0% found this document useful (0 votes)
210 views528 pages

The ESD Control Program Handbook

Uploaded by

b3ry 17
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
210 views528 pages

The ESD Control Program Handbook

Uploaded by

b3ry 17
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 528

The ESD Control Program Handbook

The ESD Control Program Handbook

Jeremy M Smallwood
Electrostatic Solutions Ltd
Southampton, UK
This edition first published 2020
© 2020 John Wiley & Sons Ltd

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in
any form or by any means, electronic, mechanical, photocopying, recording or otherwise, except as permitted by
law. Advice on how to obtain permission to reuse material from this title is available at http://www.wiley.com/go/
permissions.

The right of Jeremy M Smallwood to be identified as the author of this work has been asserted in accordance
with law.

Registered Offices
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

Editorial Office
The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK

For details of our global editorial offices, customer services, and more information about Wiley products visit us at
www.wiley.com.

Wiley also publishes its books in a variety of electronic formats and by print-on-demand. Some content that
appears in standard print versions of this book may not be available in other formats.

Limit of Liability/Disclaimer of Warranty


In view of ongoing research, equipment modifications, changes in governmental regulations, and the constant
flow of information relating to the use of experimental reagents, equipment, and devices, the reader is urged to
review and evaluate the information provided in the package insert or instructions for each chemical, piece of
equipment, reagent, or device for, among other things, any changes in the instructions or indication of usage and
for added warnings and precautions. While the publisher and authors have used their best efforts in preparing this
work, they make no representations or warranties with respect to the accuracy or completeness of the contents of
this work and specifically disclaim all warranties, including without limitation any implied warranties of
merchantability or fitness for a particular purpose. No warranty may be created or extended by sales
representatives, written sales materials or promotional statements for this work. The fact that an organization,
website, or product is referred to in this work as a citation and/or potential source of further information does not
mean that the publisher and authors endorse the information or services the organization, website, or product
may provide or recommendations it may make. This work is sold with the understanding that the publisher is not
engaged in rendering professional services. The advice and strategies contained herein may not be suitable for
your situation. You should consult with a specialist where appropriate. Further, readers should be aware that
websites listed in this work may have changed or disappeared between when this work was written and when it is
read. Neither the publisher nor authors shall be liable for any loss of profit or any other commercial damages,
including but not limited to special, incidental, consequential, or other damages.

Library of Congress Cataloging-in-Publication Data

Names: Smallwood, J. M. (Jeremy M.), author.


Title: The ESD control program handbook / Jeremy M Smallwood, Electrostatic
Solutions Ltd, Southampton, UK.
Description: First edition. | Hoboken, NJ : John Wiley & Sons, Inc., [2020]
| Series: IEEE | Includes bibliographical references and index.
Identifiers: LCCN 2020006384 (print) | LCCN 2020006385 (ebook) | ISBN
9781118311035 (hardback) | ISBN 9781118694572 (adobe pdf) | ISBN
9781118694558 (epub)
Subjects: LCSH: Electronic apparatus and appliances–Protection. | Electric
discharges. | Electrostatics.
Classification: LCC TK7870 .S5265 2020 (print) | LCC TK7870 (ebook) | DDC
621.381/044–dc23
LC record available at https://lccn.loc.gov/2020006384
LC ebook record available at https://lccn.loc.gov/2020006385

Cover Design: Wiley


Cover Image: © pinging/Getty Images

Set in 9.5/12.5pt STIXTwoText by SPi Global, Chennai, India

10 9 8 7 6 5 4 3 2 1
To Jan, who has often put up with my grumpy non-communication while I’ve been writing
this book. To Alia, now making her own life journey. To Caroline, who tragically died so
young.
To the subject of electrostatics and ESD, that has kept me occupied, perplexed and challenged
for many hours.
To the many people who have attended my courses and asked so many awkward questions
that helped me understand while trying to explain. To my fellow ESD practitioners whose
opinions and expertise give us many interesting, and sometimes heated, debates – and who
mainly agree the answer to most electrostatic questions is – “it depends.”
vii

Contents

Introduction xxvii
Foreword xxxiii
Preface xxxvii
Acknowledgments xxxix

1 Definitions and Terminology 1


1.1 Scientific Notation and SI Unit Prefixes 1
1.2 Charge, Electrostatic Fields, and Voltage 2
1.2.1 Charge 2
1.2.2 Ions 3
1.2.3 Dissipation and Neutralization of Electrostatic Charge 3
1.2.4 Voltage (Potential) 3
1.2.5 Electric or Electrostatic Field 4
1.2.6 Gauss’s Law 5
1.2.7 Electrostatic Attraction (ESA) 6
1.2.8 Permittivity 6
1.3 Electric Current 6
1.4 Electrostatic Discharge (ESD) 7
1.4.1 ESD Models 7
1.4.2 Electromagnetic Interference (EMI) 7
1.5 Earthing, Grounding, and Equipotential Bonding 7
1.6 Power and Energy 8
1.7 Resistance, Resistivity, and Conductivity 8
1.7.1 Resistance 8
1.7.2 Resistivity and Conductivity 9
1.7.2.1 Surface Resistivity and Surface Resistance 9
1.7.2.2 Volume Resistance, Volume Resistivity, and Conductivity 10
1.7.3 Insulators, Conductors, Conductive, Dissipative, and Antistatic Materials 11
1.7.4 Point-to-Point Resistance 13
1.7.5 Resistance to Ground 13
1.7.6 Combination of Resistances 13
1.8 Capacitance 14
1.9 Shielding 15
viii Contents

1.10 Dielectric Breakdown Strength 15


1.11 Relative Humidity and Dew Point 15
References 16

2 The Principles of Static Electricity and Electrostatic Discharge (ESD)


Control 17
2.1 Overview 17
2.2 Contact Charge Generation (Triboelectrification) 17
2.2.1 The Polarity and Magnitude of Charging 18
2.3 Electrostatic Charge Build-Up and Dissipation 18
2.3.1 A Simple Electrical Model of Electrostatic Charge Build-Up 19
2.3.2 Capacitance Is Variable 20
2.3.3 Charge Decay Time 22
2.3.4 Conductors and Insulators Revisited 23
2.3.5 The Effect of Relative Humidity 24
2.4 Conductors in Electrostatic Fields 25
2.4.1 Voltage on Conducting and Insulating Bodies and Surfaces 25
2.4.2 Electrostatic Field in Practical Situations 25
2.4.3 Faraday Cage 27
2.4.4 Induction: An Isolated Conductive Object Attains a Voltage When in an Electric
Field 27
2.4.5 Induction Charging: An Object Can Become Charged by Grounding It 29
2.4.6 Faraday Pail and Shielding of Charges Within a Closed Object 30
2.5 Electrostatic Discharges 30
2.5.1 ESD (Sparks) Between Conducting Objects 30
2.5.2 ESD from Insulating Surfaces 31
2.5.3 Corona Discharge 32
2.5.4 Other Types of Discharge 32
2.6 Common Electrostatic Discharge Sources 32
2.6.1 ESD from the Human Body 33
2.6.2 ESD from Charged Conductive Objects 34
2.6.3 Charged Device ESD 35
2.6.4 ESD from a Charged Board 36
2.6.5 ESD from a Charged Module 36
2.6.6 ESD from Charged Cables 37
2.7 Electronic Models of ESD 37
2.8 Electrostatic Attraction (ESA) 41
2.8.1 ESA and Particle Contamination 41
2.8.2 Neutralization of Surface Voltages by Air Ions 42
2.8.3 Ionizers 43
2.8.4 Rate of Charge Neutralization 43
2.8.5 The Region of Effective Charge Neutralization Around an Ionizer 44
2.8.6 Ionizer Balance and Charging of a Surface by an Unbalanced Ionizer 44
2.9 Electromagnetic Interference (EMI) 45
2.10 How to Avoid ESD Damage of Components 45
Contents ix

2.10.1 The Circumstances Leading to ESD Damage of a Component 45


2.10.2 Risk of ESD Damage 46
2.10.3 The Principles of ESD Control 46
References 47
Further Reading 49

3 Electrostatic Discharge–Sensitive (ESDS) Devices 51


3.1 What Are ESDS Devices? 51
3.2 Measuring ESD Susceptibility 53
3.2.1 Modeling Electrostatic Discharges 53
3.2.2 Standard ESD Susceptibility Tests 54
3.2.3 ESD Withstand Voltage 55
3.2.4 HBM Component Susceptibility Test 55
3.2.5 System Level Human Body ESD Susceptibility Test 56
3.2.6 MM Component Susceptibility Test 58
3.2.7 CDM Component Susceptibility Test 60
3.2.8 Comparison of Test Methods 62
3.2.9 Failure Criteria Used in ESD Susceptibility Tests 64
3.2.10 Transmission Line Pulse Techniques 64
3.2.11 The Relation Between ESD Withstand Voltage and ESD Damage 65
3.2.12 Trends in Component ESD Tests 66
3.3 ESD Susceptibility of Components 66
3.3.1 Introduction 66
3.3.2 Latent Failures 66
3.3.3 Built-in On-chip ESD Protection and ESD Protection Targets 68
3.3.4 ESD Sensitivity of Typical Components 70
3.3.5 Discrete Devices 71
3.3.6 The Effect of Scaling 71
3.3.7 Package Effects 71
3.4 Some Common Types of ESD Damage 72
3.4.1 Failure Mechanisms 72
3.4.2 Breakdown of Thin Dielectric Layers 73
3.4.3 MOSFETs 73
3.4.4 Susceptibility to Electrostatic Fields and Breakdown Between Closely Spaced
Conductors 74
3.4.5 Semiconductor Junctions 75
3.4.6 Field Effect Structures and Nonconductive Device Lids 76
3.4.7 Piezoelectric Crystals 76
3.4.8 LEDs and Laser Diodes 76
3.4.9 Magnetoresistive Heads 77
3.4.10 MEMS 77
3.4.11 Burnout of Device Conductors or Resistors 77
3.4.12 Passive Components 78
3.4.13 Printed Circuit Boards and Assemblies 78
3.4.14 Modules and System Components 79
x Contents

3.5 System-Level ESD 80


3.5.1 Introduction 80
3.5.2 The Relationship Between System Level Immunity and Component ESD
Withstand 80
3.5.3 Charged Cable ESD (Cable Discharge Events) 81
3.5.4 System-Efficient ESD Design (SEED) 81
References 82
Further Reading 88

4 The Seven Habits of a Highly Effective ESD Program 91


4.1 Why Habits? 91
4.2 The Basis of ESD Protection 92
4.3 What Is an ESDS Device? 92
4.4 Habit 1: Always Handle ESDS Components Within an EPA 93
4.4.1 What Is an EPA? 93
4.4.2 Defining the EPA Boundary 95
4.4.3 Marking the EPA Boundary 95
4.4.4 What Is an Insignificant Level of ESD Risk? 97
4.4.5 What Are the Sources of ESD Risk? 97
4.4.6 What ESD Protection Measures Are Needed in the EPA? 98
4.4.7 Who Will Decide What ESD Protection Measures Are Required? 98
4.5 Habit 2: Where Possible, Avoid Use of Insulators Near ESDS 99
4.5.1 What Is an Insulator? 99
4.5.2 Essential and Nonessential Insulators 100
4.5.3 Remove Nonessential Insulators from the Vicinity of ESDS 101
4.6 Habit 3: Reduce ESD Risks from Essential Insulators 104
4.6.1 What Is an Insulator? 104
4.6.2 Insulators Cannot Be Grounded 104
4.6.3 What to Do About ESD Risk from Essential Insulators 104
4.6.4 Using Ionizers to Reduce Charge Levels on Insulators 106
4.7 Habit 4: Ground Conductors, Especially People 107
4.7.1 What Is a Conductor? 107
4.7.2 Conductive, Dissipative, or Insulative? 107
4.7.3 Properties of a Conductor 108
4.7.4 Charge and Voltage Decay Time 108
4.7.5 The Importance of Material Contact Resistance in Protecting ESDS 109
4.7.5.1 Reduction of Energy Delivered from Conductor in ESD 109
4.7.5.2 Reduction of Peak Current in a Discharge 109
4.7.5.3 Specification of a Minimum Material Resistance 109
4.7.6 Safety Considerations 110
4.7.7 Elimination of ESD by Grounding and Equipotential Bonding 110
4.7.8 Understanding the Grounding (Earth) System 110
4.7.8.1 Types of Ground 110
4.7.8.2 The Grounding System 111
4.7.9 Grounding Personnel Handling ESDS Devices 111
Contents xi

4.7.9.1 Basic Requirements for Grounding Personnel Handling ESDS Devices 111
4.7.9.2 Grounding Personnel via a Wrist Strap 112
4.7.9.3 Grounding Personnel via Footwear and Flooring 112
4.7.9.4 Grounding Seated Personnel 113
4.7.10 Grounding ESD Control Equipment 113
4.7.10.1 General Considerations 113
4.7.10.2 Work Surfaces 113
4.7.10.3 Floors 116
4.7.10.4 Carts, Racks, and Other Floor Standing Work Surfaces 117
4.7.10.5 Seats 118
4.7.10.6 Tools 119
4.7.10.7 Gloves and Finger Cots 120
4.7.10.8 ESD Control Garments 121
4.7.11 What If a Conductor Cannot Be Grounded? 122
4.8 Habit 5: Protect ESDS Using ESD Packaging 122
4.8.1 Don’t Take Ordinary Packaging Materials into an EPA 122
4.8.2 The Basic Functions of ESD Packaging 123
4.8.3 Open ESD Packaging Only Within an EPA 123
4.8.4 Don’t Put Papers or Other Unsuitable Material in a Package with an ESDS
Device 124
4.9 Habit 6: Train Personnel to Know How to Use ESD Control Equipment and
Procedures 124
4.9.1 Why Train People? 124
4.9.2 Who Needs ESD Training? 125
4.9.3 What Training Do They Need? 126
4.9.4 Refresher Training 126
4.10 Habit 7: Check and Test to Make Sure Everything Is Working 126
4.10.1 Why Do We Need to Check and Test? 126
4.10.2 What Needs to Be Tested? 127
4.10.3 ESD Control Product Qualification 127
4.10.4 ESD Control Product or System Compliance Verification 127
4.10.5 Test Methods and Pass Criteria 127
4.10.6 How Often Should ESD Control Items Be Tested? 128
4.11 The Seven Habits and ESD Standards 129
4.12 Handling Very Sensitive Devices 129
4.13 Controlling Other ESD Sources 130
References 131
Further Reading 132

5 Automated Systems 133


5.1 What Makes Automated Handling and Assembly Different? 133
5.2 Conductive, Static Dissipative, and Insulative Materials 134
5.3 Safety and AHE 134
5.4 Understanding the ESD Sources and Risks 135
5.5 A Strategy for ESD Control 136
xii Contents

5.5.1 General Principles of ESD Control in AHE 136


5.5.2 The Conditions Leading to ESD Damage 136
5.5.3 Strategies for ESD Control in Automated Equipment 136
5.5.4 Qualification of ESD Control Measures 137
5.5.5 Compliance Verification of ESD Control Measures 138
5.5.6 ESD Training Implications 138
5.5.7 Modification of Existing AHE 138
5.6 Determination and Implementation of ESD Control Measures in AHE 138
5.6.1 Define the Critical Path of ESDS 138
5.6.2 Examine the Critical Path and Identify ESD Risks 139
5.6.3 Determine Appropriate ESD Control Measures 139
5.6.3.1 General Control Measures 139
5.6.4 Include ESD Control in New Equipment Specification 140
5.6.5 Document the ESD Control Measures Used in the Machine 140
5.6.6 Implement Maintenance and Compliance Verification of ESD Control
Measures 140
5.7 Materials, Techniques, and Equipment Used for ESD Control in AHE 141
5.7.1 Grounding All Conductors That Make Contact with ESDS 141
5.7.2 Isolated Conductors 142
5.7.3 Preventing Induced Voltages on ESDS Devices 142
5.7.4 Reducing Tribocharging of ESDS Devices 143
5.7.5 Using Resistive Contact Materials to Limit Charged Device ESD Current 144
5.7.6 Anodization 144
5.7.7 Bearings 145
5.7.8 Conveyor Belts 145
5.7.9 Using Ionizers to Reduce Charge Levels on ESDS Devices, Essential Insulators,
and Isolated Conductors 146
5.7.10 Vacuum Pickers 146
5.8 ESD Protective Packaging 147
5.9 Measurements in AHE 147
5.9.1 Overview of Measurements in AHE 147
5.9.2 Resistance Measurements 148
5.9.2.1 Overview of Resistance Measurements in AHE 148
5.9.2.2 Resistance Meters 148
5.9.2.3 Point-to-Point Resistance of Surfaces 149
5.9.2.4 Resistance to Ground from Surfaces and Machine Parts 149
5.9.2.5 Resistance to Ground of a Potentially Isolated Conductor 149
5.9.2.6 Measurements on ESD Protective Packaging 149
5.9.3 Electrostatic Field and Voltage Measurements 149
5.9.3.1 Voltage Measurement Instruments 149
5.9.3.2 Voltages on Small Conductors or ESDS Devices 150
5.9.3.3 Voltages on ESDS Devices 150
5.9.4 Charge Measurements 151
5.9.5 Measurement of the Voltage Decay Time and Offset Voltage Due to
Neutralization by an Ionizer 151
Contents xiii

5.9.6 ESD Current Measurements 151


5.9.7 Detection of ESD Using EMI Detectors 151
5.10 Handling Very Sensitive Components 152
References 152
Further Reading 154

6 ESD Control Standards 157


6.1 Introduction 157
6.2 The Development of ESD Control Standards 157
6.3 Who Writes the Standards? 159
6.4 The IEC and ESDA Standards 161
6.4.1 Standards Numbering 161
6.4.2 The Language of Standards 161
6.4.3 Definitions Used in Standards 162
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 164
6.5.1 Background 164
6.5.2 Documentation and Planning 165
6.5.3 Technical Basis of the ESD Control Program 165
6.5.4 Personal Safety 166
6.5.5 ESD Coordinator 166
6.5.6 Tailoring the ESD Program 166
6.5.7 The ESD Control Program Plan 166
6.5.8 Training Plan 167
6.5.9 Product Qualification Plan 167
6.5.10 Compliance Verification Plan 167
6.5.11 Test Methods 168
6.5.12 ESD Control Program Plan Technical Requirements 168
6.5.12.1 Safety 168
6.5.12.2 Grounding and Bonding Systems 171
6.5.12.3 Personnel Grounding 171
6.5.12.4 ESD-Protected Areas (EPAs) 171
6.5.12.5 Equipment Used in the EPA 172
6.5.12.6 Insulators 172
6.5.12.7 Isolated Conductors 172
6.5.12.8 Hand Electrical Soldering and Desoldering Tools 172
6.5.13 ESD Packaging 173
6.5.14 Marking 174
References 174
Further Reading 177

7 Selection, Use, Care, and Maintenance of Equipment and Materials


for ESD Control 179
7.1 Introduction 179
7.1.1 Selection and Qualification of Equipment 179
7.1.2 Use 180
xiv Contents

7.1.3 Cleaning, Care, and Maintenance of Equipment 180


7.1.4 Compliance Verification 181
7.2 ESD Control Earth (Ground) 181
7.2.1 What Does the ESD Control Earth Do? 181
7.2.2 Choosing an ESD Control Earth 181
7.2.3 Qualification of ESD Control Earth 183
7.2.4 Compliance Verification of ESD Control Earth 183
7.2.5 Common Problems with Ground Connections 183
7.3 The ESD Control Floor 183
7.3.1 What Does an ESD Control Floor Do? 183
7.3.2 Permanent ESD Control Floor Material 184
7.3.3 Semipermanent or Nonpermanent ESD Control Floor Materials 184
7.3.4 Selection of Floor Materials 185
7.3.5 Floor Material Qualification Test 185
7.3.6 Acceptance of a Floor Installation 187
7.3.7 Use of Floor Materials 187
7.3.8 Care and Maintenance of Floors 187
7.3.9 Compliance Verification Test 188
7.3.10 Common Problems 188
7.4 Earth Bonding 189
7.4.1 The Role of Earth Bonding Points 189
7.4.2 Selection of Earth Bonding Points 190
7.4.3 Qualification of Earth Bonding Points 190
7.4.4 Use of Earth Bonding Points 190
7.4.5 Compliance Verification of Earth Bonding Points 190
7.5 Personal Grounding 190
7.5.1 What Is the Purpose of Personal Grounding? 190
7.5.2 Personal Grounding and Electrical Safety 191
7.5.3 Wrist Straps 192
7.5.3.1 Conventional Wrist Strap Systems 192
7.5.3.2 Constant (Continuous) Monitor Wrist Strap Systems 193
7.5.3.3 Cordless Wrist Straps 194
7.5.3.4 Wrist Strap System Selection 194
7.5.3.5 Wrist Strap System Use 195
7.5.3.6 Wrist Strap Qualification 195
7.5.3.7 Wrist Strap Cleaning and Maintenance 195
7.5.3.8 Wrist Strap System Compliance Verification Test 195
7.5.3.9 Common Problems 197
7.5.4 Footwear and Flooring Grounding 197
7.5.4.1 The Importance of Footwear and Flooring in ESD Control 197
7.5.4.2 Types of Footwear 198
7.5.4.3 Footwear and Safety 199
7.5.4.4 Selection of Footwear 200
7.5.4.5 Qualification of Footwear 200
7.5.4.6 Use of Footwear 201
Contents xv

7.5.4.7 Compliance Verification of Footwear 201


7.5.4.8 Common Problems 202
7.5.5 Grounding via ESD Control Seating 202
7.5.6 Personal Grounding via an ESD Garment 203
7.5.6.1 Qualification of a Groundable Static Control Garment 203
7.5.6.2 Compliance Verification of ESD Control Garments 203
7.6 Work Surfaces 203
7.6.1 What Does a Work Surface Do? 203
7.6.2 Types of Work Surfaces 204
7.6.3 Selection of a Work Surface 204
7.6.4 Workstation Qualification Test 205
7.6.5 Acceptance of Work Surfaces 206
7.6.6 Cleaning and Maintenance of Work Surfaces 206
7.6.7 Compliance Verification Test of Work Surfaces 206
7.6.8 Common Problems 206
7.7 Storage Racks and Shelves 207
7.7.1 Should It be an EPA Rack or Shelf? 207
7.7.2 Selection, Care, and Maintenance of Racks, and Shelves 208
7.7.3 Qualification Test of EPA Shelves and Racks 209
7.7.4 Acceptance of Shelves and Racks 209
7.7.5 Cleaning and Maintenance of Shelves and Racks 209
7.7.6 Compliance Verification Test of Shelves and Racks 209
7.7.7 Common Problems 209
7.8 Trolleys, Carts, and Mobile Equipment 210
7.8.1 Types of Trolleys, Carts and Mobile Equipment 210
7.8.2 Selection, Care, and Maintenance of Trolleys, Carts, and Mobile
Equipment 211
7.8.3 Qualification of Trolleys, Carts, and Mobile Equipment 212
7.8.4 Compliance Verification of Trolleys, Carts and Mobile Equipment 213
7.8.5 Common Problems 213
7.9 Seats 214
7.9.1 What Is an ESD Control Seat for? 214
7.9.2 Types of ESD Seating 214
7.9.3 Selection of Seating 214
7.9.4 Qualification Test of Seating 215
7.9.5 Cleaning and Maintenance of Seating 215
7.9.6 Compliance Verification Test of Seating 215
7.9.7 Common Problems 215
7.9.8 Personal Grounding via ESD Control Seating 215
7.10 Ionizers 217
7.10.1 What Does an Ionizer Do? 217
7.10.2 Ion Sources 217
7.10.3 Types of Ionizer System 218
7.10.4 Selection of Ionizers 219
7.10.5 Qualification Test of Ionizers 220
xvi Contents

7.10.6 Cleaning and Maintenance of Ionizers 221


7.10.7 Compliance Verification Test of Ionizers 221
7.10.8 Common Problems 222
7.11 ESD Control Garments 222
7.11.1 What Does an ESD Control Garment Do? 222
7.11.2 Types of ESD Control Garments 223
7.11.3 Selection of ESD Control Garments 226
7.11.4 Qualification Test of ESD Control Garments 227
7.11.5 Use of ESD Control Garments 228
7.11.6 Cleaning and Maintenance of ESD Control Garments 228
7.11.7 Compliance Verification of ESD Control Garments 228
7.11.8 Personal Grounding via an ESD Garment 229
7.12 Hand Tools 229
7.12.1 Why Have ESD Hand Tools? 229
7.12.2 Types of Hand Tool 229
7.12.3 Qualification Test of Hand Tools 230
7.12.4 Use of Hand Tools 230
7.12.5 Compliance Verification Test of Hand Tools 231
7.12.6 Common Problems with ESD Control Hand Tools 231
7.13 Soldering or Desoldering Irons 231
7.13.1 ESD Control Issues with Soldering or Desoldering Irons 231
7.13.2 Qualification of Soldering Irons 231
7.13.3 Compliance Verification of Soldering Irons 231
7.14 Gloves and Finger Cots 232
7.14.1 Why Have Gloves and Finger Cots? 232
7.14.2 Types of Gloves and Finger Cots 232
7.14.3 Selection of Gloves or Finger Cots for ESD Control 232
7.14.4 Qualification Test of Gloves and Finger Cots 233
7.14.5 Cleaning and Maintenance of Gloves 234
7.14.6 Compliance Verification Test of Gloves and Finger Cots 234
7.14.7 Common Problems with Gloves and Finger Cots 234
7.15 Marking of ESD Control Equipment 234
References 235
Further Reading 237

8 ESD Control Packaging 239


8.1 Why Is Packaging Important in ESD Control? 239
8.2 Packaging Functions 241
8.3 ESD Control Packaging Terminology 242
8.3.1 Terminology in General Usage 242
8.4 ESD Packaging Properties 243
8.4.1 Triboelectric Charging 243
8.4.2 Surface Resistance 244
8.4.3 Volume Resistance 244
8.4.4 Electrostatic Field Shielding 245
Contents xvii

8.4.5 ESD Shielding 247


8.5 Use of ESD Protective Packaging 247
8.5.1 The Importance of ESD Packaging Properties 247
8.5.1.1 Charge Generation and Retention 247
8.5.1.2 Electrostatic Field Shielding 248
8.5.1.3 Electrostatic Discharge Shielding 248
8.5.2 Packaging Used Within the EPA 248
8.5.3 Packaging Used to Protect ESDS Outside the EPA 249
8.5.4 Packaging Used for Non-ESD Susceptible Items 249
8.5.5 Avoiding Charged Cables and Modules 250
8.6 Materials and Processes Used in ESD Protective Packaging 250
8.6.1 Introduction 250
8.6.2 Antistats, Pink Polythene, and Low-Charging Materials 250
8.6.3 Static Dissipative and Conductive Polymers 251
8.6.4 Intrinsically Conductive or Dissipative Polymers 253
8.6.5 Metallized Film 253
8.6.6 Anodized Aluminum 253
8.6.7 Vacuum Forming of Filled Polymers 253
8.6.8 Injection Molding 254
8.6.9 Embossing 254
8.6.10 Vapor Deposition 254
8.6.11 Surface Coating 254
8.6.12 Lamination 254
8.7 Types and Forms of ESD Protective Packaging 254
8.7.1 Bags 254
8.7.1.1 Pink Polythene Bags 256
8.7.1.2 Conductive (Black Polythene) Bags 256
8.7.1.3 Metalized ESD Shielding Bags 256
8.7.1.4 Moisture Barrier Bags 257
8.7.2 Bubble Wrap 258
8.7.3 Foam 258
8.7.4 Boxes, Trays, and PCB Racks 258
8.7.5 Tape and Reel 260
8.7.6 Sticks (Tubes) 261
8.7.7 Self-Adhesive Tapes and Labels 262
8.8 Packaging Standards 262
8.8.1 ESD Control and Protection Packaging Standards 262
8.8.2 Moisture Barrier Packaging Standards 264
8.8.2.1 Handling, Packing, Transport, and Use of Moisture-Sensitive Devices 264
8.8.2.2 MIL-PRF-81705 264
8.8.2.3 ESD Association ANSI/ESD S11.4 266
8.8.3 ESD Control Packaging Measurements 268
8.9 How to Select an Appropriate Packaging System 269
8.9.1 Introduction 269
8.9.2 Customer Requirements 269
xviii Contents

8.9.3 What Is the Form of the ESDS Device? 269


8.9.4 ESD Threats and ESD Susceptibility 269
8.9.5 The Intended Packaging Tasks 271
8.9.6 Evaluate the Operational Environment for the Packaging 271
8.9.7 Selecting the ESD Packaging Type and ESD Protective Functions 273
8.9.7.1 Intimate Packaging 273
8.9.7.2 Proximity Packaging 274
8.9.7.3 Packaging Systems 275
8.9.8 Testing the Packaging System 275
8.10 Marking of ESD Protective Packaging 275
References 276
Further Reading 278

9 How to Evaluate an ESD Control Program 281


9.1 Introduction 281
9.2 Evaluation of ESD Risks 281
9.2.1 Sources of ESD Risk 281
9.2.2 Evaluation of ESD Susceptibility of Components and Assemblies 282
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 283
9.3.1 Process Capability Evaluation 283
9.3.1.1 A Structured Approach to Process Evaluation 283
9.3.1.2 Evaluate the Critical ESDS Path Through the Process 283
9.3.1.3 Use of ESD Withstand Data in Process Evaluation 285
9.3.2 Human Body ESD and Manual Handling Processes 287
9.3.3 ESD Risk Due to Isolated Conductors 287
9.3.3.1 What Is an Isolated Conductor? 288
9.3.3.2 How to Deal with Isolated Conductors 288
9.3.4 Charged Device ESD Risks 290
9.3.5 Damage to Voltage-Sensitive Structures Such as a Capacitor or a MOSFET
Gate 292
9.3.6 Evaluating ESD Risk from Electrostatic Fields 293
9.3.6.1 Risk to ESDS Devices Due to Electrostatic Fields 294
9.3.7 Troubleshooting 296
9.4 Evaluating ESD Protection Needs 297
9.4.1 Standard ESD Control Precautions Do Not Necessarily Address all ESD
Risks 297
9.4.2 Evaluating Return on Investment for ESD Protection Measures 299
9.4.3 What Is the Maximum Acceptable Resistance to Ground? 299
9.4.3.1 Charging Current in a Quasicontinuous Process 299
9.4.3.2 Maximum Decay Time 300
9.4.4 Should There Be a Minimum Resistance to Ground? 300
9.4.5 ESD from Charged Tools 300
9.4.6 Use of Gloves or Finger Cots 301
9.4.7 Charged Cable ESD 301
9.4.8 Charged Board ESD 301
Contents xix

9.4.9 Charged Module or Assembly ESD 301


9.5 Evaluation of Cost Effectiveness of the ESD Control Program 302
9.5.1 The Cost of an Inadequate ESD Control Program 302
9.5.2 The Benefit Arising from of the ESD Control Program 304
9.5.3 Evaluation of the Cost of an ESD Control Program 305
9.5.4 ROI in ESD Control 305
9.5.5 Optimizing an ESD Control Program 307
9.6 Evaluation of Compliance of an ESD Control Program with a Standard 308
9.6.1 Two Steps to Compliance Evaluation 308
9.6.2 Using Checklists to Evaluate Compliance of Documentation with a
Standard 308
9.6.3 Evaluation of Compliance of a Facility with the ESD Control Program 312
9.6.4 Common Problems 314
References 314

10 How to Develop an ESD Control Program 317


10.1 What Do We Need for a Successful ESD Control Program? 317
10.1.1 The ESD Control Strategy 317
10.1.2 How to Develop an ESD Control Program 317
10.1.3 Safety and ESD Control 318
10.2 The EPA 320
10.2.1 Where Do I Need an EPA? 320
10.2.2 Boundaries and Signage 320
10.3 What Are the Sources of ESD Risk in the EPA? 320
10.4 How to Determine Appropriate ESD Measures 321
10.4.1 ESD Control Principles 321
10.4.2 Select Convenient Ways of Working 322
10.5 Documentation of ESD Procedures 323
10.5.1 What Should the Documentation Cover? 323
10.5.2 Writing an ESD Control Program Plan That Is Compliant with a Standard 324
10.5.3 Introduction Section 325
10.5.4 Scope 325
10.5.5 Terms and Definitions 325
10.5.6 Personal Safety 325
10.5.7 ESD Control Program 326
10.5.7.1 ESD Coordinator 326
10.5.7.2 Tailoring ESD Control Requirements 326
10.5.8 ESD Control Program Plan 326
10.5.9 ESD Training Plan 326
10.5.10 ESD Control Product Qualification 327
10.5.11 Compliance Verification Plan 327
10.5.12 ESD Program Technical Requirements 328
10.5.12.1 Documenting Technical Requirements 328
10.5.12.2 ESD Ground 328
10.5.12.3 Personal Grounding 329
xx Contents

10.5.13 ESD Protected Areas 330


10.5.13.1 Handling ESDS Devices and Access to the EPA 330
10.5.13.2 Insulators 331
10.5.13.3 Isolated Conductors 333
10.5.13.4 ESD Control Equipment 334
10.5.14 ESD Protective Packaging 337
10.5.15 Marking of ESD-Related Items 338
10.5.16 References 338
10.6 Evaluating ESD Protection Needs 338
10.7 Optimizing the ESD Control Program 338
10.7.1 Costs and Benefits of ESD Control 338
10.7.2 Strategies for Optimization 339
10.7.2.1 Minimization of the EPA 339
10.7.2.2 Choice of EPA Boundary 339
10.7.2.3 Minimizing Variation of the ESD Control Program 340
10.7.2.4 Standard or Tailored ESD Control 340
10.7.2.5 Design for Convenience 341
10.7.2.6 Who Audits the ESD Control Program? 341
10.8 Considerations for Specific Areas of the Facility 341
10.8.1 The Varying ESD Control Requirements of Different Areas 341
10.8.2 Goods In and Stores 341
10.8.3 Kitting 342
10.8.4 Dispatch 342
10.8.5 Test 342
10.8.6 Research & Development 343
10.9 Update and Improvement 343
References 343

11 ESD Measurements 345


11.1 Introduction 345
11.2 Standard Measurements 346
11.3 Product Qualification or Compliance Verification? 347
11.3.1 Measurement Methods for Product Qualification 347
11.3.1.1 Product Qualification of a Bench Mat 348
11.3.1.2 Example: Product Qualification of a Footwear and Floor Combination 348
11.3.2 Measurement Methods for Compliance Verification 348
11.3.2.1 Example: Compliance Verification Test of a Bench Mat 349
11.3.2.2 Example: Compliance Verification of Footwear and Flooring 349
11.4 Environmental Conditions 349
11.5 Summary of the Standard Test Methods and Their Applications 349
11.6 Measurement Equipment 350
11.6.1 Choosing a Resistance Meter for High-Resistance Measurements 350
11.6.2 Low-Resistance Meter for Soldering Iron Grounding Test 351
11.6.3 Resistance Measurement Electrodes 351
Contents xxi

11.6.4 Concentric Ring Electrodes for Packaging Surface and Volume Resistance
Measurement 352
11.6.4.1 Conversion of Surface Resistance to Surface Resistivity for a Concentric Ring
Electrode 354
11.6.4.2 Conversion of Volume Resistance to Volume Resistivity 354
11.6.5 Two-Point Probe for Packaging Surface Resistance Measurements 354
11.6.6 Footwear Test Electrode 354
11.6.7 Handheld Electrode 355
11.6.8 Tool Test Electrode 356
11.6.9 Metal Plate Electrode for Volume Resistance Measurements 356
11.6.10 Insulating Supports 357
11.6.11 ESD Ground Connectors 357
11.6.12 Electrostatic Field Meters and Voltmeters 357
11.6.12.1 Electrostatic Field Meter–Based Instruments 357
11.6.12.2 Noncontact Electrostatic Voltmeters 358
11.6.12.3 Contact Electrostatic Voltmeters 359
11.6.13 Charge Plate Monitors (CPM) 359
11.7 Common Problems with Measurements 361
11.7.1 Humidity 361
11.7.2 Accidental Measurement of Parallel Paths 361
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 362
11.8.1 Resistance to Ground 362
11.8.1.1 Resistance to Ground from a Work Surface or Floor 362
11.8.1.2 Compliance Verification of Seating in the EPA 364
11.8.1.3 Qualification of Seating 364
11.8.2 Point-to-Point Resistance 366
11.8.2.1 Point-to-Point Resistance of Work Surface 366
11.8.2.2 Point-to-Point Resistance Measurements on ESD Garments 367
11.8.2.3 Simple Point-to-Point Measurement on the Garment Fabric 368
11.8.3 Personal Grounding Equipment Tests 371
11.8.3.1 End-to-End Resistance of a Ground Cord 371
11.8.3.2 Measurement of Wrist Straps and Cords as Worn 372
11.8.3.3 Measurement of Personnel Grounding Through Footwear to Foot Plate
Electrode 374
11.8.3.4 Measurement of Resistance from Person to Ground 374
11.8.3.5 Resistance to Ground of an Earth Bonding Point 377
11.8.4 Surface Resistance of Packaging Materials 377
11.8.4.1 Surface Resistance of Packaging Measured Using a Concentric Ring
Electrode 378
11.8.4.2 Point-to-Point Resistance of Small Packaging Items 379
11.8.4.3 Point-to-Point Resistance of Packaging Using 2.5 kg Resistance Measurement
Electrodes 380
11.8.5 Volume Resistance of Packaging Materials 381
11.8.6 ESD Shielding of Bags 382
11.8.7 Evaluation of ESD Shielding of Packaging Systems 383
xxii Contents

11.8.8 Measurement of Ionizer Decay Time and Offset Voltage 383


11.8.8.1 Common Problems 385
11.8.9 Walk Test of Footwear and Flooring 386
11.8.9.1 Common Problems 386
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 387
11.9.1 Electrostatic Fields and Voltages 387
11.9.2 Measurement of Electric Fields at the Position of the ESDS 387
11.9.2.1 Common Problems 388
11.9.3 Measurement of Surface Voltages on Large Objects Using an Electrostatic Field
Meter Calibrated as a Surface Voltmeter 388
11.9.4 Measurement of Voltage on Devices or Small Conductors 390
11.9.4.1 Measurement Using a Non-contact Voltmeter 390
11.9.4.2 Measurement Using a Contact Voltmeter 392
11.9.5 Resistance of Tools 392
11.9.5.1 Resistance from Tool Tip to Handle 392
11.9.5.2 Resistance to Ground of Handheld Tool 393
11.9.6 Resistance of Soldering Irons 394
11.9.6.1 Resistance to Groundable Point of Soldering Iron Tip 394
11.9.6.2 Resistance to Ground of a Soldering Iron Tip 395
11.9.7 Resistance of Gloves or Finger Cots 396
11.9.7.1 Measurement of Resistance Through a Glove Using a Handheld Electrode 396
11.9.7.2 Testing Resistance Through a Glove Using a Wrist Strap Tester 396
11.9.7.3 Measurement of Resistance to Ground Through a Glove Using a Handheld
Electrode 399
11.9.8 Charge Decay Measurements 399
11.9.8.1 CPM Charge Decay for Tools 400
11.9.8.2 Charge Decay of Gloves and Finger Cots 403
11.9.8.3 System Test of Glove and Handheld Tool 404
11.9.9 Faraday Pail Measurement of Charge on an Object 404
11.9.9.1 The Faraday Pail 404
11.9.9.2 Measurement of Electrostatic Charging of Items Handled with Gloves or Finger
Cots Using a Faraday Pail 405
11.9.9.3 Evaluation of Charging of an Item Using a CPM Plate 407
11.9.10 ESD Event Detection 408
References 409
Further Reading 412

12 ESD Training 413


12.1 Why Do We Need ESD Training? 413
12.2 Training Planning 414
12.3 Who Needs Training? 415
12.4 Training Form and Content 417
12.4.1 Training Goals 417
12.4.2 Initial Training 418
12.4.3 Refresher Training 419
Contents xxiii

12.4.4 Training Methods 419


12.4.4.1 Video, Computer or Internet-Based Training 419
12.4.4.2 Instructor-Led Training 419
12.4.5 Supporting Information 420
12.4.6 Training Considerations 420
12.4.6.1 Preparing a Presentation 421
12.4.6.2 Presentation and Attendee Participation 421
12.4.6.3 Hands-On Learning 423
12.4.6.4 Follow-Up Assessment and Training 424
12.4.7 Public Tutorials and Courses 424
12.4.8 Qualifications and Certification 424
12.4.9 National and International ESD Groups and Electrostatics Interest
Organizations 425
12.4.10 Conferences 426
12.4.11 Books, Articles, and Online Resources 426
12.5 Electrostatic and ESD Theory 428
12.5.1 The Pros and Cons of Theory 428
12.5.2 A Technical and Nontechnical Explanation of Electrostatic Charging 428
12.6 Demonstrations of ESD Control–Related Issues 429
12.6.1 The Role of Demonstrations 429
12.6.2 Demonstrating Real ESD Damage 429
12.6.3 The Cost of ESD Damage 430
12.7 Electrostatic Demonstrations 431
12.7.1 The Value of Electrostatic Demonstrations 431
12.7.2 The Pros and Cons of Demonstrations 431
12.7.3 Useful Equipment for Demonstrations 432
12.7.4 Showing How Easy It is to Generate Electrostatic Charge 433
12.7.5 Understanding Electrostatic Fields 433
12.7.6 Understanding Charge and Voltage 434
12.7.7 Tribocharging 435
12.7.8 Production of ESD 436
12.7.9 Equipotential Bonding and Grounding 437
12.7.10 Induction Charging 438
12.7.11 ESD on Demand – The “Perpetual ESD Generator” 439
12.7.12 Body Voltage and Personal Grounding 440
12.7.13 Charge Generation and Electrostatic Field Shielding of Bags 442
12.7.14 Insulators Cannot Be Grounded 443
12.7.15 Neutralizing Charge – Charge Decay and Voltage Offset of Ionizers 444
12.8 Evaluation 446
12.8.1 The Need for Evaluation 446
12.8.2 Practical Test 446
12.8.3 Written Tests 446
12.8.4 Pass Criteria 447
References 447
Further Reading 448
xxiv Contents

13 The Future 449


13.1 General Trends 449
13.2 ESD Withstand Voltage Trends 450
13.2.1 Integrated Circuit ESD Withstand Voltage Trends 450
13.2.2 Other Component ESD Withstand Voltage Trends 454
13.2.3 Availability of ESD Withstand Voltage Data 454
13.2.4 Device ESD Withstand Test 454
13.3 ESD Control Programs and Process Control 454
13.3.1 ESD Control Program Development Strategies 454
13.3.2 A Basic ESD Control Program 455
13.3.3 Detailed ESD Control Program 456
13.3.4 Human Body ESD 456
13.3.5 ESD Between ESDS and Conductive Items 457
13.3.6 “Two-Pin” ESD From Charged Ungrounded Conductive Items 458
13.3.7 “One-Pin” ESD Between the ESDS and Another Conductive Part 458
13.3.8 Charged Board, Module, and Cable Discharge Events 459
13.3.9 Optimization 459
13.4 Standards 459
13.4.1 Impact on Future Standards 460
13.4.2 ESD Control in Automated Handling 460
13.5 ESD Control Equipment and Materials 461
13.5.1 ESD Control Materials 461
13.5.2 ESD Protective Packaging 461
13.6 ESD-Related Measurements 461
13.6.1 ESD Protective Packaging Measurements 461
13.6.2 Voltage Measurement on ESDS Devices and Ungrounded Conductors 462
13.6.3 Measurements Related to ESD Risk in Automated Handling Equipment 462
13.7 System ESD Immunity 462
13.8 Education and Training 462
References 462
Further Reading 464

A Appendix A: An Example Draft ESD Control Program 465


A.1 About This Plan 465
A.2 Description of the Example Facility 465
A.3 Test and Qualification Procedures 466
A.4 ESD Control Program Plan at XXX Ltd 466
A.4.1 Introduction 466
A.4.2 Scope 466
A.4.3 Terms and Definitions 467
A.5 Personal Safety 467
A.6 ESD Control Program 468
A.6.1 ESD Control Program Requirements 468
A.6.2 ESD Coordinator 468
A.6.3 Tailoring ESD Control Requirements 468
Contents xxv

A.7 ESD Control Program Technical Requirements 468


A.7.1 ESD Ground 468
A.7.2 Personal Grounding 469
A.7.3 ESD Protected Areas (EPA) 469
A.7.3.1 General EPA Requirements 469
A.7.3.2 Insulators 470
A.7.3.3 Isolated Conductors 470
A.7.3.4 ESD Control Equipment 471
A.7.4 ESD Protective Packaging 471
A.7.5 Marking of ESD-Related Items 471
A.7.5.1 General 471
A.7.5.2 Marking of ESD Protective Packaging 471
A.7.5.3 Marking of ESD Control Equipment Used in EPAs 471
A.8 Compliance Verification Plan 472
A.9 ESD Training Plan 472
A.9.1 General Requirements of the ESD Training Plan 472
A.9.2 Training Records 473
A.9.3 Training Content and Frequency 473
A.9.3.1 ESD Awareness Training 473
A.9.3.2 ESD Audit and Measurements Training 473
A.9.3.3 EPA Cleaning Training 473
A.9.3.4 Principles and Practice of ESD Control Training 474
A.10 ESD Control Product Qualification 474
References 475

Index 477
xxvii

Introduction

Electrostatic discharge (ESD) can damage or destroy many types of modern electronic com-
ponents, or modules or assemblies containing electrostatic discharge–susceptible (ESDS)
components.
In electronics manufacturing, sensitivity to ESD became a general concern after the
adoption of metal-oxide-semiconductor (MOS) transistor technology exacerbated by
decreasing internal physical size of semiconductor component features and the rise of
integrated circuits (ICs). The first Electrical Overstress/Electrostatic Discharge Sympo-
sium in the USA was organized in 1979 (Reliability Analysis Center 1979). The 1980
Symposium (Reliability Analysis Center 1980) shows papers on topics as diverse as theory
and practice, device failure analysis studies, failure mechanisms and modeling, design
of device protective networks, and implementing ESD controls, facility evaluation and
effective training. Standards and technical handbooks also emerged around this time. The
standards gave requirements for an ESD control program, while the technical handbooks
gave technical data and tutorial material useful for educating the user and developing the
ESD control program. ElectroStatic Attraction (ESA) of contaminant particles is a problem
for manufacturers of semiconductor devices and displays. For operating electronic systems,
ESD provides a source of electromagnetic interference (EMI) that can result in system
crash, malfunction or data corruption.
Thus, the issues of ESD in electronic components and systems give two areas of interest.
Issues of ESD control during electronic component, assembly and system manufacture are
largely concerned in preventing damage in the unpowered non-operational state and ensur-
ing that product reaches the customer in good condition without compromise to appearance
or reliability. This area can itself be further subdivided into
● Electrostatic and ESD issues affecting product yield and quality during semiconductor
wafer scale fabrication
● ESD issues affecting product yield and quality during component, assembly and system
manufacture and assembly, sometimes known as “factory issues”
● Design of semiconductor devices to withstand ESD up to target levels
ESD interference and damage during working electronic system operation is generally
viewed as part of ElectroMagnetic Compatibility (EMC) and the responsibility of a different
xxviii Introduction

community. In some areas (e.g. Europe) electronic products are subject to ESD immunity
test as part of their evaluation for fitness to be placed on the market (and qualification for
CE marking) (Williams 2007).
Nevertheless, there is some overlap and often confusion between these areas. EMI caused
by ESD in the manufacturing process can cause interference to product testers and lead
to rejection of product and hence reduction of yield. In ESD test in EMC, ESD applied to
exposed circuit connectors can lead to component or system hardware failure and may lead
to requirements for ESD robustness of components that connect to the outside world.
This book is largely concerned with the development and maintenance of an ESD control
program for protection of ESD susceptible components and assemblies during electronic
system and assembly manufacture. This is the so-called “factory issues” area of ESD control.
It is intended that the book can be used as a handbook or practical guide for personnel work-
ing in ESD control in electronics or other companies that handle unprotected ESD suscepti-
ble parts. At the same time, sufficient background information and technical explanation is
given to enable the user to understand the principles and practice of effective ESD control.
Personnel working in this field can have a wide variety of technical background and do
not necessarily have strong electronics or electrical understanding. Many will not have had
opportunity to attend courses on ESD control other than basic ESD awareness. It is sur-
prising that at the time of writing very few University electronics related courses offer any
modules on ESD control. Conversely there are as yet few industry courses available that
deal with the subject in any depth other than basic ESD awareness. Worldwide, there are
still only a very few courses and qualifications available to those who wish to obtain a good
grounding in the field.
So, I have attempted to present the subject with a minimum of theory to make it accessi-
ble to those who do not have a strong relevant theoretical background. This is balanced by
sufficient description of background theory for understanding the material presented, with
references and a bibliography of further reading for those who wish to go into the subject in
greater depth. The intention is to reveal and clarify the principles behind an area often con-
sidered a mysterious “black art.” In many ways, I have tried to write the book I would have
liked to have found when I started learning about ESD control in the electronics industry.
A widespread current approach to development of an ESD control program is to com-
ply with the requirements of an ESD control standard such as ANSI/ESD S20.20 or IEC
61340-5-1. It is often thought that this is sufficient to ensure that product ESD damage is
brought under control. While this can be successful, if applied with insufficient knowl-
edge it can lead to a program that is not well optimized or fails to address all the ESD
threats (Smallwood et al. 2014; Lin et al. 2014). With knowledge and understanding an
optimized and effective ESD control program may be achieved and maintained, often with
lower costs. Nevertheless, compliance with an ESD control standard is advantageous and
can help demonstrate, especially to customers, the seriousness with which ESD control
is treated in the facility. Development of an ESD control program in compliance with the
most widely used and respected ESD control standards 61340-5-1 and S20.20 is therefore
discussed in some length. Properly specified ESD control programs compliant with these
standards are held to be adequate to protect ESD sensitive devices with withstand voltages
down to 100 V Human Body Model (HBM), while also addressing basic ESD risks due to
charged metal objects and charged devices.
Introduction xxix

ESD susceptible components become ever more sensitive to ESD damage as time goes
on, due to component technology developments. The need for development of ESD control
programs through knowledge and understanding rather than rote application of standard
techniques will grow as a greater number of more ESD sensitive components are handled
in electronic manufacturing, assembly, and maintenance processes in the future. In parallel
with the development of ESD control techniques and standards, a massive research effort
has supported on-chip ESD protection networks aimed at reducing device ESD suscepti-
bility, with target withstand voltages of 2 kV HBM, 200 V Machine Model (MM) and 500 V
Charged Device Model (CDM) (Industry Council 2011). In the early 2000s The Industry
Council on ESD Target Levels was formed with members from IC manufacturing and elec-
tronics assembly companies, and independent consultants in the industry. In the face of
increasing difficulty in achieving the target ESD withstand levels, and the belief that mod-
ern electronic manufacturing companies have ESD control programs routinely achieving
the standard protection levels, they recommended reduction of on-chip target protection
levels to 1 kV HBM, 30 V MM and 250 V CDM (Industry Council 2011, 2010a,b). At the
same time, many discrete components and ICs exist that for various reasons do not have
on-chip ESD protection or otherwise have lower ESD withstand voltage than these levels. It
seems likely that this will be the first of many reductions in target level driven by technology
changes and the assumption that industry can handle lower ESD withstand components.
While this book is primarily intended to support the industry factory practitioner, I hope
that this book will encourage and enable Universities and Further Education organizations
to offer courses and modules on ESD control for personnel who wish to make a career in
electronics production and related fields.
This book does not attempt to address electrostatics and ESD control in semiconductor
wafers and device manufacture, or device design for ESD protection. The former may be
still as yet inadequately covered by the very few books available on the subject but is better
discussed in a book more focused on this technology area such as Welker (2006). The topic of
device design is best covered by specialist books such as Amerasekera and Duvvury (2002)
and Wang (2002).
ESD immunity of operating electronic systems is left to be treated in other books as part
of EMC issues, except for some discussion confined to areas of overlap with the ESD factory
issues topic. This field is more concerned with the design of electronic systems for immunity
to ESD than it is with ESD control (Ind. Co. White Paper 3, Johnson and Graham 1993;
Montrose 2000; Williams 2007).
While electrostatic control is used in other industries such as explosives and flammable
materials handling (the latter known in Europe as “ATEX”), these are typically governed by
other standards or regulations. They are only mentioned in this book to draw attention to
possible confusion areas and help avoid mistakes, for example in equipment specification
and sourcing.
While this book could be read “cover to cover” it is probably more likely that the reader
will “dip into” specific chapters as the need arises to learn about different topics while work-
ing in ESD control. The book has been written with this in mind. For those who wish to go
deeper into the subject, lists of references are provided with each chapter.
Every specialist subject has its own set of specialist terms or uses specific terms in spe-
cific ways. Chapter 1 defines and introduces the reader to the commonly used terminology
xxx Introduction

in ESD control. Whilst this chapter forms a general introduction to the key concepts and
terminology in the field, it is also likely to be used to revise or clarify the meaning of terms
during reading of other chapters. This is why the definitions and terminology has been
provided together in one chapter rather than being defined and explained as required dur-
ing the remainder of the book. Chapter 2 then explains in more detail the principles that
underlie ESD control work.
Chapter 3 discusses ESD susceptible devices, and how ESD susceptibility of a compo-
nent is measured. The range of ESD susceptibility of components, and current trends in
ESD susceptibility are reviewed. The topic of failure analysis as it is applied to ESD failed
components is outlined. Some case studies of ESD failures from the literature are briefly
described.
Chapter 4 describes the “seven habits of a highly effective ESD program.” This is a way
of explaining the essential activities of an effective ESD control program, that the author
has used in ESD training work for many years. If these activities are effectively and habitu-
ally implemented, it is likely that an ESD control program will be, and remain, effective. If
any one of them is neglected, it is likely the effectiveness of the ESD control program will
eventually suffer.
Most basic ESD control techniques and standards mainly address manual handling of
ESD susceptible devices, components, and assemblies. Chapter 5 extends the discussion to
ESD control in automated systems, processes and handling, which form a major part of
modern electronic manufacture.
Chapter 6 explains the approach and requirements given by the IEC 61340-5-1 and
ANSI/ESD S20.20 ESD control standards at the time of writing. These standards are
continually updated as time goes on, and so the reader is advised check for current versions
available at the time of reading.
Chapter 7 gives an overview of the equipment and furniture commonly used in ESD con-
trol and commonly specified for use in an electrostatic discharge protected area (EPA) to
control common ESD risks. The chapter explains how these often work together as part of
a system and must be specified with that in mind.
ESD protective packaging is one of the most misunderstood areas of ESD control. ESD
packaging is now available in an extraordinary range of forms from bags to boxes and bubble
wrap to tape and reel packaging for automated processes. The principles and practice of ESD
protective packaging are explained in Chapter 8. This is a deep and constantly developing
subject in itself, and this chapter can barely do more than give an introduction to it.
The thorny question of how to evaluate an ESD control program is addressed in Chapter
9 with a goal of compliance with a standard as well as effective control of ESD risks and
possible customer perceptions.
Whilst evaluating an existing ESD control program provides challenges, developing an
ESD control program from scratch provides others. Chapter 10 gives an approach to this.
ESD control product qualification and compliance verification is an essential part of an
ESD control program. Standard test methods have been developed and specified to go with
compliance with ESD control standards. These are explained in Chapter 11. The ESD con-
trol program may also need to use control measures and equipment that are not currently
specified in the standards. Some examples of test methods that may be used with these are
also given in this chapter.
Introduction xxxi

ESD Training has long been recognized as essential in maintaining effective ESD control.
Chapter 12 discusses this in more detail. It describes some demonstrations and techniques
which the author has used to help trainees understand static electricity, ESD and static
control in practice.
Finally, Chapter 13 attempts to look at where ESD control may go in the near future.

