa, however, is generally too small to obtain useful results due to the often observed nonlinearity and
space charge effects of quadrupole mass spectrometers in this range, so that it is not possible to extrapolate sensitivity to
lower pressures.

Another possibility of calibrating mass spectrometers is the use of reference leaks [94] or calibrated capillaries [95].
When the flow rate through these devices and the effective pumping speed in the chamber is known, the partial pressure
can be calculated [96]. This calibration method is particularly useful when a small trace gas pressure in a large matrix
gas pressure (this pressure can be measured by an IG) has to be established for calibration, but also if gas samples of
different species are mixed in the reservoir for the capillary by the use of CDGs at relatively high pressures.

If no calibrated leak is available, it is also possible to measure the ratio of the upstream and downstream pressures of a
leak or small conductance in situ by a secondary standard. The principal experimental setup for this kind of calibration
is shown in Fig. 12.18. CDGs, SRGs, or IGs can be used as secondary standards, dependent on the upstream pressures
which have to be established for the desired calibration pressure.

Fig. 12.18
The principal experimental setup for calibrating mass
spectrometers according to a pressure divider method.
The ratio of up- and downstream pressure of the small
conductance can be measured with the secondary
standard. It is possible to expand this setup also for
generating gas mixtures in the main chamber (see text).

< previous page page_688 next page >


< previous page page_689 next page >
Page 689

The small conductance can also consist of a leak valve; however, the stability of the leak has to be verified. If the
pressure ratio shall be independent of gas species and pressure, it must be ensured that the flow through the conductance
is in the molecular regime, since the flow through the pump orifice of much larger conductance will automatically be in
the molecular regime. As in all calibration systems, the size of the pump orifice must ensure a homogeneous pressure
distribution in the chamber.

It is possible to expand the setup of Fig. 12.18 for generating gas mixtures in the main chamber by adding a separate gas
inlet system for each gas symmetrically to the axis of the main chamber.

The accuracy of the calibration of a quadrupole mass spectrometer depends strongly on the instrument itself. Due to
their poor stability a very accurate calibrationfor example, on a primary standard with two flowmetersis very often just
not worth the effort. Uncertainties which include the instrument and instability effects smaller than 10% should not be
expected when the mass spectrometer is removed from the calibration system. Only when mass spectrometers are
calibrated in regular and frequent manner in situ is an improved accuracy possible.

12.4
Calibration of Test Leaks

The calibration of test leaks with secondary methods is treated in Chapter 6 of this book, while in this section we will
describe the calibration of test leaks with primary methods, which was also covered by recommendations of a
subcommittee of the AVS "Recommended Practices" committee (see page 678) [97].

Test leaks provide a stable gas flow, usually of helium, since leak detectors are calibrated with helium. Common types
are permeation leaks (also called diffusion leaks) and capillary leaks. The latter can be used for all gases, which do not
clog or etch the capillary, while the permeation type is only applicable to a few gases, mainly helium and hydrogen.

For users of test leaks the most convenient unit for the leak rate is Pa·liter·s1 or similar units. We have to note, however,
that this unit will not give the complete physical information if two temperatures are not stated at the same time. These
are the temperature of the permeation material or the capillary and the temperature at which the gas pressure or gas flow
was measured. To avoid confusion, it is recommended that the unit moles per second (mol·s1) is used for test leaks [98,
99].

Differentiation of the ideal gas law with respect to time t at constant temperature yields an expression that is useful in
evaluating molar leakage rates qv:

where ν is the number of moles and T is the temperature of the gas at which pressure p was measured, and R is the
universal gas constant.

Two main primary calibration methods can be deduced from this equation:

1. The pressure is held constant so that the second term is zero.

2. The volume is held constant so that the first term vanishes.

< previous page page_689 next page >


< previous page page_690 next page >
Page 690

Fig. 12.19
The principal experimental setup for primary calibrations of test
leaks. On the mass spectrometer the signal from the test leak is
compared to a similar flow rate generated and measured by the
flowmeter, which can be "switched on" and "off."

Since both methods are also used for generating and measuring gas flows in a flowmeter (Section 12.1.3), it is
obviously convenient to use a flowmeter, if available, to calibrate leaks by comparison with a known gas flow from the
flowmeter. A typical experimental setup is shown in Fig. 12.19. The signal to be compared is the helium partial pressure
reading on a mass spectrometer, which has to be installed such that identical gas flows from the flowmeter or the test
leak will give identical signals; that is, the mass spectrometer should be symmetrical to the flowmeter, and the test leak,
and/or the tubing to the both gas sources should not be too short. The flow rate from the flowmeter is adjusted close to
the one from the test leak. In practice, two points of the flow rate from the flowmeter may be adjusted slightly higher
than, one close to, and two points slightly lower than the leak rate, and the zero crossing of the five differences to the
leak rate is taken.

Leak rates are often below the lower limit of flow rate of a constant pressure flowmeter of about 1011 mol·s1 (2 × 105
Pa·liter·s1 at 23°C). In these cases the flowmeter is either operated in the constant conductance mode down to 1015
mol·s1 [31] or flow-division techniques are used to further reduce the flow rate [40].

If no flowmeter is available, a direct realization of either method (1) or (2) is necessary.

1. In the constant pressure method the gas from the leak is allowed to flow into a volume V1 in which the pressure is
measured by a secondary standard. After a while, at pressure p1 (p1 has to be negligible compared to the upstream
pressure of the leak) a second known evacuated volume V2 is added by opening a valve, similar to a static
expansion, and the time ∆t is taken for the pressure to recover to the same value p1 before the expansion (Fig. 12.20).
The flow rate qv is given by

< previous page page_690 next page >


< previous page page_691 next page >
Page 691

Fig. 12.20
Test leak calibration by the constant pressure method:
The gas from the test leak is expanded into volume V2,
and the time is measured until the pressure has
recovered to its original value.

Fig. 12.21
Variation of test leak calibration by the constant pressure
method: The volume displacer is moved such that the
signal on the mass spectrometer remains constant. The
volume speed and the measured partial pressure will
determine the leak rate.

where T is the average temperature in the volumes V1 and V2 at the time of the second measurement of p1. In case of
any significant temperature drift, p1 has to be corrected.
Instead of using a fixed volume V2, a calibrated variable volume like a pistoncylinder assembly, conveniently driven by
a stepping motor, may also be used [100] (Fig. 12.21). Table 12.6 lists the main possible errors which may be made
during a calibration of test leaks which should be accounted for.

If a permeation leak was closed with a valve for a long time, higher partial pressures will build up on the low pressure
side of the leak and change the concentration gradient in the permeation element. After opening the valve, one has to be
aware of

< previous page page_691 next page >


< previous page page_692 next page >
Page 692

Table 12.6. Possible Effects Which May Falsify Leak Rate Calibrations
Leak Calibrations
Possible problems of vacuum system:
Leak of vacuum system
Outgassing of inner surface
Adsorption of leaking gas species
Desorption from inner surface
Instability of background pressure
Temperature gradients and drifts

Problems of vacuum gauge (secondary standard):


Pumping effect
Outgassing
Instability of background signal

Problems of leak:
Leak rate in stationary equilibrium?

the time to reach equilibrium with the new boundary conditions. Capillary leaks on the other hand have quite fast equilibrium times.

2. The same precautions of Table 12.6 must be honored if the pressure rise with time is measured in a constant and known volume which should
be not too small (at least 50 cm3 [101]). It is advisable that all calibrations of leaks should start and also end with a "blind" experiment; that is,
the measurement is repeated with the leak valved off. When small, well degassed volumes with the pressure measured by the non-gas-
consuming spinning rotor gauge are used, molar flow rates as low as 1018 mol·s1 can be measured. The uncertainties of the measured leak rates
with the above methods are rising with decreasing leak rates. The uncertainties vary typically from 0.5% to 1% at 106 mol·s1 to 8% at 1014
mol·s1 [102].

Since the permeation and the conductance of a capillary are temperature-dependent, the test leaks have to be temperature-conditioned to
within ±0.1°. A typical permeation leak changes its leak rate near 23°C by 3%/°C, so that an uncertainty of 0.3% can be expected for ±0.1°.

12.5
Measurement of Pumping Speeds

The pumping speed of a pump is defined as the volume of gas removed by the pump per unit of time, for which reason the pumping speed is
also called volumetric speed. Although the pumping speed usually depends on the inlet pressure of the pump, the above definition is convenient,
because this pressure-dependence is rather small (< 30% over several decades) for many pumps in their regular operating range. Sliding vane
rotary pumps, for example, have a well-defined displacement volume, in which the sucked in gas is compressed and then exhausted at
atmospheric pressure.

< previous page page_692 next page >


< previous page page_693 next page >
Page 693
The displacement volume times the rotor frequency will give the maximum pumping speed in this case.

The volumetric speed of a pump can be determined by measuring gas flow qpV injected through the inlet port of the pump and
the pressure p in front of it at a certain temperature:

where p0 is the pressure for qpV = 0. For the application of Eq. (12.27) p should be significantly greater (about a factor of 2)
than p0.

While the quantity qpV can be reliably measured, the measurement of p is rather difficult. Due to the pump there is a strong
pressure gradient in front of the inlet port, and the movement of the gas particles is far from being isotropic or Maxwellian. An
ideal physical concept has to use a chamber of infinite volume and surface area, where the pump inlet port has a negligible
effect on the Maxwellian distribution inside [103, 104]. Since inlet ports are of significant size and practical vessels cannot
approximate the ideal chamber according to the above concept, written standards have been made which state how the
experimental system (called test dome) has to be designed and at which position and orientation relative to the inlet port the
pressure p has to be measured. The original idea of the first standards was to find a position for the measurement of p in the test
dome, so that p would have a very similar value as in an infinitely sized test dome [103105]. On the other hand, the test dome
should be similar to a practical vacuum chamber, so that the measured pumping speed is actually a useful value for a user who
is designing a vacuum system.