References

Amerasekera, A. and Duvvury, C. (2002). ESD in Silicon Integrated Circuits, 2e. Wiley. ISBN:
0 470 49871 8.
Industry Council on ESD Target Levels (2010a) White paper 2: A case for lowering
component level CDM ESD specifications and requirements. Rev. 2.0. http://www
.esdindustrycouncil.org/ic/en/documents/6-white-paper-2-a-case-for-lowering-
component-level-cdm-esd-specifications-and-requirements [Accessed: 10th May 2017]
Industry Council on ESD Target Levels (2010b) White paper 3: System Level ESD Part I:
Common Misconceptions and Recommended Basic Approaches. Rev. 1.0 http://www
.esdindustrycouncil.org/ic/en/documents/7-white-paper-3-system-level-esd-part-i-
common-misconceptions-and-recommended-basic-approaches [Accessed: 10th May
2017]
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering
component level HBM/MM ESD specifications and requirements. Rev. 3.0. Available
from: http://www.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-
lowering-component-level-hbm-mm-esd-specifications-and-requirements-pdf
[Accessed: 10th May 2017]
Johnson, H. and Graham, M. (1993). High Speed Digital Design – A Handbook of Black Magic.
Prentice Hall. ISBN: 0 13 395724 1.
Lin N, Liang Y, Wang P. (2014) Evolution of ESD process capability in future electronics
industry. In: 15th Int. Conf. Elec. Packaging Tech.
Montrose, M. (2000). Printed Circuit Board Design Techniques for EMC Compliance, 2e. Wiley
Interscience/IEEE Press. ISBN: 0 7803 5376 5.
Reliability Analysis Center (1979). Electrical Overstress/Electrostatic Discharge Symposium
Proceedings. EOS-1. Griffiss AFB, NY: Reliability Analysis Center.
Reliability Analysis Center (1980). Electrical Overstress/Electrostatic Discharge Symposium
Proceedings. EOS-2. Griffiss AFB, NY: Reliability Analysis Center.
Smallwood J., Taminnen P., Viheriaekoski T. (2014) Paper 1B.1. Optimizing investment in
ESD Control. In: Proc. EOS/ESD Symp. EOS-36.
Wang, A.Z.H. (2002). On-Chip ESD Protection for Integrated Circuits. Klewer Academic Press.
Welker, R.W., Nagarajan, R., and Newberg, C. (2006). Contamination and ESD Control in
High-Technology Manufacturing. Wiley-Interscience/IEEE Press. ISBN-10: 0 471 41452 2
ISBN-13: 978 0 471 41452 0.
Williams, T. (2007). EMC for product designers, 4e. Newnes. ISBN-13: 978-0750681704
ISBN-10: 0750681705.
xxxii Introduction

Further Reading

Danglemeyer, T. (1999). ESD Program Management, 2e. Springer. ISBN: 0-412-13671-6.


ESD Association (2014) ANSI/ESD S20:20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive
Devices). Rome, NY, EOS/ESD Association Inc.
ESD Association. (2016) ESD Association Electrostatic Discharge (ESD) Technology
roadmap, revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf [Accessed: 10th May 2017].
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
xxxiii

Foreword

I was quite flattered when Dr. Jeremy M Smallwood asked me to write this foreword
for his book. I view this as a significant honor. Dr. Smallwood (Jeremy) and I have
worked together on standards for electrostatics since the mid-1990s after the IEC formed
Technical Committee 101 – Electrostatics. Both of us had considerable time invested in
the standards process on our sides of the Atlantic: Jeremy with BSI and myself with the
ESD Association (ESDA). In the early 1990s, several things occurred: The IEC formed
Technical Committee-TC101 – Electrostatics. Most of the original members were from
the CENELEC committee that produced the electrostatics document CECC00015 along
with delegates from other non-European countries; and the ESDA became a recognized
American National Standards (ANSI) development body, thus able to officially represent
the United States to the IEC. I was appointed the first lead delegate from the United
States National Committee, and Jeremy was appointed delegate from the UK National
Committee. I had the pleasure of hosting an early IEC TC101 working group meeting
in Austin, Texas, in 1996, following our annual EOS/ESD Symposium held that year in
Orlando, Florida. That was the first time I met Jeremy. While the early years of deliberation
on standards for electrostatics were contentious at times, the committee eventually came
together to form a cohesive group and has prepared important standards, recognized
around the world. Jeremy had considerable input and influence during the formative years
of TC101, including his long-term appointment as the TC101 Chairman. Currently, he is
the lead delegate from the UK and participates actively on many working groups.
As a past president of the ESDA, I was delighted to present Jeremy with the ESD
Association’s Industry Pioneer Award in 2010 in recognition of his contribution to the
science of electrostatics, and in particular, for his role in standards development.
There are many good books on the fundamentals of electrostatics by well-known authors
and practitioners such as Professor Niels Jonassen, University of Denmark and Professor
A.D. Moore (University of Michigan). There are other books that deal with specific issues in
electrostatics and these can be found in the Bibliography of this book. However, this is one
of the few that covers all the aspects of the modern standardization process for electrostatic
control in the manufacturing of electronics and other materials sensitive to electrostatic
xxxiv Foreword

influences. This book is designed to assist a novice to the world of electrostatics as well as
the expert practitioner. I believe the book could be a useful reference to anyone who has to
deal with electrostatics in any field and in particular, electronics manufacturing operations.
The principles and standards discussed herein may be applied to any manufacturing area
and process. This book would also be useful as a text for a college level course dealing with
electronic design when the subject matter turns to reliability and sensitivity of electronic
parts and assemblies.
Dr. Smallwood uses his considerable experience in standards development and practical
experience to guide the reader through the maze and tangle often associated with applica-
tion of standards to a manufacturing process. Since electrostatics is a natural phenomenon
and around us all the time, one would think standardization is an improbable task (maybe
leaning toward impossible). The practical approach taken by Jeremy in Chapters 9 and 10
helps sort out the implementation and program management processes.
Chapter 12 on Training should be extremely useful to anyone that has to set-up, run and
maintain an ESD control program. Jeremy has provided some very useful “tips” on training
considerations, garnered from his years of experience in providing basic to advanced
classes on the “art and science” of electrostatics.
Another important section is Chapter 6 which covers standards very well. Since the early
1990s, the understanding of how to deal with the “mysteries” of electrostatics in industry
has increased dramatically, with the result being able to develop very useful and practical
standards, standard practices, advisories and operating guides. While we must appreciate
and understand that static electricity cannot be prevented, we now know how to live with
it in the manufacturing world and generally can provide mitigation techniques to resolve
most issues before serious problems occur. That being said, fires and explosions, product
damage and loss of life occur every year due to static electricity happening when least
expected or when proper procedures are not followed. Simply having a bad connection
to ground (earth) in the wrong place at the wrong time can (and does) have catastrophic
results more often than most will realize.
The first step in resolving or preventing electrostatic issues is to develop an under-
standing of the phenomenon. Chapters 1–5 provide an excellent introduction to the basic
considerations involved in electrostatics and will be a great starting point for the novice in
this subject. Even the experienced practitioner will find these chapters useful for review
and provide a reminder of what they have “forgotten.” Overall the book is written for
the non-technical person to be able to understand the electrostatic phenomenon but the
“expert” in the field will also find it useful. The extensive bibliography provides a great
resource for anyone needing details on any of the related subjects.
Foreword xxxv

I am confident that you will find this book as entertaining, enlightening and useful as I
have. Happy reading and stay safe.

David E. Swenson, President, Affinity Static Control Consulting, LLC


deswenson@affinity-esd.com
2609 Quanah Drive, Round Rock, Texas, USA 78681
Past President, ESD Association – 1998, 1999, 2008, and 2009
ESD Association Outstanding Contributions Award, 2002
Joel P. Weidendorf Memorial Award – for contribution to ESDA Standards Develop-
ment – 2004
Edward G. Weggeland Memorial Award – for contributions to the operation of the ESD
Association – 2014
xxxvii

Preface

Although the subject of ESD has been a concern to the electronics industry since the 1970s,
there have been relatively few books written on the subject from the point of view of the
person who has to put together, or evaluate, maintain, and update an effective ESD control
program. In my work as a consultant and trainer I have met many people in this position.
Some of these, when I met them, had recently had responsibility for the ESD program
thrust upon them with little or no experience of the subject. Others had some knowledge
but were confused by the array of facts and myths they had learned about the subject.
Still others had developed true expertise over a considerable experience in the subject and
developing and running their own company ESD Programs. Of course, I have learned from
them all – as a trainer and consultant I find that I learn most from trying to explain my
subject to others. Doing so often challenges my own understanding and causes me to think
things through with greater clarity.
In the process, I started trying to simplify the presentation of the principles of ESD preven-
tion. With a nod to Steven Covey’s “Seven Habits of a Highly Effective People” the “Seven
Habits of a Highly Effective ESD Program” was born. Why “Habits”? Because if these things
are done habitually when handling ESD susceptible components and assemblies, we are
well on the way to having an effective ESD control program.
In the mid 1990s, then working for ERA Technology Ltd., I started participating in devel-
opment of British Standards at BSI in Chiswick, London. Through this work I soon found
myself participating also in international standards through BSI’s participation in Interna-
tional Electrotechnical Commission (IEC) Technical Committee101 “Electrostatics” which
was formed in the mid 1990s. Standards work was an eye-opener and I found it simultane-
ously highly stimulating and very frustrating. Stimulating, because I found myself talking
with experts from around the world whose knowledge of their field could be exceptionally
deep, and their experience widely varied and practical. At BSI we could argue for hours
about technical issues, and how best to write a standard test method, that could be under-
stood and reproduced by anyone with reasonable technical ability and expected to produce
results which would agree with others who had done the same. In international standards,
we would first have to agree on a test method acceptable to the participating National Com-
mittees. There could be several to choose from already in use amongst the standards from
the participating countries. Naturally each National Committee expert would favor a partic-
ular approach, especially if it was already adopted within their country. Mostly, of course,
these would produce different results due to different conditions and methodologies.
xxxviii Preface

We would discuss these at length amongst a group of international experts whose cultures
and experiences could be very different, and English often not their mother tongue.
The resulting methodology would have to be acceptable in the very different climates,
conditions, and working practices in Japan, UK, USA, Canada, Scandinavia, France,
Germany, Italy or wherever the participating experts were from. In most cases the end
product would have to be translatable into the expert’s mother language for publication.
We found that some common ways of writing in English can be difficult to translate or may
be unclear in other languages. English phrases or words can even mean different things to
an American or a UK English speaker – leading to long discussions about the best wording
for a single sentence!
In the early 2000s the ESD Association standards were becoming widely accepted in
the electronics industry, and their use spreading from North America to other areas of the
world. Encouraged by ESDA standardization experts working with IEC TC101, the decision
to rewrite IEC 61340-5-1 with an unofficial harmonization with ANSI/ESDA S20.20 was a
landmark decision by the IEC TC101 Working Group 5 revising this standard. Subsequent
further harmonization has simplified the task of the ESD Coordinator, especially in
multinational companies.
As time goes on, the components commonly handled in electronics facilities are becom-
ing more susceptibility to ESD damage. The variety of facilities and processes in which
they are handled, stored or transported grows greater. This means that there is an increas-
ing necessity for the person developing, implementing and maintaining an ESD control
program to understand, analyze and specify effective protection against ESD risks.
Part way through writing this book, I realized I was trying to write the book I would have
liked to find when I first got into the subject of ESD control in the electronics industry. This
book aims to help the reader understand the principles and practice of ESD control to the
point where they can make the decisions needed to develop an effective and optimized ESD
control program compliant with the current ESD control standards. To do this one needs to
understand the purpose of ESD control equipment and materials, and how to specify and
test that it does the job intended. If the reader wishes to improve their knowledge further,
the references and bibliography given should give them a good starting point. Perhaps
most importantly, I hope this book will help the reader find that an initially mysterious set
of practices is actually based on sound engineering principles that they can learn to apply
with confidence.
xxxix

Acknowledgments

I would like to acknowledge my debt to all the experts in the field of ESD control and stan-
dardization who through discussions and published work have contributed to the current
state of my understanding. I am also indebted to all my clients and course attendees who
have challenged me to clarify, explain and justify effective ESD control techniques applied
in many different situations.
I would especially like to thank David E Swenson for his comments on the text and for
contributing photographs and other material, and for writing the Foreword, as well as for
many enlightening discussions over the years. Dave performed the extraordinary feat of
reading and commenting on almost all the draft Chapters at least once. This helped enor-
mously in picking up my mistakes and typographic errors, adding or clarifying important
points and generally improving my work.
Several other friends and colleagues have also very kindly read and commented on
Chapters of this book and encouraged me in this work. Special thanks are due to Rainer
Pfeifle, Charvakka Duvvury and Christian Hinz who each reviewed and commented on
various Chapters in detail. Bob Willis also contributed comments, and Charles Cawthorne
kindly provided me some photographs from his own ESD training materials. Lloyd Lawren-
son kindly allowed me to use Kaisertech facilities for some of my photography. I’m indebted
to Lisa Pimpinella of the ESD Association for arranging permission for me to include figures
from the 2016 ESD Association Electrostatic Discharge (ESD) Technology Roadmap.
Last, but definitely not least, I would like to thank my wife Jan for her good-natured tol-
erance of my absent mindedness and lack of communication when engrossed in my work,
and my daughter Alia for helping improve some of my photographs in preparation for pub-
lication in this book.
1

Definitions and Terminology

As with any specialist subject there are many terms that are the “jargon” of the subject
that can be confusing to the newcomer. There are also terms that have specific meanings
in the context of these standards but may have different meanings in common parlance.
The intention here is not to give strict and rigorous academic definitions, but to assist the
newcomer to the field to understand the following chapters.
Sometimes a range of meanings is common in different industries. For example, the terms
conductive, static dissipative, insulative or insulating, and antistatic can mean many differ-
ent things to different people from different industry areas or in the context of different
standards or electrostatic discharge (ESD) control product types. In most cases, only the
meaning common in ESD control, and in particular in the context of the IEC 61340-5-1 and
ANSI/ESD S20.20 and related standards, is emphasized here.
The task of supervising an ESD control program is often given to personnel from many
technical and educational backgrounds. For this reason, the minimum of prior technical
knowledge is assumed in this book.
Despite this, some of the terms used in this document are defined with basic mathemati-
cal relationships given where appropriate. This is because simple mathematics often helps
to clarify the subject and, in some cases, may be essential to helping the user understand
how to specify aspects of an ESD control program. In many cases, these aspects, and their
practical importance and application, are discussed further in Chapter 2.

1.1 Scientific Notation and SI Unit Prefixes

In electrostatics and ESD work, we often have to deal with very large or very small numbers.
For example, the resistance of a material may be measured to be 10 000 000 000 Ω. For conve-
nience and clarity, we use scientific notation and SI unit prefixes as shorthand for numbers
(http://physics.nist.gov/cuu/Units/prefixes.html).
In scientific notation, the number is rewritten in the form a × 10b , where a lies between 1
and 10, and b is the number of decimal places a must be shifted to get the full number. This
is probably most easily understood by examples of resistance and capacitance (Table 1.1).
Sometimes, when the number a is simply 1, it is omitted.

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
2 1 Definitions and Terminology

Table 1.1 Examples of use of scientific notation and SI prefixes.

Value Scientific notation SI prefix

150 Ω 1.5 × 102 Ω 150 Ω


4
22 000 Ω 2.2 × 10 Ω 22 kΩ
35 000 000 Ω 3.5 × 107 Ω 35 MΩ
9 9
1 000 000 000 Ω 1.0 × 10 Ω or 10 Ω 1 GΩ
1 000 000 000 000 Ω 1.0 × 1012 Ω or 1012 Ω 1 TΩ
−5
0.000 022 F 2.2 × 10 F 22 μF
0.000 000 001 F 1.0 × 10−9 F or 10−9 Ω 1 nF
−10
0.000 000 000 15 F 1.5 × 10 F 150 pF
0.000 000 000 001 F 1.0 × 10−12 F 1 pF

1.2 Charge, Electrostatic Fields, and Voltage


1.2.1 Charge
The charge is a property of elementary particles – electrons and protons – that make up the
atoms of all materials. All materials are made up of positively charged atomic nuclei and
negative electrons (Cross 1987). The charge on a proton in the nucleus is labeled positive,
and the charge on an electron is labeled negative. The charge effects of a proton are equal
and opposite to those of an electron, so if a proton and electron are together in an atom, their
effects cancel exactly. In this case, the atom is neutral. Atoms of different elements have
many protons and electrons, depending on the particular element. For example, a hydrogen
atom has 1 proton and 1 electron, and a carbon atom has 12 protons in the nucleus and 12
electrons surrounding it. An object or material is made of huge numbers of atoms and so
extraordinary numbers of electrical charges.
When talking of static electricity, it is often said that charge is “generated” in certain
circumstances. That is not so – all that happens is that a small number of negative charges
become separated from their positive companions in a material and end up in a different
place. For every negative charge appearing somewhere, there must be a positive charge
appearing somewhere else. An object is described as charged if it has a net imbalance of
the number of positive and negative charges that it contains. The electrical effects of the
charges are no longer balanced, and a net static electrical charge exists at that location. It
is this net charge imbalance that we are referring to when we talk about the charge on an
object or material.
The unit of charge is the coulomb (C). In practice, the coulomb is a rather large amount of
charge and microcoulombs (μC, 10−6 C), nanocoulomb (nC, 10−9 C), or even picocoulombs
(pC, 10−12 C) are more usual. A single electron or proton has a charge of 1.6 × 10−19 C. So,
an object having even 1nC of net charge has a large number, 6.2 × 109 , of unneutralized
electrons or protons.
1.2 Charge, Electrostatic Fields, and Voltage 3

1.2.2 Ions
Ions are very small particles having a small electrostatic charge. They are naturally present
in the air but may also be generated deliberately or accidentally around objects at high
voltage.
Charges are naturally present in atoms. Protons in the atomic nucleus have positive
charge, and electrons in the atom have negative charge. A negative ion is formed when a
particle gains one or more electrons. A positive ion is formed when a particle loses one or
more electrons. Ions may consist of free electrons, single atoms, many atoms, or molecules
(Wikipedia 2018). Sometimes these ions may become attached to larger particles.

1.2.3 Dissipation and Neutralization of Electrostatic Charge


Where there is an imbalance of charges present, there will usually be voltage differences.
Charges exert forces on each other and create an electrostatic field in which various effects
occur. Like charges repel each other, and unlike charges attract each other.
If like charges have built up in a region, they repel each other, and if they are able to
move, they will move apart and gradually spread out and dissipate. Unlike charges will be
attracted and move together.
When opposite polarity charges are sufficiently close together, their effects cancel, and
the charge is said to be neutralized.

1.2.4 Voltage (Potential)


Electric potential is defined in terms of the work done in moving a charge from one place
to another in an electric field (Cross 1987). If a charge Q is moved a distance s against a
uniform electric field E, the potential difference V between the start and end positions is

V = QEs

The energy taken to move a charge between two points is the same, no matter what route is
taken between the points. Potential difference is measured in volts (V) and is often referred
to as voltage. The unit volt (V) is equivalent to joules per coulomb. Voltage is a measure of
the potential energy at a point and is perhaps analogous to pressure in a fluid system or
height in a gravitational system.
Engineers often talk about the potential of (for example) a conductor (see Section 1.7.3
for a discussion of conductors and insulators), as a synonym to voltage. This is not strictly
correct as potential is strictly the work done in bringing a charge from infinity to the place
of measurement (Jonassen 1998).
A voltage or potential difference at a place of measurement must always be referred to
another place. In practice, the potential difference is usually quoted with reference to the
potential of the earth (also referred to as ground; see Section 1.5), which is defined for con-
venience as zero volts. If this other place is not specifically stated, it is usually ground (the
earth).
All points in space surrounding a charge have a voltage (potential) – typically this voltage
will be different from its neighboring points. For a conducting surface, if it is not initially
4 1 Definitions and Terminology

an equipotential, voltage differences cause charge (current) to flow until the voltage around
the surface is eventually equal. So, an electrically conducting surface in equilibrium is an
equipotential surface.

1.2.5 Electric or Electrostatic Field


Any charge has a region of influence around it, in which various electrostatic effects are
noticed – this region is the electric (electrostatic) field due to the charge. Charge is the
fundamental source of static electricity, and the electrostatic field shows the effect of the
charge in the world around the charge source. In this field, we find that

● Like polarity charges are repelled.


● Opposite polarity charges are attracted.
● Conductors (e.g. metals) redistribute their charges and experience a change of potential
(voltage) in response to the field.
● Particles of many materials may be attracted or repelled within the field.

Static electricity phenomena are due to these basic effects.


Dust particles, and small objects, are attracted or repelled by a field, especially if they are
themselves charged (e.g. ionized particles in the air). The force F experienced by a charge
q in an electrostatic field E is (Cross 1987)

F = qE

If equal positive and negative charge are sufficiently close together, from a distance their
electric field effects cancel, and no external field is noticed. The charges are said to be neu-
tralized.
Electrostatic fields and potentials around an object are not easy to visualize. One way of
doing so is by use of field and equipotential lines. A field line represents the path a small
charge would take, if it were free to move under the influence of the force due to the electro-
static field alone. Field lines always leave a conductor at a right angle (90∘ ) to the surface.
In Figure 1.1 a charged spherical conducting object has a potential V. Each point in the
surrounding space can also be assigned a potential, according to the work required to move
a unit charge to that position. If all the positions of equal potential are marked, an equipo-
tential line (or in three dimensions a surface) is marked out. A system of equipotential
surfaces could be marked, forming contours of potential showing the presence of the peak
in potential rather like the contours on a map showing the presence of a hill. Equipotential
lines are always at a right angle (90∘ ) to the field lines.
Equipotential lines are like contour lines on a map of an area of the earth’s surface. Height
is a form of potential energy. If a ball is released on a smooth hillside, it will roll down the
hill perpendicular to the contour lines. Similarly, if a same polarity charge (e.g. a positive
charge, next to a high positive potential) is present in the electrostatic field, it will move
away from the peak in potential, in a path perpendicular to the equipotential lines. These
paths form lines of electrostatic field. The intensity of the field is given by how close together
the field and equipotential lines are.
1.2 Charge, Electrostatic Fields, and Voltage 5

Electric field lines Equipotential


line

Q, V

Charged
Equipotential line is conductor
at right angle to field
line

Figure 1.1 Field lines and equipotential around a charged sphere.

The electric field E (vector, as it has magnitude and direction) is the gradient of voltage
V over a distance s. Electric field, therefore, has the units volts per meter (V/m).
−dV
E=
ds
In Figure 1.1 if the charged sphere is very small, it is effectively a point of charge. The
electrostatic field around the charge Q at a distance r from this point is proportional to the
charge present, according to Coulomb’s law (Cross 1987)
Q
E∝
r2
From this equation, in this case the field strength decreases rapidly with the distance from
the charge, with 1/r 2 . This is also indicated by the spreading of the field lines with distance
from the charge. Field lines can be considered to begin and end on electrostatic charges,
and so a high density of field lines at a surface implies a high charge density as well as high
electrostatic field.
For other shapes of charge patterns, the equipotentials will not in general be spherical,
and field lines will in general be curved rather than straight lines. Field lines are always
perpendicular to the equipotentials and are always perpendicular to conducting surfaces as
these are also equipotentials.

1.2.6 Gauss’s Law


In Figure 1.1, eight field lines cut the equipotential line. Each of these lines in principle
originates on a charge. So, the field lines cutting a surface is related to the net charge within
it. Gauss’s law generalizes this to state that the component of electric field perpendicular to
a surface is proportional to the charge enclosed by the surface. For further information, the
reader should refer to more academic texts such as Cross (1987).
6 1 Definitions and Terminology

1.2.7 Electrostatic Attraction (ESA)


A charge in an electrostatic field experiences a force, as described in Section 1.2.2. So, a
charged particle or object will also experience a force according to the charge it carries.
This causes charged particles and objects to be attracted or repelled by other objects, some
of which may be product or items that are required to be kept clean. This effect is known
as ESA.
A lesser known phenomenon that contributes to electrostatic attraction or repulsion is
dielectrophoresis (Cross 1987). In this case, uncharged particles can be attracted or repulsed
in a divergent or convergent electrostatic field due to differences in the permittivity of the
particle and the material in which it is immersed.

1.2.8 Permittivity
Coulomb’s law shows that the field due to a point charge is proportional to the charge and
inversely proportional to the distance from it squared (Cross 1987).
Q
E∝
r2
Permittivity (dielectric constant), 𝜀, was defined to give a convenient constant of propor-
tionality in this relation.
1 Q
E=
4𝜋𝜀 r 2
For air, the permittivity is very close to the permittivity of free space 𝜀0 (vacuum)
𝜀 = 𝜀0 = 8.8 × 10−12 Cm−1 . Other materials have different permittivity and affect field
strengths correspondingly. In general, a dielectric material has a permittivity greater than
air. This is conveniently expressed as a relative permittivity 𝜀r , and the permittivity is given
by
𝜀 = 𝜀r 𝜀0
Polymers often have relative permittivity in the range 2–3 and many other materials in the
range 2–10. Materials such as ceramics can have far higher permittivity.

1.3 Electric Current


Moving charges form electrical currents. One coulomb of charge has passed if 1 ampere has
flowed for one second.
Q = It
or for a varying current
t

∫0
Q= I dt

and so
dQ
I=
dt
1.5 Earthing, Grounding, and Equipotential Bonding 7

1.4 Electrostatic Discharge (ESD)


IEC 61340-1:2012 defines an electrostatic discharge as “transfer of charge by direct contact
or by breakdown from a material or object at a different electrical potential to its immediate
surroundings.” IEC 61340-5-1:2016a gives a slightly different definition of “Rapid transfer
of charge between bodies that are at different electrostatic potentials.”
There are various types of electrostatic discharge that are important in different fields. In
ESD in electronics handling, the main types of concern are
● Spark discharges between conducting objects or materials
● Brush discharges between a conducting object and an insulating material
● Corona discharges from sharp conducting objects and materials
Electrostatic discharges are discussed further in Chapter 2.

1.4.1 ESD Models


ESD from different sources produces very different discharge current waveforms. These can
be modeled and simulated by simple electronic circuits. Three standard ESD source cir-
cuit models, human-body model (HBM), machine model (MM), and charged device model
(CDM), have been developed and standardized for testing ESD susceptibility of electronic
components. This is discussed further in Chapter 3.

1.4.2 Electromagnetic Interference (EMI)


An ESD event can produce very large and fast-changing currents and voltages. These pro-
duce fast-changing electromagnetic fields with strong and fast-changing magnetic and elec-
tric field components and a broad frequency spectrum, sometimes extending to over GHz
frequencies. This can be radiated and conducted to be picked up by nearby electronic cir-
cuits and can cause temporary malfunction. This is known as electromagnetic interference.

1.5 Earthing, Grounding, and Equipotential Bonding


Electrostatic discharges occur because of voltage differences between the objects between
which the discharge occurs. If there were no voltage difference, then no ESD could occur.
So, one way to prevent ESD from occurring is to eliminate voltage differences between
objects. If the two objects are conductors, connecting them electrically ensures that they
are eventually at the same voltage. This must be so, as if any voltage difference were to
arise, charge (current) would flow due to the voltage difference, until there is no voltage
difference. The practice of connecting conductors together to eliminate voltage differences
is known as equipotential bonding.
If two conductors at two different voltages are brought into contact, an electrostatic dis-
charge will occur as part of the voltage equalization process. If one of the conductors is
susceptible to ESD damage, it could risk being damaged as a result. So, ESD-susceptible
parts must only make contact with other conductors, including grounded conductors, in
circumstances designed to protect against damage.
8 1 Definitions and Terminology

In many practical cases, one of the conductors concerned may already be electrically con-
nected to an electrical earth or can be conveniently connected to earth. The earth is often
defined as our 0 V reference point in electricity power distribution, electrostatics, and ESD
control. So, it is often convenient and is common practice to electrically connect all conduc-
tors to earth. Earth is also known as ground, and earthing is also known as grounding.
The terms earthing and grounding can have different meanings and requirements in dif-
ferent contexts or industries. An electrical engineer may require an earth resistance less
than an ohm. A plant engineer may earth bond two items of plant, requiring a resistance
less than 10 Ω. An electromagnetic compatibility (EMC) engineer may require an extremely
low impedance to be maintained from direct current (DC) to hundreds of MHz or even GHz.
To an ESD control practitioner, a resistance to ground <109 Ω at dc may be sufficient.
In practice in ESD control, there are various types of ground that can be used. In the ESD
standards IEC 61340-5-1:2016a and ANSI/ESD S20.20-2014, the term grounding is used to
mean any of the following:
● Connection to electrical earth (the safety earth wire of a mains electrical system)
● Connection to a functional earth (e.g. an earth rod driven into the ground)
● Connection to an equipotential bonding system

1.6 Power and Energy


Energy is the ability to do work. Physics recognizes many types of energy – heat, light,
gravitational, mechanical, and of course electrical.
Mechanical energy expended is the product of force and the distance moved. If a force
qE is applied to move a charge q over a distance s between points A and B, the work done,
W AB , is
WAB = qEs
Energy (work) expended, W, is also the product of power P and the time duration t that the
power is applied.
W = Pt
The electrical power expended is the product of voltage V and current flowing I.
P = VI
So, the electrical energy expended is
W = VIt

1.7 Resistance, Resistivity, and Conductivity

1.7.1 Resistance
Electrical resistance is the ratio between the dc voltage applied to a circuit or material and
the current flowing through it, given by Ohm’s law.
V
R=
I
1.7 Resistance, Resistivity, and Conductivity 9

1.7.2 Resistivity and Conductivity


1.7.2.1 Surface Resistivity and Surface Resistance
Surface resistivity is defined as a material surface property. It is based on the theoretical
resistance of a square of material surface with sides of unit length, with a voltage applied to
two opposing sides of the square (Figure 1.2). In theory, the current flows across the surface
of the square. For a material of surface resistivity 𝜌s with linear electrodes of width w placed
parallel on the surface a distance d apart, the surface resistance Rs measured between the
electrodes is
𝜌s d
Rs =
w
where d = w, which reduces to 𝜌s = Rs .
The unit of surface resistivity is ohms (Ω). In some industries, it is quoted as ohms per
square (Ω/sq). This reflects the property that the value of the surface resistance measured
with a square electrode pattern (d = w) is the same, no matter what the dimension of the
side of the square is.
In practice, standards exist for measuring surface resistivity using concentric ring elec-
trodes (IEC 62631-3-2 (International Electrotechnical Commission 2015), IEC 16340-2-3,
ANSI/ESD STM 11.11 (EOS/ESD Association Inc. (2015a)). This is further discussed in
Chapter 11.
Surface resistance is a resistance measured between two electrodes on the surface of a
specimen. The electrodes may be of any convenient form. Sometimes this measurement
is made using electrodes designed so that conversion from surface resistance to surface

Resistance
meter

Electrodes

Material surface

Figure 1.2 Surface resistivity definition.


10 1 Definitions and Terminology

resistivity is a simple calculation. In ESD control practice, conversion to surface resistiv-


ity is often not needed, and the surface resistance result obtained with defined standard
electrodes is used directly.

1.7.2.2 Volume Resistance, Volume Resistivity, and Conductivity


Volume resistivity is a bulk material property based on the resistance of a cube of mate-
rial with sides of unit length, with a voltage applied to two opposing faces of the cube
(Figure 1.3).
The volume resistance Rv measured through a material of volume resistivity 𝜌v using elec-
trodes of area A is given by
𝜌v t
Rv =
A
where t = A = 1, or t/A = 1, which reduces to 𝜌v = Rv .
The unit of volume resistivity is ohm meter (Ωm). The volume resistivity of a material is
often simply referred to as its resistivity.
In practice, standards exist for measuring volume resistivity using concentric ring elec-
trodes (IEC 62631-3-1 (International Electrotechnical Commission 2016c), IEC 61340-2-3
(International Electrotechnical Commission 2016b), ANSI/ESD STM 11.12 (EOS/ESD
Association Inc. 2015b))
Volume resistance, Rv , is a resistance measured between opposing faces of a material. The
electrodes may be of any convenient form. Often this measurement is made using electrodes

Resistance meter

Electrode area A

Material

Figure 1.3 Definition of volume resistivity.


1.7 Resistance, Resistivity, and Conductivity 11

designed for volume resistivity measurement so that conversion from volume resistance to
volume resistivity is a simple calculation. In ESD work, conversion to volume resistivity
is often not needed. The volume resistance obtained with defined standard electrodes is
used directly, saving the effort of calculation. Examples of surface and volume resistance
measurement methods are given in Chapter 11.
The conductivity, 𝜎, of the material is simply the inverse of its resistivity.
1
𝜎=
𝜌v
The units of conductivity are siemens per meter (Sm−1 ).
The resistivity of materials can vary by many orders of magnitude from 10−8 Ωm (e.g.
copper) to more than 1015 Ωm (e.g. mica, quartz, polytetrafluoroethylene, polyethylene).

1.7.3 Insulators, Conductors, Conductive, Dissipative, and Antistatic Materials


There is no fundamental definition of insulators and conductors in electrostatics. In real-
ity, there is a continuum of material resistivity from highly conducting (low resistance) to
highly insulating (very high resistance). Different industry areas may have differing views
on the resistance level at which a material is considered to have insulating properties.
For our purposes, a conductor is a material that allows charge to move around on the
surface or in the bulk of the material and can thereby be used to transport charge from one
place to another. An insulator (nonconductor) is a material that does not allow the charge
to move in this way.
One problem in practice is that a material that is considered “insulating” in one appli-
cation may be considered significantly conducting in electrostatics. So, for some years I
have offered the following pragmatic definitions for use in practical electrostatics and ESD
control:

● A conductor is a material that allows charge to move away quickly enough to avoid sig-
nificant electrostatic charge build up.
● An insulator is any material that is not a conductor, in other words, a material that does
not allow charge to move quickly enough to avoid charge build up.

Conductors are easily maintained at a low voltage by connecting them to earth (ground).
However, an insulator in electrostatic terms cannot be maintained at a low voltage by
installing a ground connection. The charge on the material simply does not move to the
ground connection quickly enough to be conducted away in the desired timescale.
Materials or equipment are often defined as conductors or insulators based on either their
measured resistance or a charge decay time. This is discussed further in Chapter 2.
Table 1.2 shows how the terms insulating, dissipative, conductive, and antistatic are widely
used in ESD control. Take care when using these terms, because they may be defined differ-
ently in different contexts and may mean different things to different people. When defined
in the standards, the precise definition can change as the standards evolve into new editions.
The situation becomes worse if usage of these terms in other industries and for specific
products is considered (Table 1.3). In general, these words should be considered unreliable
in meaning unless specified by standards as part of an ESD control system.
12 1 Definitions and Terminology

Table 1.2 Example of how meanings of conductive, static dissipative, insulative, and antistatic can
vary with context in ESD control in electronic manufacturing.

Meaning under Meaning under


Term Application General use 61340-5-1:2016a S20.20-2014

Conductive General Resistance <106 Ω Not defined Not defined


ESD control footwear Not defined Not defined
ESD control flooring <106 Ω Not defined Not defined
ESD protective Surface resistance Surface and
packaging <104 Ω volume resistance
<104 Ω
Static General Resistance Not defined Not defined
dissipative between 106 and
1011 Ω
ESD control footwear Not defined Not defined
ESD control flooring ≥106 Ω Not defined Not defined

≥104 and ≤1011 Ω


ESD protective Surface resistance Surface and

≥104 and <1011 Ω


packaging volume resistance

Insulative General Resistance over Not defined Not defined


1011 Ω
ESD control footwear Not defined, but Not defined, but
by implication by implication
>108 Ω resistance >109 Ω
ESD control flooring Not defined Not defined

≥1011 Ω
ESD protective Surface resistance Surface and

≥1011 Ω
packaging volume resistance

Antistatic General Widely used to Not defined The property of a


described material that
materials used in inhibits
static control; can triboelectric
mean almost charging (ESD
anything ADV1.0-2009)
ESD control footwear Note: Has defined Not defined Not defined
meaning under
ISO 20345 in
process industry
hazard work
ESD control flooring Not defined Not defined
ESD protective Not defined Materials that
packaging have reduced
amount of charge
accumulation as
compared with
standard
packaging
materials
1.7 Resistance, Resistivity, and Conductivity 13

Table 1.3 Example of how meanings of conductive, dissipative, and insulative can vary with context
in static control in other industries (IEC 60079-32-1:2013).

Object Measurement Conductive Dissipative Insulative

Material Volume resistivity (Ωm) <105 ≥105 to 109 ≥109


Clothes Surface resistance (Ω) <2.5 × 10 10
≥2.5 × 1010
Footwear Leakage resistance (Ω) <105 ≥105 to <108 ≥108
Gloves Leakage resistance (Ω) <10 5
≥10 to –<10
5 8
≥108
Floor Leakage resistance (Ω) <105 ≥105 to <108 ≥108

1.7.4 Point-to-Point Resistance


In ESD control, it is convenient to make simple measurements to evaluate the surface prop-
erties of a material or item of equipment. One simple way of evaluating a surface is to place
two electrodes on it and measure the resistance between them. The electrodes are often
cylindrical in form. This is called a point-to-point resistance measurement. Standard test
methods based on this approach are often used. Examples of point-to-point resistance test
methods are given in Chapter 11.

1.7.5 Resistance to Ground


As explained earlier, in ESD control work, voltages on conductors are often eliminated or
controlled by providing an electrical connection for the charge to pass to earth (ground). It is
often required to know the resistance from an object or surface to ground to help understand
the charge dissipation paths. This is known as resistance to ground. Examples of measure-
ment methods for this are given in Chapter 11.

1.7.6 Combination of Resistances


In practice, the resistance of a ground path may be due in part to several components. If
these are effectively in series (Figure 1.4), the effect is to add the resistance of all component
contributors R1 …Rn to get a total resistance Rtot .
Rtot = R1 + R2 … . + Rn
If resistance of ground paths is in parallel (Figure 1.5), they are combined as
1 1 1 1
= + …. +
Rtot R1 R2 Rn

Figure 1.4 Resistances in series. R1 R2 Rn


14 1 Definitions and Terminology

Figure 1.5 Resistances in parallel.

R1 R2 Rn

1.8 Capacitance

The voltage V on a conductor is related to the stored charge Q as


CV = Q
The variable C is the capacitance of the conductor. In electrostatics, any conductive object
has capacitance; it is just the relationship between the stored charge and the object’s voltage.
In practice, the capacitance of an object can vary with proximity of other conductors and
materials (see Chapter 2).
A charged capacitor stores energy. The energy W stored in a capacitance C at voltage V
is given by
W = 0.5CV 2
This can also be expressed as
W = 0.5 QV
An object in free space (with nothing in the near vicinity) still has capacitance. For a spher-
ical conductor of radius r in air or a vacuum, this capacitance C is
C = 4𝜋𝜀o 𝜀r r
In practice, the capacitance of an object may be due in part to the proximity to several
objects. If these are effectively in parallel (Figure 1.6), the effect is to add the capacitance of
all component contributors C1 …Cn to get a total capacitance Ctot as
Ctot = C1 + C2 … . + Cn
If capacitances between objects are in series (Figure 1.7), they are combined as
1 1 1 1
= + …. +
Ctot C1 C2 Cn

Figure 1.6 Capacitors in parallel.

C1 C2 Cn
1.11 Relative Humidity and Dew Point 15

Figure 1.7 Capacitors in series. C1 C2 Cn

1.9 Shielding
The term shielding is used in ESD control in a different way to other disciplines, espe-
cially EMC and radio frequency work. Shielding definitions and tests used in ESD control
are often highly specific to the standards used. Typically, the term is used to describe the
attenuation of electrostatic fields or electrostatic discharge energy applied to the outside of
a protective package, measured at the inside of the package. This is discussed further in
Chapter 8.

1.10 Dielectric Breakdown Strength


If a low voltage is applied across an insulating material, very little current will flow due to
the high resistivity of the material. If, however, the voltage is increased, a level may eventu-
ally be reached where the current suddenly increases to a high value. Typically, this current
flow may lead to formation and thermal heating of a small electrically conducting channel
through the material. For a solid material, melting or damage of a small channel through
the material may occur. This is dielectric breakdown of the material.
Typically, very high electrostatic field strengths are required for dielectric breakdown to
occur. The breakdown strength of air is, for planar parallel electrodes, around 3 MV m−1 or
about 3 kV mm−1 . For curved or sharp electrodes, it is much lower. The breakdown strength
of most insulating solids is much higher than air. For polyethylene, it is about 20 MV m−1
(IEC 61340-1 (International Electrotechnical Commission 2012)).

1.11 Relative Humidity and Dew Point


The relative humidity (rh) or dew point of the atmosphere has a large influence on electro-
static phenomena (see Section 2.3.5). At any temperature, moisture-saturated air in equilib-
rium contains a maximum amount of moisture determined by the saturated vapor pressure
of water at that temperature (Lawrence 2005). The saturated vapor pressure of water and
hence the amount of moisture in saturated air increase strongly with increasing tempera-
ture. This saturated state is defined as 100% r.h. The relative humidity of air with lower than
the saturated amount of water vapor present is given by
vapor pressure of water present
relative humidity =
saturated vapor pressure of water at the temperature
As the saturated vapor pressure increases strongly with temperature, if the amount of mois-
ture present remains the same, increasing the air temperature will result in a reduction in
relative humidity. Conversely, lowering the temperature will increase the humidity.
16 1 Definitions and Terminology

If the temperature is lowered sufficiently, the water vapor pressure eventually becomes
equal to the saturated vapor pressure, and the air becomes saturated. Any further reduction
in temperature may result in moisture condensing from the air on to surfaces in contact with
it or in fog forming. This temperature is called the dew point.

References

Cross, J.A. (1987). Electrostatics Principles, Problems and Applications. Adam Hilger. ISBN:
0-85274-589-3.
EOS/ESD Association Inc. (2014) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical
and Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive
Devices). Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015a) ANSI/ESD STM 11.11-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items – Surface Resistance Measurement of
Static Dissipative Planar Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015b) ANSI/ESD STM 11.12-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items – Volume Resistance Measurement of
Static Dissipative Planar Materials, Rome, NY, EOS/ESD Association Inc.
International Electrotechnical Commission. (2012) IEC/TR 61340-1: 2012. Electrostatics – Part
1: Electrostatic phenomena — Principles and measurements. Geneva, IEC.
International Electrotechnical Commission. (2013) PD/IEC TS 60079-32-1. Explosive
atmospheres Part 32-1. Electrostatic hazards, guidance. Geneva, IEC.
International Electrotechnical Commission. (2015) IEC 62631-3-2. Dielectric and resistive
properties of solid insulating materials - Part 3-2: Determination of resistive properties (DC
methods) - Surface resistance and surface resistivity. Geneva, IEC.
International Electrotechnical Commission. (2016a) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Electrotechnical Commission. (2016b) IEC 61340-2-3:2016. Electrostatics.
Methods of test for determining the resistance and resistivity of solid planar materials used to
avoid electrostatic charge accumulation. Section 3: Methods of test for determining the
resistance and resistivity of solid planar materials used to avoid electrostatic charging. Geneva,
IEC.
International Electrotechnical Commission. (2016c) IEC 62631-3-1. Dielectric and resistive
properties of solid insulating materials - Part 3-1: Determination of resistive properties (DC
methods) - Volume resistance and volume resistivity - General method. Geneva, IEC.
Jonassen, N. (1998). Electrostatics. Chapman & Hall. ISBN: 0 412 12861 6.
Lawrence, M.G. (2005). The relationship between relative humidity and the dewpoint
temperature in moist air. A simple conversion and applications. Bull. Am. Meteorol. Soc.:
225–233. https://doi.org/10.1175/BAMS-86-2-225 [Available from htt.s://journals.ametsoc
.org/doi/pdf/10.1175/BAMS-86-2-225. Accessed 15th Aug. 2018.].
Wikipedia (2018) Ion, viewed 17 October 2018, [Available from https://en.wikipedia.org/wiki/
Ion]
17

The Principles of Static Electricity and Electrostatic Discharge


(ESD) Control

2.1 Overview
ESD stands for electrostatic discharge or, according to some, electrostatic damage. This
chapter provides the basis of how static electricity arises and can lead to ESD in the real
world. It also provides the principles that underlie ESD control techniques and equipment
design.
Electrostatic charge can build up in a variety of ways. The charged object has an electro-
static field that could conceivably lead to an ESD event in several ways:
● Direct breakdown of sensitive parts due to high electric field
● Generation of an ESD event directly subjecting the part to discharge currents
● Generation of an ESD event subjecting a part to induced transient electric or magnetic
fields, or some other stress
At the root of any ESD event there is an object or surface that has a voltage that is different
to its surroundings. Without this voltage difference, no electric field is present, and no ESD
current can flow. Hence, the objective of ESD prevention measures has been to keep surface
voltages and electric fields to a low level, below which damaging ESD cannot occur.
A review of the explanation of electrostatic charge build-up and ESD sources included
here quickly reveals many ways in which ESD risks can be generated in the real world.
These are summarized in brief in this chapter.

2.2 Contact Charge Generation (Triboelectrification)


The first thing to state is that charge is never generated, nor is it ever destroyed. The
phenomenon that we often describe lazily as the “generation” of charge is more correctly
“separation.” Some practitioners speak of the charge being “liberated” or “set free.” The
charge is initially present in the atoms that make up all materials. There are positive
charges as protons in the atomic nucleus, and there are negative charges as electrons
around the nucleus. Normally these are present in equal numbers so that in an uncharged
atom the number of positively charged protons equals the number of negatively charged
electrons present. Static electricity arises when an imbalance is created and the local
amounts of positive and negative charge become different.
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
18 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

In practice, the amount of charge imbalance required to give strong electrostatic effects
is surprisingly small. The limit of the amount of charge that can be built up on a surface is
governed by the electrical breakdown field strength of air, around 3 × 106 V m−1 . The sur-
face charge density required to give this field is only 2.64 × 10−5 Cm−2 (Cross 1987). This is
equivalent to about 1.7 × 1014 electrons m−2 , or 8 atoms per million on the surface acquiring
or losing an electron!
One common way in which static electrical charge imbalances can arise is when two
materials make contact and then are separated. While in contact, electrons move from one
material to the other at points of contact; this material gains a net negative charge, and the
donor material gains a net positive charge. When the objects are separated, the negatively
charged object can take its charge with it, leaving an equal positive charge on the other
object. Although it is really charge separation that takes place, it is common to refer to the
“generation” of static electrical charge.

2.2.1 The Polarity and Magnitude of Charging


The polarity of charge left on a material can be positive or negative and depends on a range
of factors, especially on the other material with which it made contact. Materials may be
arranged in a table according to the polarity of charge they take in contact with other mate-
rials, called the triboelectric series (see Table 2.1).
A material in the table (e.g. aluminum) can be expected to charge positively against
another material below it in the table (e.g. polytetrafluoroethylene (PTFE)) and negatively
against a material above it (e.g. wool). The amount of charge generated is a function of the
separation of the materials on the table; aluminum and paper can be expected to charge
relatively little against each other, but polyvinylchloride (PVC) and nylon can be expected
to charge strongly against each other.
In practice, triboelectrification is a variable phenomenon and is highly dependent on sur-
face conditions, contaminants, and humidity. Small amounts of surface contaminants can
have a large effect on triboelectrification. One result is that the order of triboelectric series
is not unique. Different experiments and samples of the same materials may produce differ-
ent results especially if the experimental conditions are varied. While it could be assumed
from the triboelectric series that contact between two surfaces of the same material would
not generate charge, this is generally not what happens in practice.

2.3 Electrostatic Charge Build-Up and Dissipation


Any two materials in contact give charge separation that can lead to static electrical charge
build-up. This may or may not lead to charge and voltage build-up, depending on the cir-
cumstances.
The key to this build-up is the balance between charge generation and charge dissipation
(or neutralization). If charge is dissipated (or neutralized) more quickly than it is generated,
no static electricity builds up, and no effects are noticed. If charge is generated more quickly
than it is dissipated (or neutralized), then high voltages and static electricity effects are
quickly built up.
2.3 Electrostatic Charge Build-Up and Dissipation 19

Table 2.1 An example of a triboelectric series.

Acetate Charge to positive polarity

Glass
Mica
Human hair
Nylon
Wool
Lead
Silk
Aluminum
Paper
Cotton
Steel
Wood
Epoxy-glass
Copper
Stainless steel
Acetate rayon
Polyester polyurethane
Polyethylene
Polypropylene
PVC
Silicon
PTFE Charges to negative polarity

2.3.1 A Simple Electrical Model of Electrostatic Charge Build-Up


Static electricity can be modeled as a charge generator, and a simple electrical model can
be used to understand many practical situations (Figure 2.1).
The separation of charge is effectively a small electrical current represented by a current
source I. The capacitor C represents the charge storage properties of the system and could
be a material surface or a conducting object with a capacitance to earth. The resistance R
represents charge dissipation processes (other than ESD) and can vary from less than 1 Ω
to more than 1014 Ω for good insulators. (See Sections 1.7.3 and 2.3.4 for discussions on the
meaning of insulators and conductors.)
If the current is constant (i.e. the effect of capacitance can be neglected), it’s easy to see by
Ohm’s law that the voltage developed is highly dependent on the resistance R. If a charge
generation rate of 1 nA (1 nCs−1 ) is present, with a resistance of 109 Ω, a steady state voltage
of 1 V is produced. If, however, the resistance was 1012 Ω, a voltage of 1 kV would be pro-
duced, and for a resistance of 1014 Ω, a voltage of 100 kV would theoretically be produced!
20 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

dQ/dt = I
Charge R C
generation Resistance Capacitance
(current source)

Figure 2.1 A simple electrical model of electrostatic charge build-up.

For a charge generation rate of 1 μA, a resistance of 1010 Ω would yield a voltage of 10 kV.
In practice, electrostatic sources rarely generate charge at this rate or on a steady current
basis unless there is steady movement involved (e.g. in a conveyor system).
The rate of electrostatic charge generation is affected by many factors. Some of the key
factors are as follows:

● Relative position of the materials in the triboelectric series


● Rate of separation of contact area (high rates of movement)
● Condition of the surfaces that make contact
● Rubbing of the contacting surfaces
● Ambient humidity and temperature

The many factors involved in triboelectric charge separation make it a highly unpredictable
phenomenon.

2.3.2 Capacitance Is Variable


The capacitance C represents charge storage. There is a simple relation between charge Q
and voltage V, and capacitance is the ratio of charge and voltage.

CV = Q

Q
C=
V
In practice, capacitance is usually a variable that depends on the materials and nearby
objects and on the proximity to earth. Objects move around in daily life, and so their capac-
itance changes.
As an example, we can consider the human body. It is, in electrostatic terms, a conducting
object, being mainly composed of water, which is a conducting material. Even if we neglect
the nearby objects and earth, the human body can be approximated as a sphere that has a
similar surface area. The “free space” capacitance of a sphere is given by 4𝜋𝜀0 r, where r is
the radius and 𝜀0 is the permittivity of free space, 8.8 × 10−12 Fm−1 . Typically, a 1 m radius
sphere gives a useful approximation and has a “free space” capacitance around 110 pF.
2.3 Electrostatic Charge Build-Up and Dissipation 21

Figure 2.2 Parallel plate capacitor.


Conductor
area A d

Conductor
area A

Nearby objects and earth increase this value. In fact, the human feet, on the earth, approx-
imate two parallel plate capacitors in parallel with the free space capacitance. Each capac-
itor is made up of two electrodes: the sole of the foot and the earth. These are separated by
a layer of material (the shoe sole, typically an insulating polymer of relative permittivity 𝜀r
around 2.5). Each foot capacitance varies from moment to moment as the feet are lifted and
replaced on the ground during walking. Each foot capacitance can be modeled as a parallel
plate capacitor (Figure 2.2), with plates of area A separated by a distance d and capacitance
C given by
A
C = 𝜀0 𝜀r
d
In general, if either the area A or the distance of separation d are changed, then the capaci-
tance will change. If the charge is held constant, increasing the capacitance will decrease the
body voltage, and reducing capacitance will increase body voltage. Reducing capacitance
can be achieved by reducing the area (e.g. standing on tip-toe) or increasing the separation
distance (e.g. raising the foot from the floor).
The previous equation shows that if the charge on conductor is unchanged and the capac-
itance changes, then the voltage of the conductor changes. For example, if a person’s body
capacitance changes between 50 and 150 pF while walking and the charge on their body is
constant at 5 nC, their body voltage will vary between 100 V (at 50 pF) and 33 V (at 150 pF).
If a printed circuit board (PCB) conductor has a capacitance of 20 pF and charge 5 nC when
resting close to a large earthed machine part, its voltage will be 250 V. If its capacitance is
reduced to 5 pF when far away from this machine part, its voltage will rise to 1000 V.
It can be useful to have an idea of the approximate capacitance of everyday objects, espe-
cially when estimating the possible effect of ESD to or from such items. Table 2.2 gives some
examples (IEC 61340-1).
The variable ratio between charge and voltage behaves similarly for nonconductors.
When seated in a chair, the body generates charge on the clothes surfaces in contact with
the chair. This forms a large area of charged material with a small distance between the
charges (the two surfaces are in contact). The person’s body voltage is low in this situation
even though their clothing may be highly charged. On rising from the chair, the person can
take much of the separated charge with them. The effective “capacitance” between the body
and the chair is rapidly reduced (separation is rapidly increased), and a high body voltage
quickly results if the charge cannot dissipate to ground. It is common to feel a shock on
touching something metal after rising from a chair or car seat – voltages over 10 kV have
been measured on people after getting out of a car seat (Pirici et al. 2003; Andersson et al.
2008).
22 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Table 2.2 Approximate capacitance of typical everyday objects.

Electronic components and small assemblies 0.1–30 pF


Drinks can, small metal parts 10–20 pF
Tweezers held in hand 25 pF
Small metal containers (1–50 l), trolleys 10–100 pF
Larger metal containers (250–500 l) 50–300 pF
Human body 100–300 pF
Small signal MOSFET gate capacitance 100 pF
Power MOSFET gate-source capacitance 900–1200 pF
Car 800–1200 pF

MOSFET, Metal Oxide Silicon Field Effect Transistor

A corollary of this is that the voltage and field surrounding a charged object may be sup-
pressed by the presence of a nearby conducting object. If the capacitance of the system is
increased, the voltage is decreased.
As an example, a charged garment that fits snugly to the body has voltage suppressed
due to the proximity of its surfaces to the body. Even if the garment is highly charged, the
external field may be limited due to this. If the garment flaps open, the body and garment
surfaces move apart. “Capacitance” is reduced, and a high voltage and electrostatic field
appears outside the garment.

2.3.3 Charge Decay Time


The resistance and capacitance form a resistor-capacitor (RC) network that has a charac-
teristic time constant 𝜏.

𝜏 = RC

In a time 𝜏, the voltage will decay to about 37% of its initial value.
In the example, if the charge current is suddenly halted at time t = 0 with initial voltage
V 0 , the voltage V on the capacitor reduces as
−t
V = V0 exp
𝜏
An electrostatic field meter monitoring the material surface would measure this exponen-
tial decay of voltage. The product of a material’s resistivity 𝜌 and permittivity 𝜀0 𝜀r gives a
physical time constant for the material.

𝜏 = 𝜌𝜀0 𝜀r

This behavior has important practical implications. If we consider a situation in which the
capacitance is fixed at 100 pF (the order of magnitude of capacitance of a person) and the
charging current 100 nA, we can consider the effect of different resistances. With a resis-
tance of 1 GΩ, the voltage generated is only 100 V, and on cessation of the current, the
2.3 Electrostatic Charge Build-Up and Dissipation 23

Initial
voltage
100%

Voltage
curve

37% Decay time

Decayed
voltage

0 Time τ t

Figure 2.3 Charge or voltage decay curve.

voltage will fall to 37% of its initial value within 109 × 10−10 = 0.1 seconds. The effect of
a short duration charging current of this magnitude is unlikely to be noticed.
If the resistance is increased 10 GΩ, not only is the voltage generated increased to 1 kV,
but on cessation of the current, the voltage will take 1010 × 10−10 = 1 seconds to fall to 37%
of its initial value. The presence of this voltage may or may not be noticeable or cause a
problem, depending on the circumstance.
If the resistance is increased 100 GΩ, not only is the voltage generated increased to 10 kV,
but on cessation of the current, the voltage will take 10 seconds to fall to 37% of its initial
value. The presence of this voltage for such a long time could lead to the person experiencing
shocks on touching something or discharging to cause some problem.
In ESD control, a different definition of charge decay time is usually used in standard
measurements, and often the time for charge to reduce to one-tenth of its initial value is
measured (Figure 2.3). This value is theoretically equal to 2.3𝜏.
In practice, the charge decay time is often measured from the starting voltage down to
a certain threshold voltage, e.g. 100 V. Polymers may have time constants of many tens or
hundreds of seconds, or even days under clean dry conditions.
In practice, the simple model does not always correspond well with material behavior.
Measured charge decay curve may depart considerably from the ideal exponential, and the
measured time “constant” varies with measurement conditions. Often with high resistance
materials the decay time lengthens as the surface voltage drops and may become very long
at low voltages.

2.3.4 Conductors and Insulators Revisited


In many engineering fields, conductors are often thought of as materials such as copper or
aluminum that have very low resistance or resistivity (see Section 1.7), much less than 1 Ω.
In ESD control, materials that have a much higher resistivity than this may be thought of
as conductors. In practical electrostatic control, materials and equipment are often defined
24 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

as conductors or insulators based on either a measured resistance or a charge decay time,


or both. The model of Figure 2.1 can be used to explain this.
As charge generation rate (current I) in static electricity is often low, even a relatively
high value of leakage resistance R (Figure 2.1) may pass the current to give low voltage,
V = IR. In ESD control, a resistance of 1 MΩ (106 Ω) could be considered quite conductive
and would reduce the electrostatic voltage in the previous example to 1 V. As an example, in
a case where the charge generation currents normally experienced in practice are expected
to be no more than 1 nA, calculations can be made on this basis. Alongside this, it may be
wished to limit voltages to some level, e.g. 100 V. Given these constraints, the model and
Ohm’s law show that resistances up to V/I = 102 /10−9 = 1011 Ω would be acceptable.
In an application (e.g. electrostatic hazards avoidance in industrial processes) where
higher charge generation is expected, the allowable resistance may be considerably smaller
(IEC 60079-32-1).
A second way of looking at the matter is to decide how long a transient charge built up
on a material or object may tolerably be allowed to remain without problems occurring.
This may be evaluated in terms of the charge decay time. If a conductor has capacitance
around 10 pF, resistance to ground of 1011 Ω will give a charge decay time of one second,
and in the absence of charge generation a stored charge will reduce to only 5% of its initial
value within three seconds. In manual assembly and handling processes, this will usually
be fast enough to avoid problems. For materials, this decay time corresponds to a permittiv-
ity of 10−11 Fm−1 and resistivity of 1011 Ω. The permittivity of air is around 0.9 × 10−11 Fm−1 ,
and many plastics are around 2 × 10−11 Fm−1 . The presence of higher capacitance or mate-
rial permittivity, or a requirement for faster charge decay, may lead to a lower maximum
acceptable resistance.

2.3.5 The Effect of Relative Humidity


Water is an electrically conducting material. Moisture from the air forms a thin layer on
the surface of many materials and can contribute to their apparent electrical conductivity.
Some materials, especially natural materials such as paper, reduce by orders of magnitude
in their resistivity as relative humidity increases from dry conditions.
As material surface resistance is increased under dry conditions, electrostatic charge
build-up is often greatly enhanced. Some ESD control materials use additives to attract
moisture to a polymer surface and provide static dissipative behavior. These materials
may not work well at low humidity. As a rule of thumb, electrostatic charge build-up is
generally increased for humidity less than about 30% rh.
The external atmospheric humidity varies daily with the climate and weather, in a range
from below 10% rh (cold and dry winter conditions) to 100% rh (fog). The atmospheric
relative humidity often has a large effect on material resistance, especially for materials
that have resistance above about 1 MΩ. The effective resistance and charge decay times can
be reduced over several orders of magnitude with increasing relative humidity for some
materials.
Air relative humidity is a strong function of temperature and reduces as temperature
increases for a given moisture content. Relative humidity is approximately halved by a 10 ∘ C
rise in temperature, if no moisture is added or removed. If, as in winter, cold air is brought
2.4 Conductors in Electrostatic Fields 25

Table 2.3 The effect of humidity on typical electrostatic voltages (MIL HDBK 263).

Action Voltage observed


@ 10–20% rh @ 65–90% rh

Person walking across carpet 35 000 1 500


Person walking across vinyl floor 12 000 250
Person working at bench (not grounded) 6 000 100
Vinyl envelope 7 000 600
Polythene bag picked up from bench 20 000 1 200
Chair padded with polyurethane foam 18 000 15 000

indoors and heated, very low relative humidity can result. Hence, ESD problems can be
seasonal and occur often in winter. Even in a room where the relative humidity is con-
trolled, dry local microclimates can form where there are heat sources such as equipment,
especially if air circulation is restricted.
A view of the effect of relative humidity on static electricity in daily life is indicated by
the following typical voltages (Table 2.3) given by MIL HDBK 263 as observed at different
ambient humidities. These are indicative and cannot be used to predict voltages occurring
in real situations.

2.4 Conductors in Electrostatic Fields

2.4.1 Voltage on Conducting and Insulating Bodies and Surfaces


Like charges repel, and in a conductor where charges are free to move rapidly, charge will
rapidly move to the outer surface to minimize their proximity to each other. After charge has
redistributed, the voltage on all parts of the conductor is equal (equipotential). This must
be so – current flows due to voltage differences, until the equipotential state is achieved.
For an insulating object, charge does not flow freely and so the voltage at each point on the
surface is typically different from its neighbor. For intermediate materials, the time taken
to achieve near equipotential surface is several times the time constant.
For objects that have high resistivity and a long time constant, charge will redistribute to
equipotential if we wait long enough (and if the field source is not changing rapidly) – but
in the meantime surface voltages can be different.

2.4.2 Electrostatic Field in Practical Situations


For a small point or spherical charge with the balancing charge a long way distant, the
strength of the electric field falls off rapidly with distance r, as it is proportional to 1/r 2 . The
field line spread out radially (Figure 2.4). In many practical situations the object presenting
an electrostatic field source is too large to be considered a point source.
26 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Figure 2.4 Field lines (shown dashed)


emerging from a small point or spherical
charge.

0V

Figure 2.5 Electrostatic field between parallel plates.

For larger and different shaped objects, the field line pattern can be very different and the
fall-off in field strength can be much less rapid. In practice, the field lines start and finish
on conductors at different voltages in the region, and field lines may be more or less curved
at any region in space between them.
The electrostatic field between two large flat parallel plates, well away from the plate
edges, is uniform (Figure 2.5), and the field lines are parallel between the electrodes. Away
from the plate edges, the field E is uniform and is easily calculated from the voltage differ-
ence between the plates V and the distance between them d.
V
E=
d
If a conductor is placed in an electrostatic field, it has the effect of drawing the field to
itself with field lines always emerging at right angles to the conductor surface. In response,
charges on the conductor are redistributed until the voltage is the same all over the con-
ductor surface. One of the consequences of this is that any instrument we use to mea-
sure an electrostatic field inevitably changes the field it is measuring. Figure 2.6 shows
how this happens with the electrostatic field between a metal plate at voltage V and a
grounded electrostatic field meter. The field meter actually sees a field higher than the V/d
value. The same effect happens of course for any component, PCB or any other electrostatic
discharge–sensitive (ESDS) device brought into the field. The density of the field lines at the
2.4 Conductors in Electrostatic Fields 27

Distance d

Electrostatic field meter


0V

Metal
plate
Field E
Voltage
V

Figure 2.6 Electrostatic field between a field meter and metal plate at voltage V .

surface are related to the field strength and surface charge density induced at the surface.
A high concentration of field lines indicates a high field strength.
Field lines tend to congregate at the tip or edge of an object, and the electrostatic field
strength becomes more intense in these regions. Discharges tend to occur preferentially
from high field strength regions at sharp edges on objects. This is used to an advantage, for
example, in using sharp pins to produce intense fields and corona discharges as a source of
ions in an ionizer for charge neutralization.
For a charged insulating surface, the situation is even more complicated. A charged insu-
lator will normally have a highly variable charge density over its surface. The surface voltage
is highly dependent on the surface charge density and presence of other materials nearby.

2.4.3 Faraday Cage


For a conducting object in an electric field, charge flows until all points on the surface are
at the same voltage. This must be so, as any voltage difference would cause a current to
flow in the conductor. If the object is hollow, then the field inside the object is zero, because
an equipotential conductor surrounds it (Figure 2.7). An object placed within the hollow
conductor would therefore be shielded from the effects of an external field. This hollow
conductor is called a Faraday cage.

2.4.4 Induction: An Isolated Conductive Object Attains a Voltage When in an


Electric Field
If we consider the behavior of a conductive object as a capacitor, we can quickly understand
that the voltage on an isolated conductive object will change under the influence of a nearby
charged object and its electric field.
In Figure 2.8 an earthed electrostatic field meter is monitoring the voltage V m of a metal
plate. There is an effective capacitance Cm between the metal object and the field meter.
The metal object has no net charge and is initially at zero volts.
28 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Conductive
container surface
has equal voltage

No field
within container

Figure 2.7 Faraday cage.

Figure 2.8 Voltage developed on a


metal plate in an electric field.
–Q +Q
Cg
Electrostatic
Cm field meter

+V
Vm
Charged Metal
object plate

A positively charged object is then brought near. As it approaches, it couples to an increas-


ing amount of negative charge Q on the metal object attracted to the side nearest the pos-
itively charged object. The same amount of positive charge is repelled and appears on the
side of the metal object nearest the field meter, coupled to an equivalent negative charge on
the (grounded) field meter. The field meter sees a positive voltage on the metal object, as
voltage on the metal object increases (the capacitor Cm is charged) by an amount.
Q
Vm =
Cm
While the total amount of charge on the plate does not change, an amount –Q is attracted
toward the positively charged object to charge Cg , and an amount + Q is repelled to charge
Cm .

Q = Cm Vm = Cg (V − Vm )

Cg V
Vm =
(Cg + Cm )
2.4 Conductors in Electrostatic Fields 29

Note that the metal plate would have a capacitance and voltage with respect to ground,
even if the field meter were not present. Although there is now a voltage on the plate, the
net charge remains zero (+Q −Q), and it is not charged!
This changing voltage on an object happens in practice if any conductive object passes
through an electric field. If, for example, an integrated circuit passed into an electric field
arising from a charged garment surface, it could acquire a voltage in this way. If it were
subsequently grounded in this state, an ESD event would happen.

2.4.5 Induction Charging: An Object Can Become Charged by Grounding It


In Figure 2.8, we saw that a voltage was induced on the metal object when a charged object
was brought nearby. In that situation, we can discharge the capacitor Cm by connecting a
ground wire between the metal object and earth. The positive charges on the metal object
flow to earth. The voltage V m is then zero (Figure 2.9). Note that at the time of connection,
an ESD occurs as the charge flows to earth. This is an important phenomenon in ESD con-
trol – an ESD occurs when two conductors at different voltages make contact, e.g. when
grounding a conductor in the presence of an electrostatic field.
If the earth wire is then removed, the metal object remains at zero volts. However, it has a
net negative charge Q, as the balancing positive charge Q has flowed away. The metal plate
is now charged although the voltage on it is zero! If the charged field source object is then
taken away, the voltage on the metal object rises to a negative voltage due to its negative
charge.
−Q
Vm =
Cm
This process is called charging by induction. It can happen in practice if an object, tool,
device, or person becomes grounded temporarily when in an electrostatic field.
If a person or object can become charged and can act as a source of electrostatic field, then
a nearby device can be subjected to that field. If the device is momentarily grounded, ESD
occurs at that time, and it can become charged by induction. This can leave it in a charged
state, at risk of ESD occurring on subsequent contact with another conductor at different
voltage – grounded or not. If a grounded person moves to pick up a sensitive device within
an electrostatic field, they may cause an ESD event when they touch the ESD-sensitive
device. Practical demonstrations of these processes are given in Section 12.7.10.

Figure 2.9 An earthed metal plate in


an electric field becomes charged by
grounding. –Q
Cg
Electrostatic
Cm field meter
+Q
+V
Vm
Ground
Charged Metal wire
object plate
30 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Figure 2.10 Faraday pail.

Charge
Charge Q
Q Conductive
container

Induced voltage differences can also lead to breakdown over small gaps between nearby
conductors in a field, if the voltage difference exceeds the gap breakdown voltage. This can
also lead to ESD risks.

2.4.6 Faraday Pail and Shielding of Charges Within a Closed Object


If a charged object is placed within a closed hollow conducting object such as a box, then
the field lines from the charge couple to the surrounding conductor (Figure 2.10). The net
charge contained within the conducting box induces an equal net charge on the box.
This principle is used to measure electrostatic charge on items by placing the charged
item in a container, which is known as a Faraday pail.
If the container is grounded, then the outside world is shielded from electrostatic fields
arising from the charges. If it is not grounded, then it is itself a charged object and can be a
source of ESD.

2.5 Electrostatic Discharges


Normally air is an excellent insulator. If, however, the electrostatic field strength exceeds
about 3 MV m−1 (3 kV mm−1 ), the insulating properties of air breaks down and ESD occurs.
A large amount of stored charge can be rapidly dissipated by this event. The discharge may
be sudden, as in sparks, or it may be gradual as in corona discharge.
An understanding of ESD is important in understanding the characteristics of ESD
sources.

2.5.1 ESD (Sparks) Between Conducting Objects


The spark discharge occurs between conducting electrodes that initially have a high voltage
difference between them. Large energies (μJ to >1 J) may be dissipated in very short, or
long, times (ns to >ms) depending on discharge circuit (including the load characteristics).
Peak currents are typically greater than about 0.1 A and can exceed 100 A. The discharge
2.5 Electrostatic Discharges 31

waveform is highly dependent on the source and “load” circuit characteristics and can have
unidirectional or oscillatory waveforms (see Section 2.7).
The energy E stored in a capacitor C charged to voltage V is easily calculated using this
simple formula
E = 0.5CV 2
In the absence of significant series resistance, it is often reasonable to assume that all this
energy is transferred to the discharge.
The electrical breakdown field strength of about 3 MV m−1 is valid for normal air pressure
and rather large distances (e.g. for a gap of 10 mm and large diameter or flat electrodes,
the breakdown voltage would be about 30 kV). The relationship between breakdown field
strength and air pressure is given by Paschen’s law (Kuffel et al. 2001) and is nearly linear for
larger gaps and uniform fields. At smaller gap distances d the breakdown voltage reaches a
minimum (known as the Paschen minimum). As breakdown voltage V b is also dependent on
atmospheric pressure P, the Paschen curve is usually plotted as breakdown voltage against
the product Pd (Figure 2.11). For air, according to Paschen’s law, below about 350 V no
breakdown occurs (minimum Pd 0.55 Torr cm, or 7 μm at 1 atm), and ESD can happen only
with direct metal-to-metal contact. There is evidence that in practice discharges can occur
through small gaps below the Paschen minimum voltage, possibly due to field emission
(Wallash and Levitt 2003).

2.5.2 ESD from Insulating Surfaces


If a conductive electrode approaches a charged insulating surface, a “brush” discharge can
occur. Several contributory discharges occur on the insulating surface, radiating from a cen-
tral spark channel – the whole looks rather like an old-fashioned twig brush.
Brush discharges are less well documented than spark discharges. They typically have a
lower peak discharge current than sparks (0.01–10 A) and unidirectional waveforms with
fast rise and quasi-exponential decay (Figure 2.12) (Norberg et al. 1989; Norberg 1992;

Breakdown
voltage Vb

Paschen
Minimum minimum
breakdown
voltage Vb

Minimum Pressure × gap Pd


breakdown gap
Pdmin

Figure 2.11 The relationship between breakdown voltage and spark gap Pd (Paschen curve).
32 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

50
–0
–50 0 50 100 150 200 250 300 350 400 450
–50
–100
–150
Current (mA)

–200
–250
–300
–350
–400
–450
Time (ns)

Figure 2.12 Discharge from negatively charged (>20 kV) insulating surface.

Norberg and Lundquist 1991; Smallwood 1999; Landers 1985). The power dissipation and
energy of a brush discharge is not easy to calculate.

2.5.3 Corona Discharge


Very high electrostatic fields can occur at sharp edges or points on conductors in an elec-
trostatic field. When this field reaches or exceeds a threshold, ions can be sprayed from the
point or edge into the air, as a small continuous ion current. This effect is used in ionizers
to create a source of ionized air for neutralizing electrostatic charges.

2.5.4 Other Types of Discharge


Where an insulating surface is backed by a conducting material, and high charge levels can
be generated, a strong propagating brush discharge can occur. This type of discharge is not
usually of concern in electronic component handling, but it can be of concern as an ignition
source in industrial processes.

2.6 Common Electrostatic Discharge Sources

Any object that is at a different voltage from an ESDS device can be a source of ESD if
the object can touch the device or come close enough for a discharge to jump a small air
gap between them. The ESD that occurs may be more or less damaging or problematic
according to its characteristics. Different ESD sources produce waveforms with very differ-
ent characteristics in terms of parameters such as peak current, duration, energy and charge
transferred to the device, and frequency spectrum. Even an apparently similar source can
give widely different ESD waveforms under different circumstances. Some examples of real
ESD waveforms are given next – these may or may not be representative of ESD produced
from similar sources in other real situations, which may be highly variable.
2.6 Common Electrostatic Discharge Sources 33

2.6.1 ESD from the Human Body


The charged human body is an important source of ESD, both in device damage in manu-
facturing processes and in electromagnetic susceptibility of working systems. The body is
a conductor in electrostatic terms and can have a variable capacitance up to about 500 pF,
although considerably higher capacitance has been measured under some circumstances
(Jonassen 1998; Barnum 1991). The capacitance of the human body is dependent on its
proximity to other objects such as furniture and walls. When standing, the characteristics
of footwear and the nature of the floor are important factors.
Although the body is a conductor, it has significant resistance, and this limits the current
flow and causes ESD waveforms from the human body charged to higher voltages (more
than a few kV) to have a characteristic unidirectional wave shape (Figure 2.13). The peak
discharge current is typically in the range 0.1–10 A with duration of around 100–200 ns.
Discharges from the human body at lower voltages can have highly variable waveform and
current characteristics (Kelly et al. 1993; Bailey et al. 1991; Viheriäkoski et al. 2012). This
can significantly affect related risks of ESD damage.

60

50
Current (mA)

40

30

20

10

0
–100 0 100 200 300 400 500 600
–10 Time (ns)

1.8
1.6
1.4
1.2
Current (A)

1.0
0.8
0.6
0.4
0.2
–0.0
–50 0 50 100 150 200
–0.2
Time (ns)

Figure 2.13 Example of waveform of discharge from the author charged to 500 V and discharging
via skin of a finger (above) and small metal object (coin, below).
34 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

2.6.2 ESD from Charged Conductive Objects


When a highly conductive (e.g. metal) object is not grounded, it can gain a high voltage
either through triboelectrification or through induction in an electrostatic field. If this con-
ductor now touches another grounded conductor or device, an ESD event will occur.
The waveform of real-world ESD of this type can be highly variable depending on the
characteristics of the source and discharge path. Typically, with low resistance source and
discharge path materials, a high discharge current reaching tens of amps can occur. The
waveform is often oscillatory, with the frequency determined mainly by capacitance and
inductance of the source and discharge circuit. The waveform duration may be from a few
nanoseconds to hundreds of nanoseconds.
If there is significant resistance in the discharge circuit, the peak ESD current and dura-
tion of the discharge are reduced. (For small ESD sources, the effective resistance of the
discharge can be significant.) The number of oscillation cycles is also reduced. Eventually
with sufficient circuit resistance, a single peak may occur. In practice, discharges from small
metal items can look like charged device ESD (Figures 2.14 and 2.15).
If the resistance of the discharge circuit is sufficiently high, the peak ESD current is fur-
ther reduced, and a unidirectional waveform with fast-rising edge but long decay may occur.

2.0

1.5
Current (A)

1.0

0.5

0.0
–5 0 5 10 15 20 25 30

–0.5 Time (ns)

Figure 2.14 ESD waveform from screwdriver blade charged to +530 V. Charge transferred 0.03 nC.

3.0

2.5
Current (A)

2.0

1.5

1.0

0.5

0.0
–5 0 5 10 15 20 25 30
–0.5

–1.0 Time (ns)

Figure 2.15 ESD waveform from a160 × 180 mm metal plate charged to 550 V. Charge transferred
2.5 nC.
2.6 Common Electrostatic Discharge Sources 35

2.6.3 Charged Device ESD


When a component touches a highly conductive object (e.g. metal) at a different voltage, a
very short duration high discharge current ESD event occurs. The voltage difference may
occur if the component is charged or the object is charged, or both. The same type of dis-
charge will occur if either the component or the object is grounded.
The voltage on the device may arise from tribocharging or induced as a result of nearby
electrostatic field sources. Often field-induced voltages can give the highest voltages arising
on the device. Some examples of field-induced charged device ESD obtained in a labora-
tory experiment are given in Figure 2.16. In this experiment, the devices were slid down a
charged PVC tube onto a 1.7 Ω target plate connected to a fast digital storage oscilloscope
(500 MHz bandwidth, 2 Gs s−1 sample rate).
The fast high current peak typical of charged device ESD can be seen. The indicated peak
current and rise and fall times of the waveform peaks are probably under-represented, as
these waveforms are typically faster than the measurement system used here.

0.2

0.1

0
–5 –5 –5 10 15 20
–0.1
ESD current (A)

Time (ns)
–0.2

–0.3

–0.4

–0.5

0.4

0.2

0
–5 0 5 10 15 20
–0.2
ESD current (A)

–0.4

–0.6

–0.8

–1 Time (ns)

–1.2

Figure 2.16 ESD waveforms from charged integrated circuits: (above) 32-pin plastic-leaded chip
carrier and (below) 24-pin dual-inline package.
36 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

2.6.4 ESD from a Charged Board


PCBs often enter a production process highly charged. They can remain charged for long
periods or can become charged as they are transported or go through a handing or assembly
process. Voltages on the board up to 1000 V are not unusual, although the measured voltage
will typically change with the proximity of the PCB to other objects. A PCB can also have
high induced voltage if there is a highly charged insulator or another source of electrostatic
field nearby.
If a conductor (e.g. track or component pin) on the charged board touches a highly con-
ductive machine part (e.g. stop pin), a charged board ESD event can occur (Figure 2.17). The
PCB can have high effective capacitance, so this type of discharge can be quite energetic.

2.6.5 ESD from a Charged Module


Many products, modules, or subassemblies have an insulating plastic housing containing a
circuit board. The connections to this may be brought out to terminations at flying leads or
a connector.
The housing can become highly charged, e.g. by rubbing or removal from packaging,
inducing a high voltage on the PCB within the housing (Figure 2.18). If a connection is

0.5

0
–2 0 2 4 6 8 10 12 14 16
Time (ns)
–0.5
ESD Current (A)

–1

–1.5

10

0
–2 0 2 4 6 8 10 12 14 16
–2
Time (ns)
ESD current (A)

–4

–6

–8

–10

–12

–14

Figure 2.17 ESD waveform from a printed circuit board (above) charged to 1 kV (below) field
induced charged by insulator 40 mm away.
2.7 Electronic Models of ESD 37

20

Current (A)
1.5

1.0

0.5

0.0
–20 0 20 40 60 80 100 120
–0.5 Time (ns)

Figure 2.18 ESD waveform from a charged automotive module taken out of a polythene bag.
Charge transferred 35 nC.

6
5
4
3
2
Voltage (kV)

Time (s)
1
6
–1 0 5 10 15
–2
–3
–4
–5

Figure 2.19 Voltage on an automotive cable core as polythene packaging is removed.

made to the module in this state, a discharge can occur at the termination at which contact
is made.

2.6.6 ESD from Charged Cables


Cables and wiring looms can have significant capacitance between the wires in the cable
and between the wires and ground. This can be of the order of 100 pFm−1 . Wires in the cable
can become charged by various means such as by movement of the cable or by removal of
the cable from packaging (Figure 2.19). If the cable is connected to equipment in this state, a
charged cable ESD event can occur to the first terminal to make a connection (Figure 2.20).

2.7 Electronic Models of ESD


Many ESD sources can be simply modeled using a simple R-L-C circuit (Figure 2.21). The
values of each component vary widely between different sources and help to explain the
different types of waveforms observed.
At the heart of any ESD source is charge build-up and storage. This is represented by
the capacitance in the model C. In many cases in real life, this charge storage may be on a
conductor (e.g. metal item).
38 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

2
Current (A)

1
Time (ns)
0
–50 0 50 100 150 200 250 300
–1

–2

–3
(a)

0
Current (A)

–50 0 50 100 150 200 250 300


–1 Time (ns)

–2

–3

–4
(b)

Figure 2.20 ESD waveform from a charged automotive wiring loom cable lying against an earthed
metal plate. Positive (above) and negative (below) charging polarity.

Series resistance in discharge Series inductance in


path Rs discharge path Ls

Sorce voltage ESD impedance


VESD RESD

Cs
Capacitance Victim device
Rd

Figure 2.21 Electronic model of a simple ESD source.

The discharge is usually initiated by a breakdown of an air gap or some other insulating
medium. At low voltages, it can also be initiated by contact or near-contact between two
conductors. The discharge can itself have significant impedance RESD that can affect the
waveforms produced and the energy delivered into the victim device. Often, however, this
is negligible compared to the other impedances in the circuit, especially for larger ESD
events.
2.7 Electronic Models of ESD 39

After the discharge commences, the current flows through some circuit that includes
some elements of resistance Rs and inductance Ls . These are normally due to the resistance
and electrical properties of the materials in the current path.
In the case of ESD to a victim device, the device also has impedance, modeled in this
simple circuit by a resistance Rd . In practice, a nonlinear impedance would be more typi-
cal of a semiconductor device. The impedance of the spark channel is highly variable and
nonlinear.
For simplification, the total circuit resistance R is assumed to be linear and is the sum of
the circuit resistances.

R = Rs + RESD + Rd

The discharge current I ESD of this circuit has the form

IESD = A exp(𝛼t) − exp(𝛽t)

For derivation of the equations for this and the following equations, the reader is referred to
other texts (e.g. Agarwal and Lang 2005, https://en.wikipedia.org/wiki/RLC_circuit). This
equation has two roots α, β given by
( ) √( )
R R2 1
𝛼, 𝛽 = − ± −
2Ls 4L2s Ls Cs

The waveform shape takes very different forms depending on the circuit component val-
ues. If the total circuit resistance is large and dominates the discharge path impedance,
the waveform has a unidirectional shape, simulated in Figure 2.22 using model component
values given for human-body model ESD (see Table 3.12). This occurs when
R2 1
2

4Ls LC

I (Rd)
260mA
240mA
220mA
200mA
180mA
160mA
140mA
120mA
100mA
80mA
60mA
40mA
20mA
0mA
0ns 60ns 120ns 180ns 240ns 300ns

Figure 2.22 Simulated overdamped device current waveform IESD for dominant circuit resistance:
Rs = 1500 Ω, Rd = 10 Ω, RESD = 0 Ω, Ls = 10 000 nH, C s = 100 pF, V ESD = 500 V.
40 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

The discharge current rises rapidly to a peak I p that, when inductance is small, approaches
the value and polarity near that predicted by Ohms law.

Ip ≈ VESD∕
Rs

Thereafter, the current drops nearly exponentially with decay time approaching Rs CESD .
At the other extreme, if the circuit resistance is insignificant compared to the inductive
and capacitive impedance, the waveform is quite different. This occurs when
R2 1

4L2s LC
The waveform rises to a peak and then oscillates negative and positive about zero. The over-
all amplitude decays exponentially with time, simulated in Figure 2.23 using values given
for machine model ESD (Table 3.12).
Between the two extremes, the waveform duration decreases and is minimum around the
point of critical damped waveform, where the waveform changes between the two different
shape types. This is simulated in Figure 2.24 using model values close to those given for the
charged device model (Table 3.12). This occurs at the condition
R2 1
=
4L2s LC
Practical ESD sources often require the addition of more components (e.g. additional capac-
itors) to better represent additional charge storage (e.g. metal parts) and other features that
may be present. These may contribute further current peaks or modify the shape of the
waveform (Verhage et al. 1993).
The stored energy EESD in a conductive ESD source is given by
2
EESD = 0.5Cs VESD

I (Rd)
6.3A
5.4A
4.5A
3.6A
2.7A
1.8A
0.9A
0.0A
–0.9A
–1.8A
–2.7A
–3.6A
0ns 60ns 120ns 180ns 240ns 300ns

Figure 2.23 Simulated underdamped device ESD current waveform for the case of low circuit
resistance (dominant inductive and capacitive impedance): RESD = 10 Ω, Rd = 10 Ω, Ls = 750 nH,
C s = 200 pF, V ESD = 500 V.
2.8 Electrostatic Attraction (ESA) 41

I (Rd)
11A
10A
9A
8A
7A
6A
5A
4A
3A
2A
1A
0A
–1A
0.0ns 1.0ns 2.0ns 3.0ns 4.0ns 5.0ns

Figure 2.24 Simulated device ESD current waveform for near critical damping: RESD = 20 Ω,
Rd = 10 Ω, Ls = 2.5 nH, C s = 10 pF, V ESD = 500 V.

All this energy is dissipated in the total circuit resistance R. Only a fraction of this is the
energy dissipated in the device Ed .
EESD Rd
Ed =
R
In a source such as the charged human body that has significant resistance, most of the
stored energy is dissipated in the circuit (body) resistance, and only a small fraction is dis-
sipated in the victim device. In contrast, a charged metal object is a low-resistance ESD
source, and most of the stored energy can be dissipated in the victim device. This is one
reason why some components may be damaged by a lower voltage with a metal ESD source
compared to a charged person. In general, the likelihood of ESD damage to a component by
ESD from a source will depend on the susceptibility of the device to ESD current, voltage,
energy, or other parameter of the discharge. This is further discussed in Chapter 3.

2.8 Electrostatic Attraction (ESA)


Where there is an electrostatic field, charged particles in the vicinity will experience an
attractive or repulsive force. A lesser known effect is that uncharged particles can be
attracted or repelled in a convergent or divergent field – this is known as dielectrophoresis
(Cross 1987).
The direction of the force depends on the polarity of the charged particles and the field.
The force acts such that like polarity charges repel and unlike polarity attract. So, a positive
charge will experience a force toward a more negative potential, and vice versa.

2.8.1 ESA and Particle Contamination


These effects can be important where cleanliness of the product is essential. In a clean room
for wafer fabrication, an electrostatic field can cause charged dust particles that are present
42 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

to be transported to a wafer within the field. Particle contamination can then cause loss of
product yield (Welker et al. 2006).
Other processes in which product cleanliness is important can include

● Manufacture of flat-screen displays. Loss of even a small number of pixels due to contam-
ination can result in rejection of the product.
● Packing of consumer products where dust or particle contamination can mar the appear-
ance of the product before purchase.
● Assembly of optical systems where performance can be reduced by contamination.
● Assembly of medical systems where infection of the user may be a risk.

2.8.2 Neutralization of Surface Voltages by Air Ions


Clean air is naturally a good insulator with very few mobile charged particles present.
A small number of ions are naturally generated when air molecules are split into posi-
tive and negative ions by the action of natural radioactivity or cosmic rays (Jonassen 1985).
These ions will be repelled or attracted by surface charges due to the electrostatic field. The
ions move in the direction of the electrostatic field.
The charge migration rate and direction are dependent on the ion charge and other fac-
tors, as well as the electrostatic field strength and direction at the point in space where the
ion is located. In still air, the ion drift velocity vd is related to electrostatic field E by the ion
mobility μ.

vd = μE

The mobility of the ion is dependent on the ion size. In air, charges bind to water, nitrogen,
and other molecules or particles and form small or large ions. Small ions have mobility in
the range 1−2 × 10−4 m2 V−1 s−1 (Jonassen 1985). Large ions have mobility in the range
8 × 10−7 to 3 × 10−8 m2 V−1 s−1 .
The number of air ions present can be increased using an ionizer. These produce air
ions by various means such as corona discharge, radioactive, or X-ray ionization of the
air. Radioactive and X-ray ionization sources provide both polarity ions by splitting air
molecules into positive and negative ions.
Corona discharge sources use a high voltage applied to a sharp electrode (e.g. needle) to
produce ions of one polarity. A nearly balanced ion source can be produced by this method
by using an alternating current (AC) high voltage or two separate sources of opposite
polarity.
A charged surface produces an electrostatic field surrounding it that repels like polarity
ions and attracts opposite polarity ions. That is, a negatively charged surface repels negative
ions and attracts positive ions. A positively charged surface attracts negative ions and repels
positive ions. Opposite polarity ions will drift to the charged surface at a rate proportional
to the field strength and in numbers proportional to the ion concentration. An opposite
polarity charge on reaching the charged surface neutralizes an equal charge, reducing the
net surface charge and electrostatic field. The ion drift represents a neutralizing current,
limited by the ion concentration and field strength.
2.8 Electrostatic Attraction (ESA) 43

2.8.3 Ionizers
Ionizers are devices used to generate air ions for neutralization of surface charges on
charged materials and objects (Jonassen 1985, 1986). These are available in various types
based on passive, radioactive, electrical, and other principles of operation.
Passive ionizers rely on high electric fields developed around sharp points or edges
on earthed conductors to generate air ions by corona discharge. These will always
generate ions of the opposite polarity to the voltage producing the field at the point or
edge. Unfortunately, corona discharges do not occur below a threshold field strength,
and this means there is a threshold voltage, known as the corona inception voltage,
below which ions are not produced, and neutralization does not occur. This threshold
voltage may be several kilovolts (kV). This means that passive ionizers cannot be used
to reduce voltages to below this threshold, and they are seldom useful in ESD control in
electronics manufacture. They do find major applications in industrial processes such as
plastic film manufacturing, printing, copiers, and other processes involving insulating
materials.
Electrical ionizers also use the high electrostatic fields at points, usually needles, to
generate air ions. In this case, however, a power supply is used to raise the needle to a
voltage above the corona inception voltage to ensure that sufficient ions are produced.
A balanced ion source with equal numbers of positive and negative ions is usually required.
This can be produced either by using an AC driving voltage or by using two sets of nee-
dles, one at positive voltage and the other at negative voltage. This can produce an ion
stream that is nearly balanced, but it is difficult to produce an exact balanced ion stream
by this method.
Radioactive ionizers use a radioactive source to ionize the air by impact of radioactive par-
ticles with molecules in the air. Exactly balanced ion streams are produced by this means, as
the molecule splits into two ions, one positive and one negative ion. The rate of production
of ions is small and limited by the level of radioactivity of the ionizing material.

2.8.4 Rate of Charge Neutralization


Each ion that reaches a charged surface will add its charge to the surface. As this is usually
the opposite polarity to the surface charge, the arriving ion will normally neutralize an equal
and opposite polarity surface charge. The rate at which ions arrive at the surface depends
on the electrostatic field strength, the concentration of ions in the air, and the mobility of
the ions (Jonassen 1986). This can be thought of as an ion current flow. Different air ions
can have different mobility depending on the size, polarity, and charge of the ions.
As the charged surface neutralization process continues, surface voltages and field
strength reduce, and the drift of ions to the surface reduces with the reducing field
strength. The neutralization process continues ever more slowly until the electrostatic
fields and thus forces on the ions are insufficient to attract further charge from the air.
If the surface voltage is monitored, the voltage is seen to reduce at a reducing
quasi-exponentially decaying rate. This is used in a charge plate monitor (CPM) instru-
ment to measure the effectiveness of ionizers in neutralizing charges on surfaces. A typical
voltage decay curve obtained is shown in Figure 2.25.
44 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Offset voltage
100
0
25
–5 –100 0 5 10 15 20

–200 time (seconds)


CPM voltage (V)

–300
–400 Decay time
–500
–600
–700
–800
–900
–1000
–1100

Figure 2.25 Electrostatic voltage on a CPM plate reducing during charge neutralization showing
decay time and offset.

Charge neutralization by ionizers is typically quite a slow process and can take tens of sec-
onds or longer, depending on the electrostatic field strength and ion density in the region
surrounding the charged surface. This charge density, and the speed of charge neutraliza-
tion, can be affected by many factors.

2.8.5 The Region of Effective Charge Neutralization Around an Ionizer


The ion density in the air typically reduces with the distance from the ionizer. This is in part
because the ion cloud tends to spread out due to repulsion forces between like polarity ions.
In addition, attraction between positive and negative charges in the cloud causes them to
move toward each other and recombine to neutralize each other (Jonassen 1985).
As the ion cloud is drifting in air, movement of the air can have a significant effect on the
location of the cloud and the local ion concentration. Ions can be blown toward or away from
the region where they are wanted (e.g. by drafts from a fan or open window). Conversely, in
many situations a fan can be used to blow ions to the location they are needed to improve
charge neutralization effectiveness.

2.8.6 Ionizer Balance and Charging of a Surface by an Unbalanced Ionizer


An ion cloud is rarely accurately balanced in terms of ion charge polarity density. One con-
sequence of this is that any surface within the ion cloud tends to reach a state of excess
charge reflecting the ion density balance in the ion cloud. If, for example, the ion cloud
has excess positive ions, then surfaces within the ion cloud will tend to become positively
charged until electrostatic fields are sufficient to prevent further ion deposition. This effect
will create residual voltages on insulating surfaces and isolated conductors within the ion
cloud. With a well-balanced ionizer, this effect will be small, and the excess voltage achieved
may be only a few volts or tens of volts. This “offset voltage” is an important parameter of
an ionizer specification (Figure 2.25).
2.10 How to Avoid ESD Damage of Components 45

A poorly maintained ionizer can become seriously out of balance. This can happen in
service due to erosion or contamination of the ionizer needles. If this occurs, then sur-
faces within the ion cloud may become charged to hundreds or even thousands of volts.
Monitoring this effect over time and performing appropriate maintenance of an ionizer to
prevent the imbalance from reaching an excessive level is an important aspect of ionizer
maintenance (Simco Ion TN-003 2019).

2.9 Electromagnetic Interference (EMI)


ESDs can give waveforms that have very fast (nanosecond or subnanosecond) rise times,
high peak current levels of tens of amps, and oscillatory waveforms. Discharge current rise
rates can be of the order of 109 A s−1 , and electrostatic fields may collapse with rates of the
order of 1012 V s−1 . The ESD source may radiate or conduct strong transient electromagnetic
fields over a wide frequency band up to GHz frequencies.
These induce transients in nearby conductors, especially tracks or wires of significant
length that can act as efficient antennas at the frequencies radiated in the discharge. This
radiated or conducted interference can cause upset or malfunction to nearby equipment
such as test equipment. Giant magnetoresistive (GMR) heads have been demonstrated to
be damaged by this type of ESD transient (Wallash and Smith 1998).
EMI can also be an issue in highly automated facilities where automated test equipment
(ATE) can test large numbers of production items. EMI may cause the ATE to register a
product fail or to otherwise malfunction (Tamminen et al. 2015). This can represent a loss
of production or output.

2.10 How to Avoid ESD Damage of Components


2.10.1 The Circumstances Leading to ESD Damage of a Component
ESD damage does not occur unless there is ESD. The circumstances leading to ESD are as
follows:
● A conductor attains a high voltage through induction or triboelectrification. The conduc-
tor may be
○ A person
○ A metal or other conductive item such as a tool, machine part, or cable
○ An ESDS component
● The conductor touches, or comes sufficiently near to another conductor or ESDS device,
for ESD to occur.
● The ESD current passes through a part of the ESDS device that may be damaged by ESD.
● The discharge current, charge, energy, or voltages developed on the ESDS device must
exceed a threshold at which damage is likely to occur.
Most ESDS devices are not directly damaged by exposure to electrostatic fields. The risk is
usually due to induced voltages on the device or a nearby conductor causing ESD when the
ESDS device touches the conductor. For some high-impedance, voltage-sensitive devices,
induced voltages can cause a risk of breakdown within the device.
46 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

2.10.2 Risk of ESD Damage


The occurrence of ESD to a susceptible component does not necessarily cause a failure.
Indeed, it is likely that ESD happens before detectable damage occurs. So, ESD damage is
perhaps best thought of as a risk with a probability that may depend on many factors.

● The likelihood that ESD occurs to an ESDS device.