Table 12.7 shows existing standards for the calibration of pumps. The American Vacuum Society (AVS) replaced their former
so-called "standards" 4.1, 4.2, and 4.8 and also 5.1, 5.2 and 5.3 by new recommended procedures published in 1987 [106]

Table 12.7. The Existing Written Standards for the Measurement of Pumping Speed and Acceptance Specifications
for Several Pump Typesa
Written Standard Type of Vacuum Pump
DIN 28426, Part I,
PNEUROP, ISO 1607/1,2 Rotary plunger, sliding vane rotary
rotary piston

DIN 28426, Part II,


PNEUROP, ISO 1607/1,2 Roots

DIN 28427
PNEUROP, ISO 1608/1,2 Diffusion, vapor jet

DIN 28428, PNEUROP


Turbomolecular

DIN 28429, PNEUROP


Getter ion

PNEUROP PN5 ASRCC/5


Cryo-refrigerator
a For addresses for ordering these standards see end of References.

< previous page page_693 next page >


< previous page page_694 next page >
Page 694

and 1989 [107] in order to have agreement with the ISO, DIN, and PNEUROP standards. The location of the gauge
measuring p was moved away from the pump inlet from one-quarter to one-half of the pump inlet diameter, resulting in
a reduction of 1015% of the measured pumping speed compared to the old AVS "standard."

In all written standards or recommendations of Table 12.7 the internal diameter of the test dome (Figs. 12.22 and 12.23)
must be the same as the pump inlet diameter. The AVS procedure requires this down to 50-mm internal diameter,
whereas the other standards require it down to 100 mm, but with a well-described adapter from the test dome to the
pump if the pump diameter is < 100 mm. The top of the test dome should be rounded, conical, or inclined. This design
was chosen to ensure that any oil, when condensing on this surface, will run down the sides of the domes rather than
drop down onto the pump stack and cause erratic pressure bursts. In today's completely oil-free pumps (e.g., ion pumps
or magnetically levitated turbomolecular pumps) or nearly oil-free pumps, this is of no importance and the top may
chosen to be flat.

Depending on pressure p, two methods are defined by the standards to measure the flow rate qpV. In the molecular
regime (AVS recommendation) or p < 104 Pa (DIN/ISO/PNEUROP) the orifice flow method with the twin dome (Fig.
12.23) should be used where the pressure drop across a suitable designed orifice is measured. At higher pressures a
single-dome configuration (Fig. 12.22) is used with qpV being measured by some type of flowmeter (for details see the
specific standards). Also a flowmeter as described in Section 12.1.3 may be used. Unfortunately, it has been
Fig. 12.22
Design of a single test dome for pumping speed measurement
according to DIN/ISO/PNEUROP standards (see Table 12.7).
The diameter of the inlet flange of the pump must be D. Pressure
is measured at flange 2. A known gas-flow rate is injected into
the dome. Reprinted from PNEUROP, 1972, Vacuum Pump
Acceptance Specifications Part II, p. 8, with permission
of VDMA, Frankfurt, Germany.

< previous page page_694 next page >


< previous page page_695 next page >
Page 695

Fig. 12.23
Design of a double test dome for
pumping speed measurement
according to DIN/ISO/PNEUROP
standards (see Table 12.7). The
diameter of the inlet flange of the
pump must be D. The gas-flow rate
is measured by the pressure
difference p1 p2 across the
orifice between the domes.
Reprinted from PNEUROP, 1976,
Vacuum Pump Acceptance
Specifications Part IV, p. 11, with
permission of VDMA, Frankfurt,
Germany.

found that the orifice flow method and the flowmeter method do not agree within their uncertainties for an overlapping
pressure range [108], which is not too surprising because the molecular flow conditions are different in the single- and
twin-dome configurations. With a flowmeter of wide range down to about 105 Pa·liter·s1 as available in the National
Laboratories, it is possible to calibrate high- and ultrahigh-vacuum pumps over their entire operating range without
using the orifice flow method. This is of great convenience, since only one experimental setup and the simpler single-
dome configuration can be used. To avoid the described ambiguities between the two methods, it may be advantageous
one day to recommend only the flowmeter method in the written standards.
Although other methods for measuring pumping speeds exist, we mention the conductance modulation method [109,
110], it is recommended that only the written standards be used, even if drawbacks exist with them, in order to produce
comparable results of pumping speed.

References

1. M. S. Kaminsky and J. M. Lafferty, eds., Dictionary of Terms for Vacuum Science and Technology. AIP Press,
Woodbury, NY, 1980.

2. International Vocabulary of Basic and General Terms in Metrology. Beuth Verlag, Berlin, DIN, Deutsches Institut
für Normung (German Institute for Standardization) 1994.

< previous page page_695 next page >


< previous page page_696 next page >
Page 696

3. G. Klingenberg and F. Lüdicke, PTB-Mitt. 101, 718 (1991).

4. J. C. Legras, B. Schatz, and P. Delajoud, Bull. BNM, Bulletin Bureau Nationale de Metrologie 65, 3953 (1968).

5. P. Reity, Bull. BNM, Bulletin Bureau Nationale de Metrologie 70, 818 (1978).

6. G. N. Peggs, K. W. T. Elliott, and S. Lewis, Metrologia 15, 7785 (1977).

7. K. Jain, C. Ehrlich, J. Houck, and J. K. N. Sharma, Meas. Sci. Technol. 4, 249257 (1993).

8. C. R. Tilford, R. W. Hyland, and Y.-T. Sheng, Yi-tang, BIPM, Bureau International des Poids et Mesures Mongr.
89/1, 105113 (1989).

9. K. Jousten, Vacuum 45, 1205 (1994).

10. M. Knudsen, Ann. Phys. (Leipzig) 31, 633 (1910).

11. K. F. Poulter, J. Phys. E 10, 112 (1977).

12. R. Holanda, NASA Tech. Note NASA TN D-3100 (1965).

13. R. Holanda, NASA Tech. Note NASA TN D-5406 (1969).

14. G. Messer, PTB Jahresber., p. 203 (1972).

15. S. Schuman, Trans. Natl. Vac. Symp. 9, 463 (1962).

16. M. Bergoglio, A. Calcatelli, L. Marzola, and G. Rumanio, Vacuum 38, 887 (1988).

17. K. W. T. Elliott and P. B. Clapham, MOM (U.K., Natl Phys. Lab., Div. Mech. Opt. Metrol.) NPL Rep. MOM 28
(1978).

18. J. K. N. Sharma and Pardeep Mohan, J. Vac. Sci. Technol. A 6, 3148 (1988).

19. A. Berman and J. K. Fremerey, J. Vac. Sci. Technol. A 5, 24362439 (1987).

20. K. W. Elliott, D. M. Woodman, and R. S. Dadson, Vacuum 17, 439 (1967).

21. C. Meineke and G. Reich, J. Vac. Sci. Technol. 4, 356 (1967).

22. W. Jitschin, J. K. Migwi, and G. Grosse, Vacuum 40, 293 (1990).

23. W. Jitschin, J. K. Migwi, and G. Grosse, Vacuum 41, 1799 (1990).

24. K. Jousten and G. Rupschus, Vacuum 44, 569 (1993).

25. H. G. Bennewitz and H. D. Dohmann, Vak. Tech. 14, 8 (1965).

26. K. F. Poulter, Vacuum 28, 135 (1978).

27. G. N. Peggs, Vacuum 26, 321 (1976).

28. Standard Methods for Calibration. Part I. Pressure Reduction by Continuous Flow in the Pressure Range of 103
Torr to 107 Torr, ISO/DIS 3570/I draft. This draft never made it to an official agreed ISO standard.
29. Calibration of Vacuum Gauges within the Range of 103 mbar to 107 mbar; General Method: Pressure Reduction by
Continuous Flow, DIN 28416. Beuth Verlag, Berlin, 1976.

30. Measurement of Throughput by the Volumetric Method at Constant Pressure, DIN 28417. Beuth Verlag, Berlin,
1976.

31. K. Jousten, G. Messer, and D. Wandrey, Vacuum 44, 135 (1993).

32. K. E. McCulloh, C. R. Tilford, C. D. Ehrlich, and F. G. Long, J. Vac. Sci. Technol. A 5, 376 (1987).

33. L. Holland and C. Priestland, Vacuum 17, 461 (1967).

34. G. Grosse and G. Messer, Vacuum 20, 373376 (1970).

35. A. S. Berman, J. Appl. Phys. 36, 3356 (1965).

36. W. Liepmann, J. Fluid Mech. 10, 65 (1961).

37. J. K. N. Sharma, P. Mohan, and D. R. Sharma, J. Vac. Sci. Technol. A 8, 941 (1990).

38. P. Looney, private communication.

< previous page page_696 next page >


< previous page page_697 next page >
Page 697

39. W. Jitschin, K. Jousten, and D. Wandrey, J. Vac. Sci. Technol. A 10, 3344 (1992).

40. C. R. Tilford, S. Dittmann, and K. E. McCulloh, J. Vac. Sci. Technol. A 6, 2853 (1988).

41. H. Mittelstädt, G. Rupschus, H. Menzer, and M. Richard, Exp. Tech. Phys. 21, 449 (1973).

42. A. Calcatelli and G. Rumanio, J. Vac. Sci. Technol. A 3, 1750 (1985).

43. J. K. N. Sharma and D. R. Sharma, J. Vac. Sci. Technol. A 6, 2508 (1988).

44. A. Tison and C. R. Tilford, in RL/NIST Workshop on Moisture Measurement and Control for Microelectronices (B.
A. Moore and J. A. Carpenter, Jr., eds.), NISTIR 5241, pp. 1929. NIST, Washington, DC, 1993.