● Most devices have many possible ESD entry points. Each of these will typically have a
different ESD susceptibility.
● The likelihood that the ESD strength will exceed the ESD damage threshold of the device
at the point of discharge.

Considering these conditions shows why ESD damage is a probabilistic phenomenon. Only
a small fraction of ESD events may cause ESD damage, because

● The capacity of the ESD source to produce damage varies with the characteristics of the
source.
● The magnitude of the source voltage and energy will vary tremendously with the condi-
tions that cause it.
● The variation in discharge paths will cause great variation in characteristics of the dis-
charge such as peak current, waveform, rise and fall times and duration, and power and
energy deposited in the ESDS device.
● The susceptibility of the ESDS device to peak current, power, energy, waveform, and other
parameters of the discharge through it varies between devices.
● The ESD current path through the ESDS component may be through parts that have
different damage thresholds.
● The ESD event may not be sufficiently strong to exceed the damage threshold of its path
through the component.

So, ESD may occur many times during handling or processing a component, without ESD
damage occurring. The probability of ESD damage occurring may be one in hundreds or
thousands of ESD events. Combine this with the fact that damage will be discovered often
at a much later manufacturing stage, and may not be identified as due to ESD, it is small
wonder that many people get the impression that ESD damage is something that does not
happen to them!
Nevertheless, each time that a conductor contacts an ESDS component is an opportunity
for ESD to occur. Given sufficient ESD opportunities and insufficient ESD control, ESD
damage becomes inevitable.

2.10.3 The Principles of ESD Control


The principles of ESD control are remarkably simple. Each principle is aimed at reducing
a particular ESD risk.

● ESDS are handled only within an electrostatic discharge–protected area (EPA; see
Chapter 4) or under ESD-controlled conditions. This ensures that when ESDS devices
are handled, the ESD risks are controlled to an acceptable level.
References 47

● Outside the EPA in unprotected areas (UPAs), ESD protective packaging is used to protect
the ESDS device. The packaging is designed to prevent ESD sources outside the package
from having a significant effect on the ESDS device within the package. It also provides
a safe region within the package in which ESD risks are controlled.
Within the EPA, ESD risks are controlled by eliminating as far as is possible the sources of
ESD likely to damage the ESDS.
● Conductors that may touch an ESDS, especially metal items and people, are grounded
wherever possible. This ensures that as far as possible, electrostatic voltages on conduc-
tors are the same and near zero. This is to prevent them from becoming charged and a
source of significant ESD to ESDS devices.
● Where conductors cannot be grounded and might contact an ESDS, the voltage difference
between the conductor and the ESDS device must be reduced to a sufficiently low level
to prevent significant ESD risk.
● Electrostatic fields that may occur near ESDS devices are eliminated or reduced to a low
level. Insulators that may become charged and the source of electrostatic fields are, where
possible, removed from the vicinity of ESDS devices. This is to reduce the risk of induced
charging of isolated ESDS devices or other ungrounded conductors.
● Items that may contact an ESDS device are preferably made from materials that have
appreciable electrical resistance. If ESD occurs, this helps to reduce discharge current to
a safe level and absorb much of the discharge energy within the material rather than in
the ESDS device.
These measures do not eliminate ESD but help reduce the numbers of ESD occurring, the
magnitude of any ESD that occurs, and the likelihood of it damaging the ESDS. Reducing
the number of times contact is made with an ESDS device can also help reduce the likeli-
hood of ESD damage occurring. So, the simple measures of not handling the ESDS device
any more than necessary and reducing to minimum contact between the ESDS device and
other conductors that might be at different voltage can make a useful contribution to reduc-
ing ESD risk. If the materials are resistive rather than low resistance where they contact
ESDS devices, then a further reduction of risk of ESD damage is made. This is due to reduc-
tion of peak current in the discharge and absorption of energy by the resistance of the
material.
Suitably specified ESD protective packaging can reduce the risk of ESD damage occurring
in the UPA to an insignificant level. The packaging must be specified to address the ESD
risks appropriate to the ESDS component or item.

References

Agarwal, A. and Lang, J.H. (2005). Foundations of Analog and Digital Electronic Circuits.
Morgan Kaufmann, ISBN 1-55860-735-8.
Andersson, B., Fast, L., Holdstock, P., and Pirici, D. (2008). Charging of a person exiting a car
seat. Electrostatics 2007. J. Phys. Conf. Ser. 142: 012004.
Bailey, A.G., Smallwood, J.M., and Tomita, H. (1991). Electrical discharges from the human
body. In: Electrostatics –Inst. Phys. Conf. Se. 118 Sec. 2. Inst.Phys.
48 2 The Principles of Static Electricity and Electrostatic Discharge (ESD) Control

Barnum, J.R. (1991). Sandia’s severe human body electrostatic discharge tester (SSET). In: Proc.
EOS/ESD Symp. EOS13, 29–30. Rome, NY: EOS/ESD Association Inc.
Cross, J.A. (1987). Electrostatics Principles, Problems and Applications. Adam Hilger. ISBN
0-85274-589-3.
Department of Defense. Military Handbook. (1994) Electrostatic Discharge Control Handbook
for protection of electrical and electronic parts, assemblies and equipment (excluding
electrically initiated explosive devices) (metric). MIL HDBK-263B. Washington DC,
Department of Defense.
International Electrotechnical Commission. (2013) PD/IEC TS 60079-32-1. Explosive
atmospheres Part wp2-1. Electrostatic hazards, guidance. Geneva, IEC.
Jonassen N. (1985) The physics of air ionization. In: Proc. EOS/ESD Symp. EOS-7 Minneapolis
USA. Rome, NY, EOS/ESD Association Inc. pp. 59–66.
Jonassen, N. (1986). The physics of air ionization. In: Proc. EOS/ESD Symp. EOS-8, 35–40.
Rome, NY: EOS/ESD Association Inc.
Jonassen, N. (1998). Human body capacitance – static or dynamic concept? In: Proc. EOS/ESD
Symp. EOS-20, 111–117. Rome, NY: EOS/ESD Association Inc.
Kelly MA, Servais G E, Pfaffenbach T V. (1993) An Investigation of Human Body Electrostatic
Discharge. 19th International Symposium for Testing & Failure Analysis Los Angeles,
California, USA. Russell Township, OH, ASM International.
Kuffel, E., Zaengl, W.S., and Kuffel, J. (2001). High Voltage Engineering. Newnes ISBN 0 7506
3634 3.
Landers, E.U. (1985). Distribution of charge and fieldstrength due to discharge from insulating
surfaces. J. Electrostat. 17: 59–68.
Norberg, A. (1992). Modelling current pulse shape and energy in surface discharges. IEEE
Trans. Ind. Appl. 28 (3): 498–503.
Norberg, A. and Lundquist, S. (1991). A distributed RC transmission line model for electrostatic
discharges from insulator surfaces. In: Inst. Phys. Conf. Se. 118, 269–274. Electrostatics 1991.
Norberg, A., Szedenik, N., and Lundquist, S. (1989). On the pulse shape of discharge currents.
J. Electrostat. 23: 79–88.
Pirici, D., Rivenc, J., Lebey, T. et al. (2003). A Physical model to explain electrostatic charging in
an automotive environment: Correlation with experimental approach. In: Proc. EOS/ESD
Symp. EOS-25, 161. Rome, NY: EOS/ESD Association Inc.
Simco Ion (2019) Emitter point maintenance. Technical note TN-003. Available from: https://
technology-ionization.simco-ion.com/DesktopModules/Bring2mind/DMX/Download.aspx?
command=core_download&entryid=112&language=en-US&PortalId=0&TabId=145
[Accessed 16 April 2019]
Smallwood, J.M. (1999). Simple passive transmission line probes for electrostatic discharge
measurements. In: Inst. Phys. Conf. Se. 163, 363–366. Electrostatics 1999, Inst.Phys.
Tamminen P, Viheriäkoski T, Ukkonen L, Sydänheimo L (2015) ESD and Disturbance Cases in
Electrostatic Protected Areas. In: Proc EOS/ESD Symp. 5B.2, Rome, NY, EOS/ESD
Association Inc.
Verhage, K., Roussel, P.J., Groeseneken, G. et al. (1993). Analysis of HBM ESD Testers and
specifications using a 4th order lumped element model. In: Proc. EOS/ESD Symp, 129–137.
Rome, NY: EOS/ESD Association Inc.
Further Reading 49

Viheriäkoski T, Peltoniemi T, Tamminen T (2012) Paper 4A3. Low Level Human Body Model
ESD. In: Proc. EOS/ESD Symp. Tucson Ariz. USA. Rome, NY, EOS/ESD Association Inc.
Wallash A, Levitt L. (2003) Electrical breakdown and ESD phenomena for devices with
nanometer-to-micron gaps. In: Proc. SPIE 4980, Reliability, Testing, and Characterization of
MEMS/MOEMS II. San Jose, CA, SPIE.
Wallash A, Smith D. (1998) Paper 4B.6. Electromagnetic Interference. (EMI) damage to giant
magnetoresistive (GMR) recording heads. In: Proc. EOS/ESD Symp., Rome, NY, EOS/ESD
Association Inc. pp. 368–74.
Welker, R.W., Nagarajan, R., and Newberg, C.E. (2006). Contamination and ESD Control in
High-Technology Manufacturing. Wiley ISBN-13: 978-0-471-41452-0.

Further Reading

EOS/ESD Association Inc. (2015) ANSI/ESD STM3.1-2015. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Ionization. Rome, NY, EOS/ESD
Association Inc.
International Electrotechnical Commission. (2017) IEC 61340-4-7:2017. Electrostatics - Part 4-7:
Standard test methods for specific applications – Ionization. Geneva, IEC.
51

Electrostatic Discharge–Sensitive (ESDS) Devices

3.1 What Are ESDS Devices?

Electrostatic discharge susceptibility is defined in ESD ADV1.0-2009 as “the propensity to


be damaged by electrostatic discharge.” Various types of components can be susceptible to
damage from electrostatic fields or ESD (Figure 3.1). A component that is susceptible to ESD
damage is often called an ESD-sensitive device, often abbreviated to ESDS. Items susceptible
to electrostatic discharge are defined in ESD ADV1.0-2009 as “electrical or electronic piece
part, device, component, assembly or equipment item that has some level of electrostatic
discharge susceptibility.”
The list of ESDS technologies has grown, and is still growing, as new technologies have
developed over time. New device technologies may be inherently more or less sensitive to
ESD damage than earlier technologies. MIL HDBK 263 gave a list of types of ESDS compo-
nents and outlined damage mechanisms. Some of these and more recent technologies are
briefly reviewed in Section 3.4.
Electrical overstress (EOS) is damage to components due to their absolute maximum rat-
ings being exceeded by some means. ESD is a form of EOS, with other EOS stress sources
including lightning, electromagnetic pulse (EMP), and electrical transients that can occur
at the board or system level during test or operation. EOS is increasingly becoming a serious
concern and failure mode to electronic components (Amerasekera et al. 2002).
During an ESD event, electronic devices can be pushed outside their normal operating
range. Very high currents may flow for short times and may pass through unintended routes
in the component. Voltages may be applied, or may develop internally, that can greatly
exceed the component design ratings.
ESDS devices and systems can fail completely, partially, or temporarily in various ways
(MIL-HDBK-263B (Department of Defense 1994); Baumgartner n.d. ). Complete or catas-
trophic (also sometimes called hard) failures occur when a component or system is perma-
nently damaged by an electrostatic field or ESD. Some catastrophic failures may occur as
the cumulative effect of several ESD events. ESD damage is often difficult to distinguish
from other EOS sources in failure analysis, although the scale of damage to a component
in EOS damage is often rather greater than ESD damage.
Unpowered components and devices tend to usually suffer catastrophic failures. It is pos-
sible that components can sometimes be damaged and be weakened or suffer parametric
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
52 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Figure 3.1 ESD-sensitive devices range from individual transistors or diodes to PCBs or modules.

change. They may still pass functional tests to fail at a later stage. This is called a latent
failure.
It is not just individual components that may be susceptible to ESD. Any printed circuit
board (PCB), module, or assembly containing ESDS is likely to be itself an ESDS device if
not in some way protected against ESD. In some cases, the susceptibility to ESD may be
limited to remaining contact points such as flying leads, connectors, or limited exposed
conductors.
Fully assembled systems that include ESDS, protected within an enclosure or housing
that acts as an effective barrier to ESD, are normally no longer considered an ESDS device.
In many cases, these systems must be tested for ESD immunity during operation, using
standard ESD immunity tests required by electromagnetic compatibility (EMC) regulations
or requirements. These system ESD immunity tests are aimed at ensuring functionality
of the system is not compromised during operation by ESD occurring in the uncontrolled
operating environment or during consumer operations.
Working electronic systems tend to suffer partial or temporary (soft) failures due to ESD.
The system may recover almost immediately, or it may suffer effects such as program crash
or malfunction, spontaneous reset, or data corruption. Often effects such as program crash
or malfunction may be completely recovered by restarting the system. Data corruption
effects may lead to corrupted data remaining in the system.
Often the remaining connector terminations and other potential ESD entry points will
have been designed for immunity to expected ESD threats. Sometimes, a risk of ESD damage
3.2 Measuring ESD Susceptibility 53

can remain from strong ESD to components within the system connected to lines emerging
on connectors or accessible via user interfaces such as keyboards or touch screens. Severe
ESD to system parts such as cable discharge to connectors can sometimes result in physical,
hard damage to the system. For this reason, some ESD protection may need to be built into
the system.

3.2 Measuring ESD Susceptibility

3.2.1 Modeling Electrostatic Discharges


ESD can often be modeled using a simple electronic circuit (Figure 3.2). Additional
components may, however, be added to tailor the waveform to better represent real-world
waveforms (Wang 2002).
A capacitor C is charged to an ESD voltage V esd. The capacitor models the charge stored
on a human body, a metal object, or the device itself in a real situation. On initiating the
ESD event, the ESD current flows through a circuit resistance R, inductance L, and the load
device or spark gap. Although the circuit shows a switch, in real ESD the current flow is
usually initiated by breakdown of the insulating properties of the material through which
the discharge occurs, often air.
The ESD waveform characteristics can be matched to real ESD events by careful choice
of the capacitance, circuit resistance, and inductance. Occasionally, additional components
are required for tailoring the waveform to a real application.
This type of model is used both in ESD immunity test in EMC and in semiconductor
device ESD sensitivity tests. It has also been used in other industry areas, for example igni-
tion tests of explosives and flammable atmospheres. Typical component values are shown
in Table 3.1.
The simple model works well for many simple situations, particularly where the source
is a conductive material and a simple resistance and inductance can approximate the dis-
charge path. In practice, the ESD waveform is as much dependent on the properties of the
“load” circuit, which can be highly nonlinear, as on the source. The victim device or spark
channel may have significant impedance and may affect the resulting peak current, dura-
tion, and other ESD waveform properties, including whether the waveform is unidirectional
or oscillatory (See Section 2.7). A device may also have a nonlinear rectifying action due to
the presence of semiconductor junctions (Figure 3.3).

Figure 3.2 Simple electronic model of ESD


circuit.
R L

Device
C
and/or
Vesd spark gap
54 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Table 3.1 Typical ESD model simulation component values for common sources.

Model of ESD source Resistance R (𝛀) Capacitance C (pF) Inductance L (nH)

Human body 300–1500 100–300 Stray


Large metal object Stray 200 Stray
Charged device <10 Capacitance of device <10
under test (1–30 pF)
Charged insulator surface Spark resistance 8–11
(Norberg 1992) 2–8 kΩ

Charging resistor
100MΩ

Rs
Test
High voltage Device load
CESD under
generator 500 Ω
test or

Figure 3.3 Typical ESD generator circuit.

3.2.2 Standard ESD Susceptibility Tests


The three main ESD sensitivity test models used in ESD susceptibility testing of electronic
components are the human body model (HBM), machine model (MM), and charged device
model (CDM). These simulate particular ESD sources known to give significant ESD dam-
age in practice (Wang 2002; Amerasekeera 2002). The exact specification of the different
models varies in detail with standards and implementations.
Components are normally characterized using the HBM test during device qualification.
Components that may be assembled using automated handling techniques are normally
also characterized using the CDM test. Some manufacturers, but unfortunately not all, pub-
lish the results in the component data sheet.
Component characterization using the MM test is, at the time of writing, falling out
of practice. This is because the failure modes of MM ESD are similar to HBM, and the
withstand voltages of these tests normally correlate to a reasonable degree. Manufactur-
ers regard the MM test as giving little additional information over the HBM test (Duvvury
et al. 2012).
Typical examples of widely used component ESD susceptibility test standards are
ANSI/ESDA/JEDEC JS-001(EOS/ESD Association Inc., JEDEC 2017) and IEC60749-26
(HBM) (International Electrotechnical Commission 2013), IEC 60749-27 (MM) (Interna-
tional Electrotechnical Commission 2012), ANSI/ESDA/JEDEC JS-002 (CDM) (EOS/ESD
Association Inc., JEDEC. (n.d.)), and IEC 60749-28 (CDM) (International Electrotechnical
Commission 2017).
3.2 Measuring ESD Susceptibility 55

Table 3.2 ESD model component values to achieve specified waveforms in standards.

Model Standard Rs (𝛀) CESD (pF)

Human body model (HBM) IEC 60749-26 ANSI/ESD/JEDEC JS-001 1500 100
Machine model (MM) IEC 60749-27 ANSI/ESD/JEDEC STM 5.2 200
Human metal model (HMM) IEC 61000-4-2 ANSI/ESD S5.6 330 150

These are defined in terms of waveform characteristics when discharged into a defined
load, rather than model component values. The waveforms defined by these standards are
given in Sections 3.2.4–3.2.8. Typical model component values used to create these wave-
forms are given in Table 3.2.

3.2.3 ESD Withstand Voltage


During device ESD susceptibility test, the voltage on the capacitor is increased in steps
(stress levels), with a test performed at each stress level until damage occurs. The high-
est stress level at which damage does not occur is recorded as the ESD withstand voltage of
the device. So, a 100 V HBM device was not damaged in the HBM test with 100 V capacitor
charging voltage. It may be damaged by the next test voltage level up, depending on the
voltage increments used.

3.2.4 HBM Component Susceptibility Test


The charged human body is the most common and damaging source of ESD in manual
assembly. In daily life, people charge through normal movement, often up to several thou-
sand volts (kV). Typically, a person will not feel an ESD event unless their body voltage
exceeds about 2 kV (Brundrett 1976; Wilson 1972).
So, the primary measurement of device ESD sensitivity is by HBM ESD. In this test, a
charged 100 pF capacitor is discharged into the device via a 1500 Ω resistor. The 100 pF
capacitor simulates charged stored on the average human body, and the resistor simulates
the resistance of the human body and skin.
The ANSI/ESD/JEDEC JS-001standard defines HBM current waveforms for device test
purposes. The waveform parameters are defined in terms of peak current, rise time, and
duration (Figure 3.4), and they give required values for qualification of the HBM test equip-
ment measured with 0 and 500 Ω calibration loads (Table 3.3). With a device in the circuit,
waveforms will typically be different from those obtained with the calibration loads.
Device evaluation typically is done using three samples of the device. Each sample is
tested with negative and positive polarity ESD with a minimum pulse interval of 300 ms.
Pulses are applied to each pin in turn, with each power pin in turn grounded. The test
starts at the lowest voltage level and is increased progressively, providing the devices do not
fail. Pins that are not grounded or under test are left floating. Pulses are also applied to all
nonsupply pin pair combination.
56 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Current

Ip
0.9 Ip Ir

0.36 Ip

0.1 Ip td

tr Time

Figure 3.4 HBM waveform definition with 0 Ω calibration load.

Table 3.3 ANSI/ESDA/JEDEC JS001-2017 HBM waveform parameters with 0 Ω calibration load.

Peak current I p (A) Rise time (ns)


ESD Decay time Ringing
voltage 0𝛀 500 𝛀 0𝛀 500 𝛀 t d (ns) current
(V) load load load load 0 𝛀 load I r (A)

250 0.15–0.18 2.0–10 130–170 15% of Ip


500 0.30–0.37 2.0–10
1000 0.60–0.73 0.37–0.55 2.0–10 5.0–25
2000 1.20–1.47 2.0–10
4000 2.40–2.93 2.0–10 5.0–25

After pulses have been applied to all pin combinations at the test level, the device is tested
for failure. A failure is concluded when the device no longer meets the required parametric
and functional parameters. Devices are classified after test, according to their failure voltage
level (Table 3.4).
The peak current in an HBM discharge is mainly determined by the series 1.5 kΩ
resistance. Stray inductance and circuit resistance have relatively little effect. In this model,
the device has little effect on the waveform, providing it has low impedance compared to
the series resistance. In this case, much of the energy stored in the capacitor is dissipated
in the series resistor and not in the device.
The rather long rise time limit of 25 ns allowed equipment manufacturers to build testers
(that typically have high stray capacitance and slow the rising edge) for testing high pin
count devices.

3.2.5 System Level Human Body ESD Susceptibility Test


The ESD susceptibility of powered electronics systems has for many years been tested using
a variant of a human body ESD source model that uses a 150 pF capacitor and 330 Ω series
3.2 Measuring ESD Susceptibility 57

Table 3.4 ANSI/ESDA/JEDEC JS001-2017 HBM device


classification.

Classification ESD voltage failure range

0Z <50 V
0A 50 to <125 V
0B 125 to <250 V
1A 250 to <500 V
1B 500 to <1000 V
1C 1000 to <2000 V
2 2000 to <4000 V
3A 4000 to <8000 V
3B ≥8000 V

Current
Ipeak
0.9 Ipeak

I at 30 ns

I at 60 ns

0.1 Ipeak

tr Time
30 ns 60 ns

Figure 3.5 IEC 61000-4-2 system ESD test waveform.

resistor. A typical example is IEC 61000-4-2 (International Electrotechnical Commission


2008) (Figure 3.5). This is also known by some as the “human metal model.”
In system ESD tests these are applied to the system under test in two ways – as an air
discharge or a contact discharge applied from a handheld ESD gun. As the terminology
implies, a contact discharge is applied with the gun tip in contact with a discharge point
on the item under test. For an air discharge, the tip of the ESD gun is moved toward the
discharge point until a discharge occurs or the gun’s tip touches the system under test. The
61000-4-2 standard defines four levels for test (Table 3.5). Typical waveform parameters for
contact discharge are given in Table 3.6.
Although the 61000-4-2 system test waveform is in some ways like the HBM test wave-
form, there are significant differences. The IEC 61000-4-2 150 pF/330 Ω model discharge is
more severe than the 100 pF/1500 Ω model due to the higher stored energy and peak current
for the same charging voltage. The rise time is also significantly faster (ON Semiconductor
2010).
58 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Table 3.5 IEC 61000-4-2 test levels.

Contact discharge Air discharge


Level test voltage (kV) test voltage (kV)

1 2 2
2 4 4
3 6 8
4 8 15
x User defined User defined

Table 3.6 61000-4-2 contact discharge waveform parameters.


Source Peak Rise
Current ±30%
voltage current I p (A) time (ns)
Level (kV) ±10% ±25% @ 30 ns @ 60 ns

1 2 7.5 0.8 4 2
2 4 15 0.8 8 4
3 6 22.5 0.8 12 6
4 8 30 0.8 16i 8

The rise time is measured between 10% and 90% of the value of the first current peak.
The current is measured at 30 and 60 ns from the time at which the current reaches
10% of the first peak value.

Nevertheless, the 61 000-4-2 waveform has sometimes been used to test ESD withstand
of ESDS devices where particularly severe test conditions are required, especially for com-
ponents that have pins that are likely to emerge directly to system connectors. Recently the
IEC 61000-4-2 waveform has been adopted for component tests as the human metal model
(HMM) (ANSI/ESD S5.6 (EOS/ESD Association Inc. 2009)). This practice has arisen from
demand that components that have pins that are likely to emerge directly to system con-
nectors should be tested for susceptibility using the system test waveform (Ashton 2008;
Industry Council 2010a,b).
Other system test models exist, for example ISO 10605 “Road vehicles – Test Methods for
electrical disturbances from electrostatic discharge,” which is intended for use in evaluating
ESD susceptibility of electronics modules for vehicle use (International Organization for
Standardization 2008). This uses combinations of 150 or 330 pF capacitances with 330 or
2000 Ω resistance and test voltages up to 25 kV air discharge or 15 kV contact discharge.

3.2.6 MM Component Susceptibility Test


A cart (trolley), tool, machine part, or other metallic or highly conductive object can become
charged by the rolling action of wheels on the floor or by other contact with materials and
by movement. The MM ESD simulates a discharge between a large conductive object and
an ESD-sensitive device.
3.2 Measuring ESD Susceptibility 59

The IEC 60749-27 standard defines MM current waveforms for device test purposes. The
waveform parameters are defined in terms of peak current and period (Figures 3.6 and 3.7),
and they give required values for qualification of the MM test equipment measured with
0 and 500 Ω calibration loads (Table 3.7). With a device load, waveforms will typically be
different from those obtained with the calibration loads as the device under test adds sig-
nificant impedance and is likely to act as a nonlinear load.

Current
Ip1

tpm

t0 t1 t2 t3 Time

Ip2

Figure 3.6 IEC 60749-27 MM waveform definition with 0 Ω calibration load.

Current
Ip

I100

100 ns Time

Figure 3.7 IEC 60749-27 MM waveform definition with 500 Ω calibration load.

Table 3.7 IEC 60749-27 MM waveform parameters with 0 and 500 Ω calibration loads.

1st Peak current Ip1


Voltage 0𝛀 500 𝛀 2nd Peak current Ip1 Waveform I100 Period
(V) load load 0 𝛀 load current tpm (ns)

100 1.7 ± 15% 67–90% of Ip1 0.29 ± 15% 63–91


200 3.5 ± 15%
400 7.0 ± 15% <I100 × 4.5
800 14.0 ± 15%
60 3 Electrostatic Discharge–Sensitive (ESDS) Devices

A 200 pF capacitor simulates charge storage on the conductive object. There is no defined
additional series resistance or inductance, but in practice stray inductance and resistance
are always present in the discharge circuit. The stray inductance and circuit resistance,
including the impedance of the device subjected to the ESD, determines the peak current.
The fact that these are not well defined makes the MM ESD withstand test more prone to
variation than HBM. As the circuit resistance is low, most of the stored energy is dissipated
in the device.
As with HBM, device evaluation typically is done using three samples of the device. Each
sample is tested with negative and positive polarity ESD with a minimum pulse interval of
300 ms. The test starts at the lowest voltage level and is increased progressively, providing
the devices do not fail. Pulses are applied to each pin in turn, with each power pin in turn
grounded. Pins that are not grounded or under test are left floating. Pulses are also applied
to all nonsupply pin pair combination.
After pulses have been applied to all pin combinations at the test level, the device is tested
for failure. A failure is concluded when the device no longer meets the required parametric
and functional parameters. Devices are classified after test, according to their failure voltage
level (Table 3.8).
The MM waveform is highly dependent on the load and can in practice be unidirectional
or oscillatory, whereas the HBM waveform is largely defined by the series resistance and
is always unidirectional. The peak current, for a given ESD voltage, is typically an order of
magnitude greater for MM than for HBM.

3.2.7 CDM Component Susceptibility Test


In automated assembly and storage, ESDS devices can become charged by contact with
packaging or other materials or by induction in an electric field (see Section 4.4.7). A device
may become charged also due to accidental rubbing by an operator’s garment or by induc-
tion due to the electric field of a nearby charged material, garment, or operator. CDM is of
growing importance, especially in automated handling and assembly systems.
When the charged device is brought in contact with a grounded metal object, a very
short, high-current ESD event occurs. This is simulated by the CDM ESD test. Many devices
have been found to have high sensitivity to CDM ESD, which have very fast (nanosecond)
high-current transient waveforms. Large-area, thin, multilayer devices with high overall
device capacitance and small internal feature sizes tend to increase the CDM damage risk.
As the charge is stored on the device itself, the capacitor in the model is the capacitance of
the device under test conditions. The device capacitance is dependent on the package and

Table 3.8 IEC 60479-27 MM device classification.

Classification ESD voltage failure range

A Fails at 200 V or less


B Survives 200 but fails at 400 V
C Survives 400 V
3.2 Measuring ESD Susceptibility 61

any air gaps or dielectrics between the device and the ground plane. The series resistance
and inductance in the circuit are the resistance and inductance of the tester, of the spark,
and within the device. These determine the peak current and waveform in the discharge.
Two versions of the CDM test exist (Brodbeck and Kagerer 1998). In the field-induced
charged device model (FICDM) test, the device is placed on a metal plate that can be
raised to the test voltage. A thin layer of insulator separates the device from the plate
(Figure 3.8). After the plate has been raised to the test voltage, a metal “pogo pin” is
touched to the device leg to initiate discharge. The FICDM simulates real-world events
well but is time-consuming and expensive to perform.
The first and second peak current, rise time, full-width, and half-maximum height dura-
tion of the waveform are key parameters (Figure 3.9). The bandwidth of the measuring
oscilloscope also affects the measurement results, and tables are given for use with 1 and
6 GHz oscilloscopes.
The system is calibrated using a small or a large verification module in the place of the
device. These are metal discs, with the smaller giving a capacitance of 6.8 pF ±5% and the
larger giving 55 pF ±5%. The parameters for small (Table 3.9) and large (Table 3.10) verifi-
cation modules using a 1 GHz oscilloscope are given here as examples. The test condition

Fast digital
storage
oscilloscope
50 Ω input

50 Ω coaxial cable
Metal ground plane

Pogo pin moves to 1 Ω resistor shunts


contact device pin coaxial cable input

Thin insulating layer Device under test

High-voltage plane

Figure 3.8 ANSI/ESDA/JEDEC JS-002 field-induced CDM test arrangement.

Current
Ip
0.9 Ip

FWHM
0.5 Ip

0.1 Ip
tr Time

Ip2

Figure 3.9 ANSI/ESDA/JEDEC JS-002 field-induced charged device model calibration waveform.
62 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Table 3.9 ANSI/ESDA/JEDEC JS-002 CDM waveform parameters using small


verification module measured with 1GHz bandwidth oscilloscope.

Full width half


Test Peak maximum duration Undershoot
condition current I p (A) Rise time t r (ps) FWHM (ps) I p2

TC125 1.0–1.6 <350 325–725 <0.7 Ip


TC250 2.1–3.1
TC500 4.4–5.9
TC750 6.6–8.9
TC1000 8.8–11.9

Table 3.10 ANSI/ESDA/JEDEC JS-002 CDM waveform parameters using large


verification module and measured with 1 GHz bandwidth oscilloscope.

Full width half


Test Peak Rise time maximum duration Undershoot
condition current I p (A) t r (ps) FWHM (ps) I p2

TC125 1.9–3.2 <450 500–1000 <0.5 I p


TC250 4.2–6.3
TC500 9.1–12.3
TC750 13.7–18.5
TC1000 18.3

TCxxx denotes the voltage stress level but does not correspond to the actual voltage applied
to the field plate. This is because the plate voltage must be adjusted to give the required
peak current values for each test condition during the calibration process.
The CDM test is quite sensitive to changes in experimental conditions including atmo-
spheric humidity and contamination or oxidation of the verification module surfaces. At
higher 1 kV and above, corona discharge may limit the voltage achievable on the device
before discharge.
After test, devices are classified according to their failure level (Table 3.11).
A socketed device test (SDM) is easier to perform, but the result does not correlate well
with FICDM results. In the SDM method, the device is placed in a test socket and subjected
to discharges via a relay network. This technique has advantages for devices that have heat
sink fins or uneven packages and so do not lend themselves to FICDM test.

3.2.8 Comparison of Test Methods


The discharge current and time characteristics of the ESD models are very different and
yield different ESD withstand voltage results. The HBM, MM, and CDM ESD models can
be view as special cases of a generalized ESD model in Figure 3.2. The waveform is generated
by various combinations of the components L, C, and R (Table 3.12).
3.2 Measuring ESD Susceptibility 63

Table 3.11 ANSI/ESDA/JEDEC JS-002 CDM device


classification.

Classification Test condition failure range

C0a Fails at less than 125 V


C0b Survives 125 V but fails below 250 V
C1 Survives 250 V but fails below 500 V
C2a Survives 500 V but fails below 750 V
C2b Survives 750 V but fails below 1000 V
C3 Survives 1000 V

Table 3.12 ESD model parameters compared (Gieser and Ruge 1994) assuming a
10 Ω load in place of the device.

500 V HBM 500 V MM 500 V CDM

C (pF) 100 200 10


L (nH) 10 000 750 2.5
R (Ω) 1500 10 10
Waveform type Unidirectional Variable, oscillating Variable
Rise time (ns) <10 <10 0.1
1/e decay time (ns) 150 <100 1
Peak current (A) 0.3 6.5 14
Peak power (W) 0.9 400 2000
Dissipated energy (μJ) 0.08 13 0.63
Stored energy (μJ) 12.5 25 1.25

At a time when the HBM, CDM, and MM models were under development, Gieser and
Ruge (1994) compared the characteristics of the models (component values and character-
istic waveforms may be slightly different from current model versions).
Geiser and Ruge noted the following:

● In the HBM discharge, most of the energy stored on the capacitor is dissipated in the
series resistance.
● The series resistor and load form a voltage divider circuit. The applied voltage results in
a voltage on the device that could stress oxide layers.
● If the stray component impedance is low, then the HBM discharge current is largely deter-
mined by the series resistance.
● The peak current achieved by HBM is achieved at much lower ESD voltage in MM and
CDM. For the same ESD voltage, MM and CDM currents are about 10 times the HBM
value.
● The rise and decay times for CDM are around two orders of magnitude faster than for
HBM and MM.
64 3 Electrostatic Discharge–Sensitive (ESDS) Devices

● Assuming a resistive load, peak power dissipated in the load increases with the square
of current. For CDM, peak power is in the kW range. MM is lower, in the 500 W region.
HBM peak power is only in the watt region.
● The higher capacitance of MM stores about 20 times the energy of the CDM case for the
same test voltage.
● In practice, the real HBM, MM, and CDM waveforms achieved in any tester are influ-
enced by interactions with stray “parasitic” components, as well as the actual load
impedance, that may be far from purely resistive.

According to Smedes (2009), at 2 kV the HBM peak current is 1.3 A, and the duration about
200 ns. At 200 V MM, the peak current is 3.8 A with a pulse width around 30 ns. A 500 V
CDM ESD has peak current around 2–10 A and duration around 1 ns, depending on the
device capacitance.
Failure modes for HBM and MM are generally found in the on-chip protection cir-
cuits, whereas CDM failures are usually gate oxide damage. The duration of the CDM
pulse (<1 ns) is often less than required to trigger the protection circuits (Amerasekera
and Duvvury 1995). In CDM ESD the direction of discharge current in internal circuit
elements may be opposite to that of HBM and MM as the charge is stored internally
and discharges to the outside world. The protection circuits may not be designed for this
situation.
There is good correlation between HBM and MM damage stress levels, with the HBM
withstand voltage about 12 times greater than MM (Kelly et al. 1995). More generally HBM
withstand voltage is 10–20 times greater than MM. There is little correlation between CDM
damage stress levels and the other models.

3.2.9 Failure Criteria Used in ESD Susceptibility Tests


The selection of failure criteria in ESD susceptibility tests can have a big influence on the
failure threshold result (Gieser 2002). For ESD qualification of the product, a full functional
production test of the AC and DC parameters is necessary to ensure it meets all the product
specifications. However, this may not detect “walking wounded” latent damaged devices
that show weakness and could give early failures.

3.2.10 Transmission Line Pulse Techniques


Transmission line pulse (TLP) techniques primarily provide a means for the on-chip ESD
protection developer to characterize the performance of their ESD protection networks
(Gieser 2002; Smedes 2009; ESD Assoc. 2006). TLP allows the device to be stressed with
a repeatable short duration rectangular waveform pulse with well-defined duration and
peak current. The TLP waveform is typically 100 ns in duration and has a rise time of 10 ns,
but in some test arrangements these can be varied. A TLP waveform is rectangular rather
than the decaying shape of HBM. The stored energy is approximately the same as HBM
giving a correlation of 1.5–1.8 amps for each 1 kV stress of HBM test (Duvvury C. personal
communication). The current and voltage readings in each test are used to build a com-
plete pulsed V-I characteristic. The most reliable failure criterion has been found to be a
3.2 Measuring ESD Susceptibility 65

change in leakage current, and a leakage current measurement can be done after each test
to evaluate device damage.
Correlation between TLP, HBM, and MM results have been studied and some empirically
derived correlations established, but these may not be universally applicable.
Application of TLP to CDM and system-level ESD tests is less clear. A very fast transmis-
sion line pulse (VFTLP) method has been proposed to address this for CDM, in which the
pulse width is reduced to 1–5 ns with rise time around 200 ps. This has provided improved
results. Nevertheless, VFTLP remains a stress applied between two pins, whereas CDM is
a single-pin connection stress, and the device internal current flow is therefore different.

3.2.11 The Relation Between ESD Withstand Voltage and ESD Damage
There are essentially two main types of ESD susceptibility in devices – energy and volt-
age susceptibility (Baumgartner ESD TR50.0-03-03). Nevertheless, there are various ESD
parameters that affect the likelihood of damage.

● The ESD energy dissipated in an ESDS device


● The peak current in a discharge passing through an ESDS device
● The power developed in an ESDS device during a discharge
● The charge transferred in one or more discharges into an ESDS device
● Excess voltage developed across part of the device

These parameters may have different effects on different components, and real-world ESD
can have widely varying characteristics. Because of this, devices that have similar ESD with-
stand voltages may have different susceptibility to damage from real-world ESD. Changing
the ESD parameters can have a large impact on ESD risk even though the ESD source volt-
age is not changed. As an example, a 250 V CDM device may risk damage if it contacts a
metal part when the voltage difference between them is 250 V. The damage is often due to
the fast high-current transient that flows in the charged device ESD event causing an over-
voltage of either device internal part. If, instead of contacting metal, the device contacts
a high-resistance material, the ESD current is greatly reduced and the discharge slowed
by the material resistance. In this circumstance, the device may not be damaged even at
voltages far higher than 250 V.
In practice, it is doubtful whether it is appropriate to measure ESD withstand thresholds
in terms of ESD source voltage as this is only indirectly related to the parameter caus-
ing damage to the device. The damage sustained is often due to energy dissipation in the
device or discharge current, and protection devices divert and must withstand these cur-
rents. The current at which a device fails is often approximately equivalent for HBM and
MM (Amerasekera et al; 2002). Even in CDM the failures are due to high current levels,
which cause internal voltage stresses (Brodbeck and Kagerer 1998).
In a practical situation, the source voltage is only one of several factors affecting the ESD
current and energy in the device. Because of this, it is difficult to directly relate measured
ESD withstand voltages to ESD risk in a factory situation. Nevertheless, at the time of writ-
ing, it seems unlikely that specification of ESD susceptibility in any terms other than source
voltage will become generally accepted in the foreseeable future.
66 3 Electrostatic Discharge–Sensitive (ESDS) Devices

3.2.12 Trends in Component ESD Tests


ESD withstand test methods will need to evolve in response to technological challenges
such as increased component pin counts, reduced pin spacing, and the costs of testing
complex devices (ESD Assoc. 2016). The MM test method is already no longer recom-
mended for device qualification. The detailed HBM and CDM test methods are developing
to increase the efficiency of these tests and reduce test time and stress of the test equipment.
There is continuing pressure to test component pins that could be subject to system-level
ESD using the HMM waveform. This device test method has so far shown poor reproducibil-
ity and is under further development.
TLP testing is under development for evaluation of HBM, CDM, and system-level ESD
susceptibility (ESD Assoc. 2006, 2016).
The standardized ESD models represent only a proportion of real-world ESD sources and
threats. Other recognized threats include charged board events (CBEs) and cable discharge
events (CDEs). CDE can be regarded as a form of system-level ESD, but they often impinge
directly on devices connected directly to the connector terminals (see Section 3.5.3.)

3.3 ESD Susceptibility of Components


3.3.1 Introduction
ESD damage can happen in many ways according to the specific susceptibility of the
many different types of devices and the types and characteristics of ESD sources that they
encounter. The susceptibility of a device to ESD damage, and the damage effects, can be
very variable for different ESD sources and levels. This section attempts to provide an
informative overview to help understanding of this topic. For illustration and understand-
ing, it gives some specific examples representing a small fraction of the wealth of research
that has been done to understand ESD failures.

3.3.2 Latent Failures


McAteer (1990) defined a latent failure as one that occurs in use conditions because of ear-
lier exposure to ESD that did not result in an immediately detectable discrepancy. This topic
is one surrounded by controversy and disagreement. McAteer devoted a chapter to review
of studies on latent damage and concluded the following:
● Latent ESD failures exist.
● The presence of slight parametric shifts or softened V-I characteristics in a device does
not assure further degradation with time.
● System-level latent failures can result from parts that might be regarded as out of specifi-
cation after ESD stress.
ESD-induced oxide breakdown can distort metal-oxide silicon field effect transistor
(MOSFET) V-I characteristics, causing a reduction of transconductance or even loss
of transistor action. Loss of transconductance has been found to be associated with an
increase in supply current that can threaten battery life.
3.3 ESD Susceptibility of Components 67

Reiner (1995) suggested that a high-resistance silicon melt ball could form around the site
of gate oxide damage caused by CDM. This caused unstable leakage currents of the order
1 μA at 5 V. Subsequent heating could convert some of this material into higher conductivity
crystalline silicon and change to permanent damage.
Hellstrom (1986) analyzed ESD failures in bipolar and metal-oxide silicon (MOS) devices.
They observed latent damage through “birth,” growth, and completion to component fail-
ure.
Baumgartner commented that most experts agree that there are latent defects caused by
ESD in some technologies, but they disagree as to whether the device goes on to become a
latent failure. Latent damage is likely to occur in a narrow window of ESD strength between
no damage and complete failure. This window may be narrow, and the probability of ESD
strength lying within the window may be correspondingly small (Beal et al. 1983). Failure
analysis data suggests that latent failures are statistically insignificant. Infant failures of
components may be more likely than latent failures.
Tunnecliffe et al. (1992) subjected 32 enhancement mode and 32 depletion mode field
effect transistors (FETs) to HBM ESD of ±100 and ±200 V. Latent damage was identified, but
they concluded that the risk to reliability was small as the latent failure window is narrow. It
is perhaps more likely that the device would either be catastrophically damaged or remain
undamaged.
In contrast, Gammill and Soden (1986) found that latent failures occurred in comple-
mentary metal-oxide silicon (CMOS) integrated circuits (ICs) both as a field failure and in
a life test experiment. McKeighan et al. (1986) found that reversible charge–induced surface
inversion caused CMOS switches to fail to turn on when addressed. A factor influencing this
behavior included use of a highly insulating ceramic package. The charge-induced failures
occurred during board handling or assembly or were field induced.
Some workers have found that degraded devices could fail parametrically or functionally
when subjected to ESD (Taylor and Woodhouse 1986; Enoch et al. 1983) and recover after
further events, becoming fully functional. Shorting of a p-n junction did not recover on
further ESD event, but a dielectric short due to ESD could recover. Krakauer and Mistry
(1989) found that an ESD event lower than the threshold for catastrophic failure could cause
charge carriers to be injected into oxide to become trapped there. This could cause voltage
threshold shift and changes to MOSFET drain current.
Cook and Daniel (1993) found that ESD applied to ASIC power pins could cause latent
damage that compromised device latch-up immunity.
Anand and Crowe (1999) studied latent failures in Schottky diodes that caused loss of
microwave transceiver sensitivity. The microwave diodes they studied had a junction area
of 5–8 μm diameter. Using a reverse leakage test, they identified damaged devices that had
higher leakage current than good devices. Faulty devices were not identified by the standard
test procedure because the device specification did not include reverse leakage current.
Chen et al. (2009) found that one effect of reverse applied HBM and MM ESD on GaN
light-emitting diode (LED) devices caused an increase in leakage paths that could have a
cumulative and latent damage effect.
Smedes and Li (2003) reported that ESD can cause latent damage as well as permanent
damage to interconnect structures, reducing electrothermomigration lifetimes by a factor
of 100. At low level, ESD metal leads can melt and in cooling change their grain boundary
68 3 Electrostatic Discharge–Sensitive (ESDS) Devices

structure. This can then lead to reduction in electromigration lifetime due to an increase in
the metal resistance.
Laasch et al. (2009) found that ESD protection diodes stressed with multiple pulses suf-
fered cumulative latent damage in the form of a metal alloy front that progressed with each
pulse and eventually could short the p-n junction.
Sylvania (2009) gave examples of partial failure of an LED in a series-connected LED
“coupon” due to damage to one of the LEDs. In the examples, one of the LEDs in the string
is degraded causing reduced light output while remaining sufficiently conductive to allow
the remaining LEDs to function. The operating life of the unit may be dramatically reduced.
Dhakad et al. (2012) described detection of a functional CDM ESD failure in an advanced
CMOS IC where no obvious physical damage was present. The failing device was identified
as a single transistor that had suffered gate oxide charge trapping degradation.
Overall, these studies suggest that while latent damage has been demonstrated, it appears
to be usually a low-probability risk. Nevertheless, it can be a significant concern, espe-
cially in high-reliability applications or where a high cost or severe consequences make
a risk of failure unacceptable. They also suggest that life test studies of advanced technol-
ogy high-speed semiconductor devices stressed with HBM or CDM at levels just below their
threshold values might give some insight into their reliability in the field (Duvvury C. Pri-
vate communication).

3.3.3 Built-in On-chip ESD Protection and ESD Protection Targets


Until the 1970s most components were not susceptible to ESD damage. As component tech-
nology developed, the internal component sizes of ICs reduced. This led to them becoming
more susceptible to ESD damage. In the 1970s this became a known problem after the
introduction of large-scale integration (LSI) technology. The internal device technology
continued to increase in ESD sensitivity with reducing internal component size, and new,
more sensitive device technologies added to the problem.
To counter this trend, the semiconductor component manufacturers started to design in
ESD robustness, at least to the circuits that connect to the device pins. ESD protection net-
works began to be added to divert ESD current flow away from the more sensitive internal
components.
In the 1970s and early 1980s, the automotive industry began to implement ESD pass levels
(Industry Council on ESD Target Levels 2011, 2010b). Various companies adopted different
MM or HBM pass voltage level criteria. In response to these demands from customers, by
the mid-1980s semiconductor companies began to set internal HBM withstand voltage stan-
dard requirements, with 2 kV HBM the most common. This target for MM ESD withstand
of 200 V remained constant for more than 20 years until 2007. Over the same period, a 500 V
target for CDM withstand had developed. In 2007 the Industry Council on ESD target lev-
els published its White Paper I calling for a reduction of ESD targets to 1 kV HBM. The
correlation between HBM and MM withstand predicts that this results in MM withstand
between 30 and 200 V (Smedes 2011). This was followed by a proposal to reduce the CDM
withstand targets to 250 V in White Paper 2. This trend in reducing target ESD withstand
voltages seems set to continue into the foreseeable future.
3.3 ESD Susceptibility of Components 69

Some component pins are particularly difficult to protect with on-chip protection net-
works due to their specialist function. Examples include radio frequency (RF) and high
data rate pins or some types of analog input/output pins. ESD protection networks increase
capacitance and reduce the quality factor of RF circuits. Added capacitance can degrade
the performance of RF circuits. Simple low-capacitance protection networks must be used
in these capacitance-sensitive circuits. Higher-frequency operation generally reduces the
allowed “capacitance budget” for ESD protection. The capacitance of HBM ESD protec-
tion networks typically increases with ESD withstand voltage, so reducing the capacitance
budget can force a reduction in ESD withstand voltage of the protection network.
Not all pins on a device have the same ESD withstand voltage. For example, the Burr
Brown DAC8043 data sheet gives detailed information about the ESD performance of the
device (Burr Brown 1993). This device has mixed digital and analog functions. The digital
pins are stated as withstanding ±2500 V HBM. The analog pins are stated as withstanding
only 1000 V HBM. The data sheet goes on to say that two pins, VREF and RFB , “show some
sensitivity.” What this means is not clear, but presumably they did not survive 1000 V HBM.
On-chip ESD protection typically uses a system of clamp circuits to divert ESD current
away from the sensitive internal circuitry (Figure 3.10). The clamp components can be var-
ious types of components such as p-n or p-i-n diodes, bipolar or MOS transistors, or silicon
controlled rectifiers (SCRs) (Amerasekera et al. 2002).
Operation depends on the design clamp style used. If clamp 2 is a reverse diode, then
only clamp 1 and the power supply ESD clamp turn on for positive stress with respect to
ground. If clamp 2 is a breakdown NMOS device with characteristics shown in Figure 3.11,
then it is clamp 2 that offers protection (Duvvury C. Private communication). For a negative
transient, clamps 2 and 4 and the power supply clamp turn on.
The clamp devices typically have V-I curves that have some, or all, of the elements
shown in Figure 3.11. During an ESD event, the voltage across the device increases until
the device turns on at V t1 and the current through the device increases. If the current
reaches a level I t2 , a second breakdown and thermal failure can occur. So, ESD protection
clamp circuits are also ESD susceptible as there is limit to the current and energy they can
handle. The ESD withstand voltage in an HBM test or failure during a real ESD event is
determined by the maximum current flowing during the test rather than the source ESD
voltage.

+Ve
Output supply
ESD driver ESD
I/O clamp clamp
pin 1 3 Power
supply
Rs Rin ESD
ESD ESD Device clamp
clamp clamp input
2 Output 4 circuit
driver Ground

Figure 3.10 Typical clamp arrangement used in on chip ESD protection of an input/output (I/O)
device pin.
70 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Current Thermal Vt2 Figure 3.11 Elements of the V-I curve


failure of a typical on chip clamp circuit.
It2
Second
breakdown

Holding
voltage
Vsp
Vt1
Turn on
It1

3.3.4 ESD Sensitivity of Typical Components


The only ways to be sure of the ESD withstand voltage of a component are to measure it
or to get a manufacturer’s data on it. Unfortunately, many manufacturers do not give ESD
withstand data on their component data sheets.
Nevertheless, some attempts have been made over the years to compile generic data to
help guide ESD protection program design. Table 3.13 shows generic data based on that

Table 3.13 Range of typical HBM withstand voltages of components.

Typical HBM ESD withstand


Device type voltage range

MR heads, RF FETs, SAW devices <10–100 V


MEMS 40 V–?
MOSFETs, laser diodes, PIN diodes 100–300 V
LEDs 50 to >15 000 V
MMICs 250 to >2 000 V
Pre-1990 VLSI 400–1 000 V
Modern VLSI 1000–3 000 V
Bipolar 600–8 000 V
Linear MOS 800–4 000 V
HCMOS 1 500–3 000 V
CMOS B Series 2 000–5 000 V
Power bipolar 7 000–25 000 V
Film resistor 1 000–5 000 V
Capacitors (low capacitance and Depends on capacitance and
breakdown voltage) breakdown voltage

Source: Based on IEC 61340-5-2/TS: 1999, component data sheets, and other sources.
3.3 ESD Susceptibility of Components 71

given in IEC 61340-5-2:1998 with additions from other sources and data sheets. While this
gives a general idea of the range of HBM ESD withstand voltage that might be expected for
various technologies, it cannot be relied on for specific components.
In the 1980s and 1990s the Reliability Analysis Center published compilations of ESD
susceptibility of a range of components including discrete and passive devices (Reliability
Analysis Centre 1989a,b, 1995).