45. J. R. Roehring and J. C. Simons, Trans. Natl. Vac. Symp. 8, 511 (1962).

46. F. Feakes and F. L. Torney, Trans. Natl. Vac. Symp. 10, 257 (1963).

47. K. Jousten, Shinkuu, J. Vac. Soc. Jpn 37, 678 (1994).

48. G. Grosse and G. Messer, Vak. Tech. 30, 226 (1981).

49. Standard Method for Vacuum Gauge Calibration by Direct Comparison with a Reference Vacuum Gauge, DIN
28418. Beuth Verlag, Berlin, 1976.

50. Calibration by Direct Comparison with a Reference Gauge, ISO/DIS 3567 draft. This draft never made it to an
official agreed ISO standard. Related standard drafts are ISO 3568 (calibration of ion gauges by comparison) and 5300
(calibration of thermal conductance gauges by comparison).

51. V. V. Kuz'min, Vacuum 46, 251 (1995).

52. J. J. Sullivan, J. Vac. Sci. Technol. A 3, 1721 (1985).

53. K. F. Poulter, M-J Rodgers, P. J. Nash, T. J. Thompson, and M. P. Perkin, Vacuum 33, 311 (1983).

54. S. Chu Liang, J. Appl. Phys. 22, 148 (1951).

55. T. Takaishi and Y. Sensui, Trans. Faraday Soc. 59, 2503 (1963).

56. R. W. Hyland and R. L. Shaffer, J. Vac. Sci. Technol. A 9, 2843 (1991).

57. R. W. Hyland and C. R. Tilford, J. Vac. Sci. Technol. A 3, 1731 (1985).

58. K. F. Poulter, Vide 207, 521 (1981).

59. G. Comsa, J. K. Fremerey, B. Lindenau, G. Messer, and P. Röhl, J. Vac. Sci. Technol. 17, 642 (1980).

60. J. K. Fremerey, Vacuum 32, 685 (1982).

61. J. K. Fremerey, J. Vac. Sci. Technol. A 3, 1715 (1985).

62. B. E. Lindenau and J. K. Fremerey, J. Vac. Sci. Technol. A 9, 2737 (1991).

63. J. Setina and J. P. Looney, Vacuum 44, 577 (1993).


64. G. Messer and P. Röhl, PTB Jahresbe. p. 226 (1984).

65. P. Röhl and W. Jitschin, Vacuum 38, 507 (1988).

66. S. Dittmann, B. E. Lindenau, and C. R. Tilford, J. Vac. Sci. Technol. A 7, 3356 (1989).

67. S.-H. Choi, S. Dittmann, and C. R. Tilford, J. Vac. Sci. Technol. A 8, 4079 (1990).

68. J. Setina, Vacuum 40, 51 (1990).

69. W. Jitschin, J. Vac. Sci. Technol. A 8, 948 (1990).

70. R. T. Bayard and D. Alpert, Rev. Sci. Instrum. 21, 571 (1950).

71. W. J. Lange, J. H. Singleton, and D. P. Eriksen, J. Vac. Sci. Technol. 3, 338 (1966).

72. R. N. Peacock, N. T. Peacock, and D. S. Hauschulz, J. Vac. Sci. Technol. A 9, 1977 (1991).

< previous page page_697 next page >


< previous page page_698 next page >
Page 698

73. U. Harten, G. Grosse, W. Jitschin, and H. Gentsch, Vacuum 38, 167 (1988).

74. H. U. Becker and G. Messer, Vide, Suppl. 201, 234 (1980).

75. H. Ave, H. U. Becker, and G. Messer, PTB-Mitt. 95(1), 20 (1985).

76. H. U. Becker and G. Messer, Proc. 5th International Vacuum Congress/9th Int. Cont. on Surface Science, 9th,
Madrid, 1983, p. 84 (unpublished).

77. R. L. Summers, NASA Tech. Note NASA TN D-5285 (1969).

78. A. Filippelli, AIP Conf. Proc. 171, 236 (1988).

79. R. Holanda, J. Vac. Sci. Technol. 10, 1133 (1973).

80. K. Jousten and P. Röhl, Vacuum 46, 9 (1995).

81. C. R. Tilford, K. E. McCulloh, and H. Seung Woong, J. Vac. Sci. Technol. 20, 1140 (1982).

82. S. D. Wood and C. R. Tilford, J. Vac. Sci. Technol. A 3, 542 (1985).

83. A. R. Filippelli and P. J. Abbott, J. Vac. Sci. Technol. A 13, 2582 (1995).

84. D. Lichtman, J. Vac. Sci. Technol. A 8, 2810 (1990).

85. L. Lieszkovsky, A. R. Filipelli, and C. R. Tilford, J. Vac. Sci. Technol. A 8, 3838 (1990).

86. J. A. Koprio, Vak.-Tech. 38, 134 (1989).

87. J. D. Sankey and A. H. Bass, Vacuum 40, 309 (1990).

88. W. Grosse Bley, Vacuum 38, 103 (1988).

89. W. R. Blanchard, P. J. McCarthy, H. F. Dylla, H. LaMarche, and J. E. Simpkins, J. Vac. Sci. Technol. A 4, 1715
(1986).

90. J. R. Bennet and R. J. Elsey, Vacuum 44, 647 (1993).

91. J. K. Fremerey, J. Vac. Soc. Jpn 37, 718 (1994).

92. Y. Nakashima, K. Tsuchiya, K. Ohtoshi, M. Shoji, K. Yatsu, and T. Tamano, J. Vac. Sci. Technol. A 13, 2470
(1995).

93. J. A. Basford, N. D. Boeckmann, R. E. Elletson, A. R. Flipelli, D. H. Holkeboer, L. Lieszkovsky, and C. M. Stupak,


J. Vac. Sci. Technol. A 11, 22 (1993).

94. D. J. Santeler, J. Vac. Sci. Technol. A 5, 129 (1987).

95. R. E. Ellefson, D. Cain, and C. N. Lindsay, J. Vac. Sci. Technol. A 5, 134 (1987).

96. D. J. Santeler, J. Vac. Sci. Technol. A 5, 129 (1987).

97. C. D. Ehrlich and J. A. Basford, J. Vac. Sci. Technol. A 10, 1 (1992).


98. G. M. Solomon, J. Vac. Sci. Technol. A 4, 327 (1986).

99. C. D. Ehrlich, J. Vac. Sci. Technol. A 4, 2384 (1986).

100. S. M. Thornberg, J. Vac. Sci. Technol. A 6, 2522 (1988).

101. J. A. Basford, J. Vac. Sci. Technol. A 5, 127 (1987).

102. S. A. Tison, Vacuum 44, 1171 (1993).

103. E. Fischer and H. Mommsen, Vacuum 17, 309 (1967).

104. Feng Yu-guo and Xu Ting Wei, Vacuum 30, 377 (1980).

105. D. R. Denison and E. S. McKee, J. Vac. Sci. Technol. 11, 337 (1974).

106. M. H. Hablanian, J. Vac. Sci. Technol. A 5, 2552 (1987).

107. B. R. F. Kendall, J. Vac. Sci. Technol. A 7, 2404 (1989).

108. G. Grosse, W. Jitschin, and D. Wandrey, Vacuum 41, 2120 (1990).

109. K. Terado, T. Okano, and Y. Tuzi, J. Vac. Sci. Technol. A 7, 2397 (1989).

110. Y. Tuzi, T. Okano, and K. Terano, Vacuum 41, 2004 (1990).

< previous page page_698 next page >


< previous page page_699 next page >
Page 699

ISO standards are available from:

ISO Central Secretariat, 1 Rue de Varembé, CH-1211 Genf 20; or from the National member of ISO respectively the
publisher of National Standards.

DIN standards are available from:

Beuth Verlag GmbH, D-10772 Berlin

PNEUROP are available from:

British Compressed Air Society, PNEUROP General Secretariat 8, Leicester Street, GB-London WC 2H 7BN

< previous page page_699 next page >


< previous page page_701 next page >
Page 701

Appendix

Graphic Symbols for Vacuum Components DIN 28 401

Vacuum Pumps

Vacuum pump, general

Positive displacement pump

Positive displacement pump, oscillating

Piston vacuum pump

Diaphragm vacuum pump

Rotary positive displacement pump


Rotary positive vacuum pump

Sliding vane rotary vacuum pump

Rotary piston vacuum pump

< previous page page_701 next page >


< previous page page_702 next page >
Page 702

Liquid ring vacuum pump

Roots vacuum pump

Turbine vacuum pump, general

Radial flow pump

Axial flow pump

Gas ring vacuum pump

Turbomolecular pump

Ejector vacuum pump

Diffusion pump
Adsorption pump

Getter pump

Sublimation (evaporation) pump

< previous page page_702 next page >


< previous page page_703 next page >
Page 703

Sputter ion pump

Cryopump
Vacuum Pump Accessories

Condensate trap, general

Condensate trap with heat exchange (e.g., cooled)

Gas filter, general

Filtering apparatus, general

Baffle, general

Cooled baffle
Cold trap, general

Cold trap with coolant reservoir

Sorption trap

< previous page page_703 next page >


< previous page page_704 next page >
Page 704

Vacuum chambers

Vacuum chamber

Vacuum bell jar


Isolation devices

Shut-off device, general

Isolating valve

Right angle valve

Stop cock

Three-way stop cock

Right-angle stop cock

Gate valve
Butterfly valve

Nonreturn valve

Safety shut-off device

< previous page page_704 next page >


< previous page page_705 next page >
Page 705

Valve Modes of Operation

Manual operation

Variable leak valve

Electromagnetic operation

Hydraulic or pneumatic operation

Electric motor operation

Weight-operated

Connection and Tubes

Flange connection, general

Bolted flange connection


Small flange connection

Clamped flange connection

Threaded tube connection

< previous page page_705 next page >


< previous page page_706 next page >
Page 706

Ball-and-socket joint

Spigot-and-socket joint

Connection by taper ground joint

Change in the cross section of a duet

Intersection of two ducts with connection

Crossover of two ducts without connection

Branch-off point

Collection of ducts

Flexible connection (e.g., bellows, flexible tubing)

Linear motion leadthrough, flange-mounted


Linear motion leadthrough, without flange

Leadthrough for transmission of rotary and linear motion

Rotary transmission leadthrough

Electric current leadthrough

< previous page page_706 next page >


< previous page page_707 next page >
Page 707

Vacuum Gauges

General symbol for vacuum

Vacuum measurement, gauge head

Vacuum gauge, gauge control unit

Vacuum gauge, control unit recording

Vacuum gauge control unit with dial indicator

Vacuum gauge control unit with digital indicator

Measurement of throughput

< previous page page_707 next page >


< previous page page_708 next page >
Page 708

Table of Conversion Factors n for Pressure Unitsa


µbar Pa mbar Torr atm lbs/in2
x (N/m2)
y
µbar
1 0.1 0.001 7.50 × 104 9.87 × 107 1.45 × 105
Pa (N/m2)
10 1 0.01 7.50 × 103 9.87 × 106 1.45 × 104
mbar
1000 100 1 0.750 9.87 × 104 0.01450
Torr 1.316 × 103
1333.22 133.322 1.33322 1 0.01934
atm 1,013.25 760
1,013,250 101,325 1 14.696
lbs/in.2
68,947.6 6,894.76 68.9476 51.715 0.0680 1
a Pressure in x units = n × pressure in y units.