3.3.5 Discrete Devices


Discrete devices (individual transistors, diodes, or other simple devices) often do not have
any on-chip protection networks. While many of these devices may have relatively high ESD
withstand voltage, some can have low ESD withstand voltage. Voltage-sensitive devices
such as small signal MOSFETs can have low HBM ESD withstand voltage. RF components
are often particularly sensitive as their functionality requires small internal dimensions,
and this often leads to ESD susceptibility.

3.3.6 The Effect of Scaling


As IC technology advances, the internal component dimensions are reduced. MOS technol-
ogy components reduce in channel length, gate oxide thickness, and interconnect dimen-
sions at every technology node (Duvvury and Gauthier 2011). Thin and small dimensions
also increase the resistance of interconnects and other conductors, giving higher voltage
drop for the same current flow. This results in higher ESD sensitivity of the components
due to reduced energy, power, or current levels and oxide breakdown voltage for failure.
This makes it increasingly difficult to meet HBM and CDM target ESD withstand levels.
Clamping ESD protection circuits must maintain the voltage levels under transient ESD
conditions below the breakdown stress of the internal circuitry. This may be less than 5 V
in recent technologies.

3.3.7 Package Effects


Package differences have little effect on HBM withstand but can have considerable effect
on CDM withstand voltage (Duvvury and Gauthier 2011). This is because each package
introduces different package size and capacitance, bond wire inductance, and other factors.
The package capacitance plays an important role in determining the charged device ESD
stress.
Some modern packages such as ball grid arrays (BGAs) are much less vulnerable to
human body contact and ESD than older packages such as dual in line (DIL). Some
types with embedded and close spaced pins are difficult if not impossible to stress
with human body ESD in practice and are in any case normally handled by automated
processes. Charged device ESD is the main concern for these devices.
72 3 Electrostatic Discharge–Sensitive (ESDS) Devices

3.4 Some Common Types of ESD Damage


3.4.1 Failure Mechanisms
ESD-related failure mechanisms depend on the device type and technology. They typically
include (MIL-HDBK-263 sec 50, Analog Devices UG-311 2014, Linear Technology n.d.)

● Thermal secondary breakdown


● Metallization and interconnect or resistor melting
● Breakdown of thin dielectric layers
● Gaseous arc discharge
● Surface breakdown
● Bulk breakdown

Thermal secondary breakdown, metallization and interconnect melting, and bulk break-
down are dependent on energy dissipated in the discharge through the device.
Breakdown of thin dielectric layers, gaseous arc discharge, and surface breakdown are
due to excess voltage stress of some part of the device. In a high-resistance capacitive struc-
ture such as a MOS capacitor or MOSFET gate, conduction of sufficient charge into the
structure can raise the voltage to the breakdown level.
According to Analog Devices (2014), most ESD failures occur due to conductor or resistor
melting, dielectric damage, or junction damage or contact spiking.
Conductor or resistor melting occurs in thin metal or polysilicon interconnects and
thin-film, thick-film, or polysilicon resistors. The high current flowing due to ESD causes
local heating in the conductor or resistor material, leading to fusing and open circuit. This
type of damage is commonly due to HBM ESD because a charged person represents a
high energy ESD source. For thick- and thin-film resistors, partial melting and a change in
resistance value are possible. This can lead to parametric failure of the IC. These damage
mechanisms are largely found by HBM ESD testing rather than any vast evidence gathered
from field failures. In contrast, for CDM ESD the field failures correlate well with observed
damage for sensitive devices (Duvvury C. Private communication).
Dielectric breakdown occurs when a thin insulating dielectric layer such as silicon diox-
ide or nitride is stressed by an applied voltage that exceeds its time dependent dielectric
breakdown strength. This results in punch through of the dielectric. results. This type of
failure is typical of CDM damage because the fast rise time results in high voltages occur-
ring within the chip. The high current that flows on breakdown can lead to formation of a
melt filament of silicon.
Junction damage and contact spiking can occur when a p-n junction is subjected to
avalanche breakdown and secondary breakdown. Avalanche breakdown occurs when
the reverse breakdown voltage of a reverse biased p-n junction is exceeded. Secondary
breakdown can then occur at a point where the junction material is sufficiently hot. The
high current that flows is channeled through this site and causes high adiabatic local
heating, increasing current flow. Ultimately, the silicon can melt if the temperature exceeds
1415 ∘ C. If heating is sufficient, adjacent contact metal can melt and migrate. The junction
is shorted by resistive material.
3.4 Some Common Types of ESD Damage 73

Melted junction material has, on solidification, a modified dopant profile, crystal


structure, and electrical characteristics. This typically gives a soft reverse breakdown
characteristic. This can result in a small or large leakage current increase and change
in device parameters. Melting of contact material typically leads to failure of the
corresponding IC pin.

3.4.2 Breakdown of Thin Dielectric Layers


MOS technology is widely used to fabricate discrete MOSFETs, ICs, MOS capacitors, and
devices with metallization crossovers. Typical structures have a thin layer of insulating
oxide separating two conductive layers. The resistance across the layer is normally very
high and the leakage current between the conductors very low.
The breakdown voltage reduces with the thickness of the oxide layer. The voltage differ-
ence required to break down thin oxide layers used in modern components can be quite
small. Dielectric breakdown can occur when the voltage across the dielectric exceeds a
time-dependent dielectric breakdown value (Analog Devices 2014). The result can be a
gate-source or gate-drain short in a MOSFET or a resistive leakage path between a met-
allization track and underlying semiconductor regions. According to Analog Devices, this
is the main result of charged device ESD failures as the fast rise time is most likely to result
in excessive internal voltages. Breakdown tends to occur at high field points such as corners,
edges, or steps in the dielectric.
If breakdown occurs, a puncture defect is made in the insulating layer. If sufficient voltage
is applied across the dielectric, the breakdown field strength is exceeded, and breakdown
occurs. This type of damage is voltage dependent, requiring sufficient voltage to be applied
across the oxide layer. After breakdown, high current flows through the breakdown point
resulting in adiabatic heating of the breakdown region. If the energy in the discharge is
sufficient, a small amount of semiconductor can be melted.
The results of the breakdown may, however, depend on the energy available in the dis-
charge. Melted material can fill the puncture, short-circuiting the two conductors. This will
normally cause a detectable component failure due to a short circuit or high leakage cur-
rent. A lower energy discharge may, however, simply leave a clear puncture that may leave
the component functioning but weakened. The same point may fail in future at a lower
voltage, or an increased leakage current may result.
Dielectric breakdown damage has also been reported in microelectromechanical system
(MEMS) devices (Sangameswaran et al. 2009).

3.4.3 MOSFETs
According to Infineon (2013), there are three basic ESD failure modes in MOSFETs. These
are junction damage, gate oxide damage, and metallization burnout. The most common
failure in HBM is junction damage caused by an ESD transient of sufficient ESD energy
and duration. A damaged junction often shows high reverse bias leakage or a short circuit.
Gate oxide damage is, however, the main category of ESD damage. This occurs when
the gate is subjected to ESD causing the gate oxide to break down. This happens when
the voltage across the thin gate oxide exceeds the breakdown field strength. For thin oxide
74 3 Electrostatic Discharge–Sensitive (ESDS) Devices

layers, the breakdown voltage can be small, although for short duration ESD, a much higher
ESD voltage may be required to achieve breakdown.
Another common failure mode is that ESD may lead to trapped charge in the gate
oxide. This can cause a shift in the gate threshold voltage and functional failure. This can
sometimes anneal itself after a few hours, and so trapped charge effects are not always
permanent (Duvvury C. Private communication).
Metallization burnout can also occur, although this is often a secondary effect occurring
after initial junction or gate oxide breakdown.
A MOS gate is a capacitive structure. If the gate is charged to over the breakdown voltage
of the dielectric, breakdown occurs, and the gate is likely to be damaged. An HBM or MM
test of this type of structure simply takes the charge in one capacitor (the ESD source) and
dumps it into another (the MOS gate) (Figure 3.12). The final voltage on the gate can be
calculated if the gate capacitance is known. Hence, the MM or HBM ESD withstand can be
estimated if the MOSFET gate capacitance and breakdown voltage is known (International
Rectifier n.d, Application note AN-986).
The initial charge QESD in the ESD source capacitance CESD at a voltage V ESD is
QESD = CESD VESD
During the ESD event, this charge is shared between the source capacitance and gate
capacitance Cg in parallel. The charge Qbr required to achieve the gate breakdown
voltage V gbr is
Qbr = (CESD + Cg )Vgbr
The threshold ESD voltage at which the gate reaches this breakdown voltage can be found
when QESD equals Qbr .
CESD VESD = (CESD + Cg )Vgbr

(CESD + Cg )
VESD = Vgbr
CESD
So, the effective ESD withstand voltage is heavily dependent on the gate capacitance. If
Cg ≪ CESD , then V ESD is around the gate breakdown voltage V gbr . For small devices, this
can be a few volts or tens of volts. If Cg ≫ CESD , then V ESD is around V gbr Cg /CESD .

3.4.4 Susceptibility to Electrostatic Fields and Breakdown Between Closely


Spaced Conductors
There are relatively few components that are susceptible directly to damage from electro-
static fields. It is very common that electrostatic fields induce voltage differences between a

Figure 3.12 Charging of a MOSFET gate from


Rs an HBM ESD test source.
ESD source MOSFET
capacitance CESD under test
Vgbr
VESD
Cg
3.4 Some Common Types of ESD Damage 75

device and another object, and electrostatic field–induced charged device ESD occurs when
they touch. Control of electrostatic fields is normally essential where charged device ESD
is a concern.
Some types of components have closely spaced unpassivated conductors on a surface or
with an air gap. Components containing isolated conductors at small distances apart will
have voltage differences induced by changing electrostatic fields. These voltage differences
could produce breakdown and discharges between them. Sparking can cause damage to the
electrodes and transfer of material across the gap. Electromigration of materials under high
electrostatic fields between conductors is also possible.
Photomask reticles used in semiconductor device manufacture have this type of structure.
The mask consists of very fine metallic tracks with very fine (μm or nm) etched gaps on the
surface of a quartz substrate. Electrostatic fields, for example due to charge on the reticle
enclosure, cause voltage differences between the metallic tracks. These voltage differences
between tracks can lead to breakdown, ESDs, and damage to the reticle at voltages below
the Paschen minimum breakdown voltage (Rider and Kalkur 2008; Rider 2016). Damage to
reticles leads to damaged components in the semiconductor wafer and reduction of compo-
nent yield. Englisch et al. (1999) proposed a “canary” reticle test structure that could be used
to simulate and evaluate the effects of electrostatic field exposure during reticle handling.
Other devices such as surface acoustic wave (SAW) filters that have closely spaced elec-
trodes can also suffer damage (Mil HDBK 263B). Wallash and Honda (1997) reported that
magnetoresistive (MR) recording heads could be damaged by electrostatic fields, due to
breakdown of small gaps by voltage differences induced by the field.
Sangameswaran et al. (2009) found that MEMS capacitive switches could suffer both
dielectric and air breakdown.
Wallash and Levitt (2003) found that ESD could still occur at very small submicron gaps
less than the Paschen minimum by field emission. This could be a threat for photolitho-
graphic reticles, magnetic recording heads, MEMSs, and field emission displays.

3.4.5 Semiconductor Junctions


Semiconductor junctions are used to make bipolar transistors, diodes, junction FETs, PIN
diodes, Schottky diodes, and thyristors. P-n junctions also occur as parasitic components in
MOS technology and can be used in ESD protection networks.
An ESD current passing through a junction can cause intense local heating and even
melting of the semiconductor material. Electromigration of the materials can also occur
under the intense electrostatic fields created within the component (Analog devices 2014).
For ESD current passing through a reverse biased p-n junction, most of the power dissi-
pation occurs in the junction. The ESD event is usually of very short duration compared to
thermal time constants in device materials. The dissipation of energy in the device can usu-
ally be considered adiabatic with negligible diffusion of heat away from the heated parts. At
breakdown, the current results in a hot spot at the junction. The resistance of the semicon-
ductor material in this region reduces with increasing temperature, and this leads to more
current flowing through the spot. Current flow increasingly focuses on the hot spot, and
further heating occurs. Local temperatures in hot spots can approach or exceed the melting
76 3 Electrostatic Discharge–Sensitive (ESDS) Devices

point of the semiconductor. This can result in changes in junction characteristics or short
circuits. This is called thermal secondary breakdown.
This failure mechanism is dependent on the power dissipated in the junction. Junctions
with high breakdown voltage and low leakage current may be more susceptible to ESD
damage. Where hot spots do not develop to failure, migration of material under the elec-
tric field may cause a partially developed filament short circuit. This may increase leakage
current.
When forward biased, the junction has low voltage drop, and the power developed by
the ESD current is spread through the body of the device, and failure of the junction is less
likely.
For most junction transistors, the emitter-base junction is more susceptible than the
collector-base junction due to smaller junction dimensions. Junction FETs that have a high
gate-source breakdown voltage and low leakage can be particularly susceptible. Schottky
diodes and Schottky TTL components are more sensitive to ESD because they have thin
junctions and metal present that may be carried through the junction.
Not all p-n junctions are susceptible to ESD damage. Transient suppressor diodes, zener
diodes, power rectifier diodes, power bipolar transistors, and thyristors can be very robust
to ESD.

3.4.6 Field Effect Structures and Nonconductive Device Lids


Some types of LSI and memory device (e.g. UVEPROMS) have quartz or ceramic packages
that are highly insulating and can become highly charged (Mil HDBK 263B). Surface inver-
sion or gate threshold changes can occur due to ions deposited by ESD. This can cause
malfunction of the device that may be reversible in some cases by neutralization of the
charge on the external device surfaces.

3.4.7 Piezoelectric Crystals


Piezoelectric crystals are used in oscillator crystals, delay lines, and SAW filters. The high
voltages applied due to ESD can cause high mechanical forces that can damage or fracture
the crystal (Mil HDBK 263B).

3.4.8 LEDs and Laser Diodes


LEDs and laser diodes can be extremely sensitive to ESD damage. LEDs can suffer catas-
trophic damage so that they do not light or conduct (Sylvania 2009; Nichia 2014; Osram
n.d.). They can also suffer latent damage and degradation of reliability, remaining partly
functional. In a LED string, one LED may fail allowing the remaining LEDs to function
correctly. Nichia (2014) advised customers to test for ESD damage before assembly and
give a method of detecting damage by measurement of forward voltage when a low current
(<0.5 mA) is passed through the LED in a forward direction.
Indium gallium nitride (InGaN) blue and green LEDs have been found to be extremely
sensitive to ESD damage (Avago Technologies 2007). Damaged devices can appear dim,
3.4 Some Common Types of ESD Damage 77

dead, short, or with low forward or reverse voltage. Avago Technologies’ InGaN LEDs are
classified as Class 1x and Class 2 (HBM ESD withstand voltage 250–4000 V).
Talbot (1986) found that the ESD withstand voltage of nine various color LEDs mostly var-
ied from 4–15 kV, although two “low current” types showed damage at 100–200 V. Devices
that had been subjected to reverse breakdown often functioned normally in the forward
direction. There was a drop in light output dependent on the ESD voltage. Failure mecha-
nisms included junction burnout, nitride punch through, and metallization burnout.
Chen et al. (2009) found four different effects of reverse applied HBM and MM ESD
on GaN LED devices, affecting leakage paths in the devices. Low-level ESD below 650 V
reduced leakage but above 700 V leakage was increased by three to five orders of magnitude.
Further application of ESD caused unstable behavior. The device still glowed in forward
bias. When V-I measurements were made, the device was destroyed. They concluded that
the increase in leakage paths could have a cumulative and latent damage effect.

3.4.9 Magnetoresistive Heads


MR heads are used in hard disk drives for reading data from the disk. They are among
the most sensitive components currently in use, with ESD withstand voltages reported to
be less than 5 V HBM. ESD damage to MR heads, and ESD control during manufacturing
processes of these devices, has been the subject of intensive research over many years. There
are many papers published on these subjects, which form a perennial topic in conferences
such as the ESD Association Annual Symposium.
MR heads have susceptibility to both voltage and energy. ESD currents in the milliamp
range can change or destroy the MR sensor. Thin insulating layers in the device may be
broken down by low voltages (Wallash 1996).

3.4.10 MEMS
MEMS devices are used in modern electronics systems. Some unprotected MEMS have been
found to been highly sensitive to ESD damage (Walraven et al. 2000, 2001). Sangameswaran
et al. (2008, 2009, 2010a,b) found HBM withstand voltages as low as 40 V. Traditional elec-
trical characterization was found to overestimate ESD robustness. Capacitive switch MEMs
devices were found to suffer dielectric or air breakdown and mechanical failures in response
to ESD. ESD failures were often detectable only through mechanical tests. Atmospheric gas,
pressure, temperature, and humidity were all influential to ESD-induced breakdown.

3.4.11 Burnout of Device Conductors or Resistors


Most devices contain tracks and interconnects of metallization, polysilicon, or other
conductive material to connect internal parts and components with polysilicon, thick- or
thin-film resistors. High ESD currents can lead to intense local heating that can burn out
a conductor, rather like a fuse (Analog Devices 2014). This may occur where a wire or
interconnect dimension is reduced in cross section for some reason. Thick- or thin-film
resistors can suffer partial melt resulting in a change in resistance and parametric failure.
78 3 Electrostatic Discharge–Sensitive (ESDS) Devices

3.4.12 Passive Components


ESD damage to passive components is less well documented but is certainly possible with
some types of components. The effect of the ESD can be highly variable depending on the
resistor type. Film resistors on an insulating substrate can be susceptible to ESD damage
(MIL HDBK 263). Some types of resistor have been trimmed to the desired value using
ESD (Vishay 2011). Szwarc (2008) commented that the sensitivity of resistors varies from
hundreds of volts to a few tens of kilovolts.
Thick film resistance changes have been reported to be dependent on voltage rather than
energy. In contrast, thin-film resistors can be susceptible to the energy in the discharge,
showing only small change until an energy threshold is reached. According to MIL HDBK
263, carbon film, metal oxide, and metal film resistors at low tolerance and power ratings
can be susceptible to ESD. With a 0.05 W 0.1% part, placing the resistors in a polythene bag
and rubbing this with another polythene bags was sufficient to change the tolerance of the
resistors.
Chase (1982) found that tantalum nitride thin-film resistors used in filters in hybrid cir-
cuits could be damaged by ESD stress voltages as low as 1 kV HBM and tantalum capacitors
as low as 400 V HBM. Resistors were damaged nearly linearly in response to multiple dis-
charges. For CDM, resistors and capacitors were damaged above 2000 V. Damage to the
resistors depended on the resistor design.
As time goes on, components used in surface mount assembly PCBs have become increas-
ingly small. Taminnen et al. (2014) have found that very small 01005 resistors and capacitors
showed some sensitivity to ESD.

3.4.13 Printed Circuit Boards and Assemblies


PCBs and assemblies containing ESDS are susceptible to ESD damage due to the sensitive
devices they contain. According to the MIL-HDBK-263B, assemblies and modules contain-
ing ESDS are often as sensitive as the most sensitive component they contain. One of the
longstanding myths of ESD is that components, once assembled into PCBs, are no longer
ESD susceptible (Danglemeyer 1999). In practice, the risk of ESD damage is hard to predict,
but it, and the sensitivity of the component to ESD, may be reduced or increased (Boxleitner
1990). The ESD threat, characterized in terms of voltage, peak power, or energy dissipated
in a component may vary unpredictably over two orders of magnitude depending on the
ESD source, the discharge point, and the design of the PCB. The threat to ICs mounted on
PCBs can be significantly greater than to a device before it is mounted on a PCB.
A modern PCB may contain many ESDS components that have varying levels of suscep-
tibility. The PCB typically has many conductive tracks interconnecting the ESDS compo-
nents, all of which have inductance and capacitance. Injecting ESD current into one point
of the board can cause a rapid transient change in voltage of that point of hundreds of volts
relative to other tracks on the PCB. Transient currents flow through the track and through
any components connected to it. The resulting ESD stress to components on the PCB is dif-
ficult to predict. Boxleitner found no correlation between device ESD withstand voltages
and the immunity of the device mounted on a PCB.
Boxleitner did find a correlation between the ESD source and the level of threat. ESD from
the combined charged PCB and person holding the PCB created the greatest peak power
3.4 Some Common Types of ESD Damage 79

and energy levels in the discharge. Discharge to connector or device pins gave the greatest
threat, with the least threat for ESD to long PCB traces between devices. The threat was
worse for PCBs with no ground plane and less for PCBs with ground and power planes.
Floating PCBs represented a lesser threat than those connected to an external ground.
Often modules and assemblies are designed with protection network circuits on the con-
nector pins emerging to the outside world. These by design will give a level of protection
against ESD entering the assembly by this route. They will not, however, protect against
ESD arising from direct contact with other parts of the assembly.
Shaw and Enoch (1985) found that type 74 373 octal latch ICs failed in the voltage range
250–2500 V due to charged PCB ESD transients, compared to 1000 to >4000 V for HBM and
CDM testing. They found the charged PCB damage voltages were not related to CDM or
HBM failure levels but were dependent on the capacitance of the PCB.
Olney et al. (2003) found that ICs that are relatively robust as components could be dam-
aged by charged board ESD (see Figure 2.19). As PCB capacitance is higher than device
capacitance, the energy stored for a given voltage is much greater. They commented that
charged board ESD damage could be mistaken for EOS damage and that this should be
considered before a conclusion of EOS is drawn in failure analysis. They gave guidelines on
how to avoid charged board ESD failures.
Paasi et al. (2003) studied the behavior of charged PCBs to evaluate ESD sensitivity of
devices on the board with respect to discharge current and energy. They concluded that
energy-sensitive devices could be at risk at lower charge or voltage levels when on a PCB
than before assembly. They based their evaluation on HBM and MM device data. They
pointed out that the capacitance, voltage, and stored energy of a PCB vary with movement
thorough the manufacturing line. For a given PCB charge, a low capacitance high voltage
condition gives higher stored energy and ESD current than a high capacitance low voltage
condition and can represent a higher risk of ESD damage.
Gärtner et al. (2014) concluded that a CBE was more likely in a PCB production line than
a charged device ESD event. As peak discharge current is a critical parameter in CDM dam-
age, they used it to evaluate CBEs. They showed that the peak current is not significantly
higher than for a single device at the same voltage level. The total charge transferred is,
however, significantly higher giving a comparatively long (10–50 ns) discharge. This seemed
to represent a higher stress to devices than charged device ESD. Case studies have shown
that devices can be damaged at a voltage level at which devices would have survived in the
CDM test. Stresses encountered in the real world are typically lowered due to reduced PCB
capacitance due to greater gap between the PCB and ground. Stresses are also reduced for
field-induced stresses for a parallel external field compared to a perpendicular field.

3.4.14 Modules and System Components


Modules and system components can become damaged by ESD if this can occur to a part
that is inadequately protected. Components and modules that are designed to operate
within a larger system are often (and not always correctly) not considered ESD susceptible.
They are often, however, designed with little or no ESD protection on connector pins
because these are not intended to be exposed once assembled into the system. Before or
during assembly into the system, they may be subject to ESD to the connector pins or on
80 3 Electrostatic Discharge–Sensitive (ESDS) Devices

connection to cables. They may be assembled into the system in a facility that has little or
no ESD control.
Many system components or modules are housed within polymer enclosures or potted
with only connectors or flying leads exposed to the outside world. While these enclosures
prevent direct contact and ESD to the internal circuitry, the enclosure may become itself
charged during handling and transport, especially if packaged within plastic packaging.
The internal circuits can attain high voltages by induction especially when packaging is
removed. ESD can then occur when connection is made to the connectors or flying leads.

3.5 System-Level ESD


3.5.1 Introduction
System-level ESD issues are largely addressed as part of the topic of EMC in operating
electronic systems. In this context, it is the impact of ESD on powered and operational
equipment that is usually of concern. This subject is covered by Williams (2001) and Mon-
trose (2000).
Typical operating environments for electronic equipment are uncontrolled areas from a
static electricity view. Personnel in these areas can charge to over 10 kV through normal
activities such as walking or rising from a chair (Wilson 1972; Brundrett 1976; Smallwood
2004; Talebzadeh et al. 2015). In some environments, other types of source are possible such
as charged metal stretchers, beds, or other mobile equipment in a health care environment
(Viheriäkoski et al. 2014) or charged cables.
Electrostatic shocks can be felt by personnel if their body voltage exceeds about 2000 V
(2 kV) and they discharge to a conductive object such as a metal item or equipment or
another person. The sensitivity of the body varies from person to person and over the parts
of the body.
In many regions of the world and industries, electronic equipment must be shown to
be immune to ESD of this type expected in their operating environment. This has led to
the development of ESD tests such as the IEC 61000-4-2. Electronic equipment marketed
within Europe must pass ESD immunity tests measured using the IEC 61000-4-2 stan-
dard ESD source. Equipment is tested at 2, 4, 6, 8, or 15 kV ESD stress levels according
to the requirements of relevant equipment and market-related standards. It is assumed
that the equipment will withstand in service ESD from personnel charged up to specified
levels without serious malfunction. Allowed loss of function is defined as part of the tests
(Williams 2001).

3.5.2 The Relationship Between System Level Immunity and Component ESD
Withstand
There is, however, some overlap between the topics of system-level ESD immunity and com-
ponent ESD susceptibility. It has often been assumed, mainly erroneously, that system-level
immunity depends on component ESD withstand (Ind. Co. 2010a, 2012). This has led to
system designers placing ESD withstand requirements on components used in the system.
3.5 System-Level ESD 81

System ESD failures can be classified as hard or soft. Hard failures are those in which
physical damage has occurred that is not recoverable. Often, system-level failures are soft
failures that represent upset or temporary malfunction of the system in a way that is recov-
erable.
Ind. Co. (2010a) found that there is rarely correlation between system-level ESD immu-
nity and component HBM ESD withstand level. Component ESD withstand data is obtained
with the component in an unpowered state and represents hard failures of the component.
System failures are often soft failures and occur with the system in a powered state. The
ESD waveforms used in obtaining component ESD withstand data have significant differ-
ences with those used in system ESD immunity evaluation, and the test environment is very
different. In practice, system ESD immunity is dependent on the system design including
PCB design and on-PCB ESD protection, as well as individual component response to ESD
transients. Component ESD tests do not reflect the conditions that occur for devices during
system-level ESD events.
There may be some components, e.g. those that are directly connected to external
connector pins, that do require some ESD withstand capability. The Industry Council on
ESD Target Levels discussed this area in its White Paper 3 and proposed a system-efficient
electrostatic discharge design (SEED) approach to understanding system-level ESD needs
regarding ESD robust components.

3.5.3 Charged Cable ESD (Cable Discharge Events)


Charged cable ESD, often known as CDEs, can be considered to be a system-level ESD event,
although they often impinge directly on devices connected to connectors. These can occur
when a charged cable is plugged into an electronic system connector. Stadler et al. (2006)
found that these often give rectangular current pulses resembling TLPs with current levels
of several amps.
Stadler et al. (2017) used SPICE simulation and measurements to investigate risks due to
ESD from cables plugged into charged USB3 ports. The investigated risks included charging
of the cable shield, charging of an internal conductor while the shield was floating, and
a charged person touching the shield. Typical discharges from a 3 m cable were 20 ns in
duration and resembled discharges from coaxial cables. A discharge from a charged data
line resulted in a 2.5 A peak current. They found that not all USB cables have the shield
connected to ground or the shell of the connector. The worst case occurred when the shield
was already grounded but the data line became charged. With a charging voltage of 1000 V
the peak discharge current exceeded 13 A. This risk, however, was thought to be significant
only for some applications that were not USB compliant and had exposed data lines. They
concluded that a stress of a few amps for about 20 ns duration was sufficient for testing USB
protection.

3.5.4 System-Efficient ESD Design (SEED)


The Industry Council on ESD Target Levels has proposed a SEED approach to system design
for ESD immunity. This approach recognizes that high-component ESD withstand voltages
are not generally required for effective system ESD immunity design. Robust system ESD
82 3 Electrostatic Discharge–Sensitive (ESDS) Devices

design can be achieved by providing the interactions between the ESD stress and the com-
plete system design are understood and addressed. The SEED approach recognizes that

● System ESD immunity specification requirements are understood to be a separate issue


to component ESD withstand.
● The system ESD failure mechanisms must be understood to allow effective design of sys-
tem immunity.
● System design and component design share responsibility for system ESD protection.
● Design strategy should differentiate between system external and internal component
pins and the stresses that result to them in ESD tests.
● Placing robust ESD protection on-chip for component pins directly connected to exter-
nal connector pins may not ensure a robust system. A better design strategy may be to
use external ESD protection clamps and understand their interaction with component
internal ESD protection.
● TLP can be used to characterize on-board and on-chip ESD protection in co-design of the
system.

The SEED approach was proposed as a better system design philosophy, optimizing the
balance between system cost and performance and reducing design effort. The philosophy
and approach to implementation of ESD robust system designs and state of the art was
further described in Part 2 of White Paper 3 (Ind. Co. 2012). A more comprehensive dealing
of SEED is presented by Duvvury and Gossner (2015).

References

Amerasekera, A. and Duvvury, C. (1995). ESD in Silicon Integrated Circuits, 1e. Wiley. ISBN:
0471954810.
Amerasekera, A., Duvvury, C., Anderson, W. et al. (2002). ESD in Silicon Integrated Circuits, 2e.
Wiley. ISBN: 0 470 9871 8.
Analog Devices (2014). Reliability Handbook. UG-311 Rev. D. Available from: http://www
.analog.com/media/en/technical-documentation/user-guides/UG-311.pdf [Accessed: 10th
May 2017]
Anand, Y. and Crowe, D. (1999). Latent failures in Shottky barrier diodes. In: Proc. of EOS/ESD
Symp. EOS-21, 160–167. Rome, NY: EOS/ESD Association Inc.
Ashton, R. (2008). Reliability of IEC 61000-4-2 ESD testing on components. E E Times Available
from: http://www.eetimes.com/document.asp?doc_id=1273265 [Accessed: 10th May 2017]
Avago Technologies. (2017) Premium InGaN LEDs - Safety Handling Fundamentals ESD
Electrostatic Discharge Application Note 1142. Available from: http://www.avagotech.com/
docs/AV02-0160EN [Accessed: 10th May 2017]
Baumgartner, B. (n.d.) ESD TR50.0-03-03. Voltage and Energy Susceptible Device Concepts,
Including Latency Considerations. Rome, NY, EOS/ESD Association Inc.
Beal, J. Bowers, J. Rosse, M. (1983) A study of ESD latent defects in semiconductors. In: Proc.
EOS/ESD Symp. EOS-5. Rome, NY, EOS/ESD Association Inc.
Boxleitner, W. (1990). ESD stress on PCB mounted ICs caused by charged boards and
personnel. In: Proc. EOS/ESD Symp. EOS-12, 54–60. Rome, NY: EOS/ESD Association Inc.
References 83

Brodbeck, T. and Kagerer, A. (1998) Paper 4A.7). Influence of the device package on the results
of CDM tests – consequences for tester characterization and test procedure. In: Proc.
EOS/ESD Symp, 320–327. Rome, NY: EOS/ESD Association Inc.
Brundrett, G.W. (1976). A review of the factors influencing electrostatic shocks in offices. J.
Electrostat. 2: 295–315.
Burr Brown. (1993) DAC8043 CMOS 12-Bit serial input multiplying digital to analog converter.
www.ti.com.cn/cn/lit/ds/symlink/dac8043.pdf [Accessed: 10th May 2017]
Chase, E.W. (1982). Electrostatic discharge (ESD) damage susceptibility of thin film resistors
and capacitors. In: Proc. EOS/ESD Symp. EOS-4, 13–18. Rome, NY: EOS/ESD Association
Inc.
Chen, N.C., Wang, Y.N., Wang, Y.S. et al. (2009). Damage of light-emitting diodes induced by
high reverse-bias stress 97-B-016. J. Cryst. Growth 311: 994–997.
Cook, C. and Daniel, S. (1993). Characterisation of new failure mechanisms arising from power
pin stressing. In: Proc. EOS/ESD Symp. EOS-15, 149.
Danglemeyer, T. (1999). ESD Program Management, 2e. Clewer: Springer. ISBN: 0-412-13671-6.
Department of Defense. 1994 Military Handbook. Electrostatic discharge control handbook for
protection of electrical and electronic parts, assemblies and equipment (excluding
electrically initiated explosive devices) MIL-HDBK-263B 31st July 1994
Dhakad, H., Gossner, H., Zekert, S., Stein, B., Russ, C. (2012). Paper 3A.1. Chasing a latent
CDM ESD failure by unconventional FA methodology. Proc. EOS/ESD Symp.
Duvvury C, Gauthier R. (2011) IC Technology Scaling Effects on Component Level ESD. Ch. 6
in Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering
component level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from:
www.esdindustrycouncil.org. [Accessed: 10th May 2017]
Duvvury, C. and Gossner, H. (2015). System Level ESD Co-Design. Wiley – IEEE. ISBN:
978-1-118-86190-5.
Duvvury C., Ashton R., Righter A., Eppes D., Gossner H., Welsher T. and Tanaka M, (2012)
Discontinuing Use of the Machine Model for Device ESD Qualification. In Compliance
Magazine, July 2012. Available from: https://incompliancemag.com/article/discontinuing-
use-of-the-machine-model-for-device-esd-qualification [Accessed 6th March 2019]
Englisch, A., van Hesselt, K., Tissier, M., and Wang, K.C. (1999). CANARY: a high-sensitive
ESD test reticle design to evaluate potential risks in wafer fabs. In: Proceedings of the SPIE,
19th Annual Symposium on Photomask Technology, BACUS, vol. II, 886–892. Available from:
http://dx.doi.org/10.1117/12.373381 [Accessed: 10th May 2017].
Enoch, R.D., Shaw, R.N., and Taylor, R.G. (1983). ESD sensitivity of NMOS LSI circuits and
their failure characteristics. In: Proc. EOS/ESD Symp. EOS-5, 185–197. Rome, NY: EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2006) Trends in Semiconductor Technology and ESD Testing.
White paper II. ISBN: 1-58537-116-5
EOS/ESD Association Inc. (2009). ESD S5.6-2009. ESD Association Standard Practice for
Electrostatic Discharge Sensitivity Testing – Human Metal Model (HMM) – Component Level.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016) ESD Association Electrostatic Discharge (ESD) Technology
roadmap – revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf [Accessed: 10th May 2017]
84 3 Electrostatic Discharge–Sensitive (ESDS) Devices

EOS/ESD Association Inc., JEDEC. (2012). ANSI/ESD STM5.2-2012. ESD Association Standard
Test Method for Electrostatic Discharge (ESD) Sensitivity Testing – Machine Model
(MM) – Component Level. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc., JEDEC. (2017). ANSI/ESDA/JEDEC JS-001-2017. ESDA/JEDEC
Joint Standard for Electrostatic Discharge Sensitivity Testing – Human Body Model
(HBM) – Component Level. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc., JEDEC. (n.d.) ANSI/ESDA/JEDEC JS-002-2014. ESDA/JEDEC joint
standard for electrostatic discharge sensitivity testing – Charged Device Model (CDM) – Device
Level. ISBN: 1-58537-276-5 Rome, NY, EOS/ESD Association Inc.
Gammill, P.E. and Soden, J.M. (1986). Latent failures due to electrostatic discharge in CMOS
integrated circuits. In: Proc. EOS/ESD Symp. EOS 8, 75–80. Rome, NY: EOS/ESD Association
Inc.
Gärtner, R., Stadler, W., Niemesheim, J., Hilbricht, O. (2014) Do Devices on PCBs Really See a
Higher CDM-like ESD Risk? In: Proc. EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Gieser, H. (2002). Test Methods. In: ESD in Silicon Integrated Circuits, 2e (eds. A. Amerasekera
and C. Duvvury). Wiley. ISBN: 0 470 9871 8.
Gieser, H. and Ruge, I. (1994). Survey on electrostatic susceptibility of integrated circuits. In:
Proc. ESREF Symp, 447–455. Rome, NY: EOS/ESD Association Inc.
Hellstrom, S., Welander, A., and Eklof, P. (1986). Studies and revelation of latent ESD failures.
In: Proc. EOS/ESD Symp. EOS-8, 81–91. Rome, NY: EOS/ESD Association Inc.
Industry Council on ESD Target Levels (2010a) White paper 3: System Level ESD Part I:
Common Misconceptions and Recommended Basic Approaches. Rev. 1.0 http://www
.esdindustrycouncil.org/ic/en/documents/7-white-paper-3-system-level-esd-part-i-
common-misconceptions-and-recommended-basic-approaches [Accessed: 10th May 2017]
Industry Council on ESD Target Levels (2010b) White paper 2: A case for lowering component
level CDM ESD specifications and requirements. Rev. 2.0. http://www.esdindustrycouncil
.org/ic/en/documents/6-white-paper-2-a-case-for-lowering-component-level-cdm-esd-
specifications-and-requirements [Accessed: 10th May 2017]
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering component
level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf [Accessed: 10th May
2017]
Industry Council on ESD Target Levels (2012) White paper 3: System Level ESD Part II:
Implementation of Effective ESD Robust Designs. Rev. 1.0 http://www.esdindustrycouncil
.org/ic/en/documents/36-white-paper-3-system-level-esd-part-ii-effective-esd-robust-
designs [Accessed: 10th May 2017]
Infineon. (2013). Preventing ESD Induced Failures in Small Signal MOSFETs. Application Note
AN-2013-04 V2.0 http://www.infineon.com/dgdl/Infineon+-+Application+Note+-+
PowerMOSFETs+-+Small+Signal+-+Preventing+ESD+Induced+Failures+in+Small+
Signal+MOSFETs.pdf?fileId=db3a30433dfcb54c013dfe36f38d0295 [Accessed: 10th May
2017]
International Electrotechnical Commission. (1999) IEC 61340-5-2/TS:1999. Electrostatics – Part
5-2: Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
References 85

International Electrotechnical Commission (2008). IEC 61000-4-2. Electromagnetic


compatibility (EMC) - Part 4-2: Testing and measurement techniques - Electrostatic discharge
immunity test. Ed. 2. Geneva, IEC.
International Electrotechnical Commission (2012). IEC 60749-27. Semiconductor
devices – Mechanical and climatic test methods – Part 27: Electrostatic discharge (ESD)
sensitivity testing – Machine body model (MM). ed. 2.1, ISBN 978-2-8322-0407-8 Geneva, IEC.
International Electrotechnical Commission (2013). IEC 60749-26. Semiconductor
devices – Mechanical and climatic test methods – Part 26: Electrostatic discharge (ESD)
sensitivity testing – Human body model (HBM) Ed. 3 ISBN 978-2-83220-746-8 Geneva, IEC.
International Electrotechnical Commission (2017). IEC 60749-28. Semiconductor
devices – Mechanical and climatic test methods – Part 28: Electrostatic discharge (ESD)
sensitivity testing – Charged device model (CDM). Ed. 1. ISBN 978-2-8322-4139-4 Geneva, IEC.
International Rectifier. (n.d.). ESD Testing of MOS Gated Power Transistors. AN-986. http://
www.infineon.com/dgdl/an-986.pdf?fileId=5546d462533600a40153559f9f3a1243 [Accessed:
10th May 2017]
Kelly, M., Servais, G., Diep, T. et al. (1995). A comparison of electrostatic discharge models and
failure signatures for CMOS integrated circuit devices. In: Proc. EOS/ESD Symp, 175–185.
Rome, NY: EOS/ESD Association Inc.
Krakauer, D.B. and Mistry, K.R. (1989). On latency and the physical mechanisms underlying
gate oxide damage during ESD events in n-channel MOSFETs. In: Proc. EOS/ESD Symp.
EOS-11, 121–126. Rome, NY: EOS/ESD Association Inc.
Laasch, I., Ritter, H.M., and Werner, A. (2009). Latent damage due to multiple ESD discharges.
In: Proc. EOS/ESD Symp. EOS-31, 4A.4-1–4A.4-6. Rome, NY: EOS/ESD Association Inc.
Linear Technology (n.d.). ESD Protection Program. http://cds.linear.com/docs/en/quality/
esdprotection.pdf [Accessed: 10th May 2017]
McAteer, O. 1990 Electrostatic discharge control. MAC Services In. ISBN 0-07-044838-8
McKeighan, R.E., Dailey, W., Pang, T. et al. (1986). Reversible charge induced failure mode of
CMOS matrix switch. In: Proc. EOS/ESD Symp. EOS-8, 69. Rome, NY: EOS/ESD Association
Inc.
Montrose, M. (2000). Printed Circuit Board Design Techniques for EMC Compliance, 2e. Wiley.
ISBN: 0 7803 5376 5.
Nichia (2014) Handling of LED products. Application Note SE-AP00001B-E. Available from:
www.nichia.co.jp/specification/products/led/ApplicationNote_SE-AP00001B-E.pdf
[Accessed: 21 Feb. 2019]
Norberg, A. (1992). Modelling current pulse shape and energy in surface discharges. IEEE
Trans. Ind. App. 28 (3): 498–503.
Olney, A., Gifford, B., Guravage, J., and Righter, A. (2003). Real- world charged board model
(CBM) failures. In: Proc. EOS/ESD Symp. EOS-25, 34–43. Rome, NY, EOS/ESD Association
Inc.
ON Semiconductor (2010). Human Body Model (HBM) vs. IEC 61000−4−2. App Note
TND410/D Rev. 0, SEPT – 2010. Available from: http://www.onsemi.com/pub_link/
Collateral/TND410-D.PDF [Accessed: 10th May 2017]
Osram (n.d.) ESD protection for LED systems. Available from: https://www.osram.com/ds/
news/avoiding-damage-caused-by-electrostatic-discharge/index.jsp [Accessed: 21st Feb.
2019]
86 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Paasi, J., Salmela, H., Tamminen, P., and Smallwood, J. (2003). ESD sensitivity of devices on a
charged printed wiring board. In: Proc. EOS/ESD Symp. EOS-25, 143–150. Rome, NY:
EOS/ESD Association Inc.
Reiner, J.C. (1995). Latent gate oxide defects caused by CDM ESD. In: Proc. EOS/ESD Symp.
EOS-17, 311–321. Rome, NY: EOS/ESD Association Inc.
Reliability Analysis Centre (1989a) Electrostatic Discharge susceptibility data of microcircuit
devices Vol. I. VZAP-2 Reliability Analysis Center P.O. Box 4700 Rome, NY 13440-8200.
Reliability Analysis Centre (1989b) Electrostatic Discharge susceptibility data of
discrete/passive devices Vol. II. Reliability Analysis Center P.O. Box 4700 Rome, NY
13440-8200.
Reliability Analysis Centre (1995) Electrostatic Discharge susceptibility data of discrete/passive
devices. VZAP-95 Reliability Analysis Center 201 Mill St, Rome, NY 13440.
Rider, G.C. (2016). Electrostatic risk to reticles in the nanolithography era. J.
Micro/Nanolithogr. MEMS MOEMS 15 (2): 023501.
Rider, G. C., Kalkur, T. S. (2008) Experimental quantification of reticle electrostatic damage
below the threshold for ESD. Proc. SPIE 6922, Metrology, Inspection, and Process Control for
Microlithography XXII, 69221Y
Sangameswaran, S., De Coster, J., Linten, D. et al. (2008). ESD reliability issues in
michromechanical systems (MEMS): a case study on micromirrors. In: Proc. EOS/ESD
Symp, 3B.1-1–3B.1-9. Rome, NY: EOS/ESD Association Inc.
Sangameswaran, S., De Coster, J., Scholz, M. et al. (2009). A study of breakdown mechanisms
in electrostatic actuators using mechanical response under EOS-ESD stress. In: Proc.
EOS/ESD Symp, 3B.5-1–3B.5-8. Rome, NY: EOS/ESD Association Inc.
Sangameswaran, S., De Coster, J., Linten, D. et al. (2010a). Investigating ESD sensitivity in
electrostatic SiGe MEMS. J. Micromech. Microeng. 20 (5): 055005.
Sangameswaran, S., De Coster, J., Chermin, V. et al. (2010b). Behaviour of RF MEMS switches
under ESD stress. In: Proc. of the EOS/ESD Symp, 443–449. Rome, NY: EOS/ESD Association
Inc.
Shaw, N.R. and Enoch, R.D. (1985). An experimental investigation of ESD damage to integrated
circuits on printed circuit boards. In: Proc. EOS/ESD Symp. EOS-7, 132–140. Rome, NY:
EOS/ESD Association Inc.
Smallwood, J.M. (2004). Static electricity in the modern human environment. In:
Electromagnetic Environments and Health in Buildings (ed. D. Clements-Croome). Taylor &
Francis. ISBN: 0 415 31656 1.
Smedes, T. (2009). ESD testing of devices, ICs and systems. Microelectron. Reliab. 49: 941–945.
Smedes, T. (2011) Machine Model – Correlation between HBM and MM ESD. Ch. 3 in Industry
Council on ESD Target Levels (2011) White paper 1: A case for lowering component level
HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf [Accessed: 10th May
2017]
Smedes, T. and Li, Y. (2003). Paper 2A.6). ESD phenomena in interconnect structures. In: Proc.
EOS/ESD Symp. EOS-25, 108–115. Rome, NY: EOS/ESD Association Inc.
References 87

Stadler, W., Brodbeck, T., Gartner, R., and Gossner, H. (2006). Cable discharges into
communication interfaces. In: 2006 Electrical Overstress/Electrostatic Discharge Symposium,
144–151. IEEE.
Stadler W., Niemesheim J., Stadler A., Koch S., Gossner H. (2017) Paper 3A1. Risk Assessment
of Cable Discharge Events. In: Proc. EOS/ESD Symp. EOS-39. Rome, NY, EOS/ESD
Association Inc.
Sylvania. (2009) ESD protection for LED systems. Application note. LED093. Available from:
http://assets.sylvania.com/assets/documents/ESD.a9e90e9d-c91e-4ea3-9ebe-0be6e5570cd7
.pdf [Accessed: 30th October 2017]
Szwarc, J. (2008) ESD Sensitivity of Precision Chip Resistors Comparison between Foil and
Thin Film Chips. Vishay Available from: http://www.vishaypg.com/docs/60106/esdsensi
.pdf [Accessed: 10th May 2017]
Talbot, J.W. (1986). The effect of ESD on III-Vmaterials. In: Proc. EOS/ESD Symp. EOS-8,
238–245.
Talebzadeh, A., Patnaik, A., Moradian, M. et al. (2015, 2015). Dependence of ESD charge
voltage on humidity in data Centers: part I-test methods. ASHRAE Trans. 121: 58.
Taminnen, P., Sydänheimo, L., Ukkonen, L. (2014) Paper 9A.2 ESD Sensitivity of 01005 Chip
Resistors and Capacitors In: Proc. EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc
Taylor, R.G. and Woodhouse, J. (1986). Junction degradation and dielectric shorting: two
mechanisms for ESD recovery. In: Proc. EOS/ESD Symp. EOS-8, 92. Rome, NY, EOS/ESD
Association Inc.
Tunnecliffe, M., Dwyer, V., and Campbell, D. (1992). Parametric drift in electrostatically
damaged MOS transistors. In: Proc. EOS/ESD Symp. EOS-14, 112–120. Rome, NY, EOS/ESD
Association Inc.
Viheriäkoski, T., Kokkonen, M., Tamminen, P., Kärjä, E., Hillberg, J., Smallwood, J. (2014) 4B.2
Electrostatic Threats in Hospital Environment. In: Proc. EOS/ESD Symp. EOS 36
Vishay (2011) Resistor Sensitivity to Electrostatic Discharge (ESD). Vishay Document 63129.
Available from: http://www.vishaypg.com/docs/63129/esd_tn.pdf [Accessed: 10th May
2017]
Wallash, A.J. (1996). Field induced charged device model testing of magnetoresistive recording
heads. In: Proc. EOS/ESD Symp. EOS-18. 4B.2, 8–13.
Wallash, A. and Honda, M. (1997). Field induced breakdown ESD damage of Magnetoresistive
recording heads. In: Proc. EOS/ESD Symp. EOS-19, 382–385. Rome, NY: EOS/ESD
Association Inc.
Wallash, A., Levitt, L. (2003) Electrical breakdown and ESD phenomena for devices with
nanometer-to-micron gaps. Proc. SPIE 4980 Reliability, Testing, and Characterization of
MEMS/MOEMS II, 87 doi: http://dx.doi.org/10.1117/12.478191
Walraven, J.A., Soden, J.M., Tanner, D.M. et al. (2000). Electrostatic discharge/electrical
overstress susceptibility in MEMS: a new failure mode. In: Proceedings of the SPIE 2000, vol.
4180, 30–39.
Walraven, J. A., Soden, J. M., Cole, E. I., Tanner, D. M., Anderson, R. R. (2001). Paper 3A.6
Human Body Model, Machine Model, and Charged Device Model ESD testing of surface
micromachined microelectromechanical systems (MEMS). In: Proc. EOS/ESD Symp. EOS-23.
Rome, NY, EOS/ESD Association Inc
88 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Wang, A.Z.H. (2002). On-Chip ESD Protection for Integrated Circuits. Klewer Academic Press.
Williams, T. (2001). EMC for Product Designers, 3e. Newnes. ISBN: 0 7506 4930 5.
Wilson, N. (1972). The static behaviour of carpets. Text. Inst. Ind. 10 (8): 235.

Further Reading

Agarwal S. (2014) Understanding ESD And EOS Failures In Semiconductor Devices.


Amerasekeera, E.A. and Campbell, D.S. (1986). ESD pulse and continuous voltage breakdown
in MOS capacitor structures. In: Proc. of the EOS/ESD Symp. EOS-8, 208–213. Rome, NY:
EOS/ESD Association Inc.
Bridgewood, M.A. (1986). Breakdown mechanisms in MOS capacitors. In: Proc. of the
EOS/ESD Symp. EOS-8, 200–207. Rome, NY: EOS/ESD Association Inc.
Colvin, J. (1993). The identification and analysis of latent ESD damage on CMOS input gates.
In: Proc. of the EOS/ESD Symp, 109–116. Rome, NY: EOS/ESD Association Inc.
Electronic Design. (2017) Understanding ESD And EOS Failures In Semiconductor Devices
Available from: http://electronicdesign.com/power/understanding-esd-and-eos-failures-
semiconductor-devices [Accessed: 10th May 2017]
EOS/ESD Association Inc. (2000a). Technical Report - Transient Induced Latch-up (TLU) ESD
TR5.4-01-00. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2000b). Technical Report - Calculation of Uncertainty Associated
with Measurement of Electrostatic Discharge (ESD) Current ESD TR14.0-01-00. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2002) ESD Phenomena and the Reliability for Microelectronics ISBN:
1-58537-046-0 Available from: https://www.esda.org/assets/Uploads/documents/ESD-
Phenomena-and-the-Reliability-for-Microelectronics.pdf [Accessed: 2nd November 2017]
EOS/ESD Association Inc. (2008a). Technical Report – Determination of CMOS Latch-up
Susceptibility – Transient Latch-up – Technical Report No. 2. ESD TR5.4-02-08. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2008b). Technical Report for the Protection of Electrostatic Discharge
Susceptible Items - Transmission Line Pulse (TLP) ESD TR5.5-01-08. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2008c). Technical Report for the Protection of Electrostatic Discharge
Susceptible Items - Transmission Line Pulse - Round Robin ESD TR5.5-02-08. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2011). Technical Report For Electrostatic Discharge Sensitivity
Testing - Latch-Up Sensitivity Testing of CMOS/BiCMOS Integrated Circuits - Transient
Latch-up Testing – Component Level - Supply Transient Stimulation. ESD TR5.4-03-11. Rome,
NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2012). ESDA/JEDEC Joint Technical Report User Guide of
ANSI/ESDA/JEDEC JS-001 Human Body Model Testing of Integrated Circuits ESDA/JEDEC
JTR001-01-12. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013a). Technical Report for Electrostatic Discharge Sensitivity
Testing - Transient Latch-up Testing ESD TR5.4-04-13. Rome, NY, EOS/ESD Association Inc.
Further Reading 89

EOS/ESD Association Inc. (2013b). Technical Report for the Protection of Electrostatic Discharge
Susceptible Items – System Level Electrostatic Discharge (ESD) Simulator Verification ESD
TR14.0-02-13. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014a). Technical Report for Electrostatic Discharge (ESD) Sensitivity
Testing – Very Fast – Transmission Line Pulse (TLP) – Round Robin Analysis ESD TR5.5-03-14.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014b). Technical Report for Relevant ESD Foundry Parameters for
Seamless ESD Design and Verification Flow ESD TR22.0.01-14. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2014c). Technical Report for ESD Electronic Design Automation
Checks ESD TR18.0-01-14. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015a). Standard Practice for Electrostatic Discharge Sensitivity
Testing – Near Field Immunity Scanning - Component/Module/PCB Level ANSI/ESD
SP14.5-2015. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015b). Technical Report for ESD Process Assessment Methodologies
in Electronic Production Lines – Best Practices used in Industry ESD TR17.0-01-15. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016a). Standard Test Method for Electrostatic Discharge (ESD)
Sensitivity Testing – Transmission Line Pulse (TLP) – Component Level ANSI/ESD
STM5.5.1-2016. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016b). Technical Report for Electrostatic Discharge Sensitivity
Testing – Charged Board Event (CBE) ESD TR25.0-01-16. Rome, NY, EOS/ESD Association
Inc.
EOS/ESD Association Inc. (1999) ESD Association Technical Report - Can Static Electricity be
Measured? ESD TR50.0-01-99
International Organization for Standardization. (2008) Road vehicles – Test Methods for
electrical disturbances from electrostatic discharge. ISO 10605:2008/Amd.1:2014(en)
King, W.M. (1979). Dynamic waveform characteristics of personnel electrostatic discharge. In:
Proc. of the EOS/ESD Symp. EOS-1, 78.
Lin, D.L., Strauss, M.S., and Welsher, T.L. (1987). On the validity of ESD threshold data
obtained using commercial human-body model simulators. In: Proceedings of the 25th
International Reliability Physics Symposium, 77. IEEE.
Lin, N., Liang, Y., Wang, P., and Pelc, T. (2014). Evolution of ESD process capability in future
electronics industry. In: 15th Int. Conf. Elec. Packaging Tech, 1556–1560. IEEE.
McAteer, O.J., Twist, R.E., and Walker, R.C. (1980). Identification of latent ESD failures. In:
Proc. of the EOS/ESD Symp. EOS-2, 54–57. Rome, NY: EOS/ESD Association Inc.
McAteer, O.J., Twist, R.E., and Walker, R.C. (1982). Latent ESD failures. In: Proc. of the
EOS/ESD Symp. EOS-4, 41–48. Rome, NY: EOS/ESD Association Inc.
Paasi, J., Smallwood, J., and Salmela, H. (2003) Paper 2B4). New methods for the assessment of
ESD threats to electronic components. In: Proc. of the EOS/ESD Symp, 151–160. Rome, NY:
EOS/ESD Association Inc.
Smallwood J, Paasi J. (2003) Assessment of ESD threats to electronic devices. VTT Research
Report No BTUO45-031160
Smallwood J., Taminnen P., Viheriaekoski T. (2014) Paper 1B.1. Optimizing investment in ESD
Control. In: Proc. of EOS/ESD Symp. EOS-36. Rome, NY, EOS/ESD Association Inc.
90 3 Electrostatic Discharge–Sensitive (ESDS) Devices

Strauss, M.S., Lin, D.L., and Welsher, T.L. (1987). Variations in failure modes and cumulative
effects produced by commercial human-body model simulators. In: Proc. of EOS/ESD Symp.
EOS-9, 59–63.
Viheriäkoski T, Peltoniemi T, Tamminen T, (2012) Paper 4A3. Low Level Human Body Model
ESD. In: Proc. of EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Vinson, J.E. and Liou, J.J. (1998). Electrostatic discharge in semiconductor devices: an
overview. Proc. IEEE 86 (2): 399–420.
Voldman, S. (2009). ESD Failure Mechanisms and Models. Wiley. ISBN: 978-0-470-1137-4.
Vollman, S., Hui, D., Warriner, L. et al. (1999). Electrostatic discharge (ESD) protection in
silicon-on-insulator (SOI) CMOS thechnology with aluminium and copper interconnects in
advanced microprocessor semiconductor chips. In: Proc. of the EOS/ESD Symp. EOS-21,
105–115. Rome, NY: EOS/ESD Association Inc.
91

The Seven Habits of a Highly Effective ESD Program

4.1 Why Habits?

Habit: “settled or regular tendency or practice, especially one that is hard to give up”
(Oxford Dictionary 2017)

For effective electrostatic discharge (ESD) control, we need to set up practices that reduce
the ESD risk to our electrostatic discharge–sensitive (ESDS) devices to an acceptable level.
These practices can be ways of working but also involve using certain ESD-protective equip-
ment to reduce ESD risk. If we can establish and maintain these practices so well, they
become a habit, and then our ESD control program is likely to remain effective.
Many ESD threats occur while handling ESDS devices, for example during assembly pro-
cesses. To protect against this sort of threat, we can set up a permanent or temporary ESD
protected area (EPA) in which the ESD threats are controlled so that we can handle the
devices and assemblies conveniently and relatively free from ESD risk.
Other ESD risks occur when an ESDS device is stored or transported in an uncontrolled
unprotected area (UPA) where static electricity can build up and ESD sources arise. In this
situation, we use ESD-protective packaging to enclose and protect the ESDS device from
damage.
Of course, we need to know how effectively we are adhering to our established ESD pro-
gram practices. Also, equipment is likely to fail from time to time under the day-to-day
wear and tear it experiences during use. We need to detect when equipment fails, falls out
of specification, or requires maintenance. For these reasons, a habit of checks and tests is
required.
Finally, we need to be sure that everyone who is concerned with ESD control or must use
the provisions of the ESD control program understands what they must do and not do. They
may need to know what equipment to use and even how to check it is functioning correctly.
They need to know what procedures to follow. It can be of great benefit if they watch out
for noncompliance and correct them as they go. So, training will be needed to make sure
they have the knowledge they need to fulfill these roles.
This chapter explores these habits, the reasons for them, and how to decide what should
be included in our habitual practices. Many of them are incorporated into the design and
specifications of equipment and materials designed for use in ESD control. We can enact
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
92 4 The Seven Habits of a Highly Effective ESD Program

some aspects of the “habits” by specifying special equipment and materials that are used in
EPAs where ESD risks are carefully controlled.

4.2 The Basis of ESD Protection

There are two key strategies that form the basis of successful ESD protection practice.

● Handle unprotected ESDS devices only in an area in which the ESD risks are reduced to
an insignificant level.
● In uncontrolled (unprotected) areas, protect ESDS devices by enclosing them within
ESD-protective packaging that protects the ESDS devices from ESD risks.

These two strategies should be applied to all aspects of handling, storage, and transport
of ESDS.

4.3 What Is an ESDS Device?

ESDS devices can be of many types and forms, from minute individual semiconductor
devices such as transistors, diodes, or integrated circuits, to printed circuit boards, mod-
ules, or system components. They usually contain semiconductor devices of some sort,
although other types of device (e.g. some types of resistors and capacitors) can have some
ESD sensitivity. ESDS devices and their failure modes are discussed in more detail in
Chapter 3.
The key factor in identification of an ESDS device is understanding that the item would
be at some risk of ESD damage if handled in an UPA without precautions. An item is an
ESDS device if it satisfies two criteria.

● The item contains parts that could be damaged by ESD.


● If handled in a UPA, there is a risk that potentially damaging ESD could find a route to
the ESDS parts.

It follows that if either of these criteria is not met, the item can be considered as not
being ESDS.
In building an electronic system, the risk of ESD damage and susceptibility to ESD often
change considerably with build state. Let us take as an example the construction of a simple
product that consists of a printed circuit board (PCB) within a housing. Many of the com-
ponents that go onto the PCB are likely to be ESDS. The populated PCB is also likely to be
ESDS and should be handled as such. However, once built into and protected by its housing,
the final product may well be quite immune to normal ESD sources due to the protective
barrier provided by the housing. In many cases, electromagnetic compatibility (EMC) reg-
ulations may require testing and demonstration of ESD immunity of the working system
for market acceptance. There may, however, be some residual ESD susceptibility dependent
on its design, for example susceptibility to ESD to connector pins, e.g. from connection of
charged cables.
4.4 Habit 1: Always Handle ESDS Components Within an EPA 93

In some cases, the build stage at which the product can be considered no longer suscep-
tible to ESD damage is not clear. This will then have to be decided by some evaluation of
the ESD risks and susceptibility.
An item such as a PCB may be subject to ESD occurring to almost any part of it that may
be touched by a person, tool, or other conductive item. If the same PCB is encapsulated or
potted to become a module or subassembly, the number of ways in which ESD can occur to
it, and hence ESD risk, is much reduced as the encapsulation can form an effective barrier
to ESD to most parts of the assembly. This may not mean that the module is immune to ESD
damage or needs no ESD protection. The module may have flying leads or connectors to the
PCB within. The module may be susceptible to ESD to these leads or connections, unless
designed for immunity. Tribocharging of the module surface can lead to high induced volt-
ages occurring on the PCB within, and these can discharge if the leads or connector pins
touch something conductive.
A product that is at a built state at which it is no longer ESDS can become susceptible
to ESD damage again if modified or disassembled in some way. For example, a desktop
computer would not normally be considered ESDS once it is fully assembled to the state
at which the user would normally receive it. If, however, the covers are removed, PCBs
containing ESDS parts may become accessible to touch, and appropriate ESD control pro-
cedures should be in place while operating on these PCBs. Once the covers are replaced
and no ESDS parts are accessible, the system can again be considered not susceptible to
ESD damage.

4.4 Habit 1: Always Handle ESDS Components Within an EPA


4.4.1 What Is an EPA?
As far as ESD protection is concerned, there are two types of area: EPAs and UPAs (see
IEC61340-5-3:2015). In some companies, the EPA may be known by another name or
acronym such as safe handling area (SHA). In this book, the terminology used in current
standards is used. It doesn’t matter what they are called – it is what happens within these
areas that is important in preventing ESD damage to sensitive components.
Most areas are of course UPAs. In these areas, static electricity is uncontrolled and often
rife and omnipresent. We are not necessarily conscious of this, because we are ourselves
rather insensitive to static electricity. As we move around, we routinely develop voltages
on our bodies of hundreds of volts. We don’t feel voltages; we feel the current and energy
in a discharge. We touch things or people and discharges occur, and we are oblivious to
them. Only if these body voltages reach a few thousand volts do we start to experience these
discharges as shocks and, even then, only if we touch a substantial conductor like another
person or a metal filing cabinet. If we touch a resistive material, the discharge current is
reduced to the point where we don’t feel it.
We are even less sensitive to the voltages that arise on insulating materials like plastic
packaging and stationary materials. Voltages of several thousand volts (1000 V = 1 kV) can
arise on these materials without us noticing. As these voltages exceed 20 kV or so, small
brush discharges can arise that we might hear as an occasional clicking sound, if the condi-
tions are quiet. We might start to feel the “tickling” sensation of hairs on our skin moving
94 4 The Seven Habits of a Highly Effective ESD Program

in response to electrostatic attraction from the high electrostatic fields near these surfaces.
When taking off fleece or other clothing made of man-made fibers, we may hear the crack-
ling of small discharges, and, in low light conditions, we may see them as tiny flashes.
While we are insensitive to ESD and electrostatic fields, as we have seen, many electronic
components are not. We must protect them against electrostatic effects, either by handling
them within an EPA or by enclosing them in ESD-protective packaging. This section is
about EPAs and the habits we need to develop to make them effective.
We can make EPAs in many different forms. They can be temporary, or they can be fixed
facilities. A field service kit can be used to provide a temporary EPA for field use. A fixed
facility may be a single workstation or may enclose many workstations or a whole room or
work area (Figure 4.1). A machine, or part of a machine, where unprotected ESDS devices
are handled may need to be part of an EPA.
So, what do we need to have to make an effective EPA? There are two basic aspects.

● There must be a clear boundary.


● Within the boundary, all ESD risks must be controlled to give insignificant ESD damage
levels.

The need for a clear boundary is because we must be clear whether we are inside the
EPA or outside it. Outside the EPA boundary, the ESDS device should never be taken out

Figure 4.1 EPA and UPA.


4.4 Habit 1: Always Handle ESDS Components Within an EPA 95

of its ESD-protective packaging. To do so would be to expose it to risk of ESD. Inside the


EPA, the ESDS device may be kept within its protective packaging, or it may be taken out
to be handled or used in processes as required, because the ESD risks are controlled to a
low level.

4.4.2 Defining the EPA Boundary


The first requirement for an EPA is for it to have a clear boundary so that everyone knows
whether they are inside or outside the EPA. If we are not clear about the boundary, we
cannot be clear about whether it is safe to take an ESDS device from its ESD-protective
packaging. We will not know whether we should be taking the prescribed ESD control pre-
cautions and using ESD control equipment or not. So, lack of a clear boundary is likely to
lead to noncompliance sooner or later, as well as possible ESD risk to ESDS.
For an EPA to be effective, all processes within the EPA must be evaluated and controlled
if necessary. It is not sufficient to equip the area with common ESD control equipment
(bench mats, wrist straps, and the like) if other major sources of ESD in processes are
ignored.
It is usually beneficial to think carefully about the EPA boundary and include within the
EPA only such processes as are necessary. The less is in there, the less needs to be equipped
and evaluated for ESD control. Handling the unprotected ESDS device as little as possible
can be an effective first ESD prevention measure. Minimizing the number of workstations
and processes within the EPA can reduce the necessary expenditure on ESD control equip-
ment, as well as reducing the burden of equipment checks and maintenance. It can be a
good policy to exclude from the EPA any processes in which unprotected ESDS devices do
not need to be handled.
Against this, it is sometimes more convenient to include certain processes that may give
ESD risk within the EPA for convenience in moving between processes and operations.
This can be acceptable provided the ESDS devices are not put at risk. A common example
may be to have an office work desk area where papers may be kept and worked on. The
solution to preventing ESD risk is to make sure unprotected ESDS are never brought into
the office desk area. Conversely, the materials in the office area that might cause risk must
never be taken to workstations where unprotected ESDS devices are handled. Managing
this requires a high level of awareness and compliance in the personnel who work in
these areas. This in turn requires some training to create the intended habits. It also
requires some compliance verification to make sure the intended practice is successfully
maintained.

4.4.3 Marking the EPA Boundary


There is no single method of marking an EPA boundary. What is most important is that
the EPA boundary is recognized by, and immediately obvious to, the people entering and
working in the area. This includes personnel who are not authorized to enter the EPA,
perhaps because they have not had ESD training. Untrained personnel must clearly see
that the area should not be entered.
96 4 The Seven Habits of a Highly Effective ESD Program

It is often a good idea to restrict entry into the EPA to certain points rather than having
an extended open boundary. This can minimize the amount of boundary marking required.
A physical barrier (temporary or permanent) can go a long way to reduce unauthorized
entry into an EPA.
EPA entry points should be marked with clear signage showing personnel approaching
the entrance that they are about to enter an EPA (Figure 4.2). Signs should be obvious and
eye-catching. Signs at around eye-level may be more easily noticed than signs positioned at
low level or above an entrance.
It can also be a good idea to post signs visible to personnel leaving the EPA, warning that
they are about to leave the EPA.

Figure 4.2 Examples of marking an EPA entrance. Source: C. Cawthorne.


4.4 Habit 1: Always Handle ESDS Components Within an EPA 97

Some companies establish electronically operated turnstiles or barriers at the EPA


entrance that allow entry only to personnel who have passed personal grounding
equipment (wrist strap or footwear) tests.

4.4.4 What Is an Insignificant Level of ESD Risk?


The level of ESD risk that is considered insignificant depends on the type of product and
its market and on the possible consequences of a failure. At one end of the scale, a low-cost
consumer product might merit a low level of care and expenditure on ESD protection with
an acceptable number of ESD failures. For example, a low-cost insert for a musical greeting
card or talking toy may be a throwaway item. A certain level of failure of these may be easily
tolerable.
At the opposite extreme, a satellite must operate reliably once deployed, with no pos-
sibility of maintenance if a failure occurs. A failure would be catastrophic to the mission
and be very costly. Automotive products are often made in high quantities, and a very low
failure rate is demanded. A failure in service would be potentially catastrophic in this case
also, with possible risk of injury or death to a driver and passengers. Aerospace and military
applications are also areas where high reliability is demanded, and the consequences of a
failure are dire.
The level of ESD risk that is considered insignificant is a matter for the user to decide,
according to their view of their product and market needs.

4.4.5 What Are the Sources of ESD Risk?


There are two overall types of ESD risk that are controlled in an EPA.
● Direct ESD to or from the device
● Electrostatic fields that could lead to ESD to or from the device
Identification and evaluation of these is discussed further in Chapter 9. The sources of
ESD include the following:
● Charged personnel
● Charged metal or conductive objects or materials
● Charged devices
The sources of electrostatic fields that are of concern are normally insulating materials
such as plastics that easily become highly charged and retain their charge for long periods.
The risk of ESD damage is controlled by factors such as the following:
● The likelihood that ESD can occur to the ESDS
● The likelihood that the ESD current passes through the susceptible part of the ESDS
device
● The likelihood that ESD energy, peak current, charge transferred in the discharge, or
some other parameter exceeds a damage threshold of the ESDS device
It follows that the risk of ESD damage occurring can be reduced by reducing the risk of
ESD occurring, as well as reducing the likely strength of ESD if it should occur.
98 4 The Seven Habits of a Highly Effective ESD Program

There is a risk of ESD occurring whenever an ESDS device contacts another conductor at
a different voltage. The strength of the ESD occurring can be minimized by decreasing the
likely peak current, energy, and charge transferred in any ESD that may occur. The way we
evaluate and control ESD risk is further discussed in Habits 2–5.

4.4.6 What ESD Protection Measures Are Needed in the EPA?


The ESD protection measures needed in the EPA will depend on a wide range of considera-
tions including the processes and product, ESD susceptibility of the ESDS device, and other
factors. They commonly include use of equipment such as the following:

● Packaging for ESD protection and control


● ESD control flooring or floor mats
● Personal grounding equipment (wrist straps or ESD control footwear and flooring)
● Bench mats or work surfaces, racks, and carts
● ESD control seating
● Garments
● Gloves and finger cots
● Tools

One common way of deciding what ESD measures should be implemented is to adopt
the requirements of an ESD control standard (see Chapter 6). These normally list a range
of control measures that address the most common ESD threats. A detailed evaluation of
ESD risks is often not attempted. This approach has the advantage that it is easily achieved
and needs the minimum of expertise. The ESD control program that results may be easily
accepted by customers and found compliant with the standard by auditors.
This approach can, however, have disadvantages.

● ESD control measures may be included that address risks that do not in fact exist within
the processes and facility.
● ESD risks may exist that are not addressed by the standard ESD control measures.
● The ESDS device is unusually sensitive, having ESD withstand voltage lower than the
design withstand voltage level of the standard.

For these reasons, it may be necessary or preferable to undertake some level of evaluation
of the ESD risks in a process and facility in preparation for determining the ESD control
measures that are required. With fuller knowledge of the specific ESD risks, ESD control
measures can be specified that address these more effectively, efficiently, and completely.
Determination of ESD protection measures is further discussed in Chapter 10.

4.4.7 Who Will Decide What ESD Protection Measures Are Required?
An ESD control program that has no one leading and responsible for it is likely to fail
through lack of attention and maintenance. Someone will need to develop and implement
the ESD control program. Someone will need to document it, maintain it, test it, and train
people to work within it.
4.5 Habit 2: Where Possible, Avoid Use of Insulators Near ESDS 99

These functions may of course involve several people taking different roles and respon-
sibilities. Some companies have a committee of personnel within a site and, for a multisite
organization, on different sites, working on a company ESD program.
Nevertheless, it is advisable to have a person responsible for coordinating, implementing,
and maintaining the ESD control program at each site. In the main ESD control standards
current at the time of writing, it is a requirement to have such a person, known as the ESD
Coordinator. They do not have to do everything themselves, but they do have to make sure
it gets done. So, they must have the necessary authority, backup, and resources to fulfill
the role.
In some organizations, a committee rather than a person fulfills this role. Usually, the
tasks required in implementing and maintaining the ESD control program are delegated
wholly or in part to other people in the organization. For example, specially trained techni-
cians may do routine testing of the equipment in the EPA, and a designated trainer may do
some or all the training.

4.5 Habit 2: Where Possible, Avoid Use of Insulators Near


ESDS

4.5.1 What Is an Insulator?


What is a insulator? In this book, I define an insulator as any material or object on which
electrostatic charge cannot move away quickly enough to avoid significant electrostatic
charge buildup or voltage differences occurring. This is a pragmatic rather than academic
definition, and it reflects the way these terms tend to be used in industry in practice. It
is deliberately a definition that can lead to differences between electrical characteristics
that might be defined as insulating in different industrial contexts. This is what happens
in practice. So, in electrostatic fire and explosion hazards avoidance in industrial processes,
an insulator may be in general considered to be a material having resistance over 100 MΩ
(see IEC60079-32-1:2013). Even within the 60079-32-1 document, the word insulating has
a wide range of different definitions for different products or materials. An enclosure clas-
sified as insulating has volume resistance of 100 GΩ or greater, whereas a hose classified as
insulating has resistance greater than 1 MΩ (see Section 1.7.5).
So, charge does not move around through or on an insulator easily and may remain for a
significant time. This behavior gives the following effects:

● The voltage is unlikely to be the same at every point on the surface of the insulator. When
voltage differences occur, electrical currents will not flow quickly enough in response to
prevent the differences.
● If ESD from the insulator occurs, only a small amount of the charge and energy stored
on the insulator may be delivered during the discharge.

The apparently perplexingly flexible definition of an insulator can be explained by revis-


iting our simple model of electrostatic charge buildup (Figure 4.3). In Chapter 3 we found
that the resistance (or resistivity) had a strong effect on two important things.
100 4 The Seven Habits of a Highly Effective ESD Program

dQ/dt = I
Charge
generation R or ρ C or ε
(current
source)

Figure 4.3 A simple electrical model of electrostatic charge buildup, revisited.

● The voltage developed in response to an electrostatic charging current, governed by the


product of charging current and resistance (or material resistivity)
● The charge or voltage decay time governed by the product of resistance R (or material
resistivity 𝜌) and capacitance C (or material permittivity 𝜀 = 𝜀r 𝜀0 ).

Typically, we wish to prevent a risk by either keeping the voltage produced below a cer-
tain value at which some hazard may occur or making sure that any voltages arising are
dissipated quickly before they have chance to cause a hazard.
In an EPA where manual handling is the norm, things don’t usually happen very quickly.
If there are no strong continuous electrostatic charging mechanisms present, it will often be
sufficient to make sure that charge and voltages produced during normal contact between
materials is dissipated within a few seconds. So, materials will often be acceptable if the
decay time 𝜏 = 𝜌𝜀r 𝜀0 is of the order of a few seconds. Given that many materials have
relative permittivity 𝜀r around 2–3, and 𝜀0 = 8.8 × 10−12 Fm−1 , then setting an upper limit
of material volume resistivity of 100 GΩm (1011 Ωm) gives theoretical decay times around
1.8–2.6 seconds. This approximately correlates with the upper limits usually chosen for
ESD-protective packaging materials, although these are usually expressed as surface or vol-
ume resistance rather than resistivity (see Section 1.7).

4.5.2 Essential and Nonessential Insulators


Insulating materials can be thought of as being of two types. Essential insulators are those
that are a necessary part of the process or product. Without these, we cannot make the
product or do what we need to do with it.
All other insulators that are not a necessary part of the product or process are nonessential
insulators. These should be kept sufficiently far away from ESDS devices that ESD risk is
reduced to an acceptable level. Many organizations find it easiest to make sure that they
do not enter the EPA. If they are allowed into the EPA, then the proximity of insulators to
ESDS devices must be effectively managed. This requires careful definition of where and
how they may be used, training of personnel, and some compliance verification.
4.5 Habit 2: Where Possible, Avoid Use of Insulators Near ESDS 101

Table 4.1 Some examples of essential insulators and nonessential insulators.

Item Nonessential Essential

PCB substrate and component plastic packages Yes


Product components and parts made from plastics Yes
Plastic packaging not designed for ESD control Yes
Personal items, coffee cups, lunch boxes Yes
Parts of test jigs and fixtures Parts that do not need Parts that must be
to be made from made from
insulating materials insulating materials
Papers Papers that are not Papers that are
required to be present required to be
or used during the present or used
process during the process

Table 4.1 gives some examples of essential and nonessential insulators that can be found
in many facilities. Some types of insulator are clearly essential (e.g. the PCB and compo-
nents). Others are less easily categorized (e.g. parts of test jigs and production papers). In
some cases, these may be essential, and in others nonessential.

4.5.3 Remove Nonessential Insulators from the Vicinity of ESDS


Electrostatic fields due to charged insulators introduce a risk of ESD. This is because
they induce voltages on any isolated (ungrounded) conductors in the field. Any isolated
conductors (including any ESDS devices) within the field will be at some unknown voltage
that will in general be different from any other conductor nearby in the field (see Section
4.7 for discussion of conductors and grounding). If two conductors become sufficiently
close or touch, ESD will occur between them. If one of the conductors is part of an ESDS
device, this ESD risks causing ESD damage.
The electrostatic field from a charged object depends on the distance from that object.
In Chapter 2 we saw that a uniform field is produced between two large parallel plate
conductors. If an electrostatic field meter is inserted into an aperture in one of the plates, it
will measure the field that is produced (Figure 4.4). The electrostatic field E is easily calcu-
lated from the voltage difference between the plates V and the distance between them d.
V
E=
d
If the high-voltage plate is moved toward, or away from, the field meter plate, the electro-
static field is increased, or decreased, proportional to 1/d.
This simple equation tells us that the closer an electrostatic field source is to an ESDS
devices, the greater the electrostatic field. As electrostatic field is an indicator of ESD risk,
the ESD risk is also greater. If we set a limit of say 5 kV m−1 for the field, this limit is achieved
by different voltages at different distance (Table 4.2). The closer a voltage source is to the
ESDS device, the more concerned we will be about it and the lower the tolerable voltage
limit.
102 4 The Seven Habits of a Highly Effective ESD Program

Electrostatic field meter

Metal
plate Metal
plate E=V/d
V=1000V Field E 0V

Figure 4.4 Electrostatic field between two parallel conducting plates.

Table 4.2 Voltages and distances between plane


parallel electrodes giving the electrostatic field of
5 kV m−1 .

Voltage (V) Distance giving 5 kV m−1

10 000 2m
5 000 1m
2 500 50 cm
1 500 30 cm
500 10 cm
125 2.5 cm
50 1 cm

If the field meter plate is removed, leaving the field meter in position, the electrostatic
field lines terminate at the field meter instead of the plate (Figure 4.5). We assume that the
electrostatic field meter is constructed as an earthed conductor at 0 V. The field lines focus
on the field meter and so a higher nonuniform electrostatic field exists at the field meter in
this case. Because of this, we can no longer assume that as we change the distance between
the field meter and the plate, the field will vary with 1/d.
In practice, the assumption that the field decreases as 1/d often gives reasonable agree-
ment with experiment (Stadler et al. 2018). Many field meters are calibrated to give a reading
of surface voltage when at a set distance of 2 or 2.5 cm (1 in.) from a large flat metal sheet at
a given voltage. In Figure 4.6 the voltage reading taken with the field meter at 2 cm distance
from a flat metal plate at 1 kV. The field (voltage) readings at other distances are shown as
a percentage of the field (voltage) reading at other distances. It is interesting to note that
the field at 30 cm distance is only 7% of the value at 2 cm distance. Some standards require
4.5 Habit 2: Where Possible, Avoid Use of Insulators Near ESDS 103

Distance d

Electrostatic field meter


0V

Metal
plate
Field E

V=1000V

Figure 4.5 The electrostatic field between a field meter and a charged plate varies with distance.

Figure 4.6 Electrostatic field meter reading variation with distance from a charged metal plate at
0.5 kV, expressed as a percentage of the value measured at 2.5 cm distance.

that any insulator that has a surface voltage >2 kV measured in this way should be kept at
least 30 cm from any ESDS device. This practice ensures that the electrostatic field from the
charged insulator experienced by the ESDS device is less than 7% of its close-up value.
Alternatively, standards often give a limit to the electrostatic field at the position of the
ESDS, e.g. 5 kV m−1 . As most field meters are calibrated in terms of voltage, it is not imme-
diately obvious how to measure this. However, any field meter calibrated in terms of voltage
can be easily calibrated to find this field limit, using a metal plate raised to a set voltage to
give the field as in Figure 4.5. For example, if a field meter is set up at 0.020 m from a metal
plate, and 100 V is applied to the plate, the electrostatic field is 100/0.020 = 5000 V m−1 . The
reading observed on the meter with this setup will depend on the design operational and
calibration conditions of the meter. This does not matter, and it does not matter what the
104 4 The Seven Habits of a Highly Effective ESD Program

number is – it represents the field of 5 kV m−1 . Anything giving a reading above this value
is producing a field above 5 kV m−1 , and anything less is a field <5 kV m−1 .

4.6 Habit 3: Reduce ESD Risks from Essential Insulators


4.6.1 What Is an Insulator?
For the purposes of this book, I define an insulator as any material or object on which elec-
trostatic charge cannot move away quickly enough to avoid significant electrostatic charge
buildup or voltage differences occurring (see Section 4.5).
In ESD work, a material of resistance above about 100 GΩ (1011 Ω) is usually considered
an insulator. A material having resistance below this is usually considered a conductor (see
Section 4.7) and can be used to control and avoid electrostatic charge buildup. However,
different industries and disciplines have very different definitions or ideas about insulating
material resistance.

4.6.2 Insulators Cannot Be Grounded


Inexperienced people in ESD often believe that the charge on an insulator can be controlled
by grounding. This simply cannot work. The reason is obvious from my definition of an
insulator. The charge cannot move from the insulator to any ground wire quickly enough
to prevent static charge build up on the insulator. Sometimes, touching a ground wire to an
insulator surface can reduce the surface charge level in the vicinity of the touch point by
brush discharge (see Section 2.5.2). The charge level away from the touch point will often
be unaffected by this as the charge cannot move across the surface easily or quickly.

4.6.3 What to Do About ESD Risk from Essential Insulators


ESD risk is in many cases a result of the electrostatic field from charged materials. The risk
often arises because voltage differences arise between conductors in an electrostatic field if
they are not equipotential bonded. One of these conductors is typically the ESDS device. If
the ESDS device comes close enough to, or touches, another conductor at a different voltage,
ESD will occur between them.
The classification of insulating items is arbitrary and a matter of opinion. The same item
may be considered essential in one process and nonessential in another. For example, it may
be easy to remove papers from one workstation process, but in another it may be difficult
to proceed without them if they must be updated or signed on completion of process steps.
In most cases, the risk from insulators can be assessed by a simple process of evaluation
such as the procedure given in Section 9.3.6. Most ESDS are not inherently sensitive to
damage directly from electrostatic fields. The ESD risk is usually significant only if there are
significant electrostatic fields and there is the possibility of contact between the ESDS device
and another conductor within the field. If there is no contact with the ESDS within the field,
it may not be necessary to control the field. If the insulator is sufficiently far from the ESDS,
the electrostatic field arising at the position of the ESDS may be negligible. (Take care that
4.6 Habit 3: Reduce ESD Risks from Essential Insulators 105

the ESDS is not likely to be moved into a position where the field may be significant.) If the
insulator is not likely to be handled or moved or become charged, the risk of a field arising
may be negligible.
The ESD risk can in principle be controlled in several ways.

● Replacing the insulator with a grounded conductor


● Increasing the distance between the charged insulator and the position of the ESDS
● Containing the field from the insulator by shielding
● Preventing contact between the ESDS and other conductors within the field
● Reducing the charge and voltage level on the insulator e.g. by neutralization using an
ionizer

The ESD control measure should be chosen as appropriate to the situation. Some possible
examples are given in Table 4.3.
Electrostatic charging of the item should preferably be measured under worst-case
conditions, which usually means under low-humidity (≪30% rh) ambient atmosphere. In
practice, measurements may have to be done under ambient atmospheric conditions, as
a humidity-controlled facility is usually not available. Nevertheless, an initial evaluation
done under ambient condition at higher humidity may give useful first evaluation and
should then be followed up and repeated when the weather conditions give lower humidity.
The question then arises – what level of charging can be considered negligible (Swenson
2012)? Unfortunately, this may not be easy to answer and depends on the withstand voltages
of the ESDS being handled and other factors. For example, if it is charged device model
(CDM) damage to the ESDS device that is of concern and if the voltages produced are lower
than the CDM withstand of the ESDS device, they can be considered negligible. The voltage
induced on a conductor can never exceed the voltage of the electrostatic field source. In
practice, higher voltages may also be negligible, but evaluation of this may be more difficult.

Table 4.3 Some examples of essential insulators and possible ways of dealing with them.

Essential insulators Possible ways of dealing with them

Product components and parts Use an ionizer to reduce charge levels to acceptable value.
Parts of test jigs and fixtures that must Treat with an antistat on regular basis and reduce
be made from insulating materials charging with humidity control.
Use an ionizer to reduce charge levels to acceptable value.
Papers that are required to be present Keep in static dissipative document holders.
and used during the process If required to be removed from holders (e.g. for
annotation), do this on a separate work area a minimum
of 300 mm from workstation where ESDS devices are
handled.
Use computer-based document displays designed to be
ESD safe.
Computing equipment on the Position the equipment on a separate part of the
workstation workstation or well away from the likely position of
ESDS.
106 4 The Seven Habits of a Highly Effective ESD Program

Standards may also give requirements that can be used to evaluate fields from charged
insulators. For example, the IEC 61340-5-1:2016 standard gives requirements that the elec-
trostatic field at the position of the ESDS must be <5 kV m−1 . Also, insulators charged to
>125 V must be kept at least 2.5 cm from the ESDS, and if charged to >2 kV must be kept
>30 cm from the ESDS. If these conditions are fulfilled, the electrostatic fields and voltages
can be considered negligible for the purposes of this standard.

4.6.4 Using Ionizers to Reduce Charge Levels on Insulators


As previously explained, ionizers can be used to neutralize excess charge on surfaces. So,
they can be used to neutralize charged insulators present in the EPA. For successful control
of ESD risk, the limitations of ionizers must be understood.
Ionizers produce air ions at a given rate, and these drift to the charged surface at a rate
determined by the electrostatic field (determined by the surface voltage) and the ion mobil-
ity (see Section 2.8). The surface charge on an insulator can be neutralized only at the
rate at which these ions can arrive. The ion arrival rate and hence charge neutralization
rate decreases as the surface voltage decreases (Figure 4.7). Moreover, as the ion mobility

Figure 4.7 Ionizer charge decay curves showing decay time and offset voltage for positive (above)
and negative (below) charge neutralization.
4.7 Habit 4: Ground Conductors, Especially People 107

can be different for positive and negative ions, the charge neutralization rates can also be
different, even for a well-balanced ion stream. For a poorly balanced ion stream the differ-
ence in ion concentration will also contribute to a difference in charge neutralization rate
for each polarity. So, neutralization of one polarity can be significantly slower than the other
polarity. As a result, as Figure 4.7 shows, it can take a significant time for a surface charge to
be neutralized to a low level. In Figure 4.7, the negative polarity voltage is reduced to −100 V
in <5 seconds. The positive polarity voltage requires about double this time to reduce to
+100 V. In an assembly process, it may be necessary to wait several seconds until the charge
has been reduced to a sufficiently low level to reduce ESD risk to an acceptable level.
The voltage decay time also varies with the position of the charged object relative to the
ionizer. How it varies will depend on the type of ionizer used. This can be an important
factor in choosing an ionizer for a process role. Typically, the charge decay time will increase
as the distance from the ionizer increases. This is because the ion concentration in the air
reduces as the ions spread out by mutual repulsion of like charges, and opposite polarity
charges attract and recombine forming neutral particles.
Many ionizer types can also be quite directional in their effectiveness. For example, a
fan ionizer blows ions in one general direction with the fan airstream. Its effectiveness can
dramatically reduce outside the airstream.
Most electrical ionizers exhibit an offset voltage due to a small imbalance of the posi-
tive and negative ion density that they produce. The ionizer offset voltage does not usually
cause any problems in neutralizing insulators. Small charge levels giving a few tens of Volts
on insulators do not usually cause ESD risk except in handling extremely sensitive compo-
nents. Standards often specify a maximum acceptable offset voltage or may leave it to the
user to define an appropriate maximum for their application.

4.7 Habit 4: Ground Conductors, Especially People

4.7.1 What Is a Conductor?


What is a conductor? In this book, I define a conductor as any material or object that is not
an insulator. I define an insulator as any material or object on which electrostatic charge
cannot move away quickly enough to avoid significant electrostatic charge buildup or volt-
age differences occurring. So, a conductor is a material that allows charge to move away
quickly enough to avoid static charge buildup.
These may seem vague nonspecific definitions, but they are deliberately so. Whether a
material is considered a conductor or insulator often depends on the context or technology
area. An electrical engineer might consider a material that has resistance of 1 GΩ or above to
be an insulator. Even in electrostatics process hazards evaluation and prevention that might
be considered the case. In ESD control work, a material of this resistance is considered a
conductor and can be used to dissipate and control charge or ground conductors.

4.7.2 Conductive, Dissipative, or Insulative?


In ESD work, many people use the terms conductive, dissipative, and insulative as if they
have generally accepted definitions. Conductive is often thought to apply to materials
108 4 The Seven Habits of a Highly Effective ESD Program

having resistance <1 MΩ (106 Ω). Dissipative is thought to apply to materials having resis-
tance between 1 MΩ and 100 GΩ (1011 Ω). Insulative is thought to apply to materials having
resistance over 100 GΩ (1011 Ω). Beware – these definitions are not universal. Within the
61340-5-1 and S20.20 systems of standards, these terms have standardized definitions only
in some contexts. One is ESD-protective packaging materials (see Chapter 8), where an
insulator is a material having surface or volume resistance greater than or equal to 100 GΩ.
Even in this specific topic, the definition of a “conductive” packaging material has changed
over recent years with updating standards. Because of this lack of clarity, it is unwise to
specify materials in terms of conductive or dissipative. Instead, measurable parameters
such as an acceptable range of surface resistance should be specified.

4.7.3 Properties of a Conductor


A conductor has some important properties for ESD control. These arise from the charac-
teristic that charge can move around the conductor relatively easily.

● Under quasistatic conditions with no current flowing through the conductor, the voltage
will be the same at every point on the surface of the conductor.
● If ESD from the conductor occurs, almost all the charge and energy stored on the con-
ductor could be delivered from the conductor during the discharge

The first point arises from the fact that if a voltage difference occurred, a current would
flow until no voltage difference is present. Equilibrium could be attained relatively quickly.
In practice, the timescale in which it happens depends on the material characteristics,
namely, resistivity and permittivity.
The second point arises from the fact that when a discharge is initiated from the material,
the voltage at the point of discharge changes quickly as charge is conducted away. Voltage
differences then occur across the material that cause currents to flow until the voltage across
the material is again equalized. This can be the point at which charge stored on the material
is exhausted. Because of this, a charged conductor is often a potent source of ESD.

4.7.4 Charge and Voltage Decay Time


In theory, a material has a voltage and charge decay time characteristic given by the prod-
uct of the resistivity 𝜌 and permittivity 𝜀 (see Section 2.3.3). In the case of a conductor of
capacitance C and resistance to ground R, the charge or voltage decay time is given by the
product RC. The shorter the decay time, the more quickly the voltage reaches equilibrium
across the material. In the case of a grounded conductor or material, the time taken for the
conductor or material to approach 0 V is governed by the charge decay time characteristic.
Many materials have permittivity around 10−11 Fm−1 and so resistance up to around
100 GΩ (1011 Ω) gives decay times of the order of a second or so. In practice and especially
in manual processes, if any charge generated is dissipated within this time scale, it is
unlikely to cause any significant ESD risk.
Small conductors in the workplace (e.g. hand tool bits) may have capacitance of the order
of 10 pF. Resistance to ground as high as 100 GΩ will give charge or voltage decay times of
4.7 Habit 4: Ground Conductors, Especially People 109

the order of a second or so, which may be acceptable. Higher-capacitance items will need a
lower resistance to ground to keep charge and voltage decay times acceptably short.
This situation may be different in automated processes, partly because continuous charge
generation processes may be present. Machine movements may be much faster than in
manual processes and so shorter charge decay times may be necessary to avoid significant
charge buildup and ESD risk.

4.7.5 The Importance of Material Contact Resistance in Protecting ESDS


4.7.5.1 Reduction of Energy Delivered from Conductor in ESD
When an ESDS contacts a material and discharge occurs, the current flows through the
ESDS device and also the material. A portion of the available energy in the discharge is
absorbed in the material rather than in the ESDS device. The higher the resistance of the
material, the higher the proportion of energy absorbed in the material and the less dissi-
pated in the ESDS device. Where the energy dissipated in the ESDS device is important in
the damage mechanism, reducing this can have a useful protective function. Having the
ESDS device contact high resistance rather than low resistance material can give useful
reduction in ESD stress.

4.7.5.2 Reduction of Peak Current in a Discharge


When an ESDS device makes contact with a material and discharge occurs, the peak current
that flows is limited by the resistance and inductance in the discharge circuit. Where the
resistance is high, this can be the main factor limiting the peak current in the discharge.
Where the peak current in the discharge is important in the ESDS damage mechanism,
increasing the resistance of materials with which it may make contact can significantly
reduce the ESD stress.
This is a significant consideration in charged device ESD damage prevention. In CDM
ESD susceptibility tests, it has been found that it is the peak current in the ESD that
most often gives the device damage threshold (see Chapter 3). If the device contacts only
high-resistance materials, the peak ESD current can be effectively limited to less than the
damage threshold. Charged device ESD damage can be effectively prevented.
It is important to understand that the ESD peak current is limited by the resistance of the
material at the point of discharge. Resistance at other parts of the circuit does not have the
same effect. For example, if an ESDS is placed on a metal tray that is resting on a resistive
workstation surface, a discharge between the ESDS device and the tray is not current limited
by the resistance of the workstation surface. It is determined by the much lower impedance
of the metal and spark. A risk of charged device ESD damage could arise. Similarly, adding
resistance in a ground wire grounding a low-resistance or metal work surface would not
give protection against high-current ESD (Wallash 2007).

4.7.5.3 Specification of a Minimum Material Resistance


Where these protective functions are important, it is common to specify a minimum resis-
tance for ESD-protective materials that contact ESDS devices. Examples of this are in work
surfaces and ESD-protective packaging materials. Where charged device or other similar
ESD risks are a concern, a minimum material resistance on the order of 10 kΩ may be spec-
ified.
110 4 The Seven Habits of a Highly Effective ESD Program

4.7.6 Safety Considerations


Where personnel are working with exposed continuous voltage sources (for example, pow-
ered system supplies) there may be a risk of shocks occurring if the person touches the
power supply. In this scenario, safety can be a significant concern if high-voltage supplies
are present. This can be a good reason to specify a minimum resistance acceptable in a
ground path. Such safety issues may be subject to local or national safety regulations. If
voltage sources are not present, it may be unnecessary to specify a minimum resistance in
a ground path.
Typically, the ESD control standards do not specify a minimum resistance for safety pur-
poses. This is because the standard is often concerned only with ESD control and regards
safety as a matter for user specification. They may discuss the topic in user guides associated
with the standard.

4.7.7 Elimination of ESD by Grounding and Equipotential Bonding


ESD occurs because two objects have a sufficiently high voltage difference between them
to cause breakdown of the air gap between them (if any) and a current to flow. It follows
that if we can keep voltage differences low, we get the following benefits:
● If there is insufficient or no voltage difference, ESD cannot occur.
● If ESD does occur, the energy, peak current, charge transferred, and other potentially
damaging parameters are reduced in level.
If we connect two conductors electrically and they are at different voltages, a current will
flow between them briefly until they are at the same voltage. At this point, current flow will
stop, providing there is no externally applied voltage difference or current flow. At the initial
time of contact, ESD has of course occurred, but thereafter no ESD can occur between them
while they are at the same voltage.
Connecting two conductors to make sure that they are at the same voltage is called equipo-
tential bonding. This is the main means of preventing conductors from becoming charged
and an ESD source. In practice, we often equipotential bond all conductors to the earth.
In this case, it is called grounding or earthing the conductors. This is particularly useful in
many EPAs because they may already contain equipment that has been earthed for other
reasons such as electrical safety.

4.7.8 Understanding the Grounding (Earth) System


4.7.8.1 Types of Ground
There are various ways in which EPA grounds may be implemented in practice. The main
types are
● Equipotential bonding
● Electrical safety earth
● Functional earth
It is often thought that it is necessary to have a connection to earth for the elimination
of voltage differences and ESD sources in an EPA to be successful. This is not so – we only
4.7 Habit 4: Ground Conductors, Especially People 111

need to have equipotential bonding of the conductors in the area. It would be perfectly
possible to have successful control of voltage differences between conductors by equipo-
tential bonding in an aircraft or other situation with no contact with earth. For this reason,
in modern ESD control standards, equipotential bonding is treated equally with grounding.
The term grounding is often used to include equipotential bonding as an alternative to other
grounding methods, meaning connecting an item to the designated ESD ground.
In many EPAs, mains electrical safety earth is present, and many types of equipment are
already connected to it for electrical safety. So, it often is most convenient to use mains
electrical safety earth as the EPA earth for ESD control purposes. All items of noninsulative
EPA materials and equipment are then electrically bonded to this.
In some facilities, mains electrical safety earth may not be available or for some reason it
may be undesirable to bond to this earth. A separate “functional” ground such as an earth
rod sunk into the ground may be used.
It is normally undesirable to have two different and separate grounds present in an EPA.
If this occurs, they may be at different voltages and could become a serious source of ESD
risk. All earths (grounds) in an EPA should be electrically bonded together to make sure
that no significant voltage differences can occur between them.

4.7.8.2 The Grounding System


Reliable grounding of an item requires that a continuous electrical connection is established
and maintained between the item and ground. This ground path may rely on several items
of equipment or materials. Examples are
● A person grounded through ESD footwear and an ESD control floor
● A cart, chair, or rack grounded through an ESD control floor
● A hand tool, grounded through a person’s gloved hand and body via a wrist strap
The requirements of each part of the system must be considered as part of the grounding
system for all the items that must be grounded. Often the grounding requirements of one
system may dictate the specification of a key part of the system. For example, the resistance
requirement for a floor may be specified mainly by the need to achieve a chosen maximum
resistance from a person’s body to ground through footwear and flooring.
For reliable grounding to be achieved, the performance of each part of the system must be
maintained under all the circumstances where grounding must be maintained. This means
that each part of the system, or the system as a whole, must be tested from time to time.
This is discussed further in Section 4.10.

4.7.9 Grounding Personnel Handling ESDS Devices


4.7.9.1 Basic Requirements for Grounding Personnel Handling ESDS Devices
People generate electrostatic charge continuously as they move around the environment.
This is because their feet or clothing make and separate contact with other materials and
surfaces as they move around. Many modern ordinary shoes have soles made of insulating
materials, and outer clothes may also be made of insulating materials. Charges generated
by contact with other materials can be conducted to the body or induce voltages on the body
that can lead to ESD.
112 4 The Seven Habits of a Highly Effective ESD Program

A basic requirement for grounding personnel is to control body voltage to make sure that
it does not reach a level where damaging ESD could occur when ESDS device are touched.
Current practice is to keep body voltage lower than the human body model (HBM) with-
stand of any ESDS devices that are handled. Current standards aim to safely handle 100 V
HBM devices, and so the body voltage is controlled to less than 100 V. When devices of lower
withstand voltage are handled, body voltage may need to be maintained below a lower level.
Considering the simple electrical model of Figure 4.3 for a person, the main way in which
body voltage can be controlled is by providing a ground path with sufficiently low resistance
from the person’s body to ground through an appropriate grounding system. The two main
grounding systems used are
● Wrist strap connected to the person’s body and to an earth bonding point provided for the
purpose
● ESD control footwear when standing on an appropriately specified ESD control floor
Other systems are occasionally used, with the principles remaining the same. The max-
imum resistance from body to ground is usually specified so that under all circumstances
expected, body voltage generated is kept below a specified level.

4.7.9.2 Grounding Personnel via a Wrist Strap


For wrist strap grounding, it was established many years ago that the body voltage was
maintained below 100 V if the resistance from body to ground was less than 35 MΩ. Many
standards have therefore adopted this as an upper limit for an ESD control program han-
dling devices down to 100 V HBM.
The wrist strap system in practice consists of a wrist band in contact with the skin, a
grounding cord, and an earth bonding point connected to EPA ground. For the system to
work correctly, all the parts must function reliably. Standards often specify resistance lim-
its for the individual parts of the system that may be tested separately (e.g. from hand to
wrists strap groundable point when wrist strap is worn, or resistance to ground of the earth
bonding point) (see Section 6.5.12).