< previous page page_708 next page >


< previous page page_709 next page >
Page 709

Vapor pressure curves of common gases. (To convert Torr to Pa, multiply by 133.)
Reprinted from R. E. Honig, RCA Review 13, 567 (1962).

< previous page page_709 next page >


< previous page page_710 next page >
Page 710

Vapor pressure curves of common gases. (To convert Torr to Pa, multiply by 133.)
Reprinted from R. E. Honig, RCA Review 13, 567 (1962).

< previous page page_710 next page >


< previous page page_711 next page >
Page 711

Vapor pressure curves of solid and liquid elements. (To convert Torr to Pa, multiply by 133.)
Reprinted from R. E. Honig, RCA Review13, 567 (1962).

< previous page page_711 next page >


< previous page page_712 next page >
Page 712

Vapor pressure curves of solid and liquid elements. (To convert Torr to Pa, multiply by 133.)
Reprinted from R. E. Honig, RCA Review13, 567 (1962).

< previous page page_712 next page >


< previous page page_713 next page >
Page 713

Vapor pressure curves of solid and liquid elements. (To convert Torr to Pa, multiply by 133.)
Reprinted from R. E. Honig, RCA Review13, 567 (1962).

< previous page page_713 next page >


< previous page page_714 next page >
Page 714

General Reference Books on Vacuum Science and Technology

A. Berman, Total Pressure Measurements in Vacuum Technology. Academic Press, Orlando, FL, 1985.

R. L. Boxman, D. M. Sanders, and P. J. Martin, Handbook of Vacuum Arc Science and Technology. Noyes Publications,
Park Ridge, NJ, 1995.

A. Chambers, R. K. Fitch, and B. S. Halliday, Basic Vacuum Technology. Adam Hilger, Bristol and New York, 1989.

J. H. DeBoer, The Dynamical Character of Adsorption. Oxford University Press (Clarendon), Oxford, 1953.

S. Dushman and J. M. Lafferty, Scientific Foundations of Vacuum Technique. Wiley, New York, 1962.

J. D. Fast, Interactions of Metals and Gases. Vols.1 and 2. Macmillan, London, 1971.

S. J. Gregg, The Surface Chemistry of Solids. Chapman & Hall, London, 1965.

A. Guthrie, Vacuum Technology. Wiley, New York, 1963.

M. Hablanian, High-Vacuum Technology. Dekker, New York, 1990.

N. Harris, Modern Vacuum Practice. McGraw-Hill, 1989.

D. M. Hoffman, B. Singh, and J. Thomas, Handbook of Vacuum Technology. Academic Press, San Diego, CA, 1997.

J. H. Leck, Total and Partial Pressure Measurement in Vacuum Systems. Blackie, Glasgow and London.

R. I. Masel, Principles of Adsorption and Reactions on Solid Surfaces. Wiley, New York, 1996.

J. F. O'Hanlon, A User's Guide to Vacuum Technology, 2nd ed., Wiley, New York, 1989.

V. Ponec, Z. Knor, and S. Cerny, Adsorption on Solids. Butterworth, London, 1974.

W. Pupp, Vakuumtechnik, Grundlagen und Anwendungen. Thieme, Munich, 1972.

P. A. Redhead, J. P. Hobson, and E. V. Kornelson, The Physical Basis of Ultrahigh Vacuum. Chapman & Hall, London,
1968. (and AVS reprint 1993).

R. W. Roberts and T. A. Vanderslice, Ultrahigh Vacuum. Prentice-Hall, Englewood Cliffs, N.J.

A. Roth, Vacuum Technology. North-Holland Publn., Amsterdam, 1982.

G. L. Saksagansky, Molecular Flow in Complex Vacuum Systems. Gordon & Breach, New York, 1988.

G. L. Saksagansky, Getter and Getter-Ion Vacuum Pumps. Harwood Academic Press, New York, 1994.

B. M. Trapnell, Chemisorption. Butterworth, London, 1955.

K. M. Welch, Capture Pumping Technology. Pergamon, Oxford, 1991.

M. Wutz, H. Adam, and W. Walcher, Theory and Practice of Vacuum Technology. Vieweg, Verlagsges., Braunschweig,
1989.
< previous page page_714 next page >
< previous page page_715 next page >
Page 715

Index

Absolute reaction rate theory, 549-550

Absolute temperature, 3, 6

Absorption, 589-605

diffusion rates, 590-591

equilibrium solubility, 590

kinetics, 591

permeation, effect of desorption kinetics, 600-605

steady-state permeation, 592-595

transient permeation, 595-600

Accommodation coefficient, 47-50, 572, 579

values, 50

Activation, getter materials, 273-274

Activation energy, 607, 612, 615

diffusion, 592

permeation, 602

Adsorption, 261, 263-265, 548-588

capillarity effects, 584-588

dissociative, 576, 607

equations, 548-551

heat, 562, 566-567

immobile, 557

isosteric heat, 567-568, 570

kinetic measurements, 582-584


kinetics, 572, 574-575

mean stay time, 550-551, 615

mobile, 557

monolayer and multilayer, 616

multimolecular layer, 558

net rate, 574

nondissociative, 575, 578

physical, 551, 553, 555, 561-562, 567

precursor state, 579-580, 582

Adsorptiondesorption theory, 348-352

Adsorption isotherms, 349-351, 367, 551-567

chemisorption, 569-570

dissociative Langmuir, 554-555, 571, 580

DubininRadushkevich, 617

equation, 617

Freundlich, 566-567

Gibbs, 563, 571

HillDeBoer equation, 565

Langmuir, 552-554, 579, 617

multilayer adsorption, 617-618

observed behavior, 568-572

physical, 568

Temkin, 566, 571, 617

Adsorption lifetime, 20

Adsorptive equilibrium, 559

Alcohol lathing, 619-620

Analytical approximations, 538-541

Anode current, 322, 329

Appearance potential, 448, 450


Argon instability, 337-339

Argon shower, 330, 338

Argon treatment, 330

Atmosphere, standard, 5

Atomic beam, 27

Auger electron spectroscopy, 608-609, 612, 620

Avogadro's law, 4, 6

Avogadro's number, 6-8, 70

determination, 17-18

BaAl system, constitutional diagram, 276

Back-diffusion coefficient, 221, 223, 230

Backing pump, 248-249, 527-529

UHV/XHV, 646-647

Backscattering, 261

Backstreaming:

oil, 499, 536

roots blower, 520

Backstreaming factor, 670

Ba getters, 276-291

diffusion process, 282-284

endothermic, 277

evaporation conditions effect, 287-288

exothermic, 278

flashing, 277-278

frittable, 290-291

< previous page page_715 next page >


< previous page page_716 next page >
Page 716

Ba getters (Continued)

gas-doped, 289-290

gas-surface reactions, 288-289

high-yield, 290

interaction with gases, 278-280

low-argon, 291

sorption

characteristics, 279-284

distribution, 288

temperature-dependence, 284-286

thickness-dependence, 285-287

stages of gas sorption, 282-283

sticking probability, 281

structure, 282

total yield, 290

BaH2 system, equilibrium isotherms, 284-285

Bakeout, 338, 614, 619-621

Bar, 5

Base pressure, 511

BayardAlpert ionization gauge, 417-419, 626, 634

calibration, 683

geometric variations, 419-421

high-pressure limit, 426

ion current linearity, 435

modulated, 421-422, 630, 635

Beam:

atomic, 27
molecular, 28

Beer's law, 469

Bent-beam gauge, 629-631

Bernouilli's equation, 121

Bessel box gauge, 631-632

Betatron oscillations, 72

Blasius relation, 112

Boltzmann constant, 3

Boltzmann equation, 134-135

modified, 56

Bombing test, 493

Bourdon gauge, 382-384

BoyleMariotte law, 659

Boyle's law, 3

Brownian motion, 69

particle distribution, 17-18

BrunauerEmmettTeller model, 558-559, 561-562

Calibration:

capacitance diaphragm gauges, 676-680

comparison method, 673-676

ionization gauges, 683-686

mass spectrometers, 686-689

spinning rotor gauges, 680-683

test leaks, 689-692

Calibration chamber, 665-666, 668-669

gas flow rate, 672

Calibration leaks, 502

Capacitance, parallel plate formula, 385


Capacitance diaphragm gauges, 384-389

accuracy, 387-388

advantages and disadvantages, 388-389

calibration, 676-680

deflection of thin tensioned membrane, 386-387

sensitivity, 385-386

thermal transpiration, 388

Capacitance manometer, see Capacitance diaphragm gauges

Capacity, definition, 356

Capillarity effects, 584-588

Capillary condensation, 562

Capillary leaks, 689

Capture coefficient, see Sticking coefficient

Capture pumps, UHX/XHV, 643-646

Capture vacuum pumps, 259

Carbons, activated, 584

Catalysts, 606

Cathode fall, 317, 319

Cathodes, hot, 636-638

Cavitation, liquid ring pumps, 154, 156

Cavity ringdown spectroscopy, 470

Ceto getters, 297

Characteristic fragment ions, 449

Charles' law, 4

Chemisorption, 263-265, 551, 553

kinetics, 575-582

dissociative, 579-580

equilibrium coverage, 579

fractional coverage, 578


homonuclear diatomic molecules, 575, 579-580

potential energy curve, 575-578

second-order, 581

sticking coefficient, 580, 582-584, 586

oxygen, 610-611

Choumoff gauge, 426, 428

ClausiusClapeyron equation, 351, 567

Claw pump, 162-164

Coefficient of heat conductivity, 42

Coefficient of interdiffusion, 63-64, 65

Coefficient of self-diffusion, 62

Coefficient of slip, 38-39

Coefficient of thermal separation, 58

Coefficient of viscosity:

compared with heat conductivity, 42-43

definition, 29

at low pressures, 37-39

relation to

mean free path, 30-32

molecular diameter, 31-32

unit, 30

variation with temperature, 32-33

Cold-cathode discharge, 318, 326

Cold-cathode gauges, 427-435

calibration, 430, 433

inverted-magnetron gauge, 429-434

magnetron gauge, 431, 433

< previous page page_716 next page >


< previous page page_717 next page >
Page 717

Penning discharge, 434

theory of crossed field discharge, 431

Collision, molecular, see Molecular collisions

Collision cross section, mutual, 8

Collision frequency, 33

Collision rate, 9

Comparison method, calibration, 673-676

Compressible flow, 116-121

adiabatic flow, 117

approximation for flow through aperture, 121

choked flow, 116-117

choked pressure ratio, 116

relationship between

entry and exit velocity, 118

throughput and entry velocity, 118

speed of choked aperture, 119

temperature changes, 120

through aperture or short duct, 119-121

time to vent chamber, 120-121

Compression ratio, 496, 521-522

Condensation coefficient, 22

Condensation rate, 22, 532

Condensations, 525

Conductance, 83-84

baffles and cold traps, 197-198

correction, 670

limited value in continuum flow, 107


molecular flow, 85-86

aperture, 86

end effect, 88

long ducts, 87

short ducts, 87-88

tubes, 88-90

transitional flow, 130-131

Conductivity, permeation, 517-518

Connections, graphic symbols, 705-706

Constant pressure technique, 662

Constant temperature Pirani gauge, 406-410

Continuity equation, 522

Continuous expansion, 659, 665-670

Continuum flow, 81-82, 105-128

assumption of incompressibility, 108

choked pressure ratio, 124-126

compressible, 105, 116-121

entrance correction model, 122-124

flow obstruction corrections, 121-122

friction factor, 128

kinetic energy model, 124-126

long duct criteria, 126-128

turbulent, 112-116

viscous flow, 108-109

Correction factor, 669

Cosine law, 19-20

Coulomb scattering, charged particles, 72

Counterflow helium leak detectors, 496-499

Coverage:
adlayer, 560

equilibrium, 579-580

fractional, 553, 578

Crossover, 494-496, 529

Cryo-condensation pumps, 347-348

UHV/XHV, 645

Cryopumps, 347-364

adsorptiondesorption theory, 348-352

boiling pool, 359

capacity, 355-357

configuration, 359-363

convection heat loads, 361

cryotrapping, 352-353

impulsive heatload, 362

open-loop, 359

partial regeneration, 364-365

pumping rate, 349

pumping speeds, 353-355

refrigeration technology, 357-359

regeneration, 363-364

sorption roughing pumps, 364-368

UHV/XHV, 645

ultimate pressure, 353-355

Cryopump sets, 535-537

Cryo-sorption, hydrogen by charcoal, 351-352

Cryo-sorption pumps, 347-348

limiting pressure, 645

UHV/XHV, 645-646

Cryosurfaces:
predicting performance, 542

pump combinations, 534-535

pumping speed, 531-534

Cryotrapping, 352-353

Cycloidal mass spectrometer, 465-466

DarcyWeisbach equation, 112

de Broglie wavelength, 73

Delay times, 587-588

Desorption, 261, 263-265

adsorbed gases, 614-615

electron-stimulated, 419

kinetics, effect on permeation, 600-605

net rate, 574

over barriers, 584-585

processes, 547-548

Desorption frequency, 548, 550, 572, 575

Diaphragm pump, 169-170, 499

Diffusion, 261, 265-267

activation energy, 592

Fick's second law, 266

finite slab, 598-599

Fisk's first law, 265-266

gases in Ba film, 282-284

outgassing control, 513

rates, 590-591

semi-infinite slab, 596-598, 618, 621

thermal, see Thermal diffusion

< previous page page_717 next page >


< previous page page_718 next page >
Page 718

Diffusion coefficient, 221-222, 265-266

determination, 65

temperature dependence, 65

Diffusionejector pumps, 183-185

Diffusion of gases, kinetic theory, 62-69

Diffusion pumps, 176-183

backstreaming, 194-195

baffle design, 197-198

boiler design, 182

cold cap, 182-183

cold traps, 231

compression ratio, 181, 190

cooling water flow, 200

design, 181-183

development, 176-181

dispersion characteristics, 228-229

entrance chamber design and speed, 191-192

fluid vapor pressure, 192-194

fractionating oil, 179-180, 192

heater input, 200

jet deflection, 220

jet flow pattern, 209, 211

multistage oil, 181

nozzle design and speed, 190-191

oils, vapor-pressure data, 200-201

pump fluid, properties, 207

pumping action theory, 202


speed dependence on design, 182

speed equation, 202-204

speed factor, 191

speed measurement, 186

UHX/XHV, 642-643

ultimate pressure, 193, 221

use in manometers, 378

vapor velocity at nozzle exit, 208

Diffusion pump sets, 528-531

Diffusion rate, Brownian particles, 70

Diffusivity, 591

apparent, 604

Dispersion forces, 347, 551

Displacement, 261

Distillation rate, variation with pressure, 25

Dry helium leak detector, 499-500

Dry vacuum pumps, 159-170

claw pump, 162-164

multistage piston pumps, 167, 169

roots pump, 159-162

screw pump, 164-165

scroll pump, 165-168

two-stage diaphragm pumps, 169-170

Dynamic method, 665

Effusion law factor, 191

Effusion rate, small orifice, 24

Elastic scattering, electron, 72

Electrical matrix, correspondence with vacuum matrix, 544


Electrolytic abrasive polishing, 619

Electron:

charge, 7

secondary, 326-327

Electron cloud, 324-326

drift velocity, 324

space charge, 324

Electron emitters, 607

Electron-stimulated desorption, 419, 630

Electrostatic ion pumps, 318

EleyRideal mechanism, 607, 612

Energetic neutrals, 337

theory, 330

Energy:

activation, 550

dissociation, 576

kinetic, 3

translational

conservation, 10

distribution formula, 14

mean, 20

Energy loss:

due to thermal conduction, 44

fractional, 71

Enskog's formula, 222

Enthalpy of vaporization, 561

Entrance correction model, 122-124

Equation of state, ideal gas, 4

Equilibrium pressure, 267-269


Error function, 597, 599

Evaporation rate, 22-25

metals, gas pressure effect, 67-69

Expansion by molecular beam, 659

Expansion ratio, 662, 664

Expansion technique, 662

disadvantage, 663-664

Extractor gauge, 422, 629-630

Extreme high vacuum:

capture pumps, 643-646

definition, 625

gas-phase parameters, 627

hardware, 652

kinetic pumps, 642-643

leak detection, 647-648

measurement limits, 628-642

comparison of gauges, 641-642

gauges with long electron paths, 639-640

hot cathode effects, 636-638

residual currents, 629-636

outgassing, 648-651

pump comparison, 646-647

pumping speed, 628

Fanno line, 224

Faraday constant, 6

Faraday cup ion detection, 454

Fick's first law, 590-592, 600

< previous page page_718 next page >


< previous page page_719 next page >
Page 719

Fick's second law, 591, 595-596, 598-599, 618

Fisk's first law, 265-266

Fisk's second law, 266

Flow, of gases, see Gas flow

Flowmeter, 694-695

Flux:

diffusive, 590-591

steady-state permeation, 602

Force, unit, 4

Force constants, in repulsive force relation, 59

Forepumps, pumping speed, 519-520

Forepump sets, 519-524

Fore-vacuum pumps, 519-520

FowlerNordheim equation, 327

FranckCondon principle, 448

Free-molecule conductivity, 46-50

Free path, mean, see Mean free path

Friction factor, 112-114, 128

Gain factor, secondary electron multiplier, 456

Gas:

active, interaction with getters, 275

dissolved, permeation, 616

effusion, 22

getterable, sputter-ion pumps, 335-336

interaction with Ba film, 278-280

noncondensable permanent, 534


permeation, 516-518

properties, 83

purge, 500, 536-537

tracer, leak detection, 493

see also Kinetic theory of gases

Gas ballasting, 144-147

Gas chromatograph, 587-588

Gas constant, 6

Gas density, 3

Gas desorption, 484

Gas discharge vacuum pumps, 317-319

Gas factor, 82-83

Gas flow, 81-137

conductance, 83-84

continuity assumption, 84

desorption, 502

equation, 543

free molecular, 81

leak, 482

permeation, 484, 488

pressure reduction, 474-475

regimes versus Knudsen number and pressure, 82

throughput, 83, 667

total, 494

transitional, 82, 128-137

see also Continuum flow; Molecular flow

Gas load, 547

high-vacuum systems, 513-519

process, 518-519
time-dependence, 514-515

Gas-phase parameters, 627

Gas pressure, 18-22

Gas-surface interactions, 547-548

types, 261-262

van der Waals interactions, 555

see also Adsorption

Gas volume, striking unit area per unit time, 21

Gay-Lussac's law, 4

Getter, bulk, 596

Getter effect, 321, 329

Gettering, 262

sputter-ion pumps, 329-330

Gettering capacity, 262

Gettering rate, 262, 298

terminal, 270

Gettering speed, 264

Getter-ion pumps, UHX/XHV, 643-645

Getter materials, 269-310

activation, 273-274

applications, 313-315

basic concepts, 262-263

characteristics, 269

evaporable, 271-272

applications, 314-315

see also Ba getters; Ti sublimation getter pumps

interaction with residual gases, 275

nonevaporable, 271-272, 297-310

activation, 303-304, 308-309


applications, 314-315

binary Zr and Ti alloys, 298-301

Ceto getters, 297

compressed and sintered structures, 310-313

configurations, 310-313

diffusion rate, temperature dependence, 300, 302

gettering rates, 298

gettering speeds, 301-302

multicomponent alloys, 310

sorption characteristics, 311, 313

surface analyses, 304

ternary allows, 305-309

ZrAl alloy, 299-304

ZrCo alloys, 305

ZrFe alloys, 305

ZrNi alloys, 305

ZrTi alloys, 305

ZrV alloys, 305

ZrVFe alloy, 306-309

reactivation, 274

sorption speed and capacity, 269-271

types, 271-275

Getter pumps, evaporable and nonevaporable, UHV/XHV, 645

Getters, 262

GiffordMcMahon cycle, 348

GiffordMcMahon refrigerator, 357-358

Glow discharge, 620

< previous page page_719 next page >


< previous page page_720 next page >
Page 720

Gravimetric technique, 662

Grazing incident, 337, 339

Gyromagnetic radius, 318

Halogen sensor, 485

Heat capacity, 41

Heat conductivity, see Thermal conductivity

Heat flow, between concentric cylinders, 360

Heat load, 532

Heat of adsorption, 621

Heat of solution, 590, 592

Helium leakage rates, standards, 503

Helium leak detector, 486-487, 494

counterflow, 496-499

detection limits, 498

direct-flow, 494-496

dry, 499-500

mass spectrometer, 494

measurement range, 502

oil-free, 499-500

recalibration, 502

sensitivity, 495-497

vacuum components, 487-490

vacuum systems, 490-492

Helmer gauge, 423-424

Henry's law, 552-553, 563, 565, 568, 574-575

High-pressure ionization gauges, 426-428


High-vacuum pump sets, 524-537

cryopump sets, 535-537

with cryosurfaces, 531-535

design, 525

diffusion pump, 528-531

pumping speed, 524-525

turbomolecular, 526-528

ultimate pressure, 525

High vacuum region, 9

High-vacuum systems, 507-546

calculation methods, 537-546

analytical approximations, 538-541

numerical methods, 541-546

equivalent electrical circuit, 544-545

fore-vacuum pumps, 519-520

outgassing, 513-516

roots combinations, 520-524

see also High-vacuum pump sets

HillDeBoer equation, 565

Ho coefficient, 191

Hot-cathode gauge equation, 414-419

Hot-cathode gauges, comparison, 641-642

Hydraulic diameter, 106

Hydrocarbons, interaction with getters, 275

Hydrogen:

atomic, 607-608

flux, Arrhenius plot, 602-603

sputter-ion pumps, 336-337

Hydrogenmetal systems, pressurecompositiontemperature curves, 268


Hydrogen permeation constant, 594-595

Ideal gas law, 2-6

differentiation, 689

three-dimensional, 564

Impedance, 83

Infrared absorption measurement, partial pressure, 469-470

Inlet system, 668

Interactions, adsorbateadsorbate, 567, 568

Interatomic spacing, 556

Inverted-magnetron ionization gauge, 429-434, 639-640

Ion burial, 330-331

Ion detection, 454-456

Faraday cup, 454

microchannel plate detector, 456-457

secondary electron multiplier, 454-456

Ionization gauges, 414-441

accuracy, 435-438

BayardAlpert gauge, 417-419, 626, 634

calibration, 683

geometric variations, 419-421

high-pressure limit, 426

ion current linearity, 435

modulated, 421-422, 630, 635

bent-beam, 629-631

Bessel box, 631-632

buried collector gauges, 420

calibration, 683-686

cold-cathode gauges, 427-435


controllers, 439-441

degas mode, 685

electron-stimulated desorption, 419

equation, 415-416

extractor gauge, 422, 629-630

factors influencing calibration, 437-438

gauge constant ratios for gases, 438-440

generalized, 415

Helmer gauge, 423-424

high-pressure, 426-428

hot-cathode gauge equation, 414-419

long electron path length gauges, 424-425, 639-640

outgassing rates, 672

as reference standards, 688

reverse x-ray effect, 419

secondary electron production, 685

secondary standard hot-cathode gauges, 425-426

as secondary standards, 684-685

sensitivity, 684

stability of calibration, 436

types, 414

x-ray effect, 417

Ionization potential, 448, 450

< previous page page_720 next page >


< previous page page_721 next page >
Page 721

Ionization process, electron-impact, 448-451

Ion motion, 323-324

Ions, drift velocity, 464

Ion sources, 447-453

closed, 452-453

electron-impact ionization process, 448-451

open, 449, 451-452

Isentropic flow, 119

Isolation devices, graphic symbols, 704

Jump frequency, 590-591

Katharometer, 44, 65

Kelvin equation, 585

Kinetic energy model, 124-126

Kinetic pumps, 642-643

Kinetics, quasi-first-order, 605

Kinetic theory of gases, 1-73

Avogadro's number, 6-8

derivation of pressure relation, 2

fundamental postulates, 1

ideal gas law, 2-6

molecular collisions, 8013

Kinetic vacuum pumps, 173-254

diffusion-ejector pumps, 183-185

diffusion pumps, 176-183

see also Vapor-jet pumps


Knudsen cell, 20

Knudsen correction factor, 669

Knudsen equation, 130-131, 133

Knudsen number, 9, 51, 54, 81-82

flow regimes versus, 82

transitional flow, 128-129

Lafferty gauge, 640

Lafferty magnetron gauge, 424-425

Langmuir equation, 562

LangmuirHinshelwood mechanism, 607, 612

Langmuir isotherm equation, 332

Langmuir model, 552-554

Langmuir's film theory of heat conduction, 52-53

Leakage rate, 490, 689-690

conversion, 503-504

measurement, helium leak detectors, 486-487

minimum detectable, 500

measured, uncertainties, 692

normalized, 484, 488

quantitative measurements, 493, 502-504

standards, 502-503

uncertainty, 502

units, 482-484

conversion factors, 483

Leak detection, 481-505

future developments, 504-505

gross leaks, 485

partial pressure measurements, 486


special methods, 493

total pressure measurements, 484-486

tracer gases, 493

UHV/XHV, 647-648

Leak detector:

calibration, 503

inlet pumping speed, 501

intrinsic partial pressure sensitivity, 500

partial-flow arrangement, 490-491

partial flow connection, 491-492

pumpdown time, 488, 501

quadrupole mass spectrometer, 492

total gas flow, 491

vacuum components, 487-490

vacuum systems, 490-492

see also Mass spectrometer, leak detectors

Leak gas flow, 482

Leak localization, 488

Leaks:

calibration, 502

flow conditions, 504

temperature dependence, 503

test, calibration, 689-692

types, 482

vacuum systems, 516-517

Lennard-Jones potential, 556

L'Hôpital's Rule, 208

Light, velocity, 7

Liquid, density, molecular diameter from, 40


Liquid manometers, 378-379, 659

Liquid nitrogen trap, 494-496

Liquid ring pumps, 151-158

accessories, 158

cavitation and protection against cavitation, 154, 156

drives, 157-158

materials of construction, 157

mechanism, 151-152

operating liquid, 154

conveyance, 156-157

operating ranges, 154-155

sealing, 157

single-stage, 152-153

two-stage, 153-154

types of operation, 156

LMF mode, see Penning discharge

Long electron path length gauges, 424-425

Loschmidt number, 7

Mach angle, 218

Magnetic focusing, charged particles, 72

Magnetic ionization gauges, 639-640

Magnetic sector mass spectrometer, 460-463

resolving power, 462-463

Magnetron gauge, 424-425, 431, 433

Magnetron ionization, 639

Manometer, liquid, 378-379

Masking, leaks, 488


< previous page page_721 next page >
< previous page page_722 next page >
Page 722