4.7.9.3 Grounding Personnel via Footwear and Flooring


For many years, the 35 MΩ limit used for wrist straps was also accepted as a limit for resis-
tance to ground in grounding personnel via footwear and flooring. More recently, practice
has moved away from this for various reasons.
For a standing person, much of the charge generation leading to body voltage is due to
the contact between the footwear sole material and the floor material. Referring again to
Figure 4.3, we can see that the charge generation rates combined with resistance to ground
will have a strong effect on voltage buildup on the body. The triboelectric charge genera-
tion properties of the footwear-flooring combination have a major effect on body voltage,
and different footwear and flooring materials with identical resistance can give very dif-
ferent body voltage performance. It is possible to select a footwear-flooring combination
that gives body voltage within the required limits due to lower charge generation, despite
their achieving a higher resistance from body to ground. So, modern ESD programs often
allow this higher resistance level providing it has been demonstrated that the footwear and
flooring used achieve the required body voltage performance.
4.7 Habit 4: Ground Conductors, Especially People 113

It is important to realize that changing the footwear or flooring types for others of the
same resistance would not necessarily give the same body voltage performance as the
charge generation properties would be different. So, once a footwear-flooring combination
has been evaluated and selected, it is necessary to use only this combination, or another
that has likewise been shown to give the required performance. Unfortunately, the
resistance of the footwear and the flooring measured separately have been shown to be a
poor predictor of the footwear and flooring performance in combination (Smallwood et al.
2018).
If several floor types are to be used with the same footwear, then the performance of the
footwear must be evaluated with each type of floor. Similarly, if several types of footwear
are to be used, then each type of footwear must be evaluated in combination with the floors
with which they will be used. Quite different performance can be found with different floor
materials and at different atmospheric humidity (Figure 4.8). Most standard ESD control
programs require the maximum body voltage produced during a walk test to be less than
100 V. This may be reduced when handling components less than 100 V HBM withstand
voltage.

4.7.9.4 Grounding Seated Personnel


When seated, most people are likely to take their feet from the floor for some of the time.
When they do so, contact between footwear and flooring is broken and so grounding by this
method is unreliable. For this reason, most ESD control programs require seated personnel
to be grounded via wrist straps. ESD control seating is not usually regarded as a reliable
means of grounding personnel sitting on the seat (see Section 4.7.10.5).

4.7.10 Grounding ESD Control Equipment


4.7.10.1 General Considerations
The high resistance levels allowed for grounding conductors in ESD control may often sur-
prise engineers new to ESD work, especially electrical engineers. Achieving a resistance to
ground of around 1 GΩ, or in some cases even higher, may be sufficient to effectively ground
items for static control purposes. In practice, grounding may be provided by a material (e.g.
floor) rather than an installed permanent wire connection. As charge generation currents
are small (microamps or less), grounding wires do not have to be thick to withstand the
current. The reliability of the ground path is, however, an important consideration. This
reliability is often governed by factors such as the following:

● Robustness of the grounding system components


● Contamination of contacting surfaces (e.g. floor, footwear soles, grounding wheels on
chairs and carts)
● Human factors (e.g. deliberate or inadvertent unplugging of a grounding connector)

4.7.10.2 Work Surfaces


The purpose of providing a grounded static dissipative work surface is twofold. First, the
work surface material itself should not become charged and give electrostatic fields that
could lead to ESD risk. Second, the work surface provides a useful way of draining charge
114 4 The Seven Habits of a Highly Effective ESD Program

(a)

(b)

Figure 4.8 (a) Footwear 10 MΩ, “dissipative” floor 10 MΩ resistance to ground, 15% rh;
(b) footwear 10 MΩ, “conductive” floor 900 kΩ resistance to ground, 15% rh; (c) footwear 10 MΩ,
“standard vinyl tile” floor, 15% rh; (d) footwear 10 MΩ, “standard vinyl tile” floor, 50% rh. Body
voltage generated using ESD control footwear with different types of flooring. Source: D. E.
Swenson.

from any noninsulative material or object that is placed on it. This may include tools and
components including ESDS devices.
Any isolated (nongrounded) conductor that is placed on the work surface can be expected
to initially be at a different voltage. An ESD event will occur as the conductor approaches
4.7 Habit 4: Ground Conductors, Especially People 115

(c)

(d)

Figure 4.8 (Continued)

or touches the work surface. If the conductor is part of an ESDS, this gives a risk of charged
device ESD damage. For this reason, where the ESDS devices handled are susceptible to
charged device ESD damage, the work surface should be chosen to have a high surface resis-
tance to limit the peak ESD current. Where the ESDS devices handled are not susceptible
to charge device ESD, metal or low-resistivity surfaces may be used.
Work surfaces are usually either hardwired or connected via earth bonding plugs to the
EPA ground system. A work surface material with point-to-point resistance and resistance
from surface to ground less than 1 GΩ is normally considered adequate for ESD control in
116 4 The Seven Habits of a Highly Effective ESD Program

most EPAs. A minimum point-to-point resistance may be specified for safety or charged
device ESD damage prevention. (See Section 4.7.5.3)

4.7.10.3 Floors
ESD control floors are often provided to give a convenient way of grounding personnel as
well as carts, racks, chairs, and other free-standing equipment on the floor. The ESD control
function of the floor is often misunderstood – it must operate as a system with all the items
that it is intended to ground. So, in specifying the characteristics of a floor, it is necessary
to consider the items that will be grounded by it. Typical grounding systems using the floor
include the following:
● An operator’s body grounded through footwear and flooring
● A cart (trolley) grounded from its surface through the chassis, wheels, and floor
● A rack grounded from the shelf surface through the frame, feet, and floor
● A seat grounded from the seat surfaces through the frame, feet, or wheels and floor
The resistance to ground of the system includes the resistance of all the parts of the ground
path including the item being grounded, the floor, and the resistance of the contact between
them. So, it might be expected that if the resistance of the individual parts of the ground path
are measured, the total resistance to ground over the system would be the sum of the parts.
Unfortunately, this is not generally true, largely because the contact resistance between
the floor and the item being grounded can be higher or lower than expected and can vary
considerably. The contact resistance with the floor generally depends strongly on the area
and pressure of the contact. The contact areas and pressures between the item and the floor
can vary considerably. These pressures and areas are generally quite different from the pres-
sure and area of a measurement electrode standing on the floor. So, the contact resistance
between a shoe, wheel, or equipment foot and the floor are likely to be very different than
that of a measurement electrode. Contamination or coatings on the contacting surfaces (e.g.
dirt or polish) can also make a big difference to the effective contact resistance.
ESD control standards typically specify a maximum resistance from the floor surface to
ground of 1 GΩ. Many do not specify a lower limit, as there is no minimum resistance
requirement for ESD control purposes. In some applications, a minimum floor resistance
to ground may be desirable, e.g. for safety in the presence of high voltages. The minimum
body to ground resistance of the footwear and flooring system should be measured in qual-
ification tests with the person standing on the ESD control floor wearing the footwear with
which it will be used. This should be considered an essential part of the floor specification
process.
Similarly, to a first approximation, the resistance to ground from the surface of an item
grounded through feet or wheels through the floor can be assumed to be the sum of the
resistance from surface to feet (wheels) and the floor surface to ground. In practice, this is
often not correct for various reasons, including that the contact between feet (wheels) and
floor surface can have significant resistance that is difficult to predict.
Nevertheless, if the resistance to ground of an item that is grounded via a floor is required
to be below a certain value, the floor resistance to ground should also be below that value.
It is often best to select a lower resistance from floor surface to ground to give some margin
for added resistance from contact resistance. For example, if the resistance from body to
4.7 Habit 4: Ground Conductors, Especially People 117

ground of a person standing on a floor is required to be below 35 MΩ, the floor should be
selected to give installed resistance to ground less than 35 MΩ. In practice, selecting a target
value of, say, 10 MΩ gives some “headroom.”
It is perfectly possible to have an effective EPA that does not have an ESD control floor,
if one is not needed for grounding personnel or equipment. For example, a single worksta-
tion EPA in which a standing operator is grounded via a wrist strap may not need an ESD
control floor.
Two such workstations nearby in the same room would effectively be two separate EPAs
with uncontrolled space between them (Figure 4.9). An operator needing to transport an
ESDS from one workstation to the other would need to consider using ESD-protective pack-
aging to protect the ESDS. The operator would not be grounded when moving between the
workstations, and this would lead to ESD risk. In contrast, if an ESD floor is provided, link-
ing the two workstations, the operator can remain grounded when moving between the
workstations. The two workstations can now be part of the same EPA, and moving an ESDS
from one to the other does not take it through uncontrolled space. The ESD risk is greatly
reduced, and the handling of the ESDS is made more convenient.

4.7.10.4 Carts, Racks, and Other Floor Standing Work Surfaces


Carts, racks, and other floor standing equipment may be grounded through floor on which it
stands. Equipment that is not mobile may alternatively be connected to ground via snap-on
ground cords or permanent hardwired connections.
Carts and racks on which unprotected ESDS devices may be placed are usually considered
to be subject to the same point-to-point and resistance to ground requirement as work sur-
faces. Carts and racks that are to be grounded through the floor must usually be designed for
the purpose, as they must have a continuous connection from the shelf surfaces to ground
through the wheels or feet.
Equipment that is not designed for EPA use often includes insulating plastic joints or
other parts that may isolate shelves from the frame and supports. Wheels and feet will often
be made of insulating plastic material. In carts and racks designed for EPA use, the joints,
wheels, and feet are made of conducting materials such as conductive or dissipative plastics
or metal. It can be difficult to tell apart visually equipment or materials that are designed
for EPA use from similar item that are not.

Figure 4.9 (left) Two EPA workstations separated by a UPA, (right) joined as single EPA using ESD
control floor.
118 4 The Seven Habits of a Highly Effective ESD Program

Wheels and feet are often prone to contamination from an accumulation of dust and dirt.
This increases the resistance to ground over time and may ultimately cause it to exceed the
specified resistance requirements. Regular cleaning may be required to bring the equipment
back within specification.
Drag chains are sometimes used to ground carts, but these can be particularly prone to
dirt collection and can be unreliable.
An example of a carts is shown in Figures 7.9.

4.7.10.5 Seats
It is a common misconception that ESD control seats are intended to ground personnel
sitting in them. Although seats are sometimes used to do this, many are not designed for this
purpose. They are generally designed to prevent the seat from becoming highly charged and
an ESD risk (Figure 4.10). ESD control seats also reduce the charging of personnel sitting
on the seat.
Like carts and racks, seats designed for EPA use must have a continuous conducting path
from the various parts through the legs and feet or wheels to the floor.
Wheels and feet are often prone to contamination from an accumulation of dust and dirt.
This may increase the resistance to ground over time and may ultimately cause it to exceed
the specified resistance requirements. Regular cleaning may be required to bring the equip-
ment back within specification.

Figure 4.10 An ordinary seat can give


high electrostatic fields, in this case
showing 14 kV surface voltage.
4.7 Habit 4: Ground Conductors, Especially People 119

4.7.10.6 Tools
Ordinary tools often have insulating handles and metal bits or other parts that are electri-
cally isolated. These can reach a high voltage by tribocharging or induction and cause risk
of ESD to any ESDS devices that they may contact.
Hand tools designed for EPA use typically have noninsulative handles that electrically
connect the bit and any other metal parts to the user’s hand. Any charge developed on the
tool is dissipated safely to the user’s hand and via their body and personal grounding to
ground.
If a glove is worn, the glove must also be noninsulating to allow the tool to be grounded
through the glove to the operator’s hand (Figure 4.11).
Typically, the important characteristics for an ESD-protective tool include the resistivity
of the material that may contact the ESDS devices and the resistance to ground via the nor-
mal ground path. A surprisingly high resistance to ground may be tolerable. For example,
a tool that has capacitance 10 pF would have a charge decay time around one second if
grounded with a resistance to ground of 1011 Ω. This would in most cases be sufficient to
ensure insignificant ESD risk.
If low CDM ESD withstand voltage devices are to be handled, the resistivity of the mate-
rial of any part of a tool that may contact an ESDS may need to be sufficiently high to limit
ESD currents during any contact discharge (Wallash 2007).

Figure 4.11 A hand tool designed for EPA use is grounded via the user’s hand. If a glove is used, it
must also be noninsulative to maintain the ground path.
120 4 The Seven Habits of a Highly Effective ESD Program

Figure 4.12 Gloves and finger cots.

4.7.10.7 Gloves and Finger Cots


In some operations or processes, the operator may need to wear gloves or finger cots to pro-
tect their hands or the components or product while being handled (Figure 4.12). Gloves
are often required to protect the hands from chemicals, and oven gloves may be required
to protect from heat while unloading product from an oven. If these are made from insu-
lating materials, ESDS devices or tools held in the hand may risk becoming charged and a
source of ESD to themselves or other ESDS devices. So, it is often necessary to consider the
electrostatic properties of the gloves and finger cots used in these processes.
In some cases, gloves may be necessary to protect the operator for safety reasons, for
example from heat or chemicals or high voltages, rather than to protect the product. In
this case, the gloves are personal protective equipment (PPE). In Europe, use of PPE is sub-
ject to directives, and compliance with these is a legal requirement. Such safety and legal
requirement always take precedence over ESD control, although often with careful thought,
ways of working can be devised that meet legal requirements and minimize ESD risks.
When gloves are used while handling ESDS devices, the main ESD risks are
● The glove material may charge and give electrostatic fields very close to any ESDS device
handled.
● The glove may act as an insulating barrier to prevent grounding of any tools, objects, or
ESDS devices held in the hand.
● Touching an ESDS device with the gloved hand may result in the ESDS device becoming
charged by triboelectrification.
Different materials will typically give different performance regarding these concerns.
The risk of electrostatic fields can be eliminated, and grounding of handheld items can
be ensured by use of static dissipative materials in the glove. This, however, will not
4.7 Habit 4: Ground Conductors, Especially People 121

eliminate the possibility that an ESDS device may become charged when handled due to
tribocharging. The amount of charging of any item handled will depend on the choice
of material used in the outer surface of the glove that comes into contact with the items
handled.

4.7.10.8 ESD Control Garments


Many ordinary outer garment materials can easily become highly charged. These can be a
source of strong electrostatic fields that contribute to ESD risks. Many ESD programs use
ESD control garments to provide an outer clothing layer that does not give external electro-
static fields and acts to contain the electrostatic fields from clothing within (Figure 4.13).
It follows that these garments should cover all the clothing within. This is particularly
important for the sleeves and front of the garment as these are normally the areas closest
to the ESDS device. It also follows that coats should be fully fastened, as the underlying
clothing is exposed by unfastened areas.

Figure 4.13 (left) ESD control garments should cover the clothing of the arms and body, (right)
incorrectly worn.
122 4 The Seven Habits of a Highly Effective ESD Program

Some types of garments are designed to maintain contact with the wearer’s skin, usually
at the wrists. This can be an effective way of grounding the garment material via the wearer’s
body (which must always be grounded).
An ESD garment should not contain any exposed ungrounded conductors of any signifi-
cant size and capacitance that could possibly become charged and a source of ESD risk.

4.7.11 What If a Conductor Cannot Be Grounded?


Grounding a conductor is by far the best way of preventing it from becoming charged to a
voltage that might cause ESD risk. If this is for some reason impractical and the conductor
is likely to contact an ESDS, then it may be necessary to find some other way of reducing
any voltage arising.
One possibility is to use an ionizer to neutralize charge on the conductor. This can work
but suffers some disadvantages compared to grounding. Neutralization can be slow and
may not adequately control charge produced by triboelectrification or voltage changes
induced by fast-changing electrostatic fields nearby. Furthermore, the conductor will
become charged to the ionizer offset voltage. For very sensitive components, this residual
charge may be sufficient to cause ESD risk. The ionizer offset voltage will typically increase
as time goes on, unless the ionizer is properly maintained. These issues are discussed
further in Section 4.6.4.

4.8 Habit 5: Protect ESDS Using ESD Packaging


4.8.1 Don’t Take Ordinary Packaging Materials into an EPA
Ordinary packaging materials such as paper, cardboard, polythene bags, and bubble wrap
and polystyrene foam are designed for physical functions such as containment and physical
protection of items. They are often made from highly insulating materials such as plastic
or of materials that have unknown electrostatic charging and charge dissipation proper-
ties. These properties may vary widely between apparently similar materials (e.g. different
papers or cardboards) and may be highly dependent on atmospheric humidity conditions.
These materials are often static generators or can isolate other items from ground, or
their properties may at best be unknown and dependent on the weather and climate. In this
case, so will be the ESD risk arising from their presence near an unprotected ESDS device.
They are likely to have little or no ability to protect ESDS devices against ESD from external
sources such as charged personnel or objects. If present in an EPA, electrostatic control is
severely compromised. For this reason, ordinary packaging materials are best kept out of
an EPA, or at least well away from any workstation or process in which ESDS devices are
handled unprotected.
It follows that special ESD-protective packaging is needed for two reasons – first to pro-
tect ESDS from ESD risks when in an uncontrolled area and second to provide packaging
materials that can be used within an EPA without compromising electrostatic control or
introducing ESD risk.
4.8 Habit 5: Protect ESDS Using ESD Packaging 123

4.8.2 The Basic Functions of ESD Packaging


ESD packaging can have several ESD control functions. First, it may be required to protect
any ESDS device within from external electrostatic fields or direct ESD from sources in
an external uncontrolled environment. Second, it must not itself provide ESD risk to the
ESDS devices within. Third, it will be necessary for the packaging to go into an EPA, and it
must not cause ESD risk to any unprotected ESDS devices in the EPA. Last, the internal sur-
faces may be required to prevent discharging of on-board batteries on a circuit board within
the package. So, ESD packaging is typically designed with one or more of the following
functions:
● Inner and outer surfaces minimizing electrostatic charging of the surface and any ESDS
device in contact with it
● The surfaces dissipating electrostatic charge
● Surfaces in contact with ESDS devices having high enough resistance to minimize dis-
charging of on-board batteries
● Shielding against electrostatic fields
● A barrier to direct ESD
These functions, and ESD packaging, are discussed further in Chapter 8. ESD-protective
packaging materials can have one or more or all of these properties. The properties are
in some cases relatively independent and in some cases highly dependent on atmospheric
humidity.

4.8.3 Open ESD Packaging Only Within an EPA


At the beginning of this chapter we said that a key part of our ESD control strategy is that
in UPAs, ESDS are protected by enclosing them within ESD-protective packaging. Unpro-
tected ESDS should be handled only in an EPA. It follows that ESD-protective packaging
should never be opened in an UPA.
Typical processes in which this principle is important are the goods in and stores areas
and during transport. At goods in, shipments of components, assemblies, and materials
typically arrive packaged within ordinary packaging materials for physical protection
during transport. Within this ordinary packaging there may be an ESDS device within
ESD-protective packaging, or there may be some non-ESD-sensitive item.
ESD-protective packaging is normally marked on the outside with symbols or warnings
that it contains ESDS devices. At goods in, ordinary packaging may be stripped off only
until these markings identify ESD-protective packaging. Typical markings in current use
are discussed in Chapter 8.
ESD-protective packaging should never be opened in an UPA. If further opening of the
ESD-protective packaging is required to access the ESDS, e.g. for inspection, the package
must be taken into an EPA.
Ordinary packaging materials should not be taken into an EPA, as the act of removing
ordinary packaging materials is likely to charge it highly. This would then cause ESD risk
to any ESDS nearby.
In a stores area, non-ESDS items are often stored adjacent to ESDS items that are
protected within ESD-protective packaging. There is no problem in doing this, providing
124 4 The Seven Habits of a Highly Effective ESD Program

the ESD-protective packaging is not opened in an UPA. If it is necessary to open the


ESD-protective packaging to inspect, count, or remove components for a kit, this must be
done in an EPA.
Kits of parts are often made up for product assembly processes. Where a kit is made up
to go into an EPA, then every item that goes into the kit must be compatible with EPA
requirements, even though the assembly operation may be done in an UPA. If this is not
done, then the kit will introduce noncompliant items or materials and ESD risk into the
EPA. So, the tote box containing the kit must itself be of ESD packaging material. ESDS
devices will clearly be packaged in ESD-protective packaging. However, there may be many
non-ESDS parts, assemblies, or materials also included in the kit. If these must be organized
or protected in bags or other packaging, it must be ESD packaging suitable for EPA use. This
packaging does not need to have an ESD-protective function to its contents, but it must not
cause ESD risk to unprotected ESDS devices in the vicinity.

4.8.4 Don’t Put Papers or Other Unsuitable Material in a Package with an


ESDS Device
The materials of the inner surfaces of ESD packaging that contact ESDS devices have care-
fully designed and controlled properties to minimize ESD risks. In contrast, most grades of
paper have unknown and poorly controlled electrostatic properties and become insulative
at low humidity (see Chapter 8). It therefore is nonsensical to compromise the ESD risk
control of ESD-protective packaging by including papers in the package in contact with the
ESDS devices.

4.9 Habit 6: Train Personnel to Know How to Use ESD Control


Equipment and Procedures

4.9.1 Why Train People?


The greatest ESD risk in manual handling and assembly processes is often from ungrounded
personnel handling ESDS. Untrained personnel are likely to make actions that cause ESD
risk, such as opening ESD-protective packaging and handling the contents outside an EPA.
Untrained personnel are unlikely to be aware of the procedures and equipment to be used
inside an EPA and how to use them correctly. If aware of ESD risks and control measures,
they are likely to misunderstand and undervalue them. They may not know the differences
between ESD packaging, tools, and equipment compared to ordinary packaging, tools, and
equipment. Because of this, they are likely to bring noncompliant packaging, tools, and
equipment into an EPA.
Conversely, a trained and aware person can avoid these mistakes. Furthermore, they are
likely to recognize noncompliant packaging, tools, and equipment and can be trained to
remove them from areas where they cause ESD risk. They can be trained to perform simple
tests on essential equipment such as wrist straps and footwear and take corrective action
when they are found faulty.
It is probably no exaggeration to say that untrained personnel can be the greatest ESD
risk, and they can be converted into the first line of defense against ESD damage by effective
4.9 Habit 6: Train Personnel to Know How to Use ESD Control Equipment and Procedures 125

ESD training. So, ESD training can be an excellent investment. ESD training is discussed in
greater depth in Chapter 12. An overview is given here.
For those involved in implementing ESD control, training and education can elevate the
topic from “magic” to application of sound engineering principles. This helps them lay the
foundation of an effective and efficient ESD control program through understanding ESD
controls and systems.

4.9.2 Who Needs ESD Training?


The biggest ESD risk in manual handling and assembly operations is from charged
ungrounded personnel touching the ESDS. It follows that it is essential that personnel
working in an EPA reliably test and use their personal grounding equipment (wrist straps
and footwear). Furthermore, personnel working in the EPA can be trained to prevent
noncompliant materials and equipment from inadvertently being brought into the EPA and
to recognize and correct common ESD control problems and noncompliances. Carefully
trained personnel working in the EPA can, instead of being the greatest ESD risk, become
an important part of the first line of defense against ESD damage.
Personnel who enter an EPA need to have ESD training relevant to their activities, before
they enter the EPA. This can vary from minimal (e.g. for visitors) to extensive according
to the person’s role. As a minimum requirement, all personnel who handle ESDS must
have training in use of the ESD control equipment and materials they will use. It may be
beneficial for other personnel to have some ESD training, and some examples follow.
Managers may visit the EPA for a variety of reasons including showing visitors around.
They may also have budget responsibility for expenditure related to ESD control. They will
therefore need a working knowledge of ESD issues and the value of ESD control practice,
relevant to their responsibilities. They may need sufficient knowledge of use of ESD control
equipment (e.g. use of foot grounders, ESD garments) and EPA practices (such as to refrain
from touching ESDS) for themselves and for escorting visitors.
Audit and test personnel are likely to need in-depth knowledge of ESD test and audit
practice to test and audit the facility for compliance verification.
The ESD coordinator and other personnel responsible for the ESD program may need
continuing development and update of their knowledge and skills. This may include how
to specify, use, and evaluate ESD control materials and equipment, as well as comply with
standards.
Purchasing personnel may not need to enter a EPA, but they may need to source ESD
control equipment and materials as well as ESDS components for production requirements.
They may need an awareness of ESD control practice to support these responsibilities.
Some subcontractors may need to enter the EPA to fulfill their contracts. These will need
to have training in compliance with the company ESD control procedures. Other subcon-
tractors may supply or process ESDS product or components. They will need to understand
any ESD control requirements that are placed on them in handling the product or compo-
nents they handle.
If cleaners are required to clean within the EPA, they will need instruction on the cleaning
products, materials, and processes they are to use. They may also need instruction not to
126 4 The Seven Habits of a Highly Effective ESD Program

touch certain items or to avoid certain actions such as placing cleaning equipment in certain
places or unplugging ground points to plug in cleaning equipment.
Visitors may need a simple set of instructions as to what they may or may not do, as well
as instruction on using personal grounding equipment or garments.
Maintenance and facilities personnel may need specific instructions on how to undertake
maintenance and other work in EPAs without causing ESD risk. There may be contention
between ESD requirements and safety requirements (e.g. electrician’s use of insulating
footwear) that requires specific instructions to resolve. Safety requirements should always
take precedence over ESD requirements.

4.9.3 What Training Do They Need?


The appropriate training content for personnel from different job roles is likely to vary
according to their role and activities. This is discussed further in Chapter 12.
The training given should be clearly applicable to the situation in which the trainee
works. General commercially available ESD training materials can be useful but can
sometimes appear to be of little relevance if prepared for a situation very different from the
trainee’s experience and workplace.

4.9.4 Refresher Training


Refresher training is needed to reinforce understanding of ESD control practices and
remind them of aspects they may have forgotten. It may also be necessary to give updates
on changes in equipment, materials, or procedures used. These can occur due to changes
in production techniques or processes, standards, or organization practices.
Refresher training can be an opportunity to redress aspects of ESD control that have been
found to go wrong or be misunderstood in practice, e.g. incorrect use of personal grounding
equipment or garments, or lack of discrimination between compliant and noncompliant
packaging materials.

4.10 Habit 7: Check and Test to Make Sure Everything Is


Working

4.10.1 Why Do We Need to Check and Test?


There are two main reasons to check and test equipment and materials. Before approval for
use, they need to be specified and tested according to the specification to make sure that
they will work as intended for the duration of their intended working life. After installation
and commissioning, they need to be tested to ensure they continue to work as intended.
Maintenance of equipment is necessary for prevention of equipment failures and as part of
remedial actions.
Other checks may be needed of practice or procedure (e.g. correct wearing of wrist straps
or foot straps).
Compliance verification is further discussed in greater depth in Chapter 9. An overview
is given here. The test methods commonly used are discussed in Chapter 11.
4.10 Habit 7: Check and Test to Make Sure Everything Is Working 127

4.10.2 What Needs to Be Tested?


Everything on which effective ESD control depends should be tested in some way. Each item
or aspect to be tested needs to have defined a test methodology, pass criteria, a frequency of
testing, and procedures for recording, disseminating, and keeping results.
Many of these tests will be of the nature of a measurement made of equipment, material,
system, or installation characteristics. Examples of this are the resistance to ground from
the surface of a floor or bench top and the resistance from a person’s body to the end of a
wrist band cord while the wrist band is worn.

4.10.3 ESD Control Product Qualification


Any ESD control equipment should be tested before selection for use to ensure that it will
do the job that it is intended to do. Suitable tests methods must be found, and pass criteria
decided.
In practice, for common ESD control equipment, suitable tests and pass criteria are given
by the ESD control standards. ESD control product qualification can then be done using
data sheet information or tests according to these standards.
ESD control equipment product qualification tests are often done only once during the
process of selecting the product for use. Different tests are preferably done in a controlled
(usually dry) atmosphere to check operation under expected “worst-case” conditions. More
tests are often done to evaluate the equipment or material in some depth. The tests should
aim to evaluate the suitability of the equipment or material over its intended life. The equip-
ment or material is often tested in isolation, i.e. when not yet installed as part of the ESD
control system.
In contrast, compliance verification tests should be simple and efficient in demonstrating
the correct functioning of the ESD control equipment or material as part of the ESD control
system.

4.10.4 ESD Control Product or System Compliance Verification


Once ESD control equipment is in use, regular testing is necessary to identify failed equip-
ment or systems. Simple tests are usually used to test the operation of equipment in situ in
the workplace. Often this forms a “system test.” Examples include testing of the grounding
of a chair through flooring using a “resistance to ground” test or testing of a wrist strap or
ESD control footwear while worn by the operator.
These tests are normally done under normal operating atmospheric conditions in the
workplace.

4.10.5 Test Methods and Pass Criteria


For commonly used ESD control equipment and materials, standard test methods and pass
criteria are often provided in ESD control standards. These have the advantage of being
widely recognized and accepted. Adoption of these methods and criteria will facilitate easy
compliance with the relevant standard, if this is required.
128 4 The Seven Habits of a Highly Effective ESD Program

For nonstandardized items, test methods and criteria must be defined during the ESD
program planning process. Nonstandard test methods and pass criteria can also be used
for items that are covered by standard methods and criteria. Modern ESD control stan-
dards such as IEC 61340-5-1 and ANSI/ESD S20.20 allow nonstandard test methods to be
used, providing correlation of the results of the nonstandard test method with the standard
test method are demonstrated (see Chapter 6). In most cases, it may be easier to adopt the
standard test method.
Different test methods will usually give different results for various reasons. So, use of a
different test method will usually require changing the pass criteria. A specified pass cri-
terion should be applied only to results obtained using the test method with which it was
intended to be used or a test method that has been demonstrated to reliably give the same
results.

4.10.6 How Often Should ESD Control Items Be Tested?


The frequency of testing should be chosen to reflect the reliability and importance of the
ESD control item or system concerned. Factors that may be considered in deciding this
include

● The likely consequence and ESD risk to ESDS products associated with failure
● Likely permanence and ruggedness of the item
● Risk of performance change through contamination or other factors
● Failure history found by experience
● Risk of accidental human intervention (e.g. unplugging ground cord)
● Known item lifetime limitations

Some examples of how this might be applied follow:

● A wrist strap is essential for grounding a seated person who handles ESDS devices. The
consequences of failure would be that the person would be ungrounded and a serious ESD
risk to the ESDS handled. Because of this, wrist straps are normally tested each working
day before use. Some organizations that handle high-cost items or require low ESD failure
rates test even more frequently. Some organizations use an automated test system (e.g.
turnstile access control) to exclude personnel from the EPA unless they pass wrist strap
and ESD footwear tests.
● An ESD control floor that has been installed for some years in a clean area and has been
regularly tested and found to give stable performance might have the frequency of testing
reduced accordingly. A newly installed floor, or one in an area subject to dirt or other
contamination, might need to be tested frequently to monitor performance changes. This
would be especially important if it were used to ground personnel handling high-value
ESDS devices or where a low failure rate is required.
● A permanently installed workstation surface mat that has hardwired ground connections
and is in a clean process area might need relatively infrequent testing. A workstation mat
that is grounded by plug-in connectors to a mains electrical point might need more fre-
quent testing to ensure it has not been inadvertently disconnected. Similarly, workstation
surface in a process that has a high risk of contamination would need frequent testing.
4.12 Handling Very Sensitive Devices 129

● An ESD control chair grounded through wheels to the floor might need frequent testing
due to the risk of grounding failure due to contamination buildup on the wheels. This
would be especially true in a dusty area. In a clean room facility, a longer time between
tests could be adequate.

4.11 The Seven Habits and ESD Standards

It is not surprising that many of the aspects of ESD control described in this chapter form
the basis of modern ESD control standards. These have been crystallized over many years
in systems of requirements for ESD control equipment and procedures. The requirements
of two ESD control standards, IEC 61340-5-1 and ANSI/ESD S20.20, prevalent at the time
of writing, are discussed in Chapter 6.
The ESD control standards evolve over time, reflecting developments in ESD control
practice. When preparing for compliance with a standard, obtain copies and refer to the
standards for up-to-date compliance requirements.

4.12 Handling Very Sensitive Devices

At the time of writing, standard ESD control programs are designed to handle devices
with ESD withstand voltage down to 100 V HBM and 200 V CDM (IEC 61340-5-1:2016,
ANSI/ESD S20.20-2014). For devices down to about 500 V HBM and 250 V CDM (Industry
Council on ESD Target Levels 2011), it is relatively easy to implement an effective ESD
control program using standard ESD control measures alone. These basic ESD control
measures can be summarized as

● Handle unprotected ESDS devices only in an EPA.


● In uncontrolled areas, protect ESDS devices using ESD-protective packaging.
● Keep voltages on personnel less than 100 V using personal grounding.
● Keep conductors at the same voltage using grounding.
● Reduce electrostatic fields in the EPA to a minimum by removing nonessential insulators
and controlling the fields from essential insulators.
● Avoid contact between the ESDS and highly conductive materials.

As time goes on, more organizations are handling components and assemblies with ESD
withstand below this level and some below 100 V HBM. These very sensitive components
are sometimes known as “Class 0” devices. The 2016 ESD Roadmap (EOS/ESD Association
2016) predicts that the proportion of components in this range is likely to increase into
the future (see Chapter 13). Handling these components requires a greater level of care
and may require additional nonstandard ESD control measures. Multiple redundant control
measures may also be used to ensure that control is maintained if a single measure fails.
When handling low ESD withstand voltage devices, specific ESD risks may need to be
evaluated and addressed with specific control techniques. The technical requirements of
standard ESD control programs may need to be tightened.
130 4 The Seven Habits of a Highly Effective ESD Program

As an example, for handling 50 V HBM components, the body voltage developed on per-
sonnel should be limited to 50 V. In addition, it becomes more important that the body
voltage developed during normal activities is verified (by ESD control product qualification
and/or compliance verification). This means that the wrist band cords may need to have
specification of a reduced upper limit of resistance. It becomes even more important to
ensure that body voltage developed when grounded by ESD control footwear and flooring
remains below 50 V under all working conditions. For reduced CDM withstand voltage,
the electrostatic fields and voltages developed on insulators near the ESDS, and voltages
developed on the ESDS itself, may need to be controlled to a lower voltage level.
For handling low CDM ESD withstand devices, it becomes important that voltages devel-
oped on the device by induction, conduction, or triboelectrification remain low. At the same
time, the possibility of contact between the ESDS and highly conductive materials should be
eliminated where possible. Materials contacting the ESDS should, where possible, be static
dissipative (>10 kΩ surface resistance as defined in current ESD packaging standards; see
Section 8.5).

4.13 Controlling Other ESD Sources

Standard ESD control programs largely address the most common ESD sources, namely,
charged personnel, charged metal objects, and charge devices. Other ESD sources occur
that are not necessarily well controlled by standard ESD control measures.
PCBs, assemblies, and modules can become charged and then discharge when they touch
another conductor. The ESD susceptibility of these items is not usually tested. Some studies
have found that damage previously thought to be due to electrical overstress can result from
charged board ESD or charged cable ESD (Olney et al. 2003). Cables, PCBs, and assemblies
can have rather high capacitance due to their size and so can store relatively high energy
prior to discharge.
Charged cable ESD can arise when a cable is connected to a PCB, module, or assembly
at a different voltage. This can occur if the cable is charged or the ESDS is charged or both.
Cables can become charged by triboelectrification by contact with external surfaces (work
surfaces or equipment) or packaging. High voltages can be induced on cable conductors
by nearby electrostatic field sources. ESD can occur to connector pins when the cable is
connected to an ESDS. A similar discharge occurs if the ESDS is charged when connected
to the cable.
PCBs or assemblies are often potted in resin or enclosed within a plastic housing. Potted
modules sometimes have flying lead terminations for power and I/O connections. When
the housing becomes charged, high voltages can be induced on the PCB or assembly within.
Similarly, electrostatic fields from nearby charged insulators can induce high voltages on
the PCB or assembly. A discharge can occur when a cable is connected or a flying lead
touches a metal item.
References 131

Evaluation and control of these ESD risks often requires analysis of the circumstances
and evaluation of the possible ESD susceptibility of the ESDS. Simple specific ESD control
measures can often be chosen to address specific ESD risks.

References

EOS/ESD Association Inc. (2014) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016) ESD Association Electrostatic Discharge (ESD) Technology
roadmap – revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf [Accessed: 10th May 2017] Rome, NY, EOS/ESD
Association Inc.
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering component
level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf [Accessed: 10th May
2017] Industry Council on ESD Target Levels
International Electrotechnical Commission. (2013) PD/IEC TS 60079-32-1. Explosive
atmospheres Part 32-1. Electrostatic hazards, guidance. ISBN 978-2-8322-1055-0, Geneva, IEC.
International Electrotechnical Commission. (2015) IEC 61340-5-3:2015. Electrostatics - Part 5-3:
Protection of electronic devices from electrostatic phenomena - Properties and requirements
classification for packaging intended for electrostatic discharge sensitive devices. Geneva, IEC.
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
Olney, A., Gifford, B., Guravage, J., and Righter, A. (2003). EOS-25. Real-world charged board
model (CBM) failures. In: Proc. EOS/ESD Symp., 34–43. Rome, NY: EOS/ESD Association
Inc.
Oxford Dictionary. Available from: https://en.oxforddictionaries.com/definition/habit
[Accessed: 12th May 2017]
Smallwood, J., Swenson, D.E., and Viheriäkoski, T. (2018). Paper 1B.1. Relationship between
footwear resistance and personal grounding through footwear and flooring. In: Proc.
EOS/ESD Symp. EOS-40. Rome, NY: EOS/ESD Association Inc.
Stadler, W., Niemesheim, J., Seidl, S. et al. (2018). The risks of electric fields for ESD sensitive
devices. Paper 1B.4. In: Proc. EOS/ESD Symp. EOS-40. Rome, NY: EOS/ESD Association Inc.
Swenson, D.E. (2012). Electrical fields: What to worry about? Paper 3B.6. In: Proc. EOS/ESD
Symp. EOS-34. Rome, NY: EOS/ESD Association Inc.
132 4 The Seven Habits of a Highly Effective ESD Program

Wallash, A. (2007). A study of “Soft Grounding” of tools for ESD/EOS/EMI control. 2B8-1. In:
Proc. EOS/ESD Symposium EOS-07, 152–157. Rome, NY: EOS/ESD Association Inc.

Further Reading

EOS/ESD Association Inc. (2016) ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
International Electrotechnical Commission. (2018) IEC TR 61340-5-2. Electrostatics – Part 5-2:
Protection of Electronic Devices from Electrostatic Phenomena - User Guide. ISBN
978-2-8322-5445-5 Geneva, IEC.
133

Automated Systems

5.1 What Makes Automated Handling and Assembly


Different?
The American National Standards (ANSI)/Electrostatic Discharge (ESD) SP10.1 defines
automated handling equipment (AHE) as “any form of self-sequencing machinery that
manipulates or transports product in any form; e.g. wafers, packaged devices, paper,
textiles, etc.” One thing that makes automated assembly and handling different is that
the ESD risks are mainly due to contact with machines and machine parts or part-built
product assemblies. ESD risks due to manual handling are restricted to parts of the process
where manual handling is performed. Manual handling areas should be set up according
to the usual ESD control procedures.
Many people assume that they do not have to be concerned about ESD control within
the automated equipment, as they assume it has been taken care of by the machine
manufacturer. This can be incorrect for several reasons (Yan et al. 2009; Paasi 2004). First,
although an ESD-aware manufacturer may have built-in basic ESD control, they may not
have anticipated all ESD risks. Equipment manufacturers often include control measures,
especially ionizers, with little apparent understanding of their operation and effectiveness.
As an example, automated equipment moves quickly during operation. Ionizers used to
neutralize charge may not have sufficient time to achieve this (Tan 1993).
In some cases, different materials such as anodized coatings are used for ESD control
than in manual processes (see Section 5.6.3). These may have unusual or variable
characteristics.
Second, due to the lack of manual handling, the usefulness of human body model (HBM)
ESD withstand data is limited. Most ESD risks are instead due to contact between the
ESD-sensitive (ESDS) devices and metal or nonmetal conductors. So, primarily charged
device model (CDM) or sometimes machine model (MM) ESD withstand voltage data are
more likely to be relevant. (At the time of writing, measurement of MM ESD withstand
voltage of ESDS is falling out of practice.) The ESD withstand voltages of components are
reducing year by year, so yesterday’s ESD control may not be sufficient for today’s devices
and are unlikely to be sufficient for tomorrow’s components (ESDA 2016a; Koh et al. 2013).

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
134 5 Automated Systems

So, if the machines used are not completely up to date, there is a chance that the ESD
control may not be up to the standard required for handling the most sensitive of current
components.
The ESDS devices of concern may be printed circuit boards (PCBs) as well as devices.
Charged board ESD has been increasingly recognized as a significant source of damage. In
recent years, it has been realized that damage mistaken for electrical overstress (EOS) has
actually been caused by charged board ESD (Olney et al. 2003).
Third, adaption or modification of machines or processes can easily inadvertently intro-
duce ESD risks. As in manual processes, failures of ESD control equipment can also result
in ESD risks. Standards are not yet available for ESD control in automated equipment,
although the ANSI/ESD SP10.1 gives some guidance on standard practices. In an auto-
mate process, automated ESD generators can arise that can provide automated damage of
devices! If this occurs, the ESDS damage rate can be high.
The automated environment also brings some special challenges. Many components
designed for automated handling are extremely small. The packaging for these has
correspondingly small features, and this can make it difficult or impossible to test the
packaging with standard test methods. Some components are so small that electrostatic
charging can result in ejection of the component from the packaging in an uncontrolled
manner by electrostatic forces (see Section 5.8).
Access into automated processes during operation may be highly restricted. So, it
can be difficult to observe and diagnose potential ESD issues and make measurements
during operation of the process. Special test and measurement methods may be required,
either due to the form of the items measured or due to the restrictions of the measure-
ment environment in AHE. Stepped process simulation can help enable measurements
to be made.

5.2 Conductive, Static Dissipative, and Insulative Materials


It is difficult to write this chapter without making reference to materials categorized by their
resistance as conductive, static dissipative, or insulative. It is important to realize that there
is in general no fixed definition of these terms. In the case of the International Electrotechni-
cal Commission (IEC) 61340-5-1 and ANSI/ESD S20.20, these terms are specifically defined
in terms of surface and volume resistance ranges for ESD-protective packaging materials
(see Chapter 8). These definitions are used also in this chapter.
In this discussion, a conductor may be any item made from conductive or static dissipative
materials – in other words, any material that is not an insulator.

5.3 Safety and AHE


Working with AHE may bring specific safety issues relating to the equipment and processes.
These may include risks due to moving machinery, electrical hazards, or high process
or equipment temperatures. These must always be approached with due consideration
of any hazards arising in the process, as well as any local regulations and practices that
may apply.
5.4 Understanding the ESD Sources and Risks 135

AHE systems often work within interlocked protective enclosures. Making observations
or measurements on operating systems can be difficult.

5.4 Understanding the ESD Sources and Risks


The main ESD sources in an automated production environment include
● Charged devices making contact with low-resistance conductors.
● Charged metal objects and machine parts making contact with ESDS.
● Charged PCB (or subassemblies or modules) making contact with low-resistance
conductors.
● Charged personnel (in manual handling parts of the process) making contact with ESDS.
● For a few types of ESDS, electrostatic fields alone could give ESD risk. This is likely to be
a concern particularly with voltage-sensitive devices in high impedance circuits and may
be unusual.
In principle, the ESD risk could be understood by knowing the susceptibility of each type
of ESDS to each ESD source. The risk could then be controlled by maintaining the param-
eters of ESD from each source below thresholds at which ESD damage are known to occur.
To this end, HBM, MM, and CDM ESD susceptibility tests have been developed to allow
reproducible measurement of the ESD susceptibility of components (see Chapter 3). The
component ESD susceptibility is quoted as an ESD withstand voltage, which is the highest
test voltage in the ESD source that did not result in damage to the component.
At first sight then, a strategy for risk evaluation and management might seem clear – if
the real-world voltages in ESD sources can be kept below the withstand voltage levels, no
ESD damage should be possible. Specification of ESD control measures would then be a
matter of defining controls that can maintain the source voltages below the risk threshold
defined by component ESD withstand voltages.
While this simple view and strategy is partially valid, it represents a significant oversim-
plification (see Section 9.3.1). First, component susceptibilities are not in fact normally
damaged directly by the ESD source voltage, but to other parameters such as the peak
discharge current or charge power or energy transferred to the device in the discharge.
These parameters are related to the source voltage by circuit parameters such as induc-
tance, resistance, and source capacitance. Real-world sources are unlikely to have the same
capacitance, resistance, and inductance as the HBM, MM, and CDM models. This means
that HBM, MM, and CDM discharges do not happen in the real world – they happen only
in the component susceptibility test equipment. Real-world ESD is simply a variety of dis-
charges from the human body, a metal part, a device, or other source. Each source has very
different ESD waveform characteristics from the models and would yield a different with-
stand voltage (ESDA 2015; Gaertner and Stadler 2012). An ESDS might be more or less
susceptible to damage from real sources than it is to HBM, MM, or CDM ESD. A simple
action such as changing the position of an ESDS can change a relevant parameter such as
its capacitance as well as the voltage and hence the ESD damage susceptibility. In specifi-
cation of ESD controls and evaluation of ESD risks, it remains a major challenge to relate
real-world risks to component ESD withstand data.
136 5 Automated Systems

5.5 A Strategy for ESD Control


5.5.1 General Principles of ESD Control in AHE
The general principles of ESD control in AHE are not very different from in manual
processes. Many ESD problems can be avoided by applying the recommendations of
standards and textbooks, although in practice these may need some modification for use
with AHE.

5.5.2 The Conditions Leading to ESD Damage


Review of the ESD sources shows that they have some simple commonality.

● ESD can usually occur only where contact occurs between an ESDS device and a conduc-
tor. The only exception is for ESDS devices where electrostatic fields could cause voltage
differences that can cause damage within the ESDS device.
● ESD can occur only where there is a sufficient voltage difference between the ESDS
device and the contacting conductor. The ESD risk increases with the voltage difference.
The voltage difference that is “sufficient” is dependent on the circumstances and the
component sensitivity.

The first point focuses our attention on parts of a process where an ESDS device makes
contact with other conductors. These could be personnel, machine parts, tools, or other
ESDS devices. Sufficiently far away from the vicinity of the ESDS device, where the risk of
contact and the influence of electrostatic fields on it are insignificant, we do not usually
need to take ESD control measures. It also makes clear that if we can reduce the number of
occasions on which the ESDS makes contact with a conductor, we will reduce the number
of opportunities for ESD to occur. This can in some circumstances reduce ESD risk and the
burden of controlling it.
The second point highlights a need to control voltages on the ESDS device and on
conductors that could contact the ESDS. The voltage difference between the ESDS device
and any conductor that contacts it must be minimized, or at least kept below a risk
threshold level. These voltage differences can be from two causes – triboelectrification
(contact charging) and induction (voltages induced by nearby electrostatic fields)
(see Chapter 2).

5.5.3 Strategies for ESD Control in Automated Equipment


From the previous analysis, we can see that we need concern ourselves only with the imme-
diate region of the process in which an unprotected ESDS device can be present and the near
vicinity of the ESDS device. So, the first task is to plot the path of ESD through the process.
ESD control must be applied in a region around this critical path (Paasi 2004). Jacob et al.
(2012) refer to this as the zone of processing, identifying this as the region in which ESDS
devices are loaded, transported, or processed. SP10.1 recommends that in a critical path
region within 15 cm of the ESDS device’s critical path, all conductive machine parts should
be grounded, and all insulators rendered static safe.
5.5 A Strategy for ESD Control 137

Within the critical region, ESD risks should be analyzed using visual observations,
measurements, statistical failure data, or any other available relevant techniques and
information. Critical parts of the process are where ESDS devices make contact with other
items, especially when under the influence of electrostatic fields. Where contact is made,
voltage differences between the ESDS and the contacting item must be controlled. Ideally
ESDS devices should make contact only with static dissipative (see Section 5.2) materials.
Chapter 9 looks in more detail at evaluation of ESD risks.

● Conductors that make contact with ESDS devices must be grounded, where possible.
● Conductors that cannot make contact with ESDS devices need not be grounded.
● If a conductor makes contact with ESDS devices and it cannot be grounded, the ESD risk
must be evaluated. Other ESD controls must be devised if necessary.

Once the ESD risk points have been identified, they can be prioritized and suitable ESD
control measures developed. The control measures will need to be documented and a com-
pliance verification program developed. This may require test methods and equipment to
be defined with suitable pass criteria and frequency of testing.
Paasi (2004) suggested it is useful to think in terms of “intimate” and “proximity” mate-
rials in AHE systems, as defined in ESD-protective packaging materials. From this view,
intimate parts, materials, and surfaces are those that could come into direct contact with
ESDS devices or PCBs during normal operation. Intimate parts of AHE include conveyor
belts and rollers, supports, racks, and other parts that come into direct contact with the
ESDS device. Tape and reel packaging is also of concern. Materials that come into contact
with ESDS devices should be selected to minimize tribocharging of the ESDS device being
handled.
Proximity parts, materials, and surfaces are in the critical region surrounding intimate
parts, containing objects, materials, and surfaces that do not normally come into contact
with the ESDS device. By analogy with packaging, intimate and proximity items should, if
possible, not be made of insulating materials and should be grounded. This is to prevent
electrostatic fields due to charged materials inducing voltages on ESDS devices within the
critical path.
The materials, techniques, and equipment used to achieve these objectives are discussed
in greater detail in Section 5.7.

5.5.4 Qualification of ESD Control Measures


Once ESD risk points have been identified and quantified through measurements, suit-
able ESD control measures should be defined. The effectiveness of these control measures
should then be qualified and verified using measurements.
ESD control measures in AHE will often need to be developed specifically for the ESD risk
that has been identified. So, the qualification process will be part of the process of specifying
the control measure and checking that it works. Test methods, and pass criteria specific to
the control measures, will need to be selected to demonstrate that the control measures
work as intended. Sometimes these can be based on existing standard tests, but sometimes
nonstandard tests will need to be defined.
138 5 Automated Systems

5.5.5 Compliance Verification of ESD Control Measures


Most ESD control measures can fail for various reasons, so it is necessary to verify them on
a regular basis. As with ESD control measures in manual processes, suitable test methods,
pass criteria, and a frequency of testing should be defined and documented in a compliance
verification program.

5.5.6 ESD Training Implications


While participation of personnel in automated processes is minimized, the need for training
of different types should be considered.
Personnel who evaluate the AHE for ESD risks and develop ESD control measures will
need an excellent level of expertise to successfully provide these functions. Standards and
guidance documents are also regularly published or updated, and new technologies are
developed. Understanding of ESD risks is continually improving, and the susceptibility of
ESDS to these risks changes with technology development. ESD coordinators and personnel
who are active in these areas will therefore need provision for developing and updating their
skills and knowledge.
ESD control measures incorporated into AHE will often need special tests to be defined
that may need specific setup and procedures for each type of equipment. Personnel who
perform audit measurements will therefore need training in these test and measurement
techniques.