Mass analysis, 456-467

cycloidal mass spectrometer, 465-466

magnetic sector mass spectrometer, 460-463

omegatron, 466-467

quadrupole mass spectrometer, 456-460

time-of-flight analyzer mass spectrometer, 464-465

Mass spectra, 470-471

fragmentation ratio, 477

Mass spectrometer, 608-609, 612

calibration, 686-689

double-focusing, 633

leak detectors, 493-500

counterflow helium leak detectors, 496-498

advanced, 498-499

direct-flow helium leak detectors, 494-496

helium, 494

maximum tolerable inlet, 501-502

maximum tolerable total gas flow, 502

oil-free and dry helium leak detectors, 499-500

other tracer gases, 504-505

specifications, 500-502

quadrupole, 633-634

calibration, 689

sensitivity, 686

sniffing leak detectors, 493

Mathieu equations, 459

MaxwellBoltzmann distribution laws, 12-14


application, 15, 18

MaxwellLoschmidt method, 65-67

Maxwell's distribution law, 210

Mbar, 5

McLeod gauge, 379-381

linear mode, 381

quadratic mode, 380

sources of error, 381

Mean free path, 8, 26-28, 586-587

electrons, 41

experimental determination, 27

homogeneous maxwellian gas, 9

molecular collisions within unidirectional beam, 72

relation to coefficient of viscosity, 30-31

StefanMaxwell formula, 63

temperature effect, 32

values, 34

at very low pressures, 21

Mean stay time, adsorption, 615

Membrane, thin tensioned, deflection, 386-387

Mercury vapor, mean free path and molecular diameter, 34

Mercury vapor pumps, 177-178

Metals:

classification by adsorption properties, 574

evaporation rate, gas pressure effect, 67-69

Microchannel plate detector, 456-457

Molar flow rate, 6

Molar volume, 4

Molecular beam expansion, 670-673


Molecular beam scattering, 582-583

Molecular collisions, 8013

with electrons, cross section, 41

frequency, 33, 36

orthogonal, 11-12

striking surface, 18-22

velocity vectors, 9-10

at very low pressures, 21

Molecular diameter:

from density of solid or liquid, 40

gas mixtures, 64

relation to coefficient of viscosity, 31-32

values, 34, 40

from van der Waals' constant, 39

Molecular drag pump, 233-238, 499

backing pumps, 248-249

combined with turbomolecular pumps, 247-248

design considerations, 237

early, 233-234

maximum compression, 235-236

maximum pumping speed, 236

performance data of commercial pumps, 237-238

theoretical considerations and performance data, 234-237

Molecular flow, 85-105

angular distribution of molecules, 85-86

beaming effect, 85

conductance, 85-86

aperture, 86

end effect, 88
long ducts, 87

short ducts, 87-88

tubes, 88-90

time constant in unsteady flow, 103

unsteady flow cases, 102-105

Molecular flux, incident, 21

Molecular impingement rate, 548

Molecular mass, determination, 22

Molecular velocity, 12-13, 16-17

Molecules:

free paths, 26-28

masses, velocities, and rates of incidence, 16

number per monolayer, 36, 41

number per unit volume, 7

random motions, 69-71

Molthan angular distribution function, 212

Momentum, linear, conservation, 10

Multistage claw pumps, 162-164

Mutual diffusion coefficient, 63

Newton, 5

Noble gases:

interaction with getters, 275

sputter-ion pumps, 337-338

Nondestructive testing, 504

< previous page page_722 next page >


< previous page page_723 next page >
Page 723

Numerical methods, 541-546

dedicated software, 541-542

network approach, 542-546

Ohm's law, 543

Oil backstreaming, 499, 536

Oil contamination, 499

Oil-free helium leak detector, 499-500

Oil-sealed vacuum pumps, 143-149

accessories, 149

design, 143-144

gas ballast, 144-147

oil suckback, 148

power requirements and system protection, 148-149

pump oil, 147-148

Omegatron, 466-467

Optical measurements, partial pressures, 467-470

Orbitron pump, 645

Orifice flow method, 665

Oscillating quartz crystal viscosity gauge, 402-403

Outgassing, 513-516, 614-621

bakeout, 620-621

desorption of adsorbed gases, 614-615

dissolved gases, 616

flow rate, 538

in situ surface treatments, 620

mitigation, 619
pumpdown curves, 616-618

rate, 596-598, 672

rate reduction, UHV/XHV, 648-651

surface treatments during construction, 619-620

UHV/XHV, 648-651

Partial-flow factor, 490

Partial pressure analysis, 447-477

calibration of analyzers, 475-477

computer control, data acquisition, and presentation, 470-471

infrared absorption measurement, 469-470

ion detection, 454-456

ion sources, 447-453

laser multiphoton ionization, 468-469

optical measurement, 467-470

residual gas analysis, 471-474

vacuum process analysis, 474-475

see also Mass analysis

Partial pressure analyzers, calibration, 687

Partial pressure gauge, 486

Particle beams, scattering, 71-73

Particle distribution, at levels in gravitational field, 18

Partition functions, 550

Pascal, 5

Penning discharge, 318-329, 650

configuration, 343

discharge modes, 320

electron cloud, 324-326

electron transit and collision parameters, 434


ion motion, 323-324

pump sensitivity, 321-323

secondary electrons, 326-327

sputtering, 329

transition from HMF to HP mode, 327-328

transition from LMF to HMF mode, 328-329

Penning gauge, 427-429

Permeability, 484, 593-595

Permeability constant, 594-595

Permeation, 516-518, 615

dissolved gases, 616

effect of desorption kinetics, 600-605

kinetics, 591

steady-state, 592-595

surface contaminants, 604-605

transient, 595-600

Permeation conductivity, 517-518

Permeation gas flow, 484, 488

Permeation leaks, 689

Photoionization measurement, partial pressure, 468-469

Physiosorption, 551

Physisorption, 263-265

Pirani gauge, 406-410

comparison with thermocouple gauge, 412

Pirani leak detector, 410

Piston gauges, 659-661

Piston pressure balance gauge, 381-383

Piston pump, 167, 169

Plasma coating, anode, 330


Plasma processing, 620

Poisson's equation, 324, 327

Polymer materials, outgassing, 514-515

Positive displacement pump, 141-170

dry vacuum pumps, 159-170

liquid ring pumps, 151-158

oil-sealed, 143

Potential energy, 600-601

curve, 575-578

well, 549

Potential well, 556-557, 576, 578, 589, 600

PrandtlMeyer formula, 214, 216

PrandtlMeyer ratio, 216-217

Pressure:

allowable in synchrotron, 73

definition, 18, 657

effect on rates of evaporation of metals, 67-69

flow regimes versus, 82

kinetic theory relation, 2

limiting ratio, 53

measurement, 21

ranges, 658

time-dependence, 509

units, 5, 657

see also Partial pressure analysis

< previous page page_723 next page >


< previous page page_724 next page >
Page 724

Pressure gauges, total, 484

Pressure units:

conversion factors, 377, 707

vacuum measurements, 377

Process gas loads, vacuum systems, 518-519

Process pressure, vacuum systems, 511-513

Pumpdown, UHV range without bakeout, 649

Pumpdown curve, 510, 513, 616-618

Pumpdown equations, 508-511

Pumpdown time, 538-539

Pump fluid:

properties, 207

vapor pressure, 192-194

Pumping speed, see Volume throughput

Pumping system, 669-670

Pump orifice, 669

Pumps:

booster, 183-184

diffusion, 176-183

diffusion-ejector, 183-185

dry vacuum, 159-170

liquid ring, 151-158

oil-sealed vacuum, 143-149

see also specific pumps

Quadrupole mass spectrometer, 447, 456-460, 486, 631-632

closed ion source, 452-453


filter structures, 457-458

leak detection, 492

Mathieu equations, 459

operating line, 460

potential difference, 452

quadrupole potential, 457

resolving power, 459

Quadrupole potential, 457

Quartz helix Bourdon gauge, 383-384

Random motions, molecules, 69-71

Rate constants, 579, 581

Rayleigh line, 224

Reactivation, getter materials, 274

Readsorption, 616

Reference gauge, 673

Reference standard, 658, 676

Refrigeration, technology, 357-359

Refrigeration loss, 359

Refrigerator cryosurface combination, 533

Regeneration:

cryopumps, 363-364

partial, 364-365

Regenerative drag pumps, 251-254

Relaxation time, 21

Repulsive force constant, 60-61

Residual currents, 629-636

Residual gas analysis, 471-474

diagnosing leaks and contamination, 472


filament selection, 472

Residual gas analyzer, see Quadrupole mass spectrometer

Reverse x-ray effect, 419

Reynolds number, 105

in terms of throughput, 106

transitional flow, 128-129

transition from viscous laminar to turbulent flow, 106

units conversions, 107

Roots blower, 520

efficiency, 521

predicting performance, 542

pumping speed, 522-523

Roots combinations, 520-524

Roots pump, 159-162

Rutherford formula, scattering cross section, 72

Saturation time, 335

Scattering, particle beams, 71-73

Screw pump, 164-165

Scroll pump, 167-169, 499

Secondary electron multiplier detection, 454-456

Secondary standard hot-cathode gauges, 425-426

Semimetals, classification by adsorption properties, 574

Sensitivity, 475-477

Separation of gases, by thermal diffusion, 60

Sieverts' equation, 303, 307

Sieverts' law, 267-268

Sieverts' plots, H2, 299-300

SI system, pressure units, 377


Slip theory, 129

Sniffing device, 492

Sniffing technique, 505

Solid, density, molecular diameter from, 40

Solubility, 267-268

equilibrium, 590

Sorption:

Ba film properties, 279-284

processes, 547-548

titanium sublimation getter pumps, 292-295

Sorption capacity, 269-271

Sorption roughing pumps, 364-368

Sorption speed, 269-271

Sound, velocity, relation with molecular velocities, 16-17

Specific heat:

molar, ratio, 17

molecular, 42-43

ratio, 42-43

Spectrometer, linear response, 475

Speed factor, 191

< previous page page_724 next page >


< previous page page_725 next page >
Page 725

Spinning rotor gauge, 391-402

accommodation coefficient, calibration, 397-398

advantages and disadvantages, 401-402

calibration, 680-683

commercial, 394-397

fluctuations due to timing errors, 396

gassurface interaction assumptions, 391-392

head, 394-395

residual drag changes, 398-400

secondary or transfer standard, 399-401

stability, 397-400

theory, 391-394

Sputtering, 329

Sputtering pattern, 323, 337

Sputtering rate, 339

Sputter-ion pumps, 317

argon instability, 337-339

bakeout, 338

''built-in", 334, 341

diode-type pump, 338-339

discharge current as function of nitrogen pressure, 322

distributed, 332

gas discharge vacuum pumps, 317-319

getterable gases, 335-336

gettering, 329-330

hydrogen, 336-337

ion burial, 330-331


life, 329

magnetron-type, 345-346

magnet system, 345-346

memory effect, 343-344

modes, 320-321

noble gases, 337-338

Penning discharge, 318-329

starting pressure, 342-343

trapped electron density, 333

triode-type, 331, 340-341

types, 338-342

UHX/XHV, 643-644

ultimate pressure, 343-344

Sputter yield, 329-330

Standard

primary, 658-673

definition, 658

secondary

definition, 676

see also Reference standard

Standard atmosphere, 5

Static expansion, 659, 661-665

uncertainties, 665-666

Stay time, adsorption, 588

StefanMaxwell formula, mean free path, 63

Sticking coefficient, 22, 264, 334, 349

chemisorption, 580, 582-584, 586

Sticking probability, 264, 281

Surface chemical reactions, 606-614


classes, 606

cleanoff, 609

continuous expansion method, 665-670

EleyRideal mechanism, 607, 612

LangmuirHinshelwood mechanism, 607, 612

liquid manometers, 659

molecular beam expansion, 670-673

piston gauges, 659-661

standard, primary, 658

static expansion, 661-665

structure-sensitive or structure-insensitive, 606-607

water cycle, 612-614

Surface coverage, 264

Surface defects, 566

Surface diffusion, 587

Surface heterogeneity, 566-568

Surface lifetime, 578, 581, 615

mass spectrometric molecular beam techniques, 582-583

mean, 550, 587

Surface machining processes, 619-620

Surface reactions, 261

titanium sublimation getter pumps, 295

Surface roughness, 619

Surface tension, 563

Sutherland constant, 32-33

Tantalum, evaporation rate and vapor pressure, 23

Temperature:

absolute, 3, 6
definition, 2

discontinuity, 50

two-dimensional critical, 565

Thermal conductivity, 41-44

compared with coefficients of viscosity, 42-43

at low pressures, 44-53

free-molecule, 46-50

temperature discontinuity, 50-53

molecular, values, 48

variation with pressure, 45-46

Thermal conductivity gauges, 403-414

ambient temperature compensation, 411

applications, 413-414

calibration, 406-408, 410-411

comparison of Pirani and thermocouple gauges, 412

constant temperature Pirani gauge, 406-410

convection enhanced, 410-411

energy loss mechanisms, 403-404

energy transfer by radiation, 405

integrated transducers, 412-413

lowest useful pressure, 408-409

precautions for use, 414

stability, 408-409, 412

theory, 404-406

< previous page page_725 next page >


< previous page page_726 next page >
Page 726

Thermal conductivity gauges (Continued)

thermistor Pirani gauges, 412-413

upper pressure limit, 410-411

Thermal creep, 55

Thermal diffusion, 57-62

Thermal radiation, 532

Thermal transpiration, 53-57, 354

capacitance diaphragm gauges, 388

effect in

mass spectrometer inlet, 57

variable capacitance diaphragm transducers, 57

effect on vacuum microbalances, 54

variation with Knudsen number, 43-55

Thermal transpiration effect, 678, 684

Thermistor Pirani gauges, 412-413

Thermocouple gauge, comparison with Pirani gauge, 412

Thermomolecular flow, 53-57

Throttle valve, 495-496

Throughput, 83

units, 6

vapor-jet pumps, 198-202

Time constant, 500-501

diffusion time, 501

electrical, 501

gas transport, 501

vacuum, 511-512

Time-of-flight analyzer mass spectrometer, 464-465


Titanium sublimation getter pumps, 291-297

displacement phenomena, 295

peeling, 295

sorption

characteristics, 292

dependence on thickness, 293-295

temperature-dependence, 292-294

surface reactions, 295

types, 295-297

Torr, 5

Total pressure measurements, 484-486

Townsend avalanche, 326

Tracer gas, 486

leak detection, 493

rejection levels, 503-504

Transducers, 376

integrated, 412-413

Transitional flow, 128-137

conductance, 130-131, 133

Knudsen equation, 130-131, 133

Knudsen number, 128-129

leakage through small hole, 131

long duct criterion, 134-135

long ducts, 129-134

Reynolds number, 128-129

simple approximation, 133

slip theory, 129

through apertures and short ducts, 135-137

Transmission probability, 86
component combinations, 96-105

addition theorem, 100

entrance correction, 99

Oatley method, 98-99

overall conductance, 101

pump connected to chamber, 99-100

pump connected to chamber via second chamber, 102

pump connected to chamber via tubes or components, 102

series arrangement of different diameter tubes, 100-102

cross sections intermediate between rectangular and elliptical, 93

cylindrical annulus, 93-95

elliptical cross section, 92-93

other shapes, 94, 96-98

rectangular cross section, 90-92

triangular section, 94

tubes, 88-90

Trochoidal mass analyzer, 465-466

Tubes, graphic symbols, 705-706

Tungsten, evaporation rate and vapor pressure, 23, 68

Turbomolecular pumps, 238-247

backing pumps, 248-249

baking, 244

balancing and vibration, 243

bearing lubricant, 242

in combination with other pumps, 245

combined with molecular drag pumps, 247-248

cooling, 244

design, 241-243

drive systems, 243


magnetic rotor suspension, 242

operation in magnetic fields, 244-245

performance data of commercial pumps, 245-247

pumping corrosive gases, 245

pumping speed, 240-241, 246-247

pumping toxic or radioactive gases, 245

rotor and stator geometry, 241-242

rotor materials, 243

rotor suspension, 242

theoretical considerations and performance data, 239-241

UHX/XHV, 643

venting, 243-244

Turbomolecular pump sets, 526-528

Turbulent flow, 112-116

entrance correction model, 123-124

entry length, 116

friction factor, 112, 114

kinetic energy, model, 125-126

shape factor, 112

< previous page page_726 next page >


< previous page page_727 next page >
Page 727

Ultimate pressure, 511

Ultrahigh vacuum:

capture pumps, 643-646

definition, 625

gas-phase parameters, 627

hardware, 652

history, 625-626

kinetic pumps, 642-643

leak detection, 647-648

measurement limits, 628-642

comparison of gauges, 641-642

gauges with long electron paths, 639-640

hot cathode effects, 636-638

residual currents, 629-636

outgassing, 648-651

pump comparison, 646-647

pumping speed, 628

Ultrahigh-vacuum molecular beam scattering system, 582-583

Universal gas constant, 6

Vacuum chambers, graphic symbols, 704

Vacuum components:

graphic symbols, 701-707

helium leak detection, 487-490

Vacuum degassing, 619

Vacuum gauge constant, 684


Vacuum gauges, 375-376

Bourdon gauge, 382-384

calibration, 676-686

capacitance diaphragm gauges, 384-389

direct gauges, 375

graphic symbols, 707

indirect gauges, 375

liquid manometers, 378-379

McLeod gauge, 379-381

piston pressure balance gauge, 381-383

stability, 680

thermal conductivity gauges, 403-414

see also Ionization gauges; Viscosity gauges

Vacuum matrix, correspondence with electrical matrix, 544

Vacuum measurements, pressure units, 377

Vacuum pumps, graphic symbols, 701-703

Vacuum systems:

calculations, 507-513

helium leak detection, 490-492

leaks, 516-517

process gas loads, 518-519

process pressure, 511-513

pumpdown equations, 508-511

with two chambers, 543

see also High-vacuum systems

Vacuum time constant, 511-512

Valve modes of operation, graphic symbols, 705

van der Waals equation, 39-40

van der Waals forces, 347, 551


Vapor-jet pumps, 185-202

back-diffusion of gases from fore-vacuum, 221-223

baffles and traps, 195-198

boiler pressure dependence, 199

breakdown, 187-188

DeLaval nozzle, 225

equation of continuity for gas flow, 186

flow pattern, 205-221

deflection from nonexpanding verticle nozzle, 219

forepressure below limiting value, 209

Gaede's alpha formula, 214

Mach angle, 218

maximum discharge rate, 206-207

Molthan angular distribution function, 212

specific heat ratio, 207

spherical surface element, 210-211

static vapor pressure, 216-217

thermal velocity magnitude, 210

heater input and water cooling, 199-201

impact pressure, 224-227

limiting forepressure, 187-190, 204-205

formula, 228, 230

nozzle design, 219, 225

performance in forepressure breakdown region, 223-231

pump fluid, backstreaming and back migration, 194-198

pumping speed, 185-187

shock boundary, 204, 227-228

speed for various gases, 192

speed measurement, 190


throughput, 198-202

ultimate pressure, 192-194

vapor pressure of condensed fluid, 231

Vapor pressure, 22-25

curves

common gases, 708-709

solid and liquid elements, 710-712

determination, 22

equilibrium, 561, 585

gases as function of temperature, 352

temperatures, 645-646

water, 531-532

Velocity, molecular, root-mean-square, 3

Viscosity:

kinematic, 35

at low pressures, 37-39

Viscosity gauges, 389-403

oscillating quartz crystal, 402-403

precautions for use, 401

spinning rotor gauge, 391-402

types, 390

Viscositytemperature functions, constants, 35

Viscous flow, 82

< previous page page_727 next page >


< previous page page_728
Page 728

Viscous laminar flow, 108-111

artificial conductances, 110

circular cross-sectional tube, 110-111

concentric cylindrical annulus, 111

eccentric cylindrical annulus, 111

entrance correction model, 123

entry length, 108

friction factor, 113

kinetic energy model, 124-125

rectangular cross section, 111

Volume throughput, 331-335, 616

approximation, 538

cryosurfaces, 531-534

effective, 508-509, 515, 542

measurement, 692-695

noble gas, 330-331, 337

saturated, 335

time dependence, 335

Volumetric speed, see Pumping speed

Water cooling, vapor-jet pumps, 199-200

Water vapor, mean free path and molecular diameter, 34

X-ray effect, 417, 626

Young and Hession triggered discharge gauge, 428-430


Z

Zeolites, 584

< previous page page_728

You might also like