5.5.7 Modification of Existing AHE


ESD risks can often be created inadvertently when equipment is modified (Paasi 2004).
Analysis of ESD risks should be made as part of the process of design of any modifications so
that ESD control measures can be specified and built into the modifications. Qualification
and compliance verification tests may be needed.

5.6 Determination and Implementation of ESD Control


Measures in AHE
5.6.1 Define the Critical Path of ESDS
The first step is to define the critical path (or paths) that ESDS takes through the AHE.
These may include introduction of part-populated assemblies into the process as well as
parts of the process where ESDS devices are loaded for mounting on an assembly. The size
of the critical region around that path should then be decided. Electrostatic Discharge Asso-
ciation (ESDA) SP10.1 recommends a region of 15 cm around the critical path, but if it is
feasible, a larger region up to 30 cm around the critical path could be considered for control
of insulators. Further away from this, it is unlikely that significant risks are posed unless
items can become very highly charged, as the electrostatic fields are much reduced over the
distance (see Section 4.5.3).
5.6 Determination and Implementation of ESD Control Measures in AHE 139

5.6.2 Examine the Critical Path and Identify ESD Risks


The critical path should then be examined visually from the point(s) at which ESDS enter
the process to the point(s) at which they leave the process. It is helpful to observe the ESDS
traverse the critical paths. The objective is to identify ESD risks in terms of
● Points at which ESDS contact conductors including machine parts or personnel
● Places where potentially charged insulators are within the critical region
These points must be evaluated for ESD risk and, if necessary, addressed with ESD control
measures.
One of the main ESD risks in AHE is charge device or charged board ESD. These can
occur where a charged device or PCB makes contact with a low-resistance conductor or a
charged ungrounded conductor makes contact with an ESDS device. There are four princi-
pal strategies for controlling these risks.
● Keep tribocharging of devices and PCBs below a risk threshold level by controlling con-
tact between ESDS and other materials.
● Keep induced voltages on devices and PCB conductors below a risk threshold level by
controlling exposure of ESDS to electrostatic fields.
● Reduce the potentially damaging effect of ESD by preventing contact between ESDS
devices and low-resistance conductors.
● Ensure that ESDS devices where possible make contact only with grounded conductors.
If the ESDS device makes contact with a conductor that cannot be grounded, the voltage
difference between the ESDS device and the conductor should be controlled.

5.6.3 Determine Appropriate ESD Control Measures


5.6.3.1 General Control Measures
The following guidelines for typical ESD control measures are based on recommendations
of ESD Association SP10.1 (ESDA 2016b) and should be applied within the critical region.
The definitions of conductive and static dissipative used for packaging materials are used
here (see Section 8.3.1, 8.8.1 and Table 8.3);
● All conductors should be grounded.
● Where operators may need to be grounded for participation in the process, designated
wrist strap grounding points should be provided.
● All machine parts that may make contact with ESDS leads should have static dissipative
surface and be grounded to reduce charged device ESD risk.
● All machine parts that are separated from the machine chassis by bearings should be
grounded by some means that remains reliable during movement (SP10.1 recommends
resistance to ground <1 MΩ).
● Surfaces that ESDS could be placed upon should meet the requirements for EPA work-
station surfaces in the ESD program.
● Pneumatic and electrical lines should be constrained to avoid rubbing and tribocharging
effects. Pneumatic lines within the critical region, if possible, should not be insulative and
should be grounded. Alternatively, they could be shielded with a grounded metal braid
shield.
140 5 Automated Systems

● Wire bundles in the critical region should be shielded, e.g. with a grounded metal braid.
● Device pickup mechanisms such as vacuum cups, nozzles, and grippers should be made
of conducting materials and should be grounded. The contact area and velocity should
be minimized to reduce tribocharging of device packages.
● ESD grounding points should be directly connected to the equipment ground. SP10.1
recommends a resistance to ground of 1 Ω or less. Lower resistance might be necessary if
high currents (e.g. for motors) or fault currents might be carried.
● Conductors that are relied upon to provide a ground path must be sufficiently robust to
prevent accidental disconnection. The conductors should be braided wire where possible.
● For anodized surfaces, ensure the underlying substrate is grounded to the machine earth.

Methods of grounding moving parts suggested by SP10.1 include flexible conductors such as
braided cables, metal bushes, graphite or beryllium copper commutators, and conductive
greases within the bearings. One problem with moving parts is that resistance to ground
measurements made when stationary may not represent the resistance to ground when
moving. Oil films and other effects can result in intermittent loss of contact that leads to
charge/discharge behavior.

5.6.4 Include ESD Control in New Equipment Specification


The ESD control development process should preferably start during equipment design and
development (Paasi 2004). The equipment vendor should provide documentation of any
necessary ESD control measurements and acceptance values.
ESD control should be included as part of any new automated equipment purchase spec-
ifications. It may be inadequate to assume that the equipment manufacturer has built in
adequate ESD control (Yan et al. 2009; Tan 1993; Millar and Smallwood 2010), especially if
low ESD withstand voltage devices are handled.
Whenever equipment is modified, the modified equipment should be carefully evaluated
to make sure that ESD risks have not been inadvertently introduced. ESD control measures
should be qualified using tests appropriate to their function and the intended conditions of
operation.

5.6.5 Document the ESD Control Measures Used in the Machine


Once ESD control measures have been identified and specified, these should be docu-
mented. This is analogous to writing an ESD Control Program Plan for a manual handling
or assembly process. If the ESD control measures are not adequately documented, it’s likely
that they will be omitted, fail, or be inadvertently rendered ineffective by some means in
the future, reintroducing undetected ESD risk.

5.6.6 Implement Maintenance and Compliance Verification of ESD Control


Measures
As in ESD control in manual handling processes, ESD control measures in AHE can
fail or be compromised for many reasons. As examples, Millar and Smallwood (2010)
5.7 Materials, Techniques, and Equipment Used for ESD Control in AHE 141

found that all but one of eight YAC handler systems they examined had multiple failures
of internal ground wires. They found that anodized soak boats suffered wear of the
anodizing that allowed metal-to-metal contact between ESDS devices and soak boat
substrate metal.
As with manual processes, all ESD control measures must be checked at regular intervals
to detect any failures or damage that could cause ESD risk. This means that an inspection
regime must be defined, with checks and tests performed at appropriate intervals designed
to detect the failures. In some cases, it may be necessary to implement maintenance proce-
dures to prevent failures.
Tests may be required using standard or custom measurements, with suitably chosen pass
criteria. This compliance verification test program must be documented and established as
part of the ESD control procedures.

5.7 Materials, Techniques, and Equipment Used for ESD


Control in AHE
5.7.1 Grounding All Conductors That Make Contact with ESDS
As in manual handling of ESDS devices, all conductors that make contact with ESDS
devices should be grounded if at all possible. Allowing contact between ESDS devices and
an isolated (ungrounded) conductor should be a last resort adopted only where grounding
is very difficult. The ground path must be continuous at all times during operation when
ESDS devices might make contact with the conductor. A break in the ground path at any
time can allow charge to build up on the conductor, which then becomes a potential source
of ESD.
The conductor may be a conductive or static dissipative material (see Section 5.2), and
the resistance to ground that can be achieved may depend on the materials used. Items
that can be subject to charge generation due to movement (e.g. conveyor belts) should have
sufficiently low resistance to ground (at all points) to prevent significant voltage buildup
on them. The voltage buildup and resistance to ground should if possible be verified under
operating conditions.
Grounding can, of course, be achieved either through a wired connection or through an
electrically conducting material path. The ground path may include moving items such as
bearings or sliding surfaces, providing the ground connection can be shown to be reliable
under operating conditions.
In AHE, ground wires may often be used to link two parts that repeatedly move relative
to each other. In this situation, ensuring the reliability of the connection can be a challenge.
Ordinary wires may have short lifetime due to metal fatigue when repeatedly flexed. Braided
wires may give a longer lifetime. Regular testing of the connection may be required to detect
any failures occurring.
A maximum resistance to ground is normally specified for grounding of conductors.
This gives a simple measurable parameter for verification of grounding. A maximum
resistance to ground can be specified from two views. If the conductive item is subject
to charging from a current source, e.g. due to machine motion, the maximum resistance
142 5 Automated Systems

to ground can be specified from the view of maintaining low voltage under condition
of maximum charge generation current flow. For circumstances where current flow is
low, maximum resistance to ground can be derived from charge and voltage decay time
considerations.
In some cases, a minimum resistance to ground may also be specified for safe limitation of
current under a fault condition, or to limit ESD current when charged ESDS devices contact
the grounded material. For safety considerations, adding discrete resistance on the ground
path may be acceptable. For reduction of charged device ESD current, the resistance must
be inherent in the contacting material (see Section 4.7.5.2 and 5.7.5).

5.7.2 Isolated Conductors


Conductors in the critical region should always be grounded if possible. In AHE, it can
sometimes be difficult to achieve grounding reliably due to moving parts preventing reli-
able connections. Ungrounded conductors can be evaluated and dealt with as described in
Section 9.3.3.

5.7.3 Preventing Induced Voltages on ESDS Devices


Electrostatic fields in the vicinity of ESDS devices can induce voltages on the ESDS devices.
If these voltages are present when the ESDS devices makes contact with another conduc-
tor, then an ESD will occur. The concern here is charged device or charged board ESD. It
is therefore particularly important to control electrostatic fields in the critical region near
positions where the ESDS devices make contact with other conductors. Low ESD with-
stand voltage devices will require voltage sources to be controlled below a lower voltage
or field limit. The closer an electrostatic field source is to the ESDS devices at the con-
tact zone, the lower the field or source voltage that can be tolerated (see Section 4.5.2).
The sources of electrostatic fields that are often of most concern are charged insulators.
Nevertheless, any significant voltages sources could be of concern, particularly if close to
the contact zone. These could include fields from nearby wiring, pipes, or other machine
components.
As with essential insulators in manual processes, the first step in treating an essential
insulator within the ESDS critical path is to evaluate the ESD risk. If no contact is made
between ESDS and other conductors in the vicinity of the insulator, there may be little ESD
risk even if the insulator becomes significantly charged (Gaertner 2007; Yan et al. 2009).
Nevertheless, it may be preferred to reduce ESD risk by reducing the electrostatic field expe-
rienced by the ESDS device.
If an insulator can become charged but is not near a part of the process where contact
between the ESDS device and a conductor can occur, then no control is required unless the
ESDS device is known to be susceptible to electrostatic fields. In practice, few ESDS devices
may have this type of susceptibility. If the insulator cannot become charged, then no action
is needed, except to verify that the insulator really cannot become charged under the full
range of operating conditions. This may require making measurements under low-humidity
conditions.
5.7 Materials, Techniques, and Equipment Used for ESD Control in AHE 143

Where an insulator can become charged and could make contact with a conductor, an
ESD control measure is likely to be essential. Control techniques could include, for example
(Paasi 2004, SP.10), the following:
● Replacement or coating of the insulator with a conducting material, and grounding it
● Prevention of charging by some means (e.g. prevention of rubbing)
● Relocation outside the contact zone
● Shielding or suppression of the field from the insulator by positioning a grounded con-
ductor between it and the ESDS device
● Use of an ionizer to neutralize charge
● Use of topical antistats to reduce charging
Replacing the insulator with a grounded conductor, moving it outside the critical contact
zone, or preventing charging can be among the most effective solutions, if they are possi-
ble. Use of an ionizer is discussed in Section 5.7.9. Prevention of insulating surfaces from
rubbing can be a useful low-cost way of preventing an insulator from becoming charged.
This may require adjustment of the position of adjacent parts of equipment. It can be a par-
ticularly useful technique with cables, pneumatic lines, or other flexible insulating lines.
Perhaps the easiest but least reliable technique is to use a topical antistat to reduce electro-
static charging. These often effect a temporary solution but can quickly become ineffective.
Tan (1993) used antistats to confirm that electrostatic charging of insulating spacers was the
cause of some charged device ESD damage. He found that the lifetime of the antistat was
much shorter in a varying (hot and cold) temperature environment than the accelerated
life test had suggested, lasting only one day in practice. Often antistats rely on atmospheric
humidity for their action, and so they can become ineffective under low-humidity con-
ditions. In an enclosed space within AHE with heat sources present, local low-humidity
regions can easily be established.
The effect of electrostatic field sources can also be reduced or eliminated by shielding the
ESDS device from the field. This may be done by inserting a grounded conductor between
the field source and the ESDS critical path. Tan (1993) gives several examples of using metal
flanges, washers, copper tubes, or self-adhesive metal tapes for shielding purposes.

5.7.4 Reducing Tribocharging of ESDS Devices


It is often thought that an ESDS will not be charged by contact with a conducting material.
This is not so – any contact between materials will separate electrostatic charge to some
extent. Each of the contacting materials becomes charged by an equal amount and oppo-
site polarity. If one of the surfaces is an insulating material (e.g. device plastic package), it
will become charged. An insulating device package can be charged by contact with a con-
ducting packaging material or a vacuum cup (Yan et al. 2009). Use of conducting materials
for these contacting items merely allows the charge produced on them to dissipate, pro-
viding of course there is a route for the charge to be conducted away, i.e. the conducting
material is grounded.
Tan (1993) gives an example of damage caused by tribocharging of a plastic device pack-
age body due to contact with a steel tool in a trim and form process. Even with the tools
grounded, voltages up to 1100 V were measured on the devices.
144 5 Automated Systems

It can be difficult to prevent charging by contact in this way. Choice of contacting mate-
rials can make a difference to the charge generation levels. Often the charge accumulated
on the ESDS must be neutralized to bring it to an acceptable level before contact is made
with another conductor.
Kim et al. (2012) found that triboelectrically generated charge on liquid crystal display
(LCD) screens during production could not be adequately controlled by ionization (see
Section 5.7.9). The glass panel substrate generated high charge levels and held charge for
long periods. More than 40 material contact and separation actions occurred during the
production process, including friction from photoresist coatings applied to the glass sur-
face, deionized water spray rinses used to clean the surface, contact with rollers or belts,
and pressure and separation of glass panels from vacuum chucks. Control measures they
used to reduce ESD risks included increasing surface roughness and minimizing separa-
tion speed and vacuum pressure to control the charge generation of glass substrates. They
found that using insulative instead of dissipative or conductive lift pin materials prevented
ESD events. Specification of adequate separation distance and insulation thickness was also
used to prevent air discharge between the glass and metal objects.

5.7.5 Using Resistive Contact Materials to Limit Charged Device ESD Current
Charged device ESD risk is greatest when the charged device makes contact with a low resis-
tance conductor. This results in a high peak current discharge (so called hard discharge).
One means of reducing this risk is to ensure that the device contacts a resistive material,
rather than a highly conductive material. In this case, the peak discharge current can be lim-
ited by including sufficient contacting material resistance. Placing a resistor in the ground
path is ineffective for this purpose (see Section 4.7.5.2).
There is some debate as to whether direct current (DC) resistance measurements on poly-
meric ESD control materials provide a realistic guide to ESD risk (Viheriäkoski et al. 2017).
They found the “ESD resistance” due to frequency dependent impedance of the materials
could give much lower ESD current and charged device ESD risk than would be expected
from DC resistance measurements.

5.7.6 Anodization
Anodization is often used as a passivation and protection of machine surfaces. Although
considered insulative by some, it can be used to provide a noninsulating surface coating
on an aluminum alloy substrate (Bellmore 2001). The anodized layer may be 5–40 μm
(0.0002–0.0015 in.) thick. Different anodizing processes can produce a wide range of
surface resistance from <1 MΩ to over 100 GΩ.
If done after screw holds and threads are made, anodization can prevent good contact
between metallic machine parts (Yan et al. 2009). The resistance through the joints may
be increased to several ohms. Thick anodization can result in higher resistances. It can be
difficult to remedy this after the machine has been constructed.
Anodization can be used to provide a hard-wearing static dissipative surface coating to
reduce the risk of charged device ESD damage when device leads touch the surface (Yan
et al. 2009; Smallwood and Millar 2010; Millar and Smallwood 2010).
5.7 Materials, Techniques, and Equipment Used for ESD Control in AHE 145

Millar and Smallwood (2010) found that anodized soak boats experienced wear or dam-
age to the anodized layer. Direct contact could then be occasionally made between the leads
of devices in the boats and the underlying metal soak boat material. Smallwood and Millar
(2010) measured anodized soak boat surface resistance using standard and nonstandard
electrodes as well as resistance from the boat metal substrate through the surface and a
supporting dissipative rail to ground.
The influence of applied test voltage was also measured. They found that the resistance
of the anodized layer varied dramatically with the applied voltage and test electrodes used.
At 1000 V the resistance through the anodized layer was <1 Ω, but at 100 V this increased
to over 20 GΩ. Surface resistance from point to point was <1 GΩ using ESD S4.1 electrodes,
but 30–96 GΩ using an S11.13 2-pin measurement. Other methods also gave varying results.
As variation in resistance could be expected to cause variations in charge decay time of
the boat standing on the rail, this was also measured. The charge decay time varied with
the remaining voltage on the boat, increasing as the voltage reduced. At 1000 V, the voltage
disappeared immediately when the applied voltage source was removed. Decay times could
be as long as 20 seconds at 100 V applied voltage.
Their results showed that large soft electrodes with high contact pressure such as S4.1
gave lower resistance values. The surface may have small regions of high conductivity,
damage, or pin holes that show low resistance with these electrodes. Small area electrodes
with hard surfaces are less likely to make good contact through low-resistance or damaged
regions.
These results showed that when measuring the performance of an anodized layer, it is
important to use test voltage and electrodes that simulate the working condition as far as is
possible.

5.7.7 Bearings
Many moving machine parts are supported on bearings. The grounding of these machine
parts is often through the bearing. If measured when stationary, the resistance through
the bearing may be low. When the machine is running, the resistance through the bearing
may be much higher or intermittent due to the oil film that separates the moving bearing
surfaces. This problem can be avoided by fitting a bearing that is lubricated by a conducting
(noninsulative) lubricant. Alternatively, a parallel ground path can be established through
a commutator or other mechanism.

5.7.8 Conveyor Belts


Conveyor belts by their nature are often supported on wheels or rollers, which are
themselves supported on bearings. If made of insulating materials, a belt will become
charged.
Conveyor belts in the critical zone should be made of static dissipative materials and
grounded. Grounding of the belt may not be easy as it may rely on conduction through the
belt and bearings or other moving contacts. The resistance of the belt and belt resistance to
ground may need to be set to a level comparable with work surfaces. They must at least be
sufficiently low to prevent significant charging under operating conditions.
146 5 Automated Systems

5.7.9 Using Ionizers to Reduce Charge Levels on ESDS Devices, Essential


Insulators, and Isolated Conductors
Ionizers can be used to reduce the charge levels on essential insulators, ESDS devices, or
isolated conductors in the critical path. To be effective, ionizers must be correctly positioned
to reduce voltages before the point at which control is required and have sufficient time to
neutralize the charge. This is typically before a point where the ESDS will contact another
conductor.
The limitations of ionizers must be borne in mind when specifying this control measure.
● Ionizers take time to neutralize charge and reduce charge levels below the required
threshold.
● Neutralization brings the final voltage to the offset voltage of the ionizer. This must be
below the risk threshold for the device.
● Effects such as air flow can change the operational characteristics of charge neutraliza-
tion in a process.
Unfortunately, equipment manufacturers may not have sufficient understanding to specify
and fit ionizers that work effectively. Automated equipment moves quickly during oper-
ation. Ionizers are often used to neutralize charge but require time to achieve this (see
Sections 2.8.2–2.8.6, 4.6.3, Tan 1993). Their effectiveness may be impaired by speed of oper-
ation compared to charge neutralization rate, and factors such as air flow can also affect
their performance.
The ESD TR10.0-01-02 discusses use of ionizers in AHE. In a fast-moving process, charge
neutralization time must be correspondingly short. The neutralization time will depend on
the characteristics of the ionizer, the position of the item to be neutralized in the region
around the ionizer, and other factors such as air flow.
Yan et al. (2009) reported a case in which an ionizer fitted by the manufacturer as an
“ESD option” failed to prevent ESD damage. The ionizer was unable to neutralize charge
on a device after release from a suction cup and before the device solder balls contacted a
conductive transport boat. The problem was instead solved by anodizing the metal surface
to give a static dissipative coating layer. The rate of charge neutralization drops as the elec-
trostatic field and voltage on the item to be neutralized drops. It therefore takes longer to
reduce voltages to a lower level.
One useful characteristic of charge neutralization by ionizers is that given sufficient time,
all items neutralized tend to move toward the same voltage level – the offset voltage of the
ionizer. If the task is to reduce the voltage difference between a charged isolated conductor
and a charged ESDS that will make contact with it, the ionizer will move both items toward
the same nonzero voltage. So, given sufficient time, the voltage difference between the items
can be much less than the offset voltage of the ionizer.
The effectiveness of an ionizer in neutralizing charge on items should be verified in the
process and circumstance in which it is used.

5.7.10 Vacuum Pickers


One area where it can be critical to reduce device charging is on vacuum pickers (Yan et al.
2009). The amount of charge on the device is critical at the time that it is released to make
5.9 Measurements in AHE 147

contact with another conductor. Where a device has an insulating package material, the
package material that makes contact with the vacuum cup will become charged. If the vac-
uum cup material is also made of insulating material or is not grounded, charge will build
up on the cup during successive operations. So, the cup should be made of a static dissipa-
tive material and grounded. Checking of the grounding of cups should be specified as part
of compliance verification planning.
Nevertheless, some contact charging of the device package cannot be avoided (Yan et al.
2009). The residual charge on the device does not produce voltage on the device until separa-
tion between the cup and the device occurs. This may occur only when the device is a small
distance from making contact with a conductor. Even highly efficient ionizers are unlikely
to have much effect in the small time between release of the device and contact with the
conductor. In this situation, specification of resistive contacting materials, if possible, may
be necessary to avoid damaging ESD.

5.8 ESD Protective Packaging

ESD-protective packaging used in automated equipment often takes special forms such as
tape and reel or JEDEC trays (see Chapter 8). These may be specifically designed to handle
the devices used in the processes. Their shape and form may pose difficulties in measure-
ment for verification of the packaging (see Section 5.9.2).
Components are supplied in tape and reel for placement on PCBs in a pick and place
machine. ESDS devices should be supplied in ESD-protective packaging, but non-ESDS
components are often supplied in tape and reel made from insulating materials. If ESDS and
non-ESDS components are placed in the same operation, electrostatic fields from insulating
tape and reel packaging can become a risk to the ESDS devices. One way of reducing this
risk may be to separate ESDS and non-ESDS components in the placement machine.
Very small devices and their packaging can become sufficiently highly charged that
they may be ejected from the packaging in an uncontrolled manner by electrostatic forces
(Swenson 2018).

5.9 Measurements in AHE

5.9.1 Overview of Measurements in AHE


Measurements in AHE typically include

● Resistance to ground from parts of equipment, surfaces, conveyor belts, or other items
● Electrostatic field and voltage measurements on essential insulators or ungrounded con-
ductors
● Voltage measurements on ESDS devices or small conductors
● Charge measurements on conductors or insulators
● Voltage decay time and offset voltage due to neutralization by an ionizer
● Detection of ESD using radiated electromagnetic interference (EMI) from ESD
148 5 Automated Systems

In addition, measurements of ESD current are being explored by some as a means of


understanding charged device ESD risk.
The principles and practice of measurement given in Chapter 11 should in general be
applied but may need some modification for use in AHE. Standard electrodes and test volt-
ages may not be appropriate for use in automated equipment or packaging used with it.
Measurements on machines are usually made in the ambient conditions in which they
will be operating. If the conditions are variable, it is useful to make measurements under
minimum humidity conditions. These will often be conditions under which maximum elec-
trostatic charging will occur. Charging may often increase also with operating speeds. This
may make it difficult to make measurements under likely worst-case operating conditions.
Test arrangements will often need to be improvised according to the machinery and con-
ditions. It can be challenging to interpret the results.
Specific guidance on making measurements in AHE is given in ESD SP10.1.
Testing will normally be done in the operating environment. Testing under operating
conditions is often preferable but may not be possible due to safety considerations or the
difficulty in connecting instrumentation to moving machine parts.

5.9.2 Resistance Measurements


5.9.2.1 Overview of Resistance Measurements in AHE
Typical resistance measurements will include
● Point-to-point resistance of surfaces
● Resistance to ground from surfaces and machine parts
Different resistance measurement methods usually give different results (Smallwood
and Millar 2010; Smallwood 2017; Smallwood 2018). Large standard electrodes may not
be usable due to lack of large planar surfaces. Large soft electrodes with high contact
pressure such as S4.1 gave lower resistance values. The surface may have small regions of
high conductivity, damage, or pin holes that show low resistance with these electrodes.
Small electrodes often give considerably higher results than large electrodes. Metal-faced
electrodes often give higher results in contact with higher resistance hard materials than
electrodes faced with conductive rubber.
Small-area electrodes with hard surfaces are less likely to make good contact through
low resistance or damaged regions in anodized layers or highly variable ESD control mate-
rials. Where possible, it is perhaps best to choose measurement electrodes that simulate the
actual working situation. This may require some ingenuity in devising a suitable nonstan-
dard electrode.
Smallwood and Millar (2010) also found that test voltage could change the resistance
results by several orders of magnitude. Low voltages typically give high resistance results
(see Section 5.7.6). Low test voltages may be needed to make meaningful measurements
where grounding is required to control voltages on conductors to low levels.

5.9.2.2 Resistance Meters


When measuring resistance to ground of metal machine parts, a suitable multimeter can
be used. For measurements of ESD control materials, a 10 V/100 V meter should be used as
per other resistance measurements on ESD control equipment (see Section 11.6).
5.9 Measurements in AHE 149

5.9.2.3 Point-to-Point Resistance of Surfaces


Point-to-point resistance measurements may be used to evaluate the surface resistance of
surfaces that the ESDS may contact in passing through the equipment. To reduce charged
device ESD risk, it is often preferable to use a static dissipative contacting material rather
than a low-resistance material such as metal (see Section 5.7.5).

5.9.2.4 Resistance to Ground from Surfaces and Machine Parts


Perhaps the most common measurement is of resistance to ground from a surface or
machine part. The resistance should be measured from all designated ESD ground points
to the machine chassis and earth. The resistance from all conductors within the machine
critical path should also be determined.
Resistance from machine parts to machine earth is advised by SP10.1 to be <1 Ω. This can
be measured using a multimeter that has capability to measure resistance down to 0.1 Ω.
For moving parts, a stationary test may not represent performance during operation
in motion. This is because oil films in bearings and other factors may cause intermittent
connection.

5.9.2.5 Resistance to Ground of a Potentially Isolated Conductor


The resistance to ground of a conductor can be measured using a 10 V/100 V high-resistance
meter to determine whether it is isolated. Making a reliable connection to the conductor for
measurement can be a challenge.

5.9.2.6 Measurements on ESD Protective Packaging


Standard concentric ring resistance test electrodes are suited for measurement of large pla-
nar surfaces (see Chapter 11). A standard two-pin electrode is described in IEC 61340-2-3
and ESD S11.13 that is suited for measurements on smaller features or moderately curved
surfaces. Many of the ESD-protective packaging designs used with small ESDS components
in automated processes, such as component tapes, have small features that cannot easily
be measured using these electrodes. At the time of writing, this is an area of current
development in standardization. Some nonstandard electrodes are available from some
suppliers. Different electrodes will in general give different results in resistance tests
(Smallwood 2018).

5.9.3 Electrostatic Field and Voltage Measurements


5.9.3.1 Voltage Measurement Instruments
Many low-cost electrostatic “voltmeters” are in fact electrostatic field meters calibrated in
terms of voltage to make readings from large flat surfaces (see Section 11.6.12). They give
correct surface voltage readings only with a large flat conducting surface held at the correct
calibration distance from the meter. With a large flat insulating surface, these meters give a
reading that is a net result of the varying surface voltage across the surface. Nevertheless, the
results of these measurements on large surfaces are useful in that they are a valid indicator
of ESD risk due to net electrostatic fields in the vicinity.
Field meter–based instruments are not usually suitable for making voltage measurements
on small objects. Field meter–based instruments can sometimes be used with smaller and
150 5 Automated Systems

nonflat surfaces by calibrating the meter in use with a target of the same shape and form
raised to a known voltage.
Voltages on ESDS devices or other moving items typically change with the position of the
item due to changes in capacitance and charge levels. If the item is moving quickly or in a
confined space, it can be difficult to mount a voltmeter in a way that can measure the voltage
on it. In a fast-moving system, the response speed of the voltmeter needs to be fast enough
to follow the voltage changes accurately. A data logger or storage oscilloscope operating in
roll mode may be needed to record voltage changes as they happen. In practice, it is likely
to be the voltage on the conductor immediately before contact with another conductor that
is relevant for evaluating charged device ESD risk.

5.9.3.2 Voltages on Small Conductors or ESDS Devices


When evaluating ESD risk from charged devices or small ungrounded metal objects contact-
ing ESDS devices, it may be necessary to make voltage measurements on small conductors
or ESDS pins. This can be challenging, especially for very small items. Voltages on these
items can be made using noncontact or contact electrostatic voltmeters with very high input
impedance and low input capacitance (see Section 11.6.12 and 11.9.4).

5.9.3.3 Voltages on ESDS Devices


It is often not practical to measure charge on ESDS directly, so voltage or field measurements
are used instead (Yan et al. 2009). Voltage measurements on ESDS devices can be done with
contacting or noncontacting electrostatic voltmeters.
Tribocharging occurs whenever the ESDS is package contacted by another material.
Charging of the conductors within the device can occur via contact with a charge source
via the device pins. Any charge on the device package, or nearby electrostatic field sources,
will induce voltages on the device conductors. The device voltage typically increases
dramatically when it is separated from a surface against which it has become tribocharged.
Tribocharging of the device cannot be avoided, because contact between the device and
other materials such as packaging or handler part is inevitable. ESDS packages are often
made of polymer or ceramic insulating materials, and tribocharging occurs on insulating
ESDS package parts even when the contacting materials are conductive or static dissipative.
The voltage on a device passing through a handling system is not constant, even if the
charge on the device is unchanging. Voltages on noninsulative parts vary with charge level,
proximity to electrostatic field sources, and conductor capacitance. The voltage V on a con-
ductor is related to its charge Q but changes with the capacitance C of the conductor as
CV = Q
Capacitance varies with the orientation and proximity to other materials, especially con-
ductors. So, the ESDS voltage is constantly changing as the device moves through a pro-
cess. SP10.1 describes a method of calibrating voltage measurements with a known voltage
applied to a conductor attached to a device.
Discharge occurs when the device pins touch another conductor at a different voltage. So,
it is the voltage on the device immediately before contact is made that is of most interest. At
this point, the voltage may have reduced considerably from a value when the ESDS device
is far from the contact point, due to capacitance changes.
5.9 Measurements in AHE 151

Voltage levels on insulative parts of an item are even more variable as the charge density
on the surface of the insulator varies from point to point across the surface. There is no
single value that can be ascribed to the surface of an insulator such as a device package
material. Voltages measured are typically the net effect of charge density over a region on
the surface and will change with proximity to conductors including any voltage measuring
equipment used. This can be a useful measurement in ESD risk evaluation as in practice it
is often also the net effect of charges inducing a voltage on a conductor that causes ESD risk.

5.9.4 Charge Measurements


The charge on an item is sometimes measured either directly, e.g. using a Faraday pail or
transferred charge measurement, or indirectly using field or voltage measurements (Paasi
2004). In some cases, the charge measured on a device can be compared with an estimated
charge threshold for failure calculated from CDM ESD withstand data.

5.9.5 Measurement of the Voltage Decay Time and Offset Voltage Due
to Neutralization by an Ionizer
Standard measurement of the effectiveness of ionizers is done using a charged plate monitor
(CPM) in which the measurement plate is 150 × 150 mm (see Section 11.6.13). The positive
and negative polarity charge decay time and offset voltage are measured. This size of plate
approximates the size of a semiconductor wafer or PCB. Many devices or machine parts
that are to be neutralized in AHE are much smaller than this. It is questionable whether a
CPM plate this size is representative of charge neutralization on small devices. Many non-
standard CPM instruments are available that have smaller plates a few tens of millimeters
in dimensions. It is possible that these give more representative indication of neutralization
of similarly sized devices.

5.9.6 ESD Current Measurements


ESD current measurements are being explored by some as a means of evaluating charged
device ESD risk (Bellmore 2004; Tamminen et al. 2017a,b). This is a difficult type of mea-
surement that shows some promise but has not yet been shown to be achievable in practice.

5.9.7 Detection of ESD Using EMI Detectors


ESD occurring within operating equipment can be detected from its electromagnetic
radiation (in other words, EMI) using a suitable detector (see Section 11.8.10) (Millar and
Smallwood 2010). Some detectors are designed to differentiate between HBM and CDM
ESD, but in real situations ESD usually has variable characteristics somewhat different
from these models. ESD detectors will detect any ESD in the vicinity, including those
that have nothing to do with ESD threat to semiconductor devices, such as operating
contactors. So, it is necessary to critically evaluate any ESD detected and determine its
relevance to ESD threat. Sometimes this can be done by direct visual observation. If an
ESD is detected at a time that a device makes contact with a conductive item, ESD risk is
indicated.
152 5 Automated Systems

5.10 Handling Very Sensitive Components

As time goes on, devices having low HBM and CDM ESD withstand voltages are becoming
more common (ESDA 2016a). Some have HBM withstand voltage less than 100 V HBM and
200 V CDM withstand voltage.
The ESD TR10.0-01-02 discusses measurement and control issues in handling very
sensitive devices in AHE. This document defines hard ground as a having resistance
to ground of 1 Ω or less, and soft ground as between 1 kΩ and 1 GΩ. Machine parts
typically require hard grounding, but an ESDS device making contact with this low
resistance may risk charged device ESD damage. AHE parts that may contact ESDS
may require soft grounding. Charged device ESD damage is typically due to high
ESD peak current flow, and it is the resistance (or impedance, see Viheriäkoski et al.
2017) of the material at the point of contact that typically limits ESD current in the
discharge.
When handling low CDM ESD withstand voltage devices in AHE, the voltages in the
critical region and on the device need to be kept to a low level. This means that if ionizers are
used for voltage reduction, the ionizer may need to have greater capacity to reduce voltages
in a shorter time, as well as having low-offset voltage (see Section 5.7.9).
CPMs used for characterizing performance of ionizers should be as representative as
possible of the situation that they are simulating. In the AHE environment, this may
mean use of a small low capacitance CPM plate. For simulating neutralization of low
CDM ESD withstand voltage devices, it may be better to measure decay time over a
lower voltage range (e.g. 200–20 V) than the standard CPM measurement (typically
1000–100 V).

References

Bellmore, D. (2001). Anodized aluminium alloys – insulators or not? In: Proc EOS/ESD Symp.
EOS-23, 141–148. Rome, NY: EOS/ESD Association Inc.
Bellmore, D.G. (2004). Paper 4A.6. Characterizing automated handling equipment using
discharge current measurements. In: Proc EOS/ESD Symp. EOS-26. Rome, NY: EOS/ESD
Association Inc.
ESD Association. (2015). ESD TR17.0-01-15. Technical Report for ESD Process Assessment
Methodologies in Electronic Production Lines – Best Practices used in Industry, Rome, NY,
EOS/ESD Association Inc.
ESD Association. (2016a) ESD Association Electrostatic Discharge (ESD) Technology
roadmap – revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf [Accessed: 10th May 2017] Rome, NY, EOS/ESD
Association Inc.
ESD Association (2016b) ANSI/ESD SP10.1-2016. Standard practice for protection of
Electrostatic Discharge Susceptible Items – Automated handling Equipment (AHE), Rome, NY,
EOS/ESD Association Inc.
References 153

Gaertner, R. (2007). Paper 3B.1. Do we expect ESD-failures in an EPA designed according to


international standards? The need for a process related risk analysis. In: Proc. EOS/ESD
Symp. EOS-29, 192–197. Rome, NY: EOS/ESD Association Inc.
Gaertner, R. and Stadler, W. (2012). Paper 3B.5. Is there a correlation between ESD
qualification values and the voltages measured in the field? In: Proc. EOS/ESD Symp.
EOS-34. Rome, NY: EOS/ESD Association Inc.
Jacob, P., Gärtner, R., Gieser, H. et al. (2012). Paper 3B.8. ESD risk evaluation of automated
semiconductor process equipment – A new guideline of the German ESD Forum e.V. In:
Proc. EOS/ESD Symp. EOS-34. Rome, NY: EOS/ESD Association Inc.
Kim, D.-S., Lim, C.-B., Oh, D.-S. et al. (2012). Paper 2B.1. Minimizing electrostatic charge
generation and ESD Event in TFT-LCD production equipment. In: Proc. EOS/ESD Symp.
EOS-34. Rome, NY: EOS/ESD Association Inc.
Koh, L.H., Goh, Y., and Lim, S.H. (2013). Reliability assessment of high temperature automated
handling equipment retrofit for CDM mitigation. In: Proc. EOS/ESD Symposium EOS-35,
43–48. Rome, NY: EOS/ESD Association Inc.
Proc. EOS/ESD Symp. EOS-32Millar, S. and Smallwood, J.M. (2010). Paper 3B.2. CDM Damage
due to Automated Handling Equipment. In: , 217–223. Rome, NY: EOS/ESD Association Inc.
Olney, A., Gifford, B., Guravage, J., and Righter, A. (2003). Real-world charged board model
(CBM) failures. In: Proc. EOS/ESD Symp. EOS-25, 34–43.
Paasi, J. (2004). ESD control in automated handling. In: 6th International ESD Workshop in
Dresden, Germany, September 7–8, 2004. Rome, NY: EOS/ESD Association Inc.
Smallwood, J. (2017). A practical comparison of surface resistance test electrodes. J. Electrostat.
88: 127–133.
Smallwood, J. (2018). Paper 4B3. Comparison of surface and volume resistance measurements
made with standard and non-standard electrodes. In: Proc. EOS/ESD Symp. EOS-40. Rome,
NY: EOS/ESD Association Inc.
Smallwood, J.M. and Millar, S. (2010). Paper 3B4. Comparison of methods of evaluation of
charge dissipation from AHE soak boats. In: Proc. EOS/ESD Symp. EOS-32, 233–238. Rome,
NY: EOS/ESD Association Inc.
Swenson D.E. 2018. Private communication.
Tamminen, P., Smallwood, J., and Stadler, W. (2017a). Paper 1B.4. Charged device discharge
measurement methods in electronics manufacturing. In: Proc. EOS/ESD Symp. EOS-39.
Rome, NY: EOS/ESD Association Inc.
Tamminen, P., Smallwood, J., and Stadler, W. (2017b). Paper 4B.2. The main parameters
affecting charged device discharge waveforms in a CDM qualification and manufacturing.
In: Proc. EOS/ESD Symp. EOS-39. Rome, NY: EOS/ESD Association Inc.
Tan, W.H. (1993). Minimizing ESD hazards in IC test handlers and automated trim/form
machines. In: Proc. EOS/ESD Symp. EOS-15, 57–64. Rome, NY: EOS/ESD Association Inc.
Viheriäkoski, T., Kärjä, E., Gärtner, R., and Tamminen, P. (2017). Paper 4B.3. Electrostatic
discharge characteristics of conductive polymers. In: Proc. EOS/ESD Symp. EOS-39. Rome,
NY: EOS/ESD Association Inc.
Yan, K.P., Gaertner, R., Wong, C.Y., and Ong, C.T. (2009). Automatic handling equipment-The
role of equipment maker on ESD protection. In: Proc. EOS/SED Symp. EOS-31,
1B.2-1–1B.2-6. Rome, NY: EOS/ESD Association Inc.
154 5 Automated Systems

Further Reading

Bellmore, D.G. and Bernier, J. (2005). Characterizing automated handling equipment using
discharge current measurements II. In: Proc. EOS/ESD Symp., 195. Rome, NY: EOS/ESD
Association Inc.
Dangelmeyer, T. (1999). ESD Program Management, 2e. Clewer. ISBN: 0-412-13671-6.
EOS/ESD Association Inc. (2016). ESD TR20.20-2016, ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
ESD Association. (2002). ESD TR10.0-01-02. Technical Report - Measurement and ESD Control
Issues for Automated Equipment Handling of ESD Sensitive Devices Below 100 Volts, Rome,
NY, EOS/ESD Association Inc.
ESD Association. (2006). ANSI/ESD STM4.1-2006. ESD Association Standard for the Protection
of Electrostatic Discharge Susceptible Items – Worksurfaces - Resistance Measurements, Rome,
NY, EOS/ESD Association Inc.
ESD Association. (2014). ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices),
Rome, NY, EOS/ESD Association Inc.
ESD Association. (2015) ANSI/ESD STM11.13-2015. ESD Association Standard Test Method for
the Protection of Electrostatic Discharge Susceptible Items – Two-Point Resistance
Measurement, Rome, NY, EOS/ESD Association Inc.
Halperin, S., Gibson, R., and Kinnear, J. (2008). Paper 2B-21. Process capability & transitional
analysis. In: Proc. EOS/ESD Symp. EOS-30. Rome, NY: EOS/ESD Association Inc.
International Electrotechnical Commission. (2015) IEC 61340-5-3:2015. Electrostatics.
Protection of electronic devices from electrostatic phenomena. Properties and requirements
classifications for packaging intended for electrostatic discharge sensitive devices, International
Electrotechnical Commission, Geneva.
International Electrotechnical Commission. (2016a). IEC 61340-2-3:2016. Electrostatics. Part
2-3: Methods of test for determining the resistance and resistivity of solid materials used to avoid
electrostatic charging, International Electrotechnical Commission, Geneva.
International Electrotechnical Commission (2016b) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements,
International Electrotechnical Commission, Geneva.
International Electrotechnical Commission. (2018) IEC TR 61340-5-2:2018. Electrostatics – Part
5-2: Protection of electronic devices from electrostatic phenomena - User guide. International
Electrotechnical Commission, Geneva.
Kietzer, G. (2012). Paper 2B.2. ESD risks in the electronics manufacturing. In: Proc. EOS/ESD
Symp. EOS-34, 202. Rome, NY: EOS/ESD Association Inc.
Kim, D.-S., Lim, C.-B., Yoon, S.-H. et al. (2013). Paper 7B.1. Electrostatic control and its analysis
of roller transferring processes in FPD manufacturing. In: Proc. EOS/ESD Symp. EOS-35.
Rome, NY: EOS/ESD Association Inc.
Koh, L.H., Goh, Y., and Lim, S.H. (2013). Paper 1B.3. Reliability assessment of high
temperature automated handling equipment retrofit for CDM mitigation. In: Proc. EOS/ESD
Symp. EOS-35. Rome, NY: EOS/ESD Association Inc.
Further Reading 155

Koh, L.H., Goh, Y.H., and Wong, W.F. (2017). Paper 1B.2. ESD risk assessment considerations
for automated handling equipment. In: Proc. EOS/ESD Symp. EOS-39. Rome, NY: EOS/ESD
Association Inc.
Paasi, J., Tamminen, P., Kalliohaka, T. et al. (2002). ESD control tools for surface mount
technology and final assembly lines. In: Proc. EOS/ESD Symp. EOS-24, 250–256. Rome, NY:
EOS/ESD Association Inc.
Paasi, J., Tamminen, P., Salmela, H. et al. (2005). ESD control in automated placement process.
In: Proc. EOS/ESD Symposium EOS-27, 203. Rome, NY: EOS/ESD Association Inc.
Steinman, A. (2010). Paper 3B3. Measurements to establish process ESD compatibility. In: Proc.
EOS/ESD Symp. EOS-32. Rome, NY: EOS/ESD Association Inc.
Steinman, A. (2012). Paper 2B.4. Process ESD capability measurements. In: Proc. EOS/ESD
Symp. EOS-34. Rome, NY: EOS/ESD Association Inc.
Steinman, A. (2014). Paper 1B.3. Measuring handler CDM stress provides guidance for factory
static controls. In: Proc. EOS/ESD Symp. EOS-36. Rome, NY: EOS/ESD Association Inc.
Tamminen, P. and Viheriäkoski, T. (2007). Paper 3B.3. Characterization of ESD risks in an
assembly process by using component-level CDM withstand voltage. In: Proc. EOS/ESD
Symp. EOS-29, 202–211. Rome, NY: EOS/ESD Association Inc.
Tamminen, P. and Viheriäkoski, T. (2011). Product specific ESD risk analysis. In: Proc.
EOS/ESD Symp. EOS-33, 97. Rome, NY: EOS/ESD Association Inc.
Welker, R.W., Nagarajan, R., and Newberg, C. (2006). Contamination and ESD Control in
High-Technology Manufacturing. Wiley-Interscience/IEEE Press. ISBN-10: 0 471 41452 2,
ISBN-13: 978 0 471 41452 0.
Yan, K.P., Gaertner, R., and Wong, C.Y. (2010). Paper 3B.1. ESD protection program at
electronics industry – areas for improvement. In: Proc. EOS/ESD Symp. EOS-32. Rome, NY:
EOS/ESD Association Inc.
Yan, K.P., Gaertner, R., and Wong, C.Y. (2013). Semiconductor back end manufacturing
process – ESD capability analysis. In: Proc. EOS/ESD Symposium EOS-35, 30. Rome, NY:
EOS/ESD Association Inc.
157

ESD Control Standards

6.1 Introduction
The facilities equipment and materials needed for use in electrostatic discharge (ESD) pre-
vention are specified in various standards worldwide. The two main standards discussed in
this book are IEC 61340-5-1 and ESD Association ANSI/ESD S20.20. IEC 61340-5-1 has also
been adopted as a national standard in many countries, in Europe becoming the European
Norm EN61340-5-1. Individual countries may have their own versions, and, in the United
Kingdom, this is BS EN 61340-5-1.
For clarity, the 61340-5 series terminology is used, although the ESD Association (ESDA)
series documents are in many cases nearly identical and are cross-referenced. In many
cases, International Electrotechnical Commission (IEC) documents were based on ESDA
standards that were published earlier. Where appropriate, some differences with the ESDA
standards are noted.

6.2 The Development of ESD Control Standards


Before the days when ESD control became necessary in electronics manufacture, ESD
control practices were already in use in gunpowder and explosives handling and some
industrial processes where flammable materials were handled. Until the 1970s there was
little need for ESD control in electronics manufacture and component handling (Ind. Co.
2011). ESD damage started to show as a problem in the late 1970s with the introduction
of large-scale integration (LSI). The first ESD control programs were set up by industry
companies, with little standardization or sharing of information on the subject.
One early standard was the US Military MIL-STD-1686 (Department of Defense (1980a),
released with the handbook MIL-HDBK-263 (Department of Defense (1980b). Companies
that supplied electronics to the military were required to comply with this standard, but
others followed in-house ESD control procedures.
Some countries produced national standards such as the BS 5783:1984 in the United
Kingdom. A European standard CECC00015 was introduced in 1991, providing a common
standard for countries in the CENELEC Electronics Components Committee (CECC)
and superseding BS5783 as the BS EN 100015-1:1992. At the time, CENELEC members
included national electrotechnical committees of Austria, Belgium, Denmark, Finland,
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
158 6 ESD Control Standards

France, Germany, Greece, Iceland, Ireland, Italy, Luxembourg, Netherlands, Norway,


Portugal, Spain, Sweden, Switzerland, and the United Kingdom. The EN 100015 had
four parts (British Standards Institute 1992, 1994a,b,c). BS EN100015-1 gave the basic
requirements for compliance, setting up ESD-protected areas (EPAs) and ESD control
packaging. BS EN100015-2 gave additional requirements for low-humidity conditions less
than 20% rh. Part 3 of the standard gave requirements for clean room conditions, and Part
4 for high-voltage areas. All these parts of EN 100015 were in turn superseded by the first
version of EN61340-5-1.
Materials and equipment used in ESD control initially did not have standard test meth-
ods, procedures, or equipment to measure their properties. The variety of test methods and
equipment used led to differences in results.
The ESD Association was formed in the early 1980s and commenced standards devel-
opment for ESD control materials and equipment such as flooring, wrist straps, and work
surfaces. Suppliers and users could use these to verify the properties and functioning of ESD
control materials and equipment on a commonly recognized basis.
In 1995, the ESDA was given the task of replacing MIL-STD-1686 with an industry stan-
dard, leading to the development of ANSI/ESD S20.20-1999. Meanwhile in the 1990s the
IEC had set up IEC Technical Committee 101 Electrostatics, to produce internationally
applicable standards for test methods for evaluation of the generation, retention, and dis-
sipation of ESD, ascertaining the effect of ESDs and methods of simulation of electrostatic
phenomena. The scope of TC101 was later extended to include design and implementation
of handling areas or procedures, equipment, and materials used to reduce or eliminate elec-
trostatic hazards or undesirable effects and exclusions. Much of TC101’s work has focused
on electronics industry ESD control, although more generally applicable standards (e.g. for
measurement of resistance of static control materials, properties of flooring, and footwear
and flooring in combination) were developed for use in a wide range of applications. These
were based on existing ESD Association, European, Asian, or industry standards. Electro-
static control equipment and material for other industries (e.g. properties of flexible inter-
mediate bulk containers used in the chemical industry) are also developed by IEC TC101.
The first IEC 61340-5-1 published in 1998 was developed from the CECC100015 and other
standards and had a less flexible approach compared to the ANSI/ESD S20.20 standard
(EOS/ESD Association Inc 2014c). The first decade of the twenty-first century saw the glob-
alization and increasing diversity of electronic industry processes and environments and
increasing spread of the influence and use of the ANSI/ESD S20.20 worldwide. In response,
the second edition 61340-5-1 adopted an approach based on ANSI/ESD S20.20 in the spirit
of worldwide harmonization of ESD control standards for the electronics industry, albeit
with some remaining differences. These have largely been eliminated with the publication
of the 2016 edition of 61340-5-1 (International Electrotechnical Commission. 2016c). It is
now widely considered that the 61340-5-1 and ANSI/ESD S20.20 standards are technically
equivalent.
Some earlier standards gave guidance as well as requirements for compliance in the same
documents. This was often confusing for the user, and many interpreted the guidance as
aspects that must be implemented for compliance. Later standards removed guidance into
6.3 Who Writes the Standards? 159

separate documents IEC 61340-5-2 and ESD TR20.20 (International Electrotechnical Com-
mission 2018; EOS/ESD Association Inc. 2016b). These user guides contained no require-
ments for compliance but gave an increasing amount of useful information and guidance
to help develop and maintain an effective ESD control program.
These documents that do not contain requirements for compliance, but often contain
guidance or information of a different type, are called “Technical Reports.” In the IEC and
ESDA systems they are designated by “TR” in the document number, for example ESD
TR20.20 or IEC TR 61340-5-2. They can be extremely useful sources of information of many
types. The references at the end of this chapter are a list of many of these.

6.3 Who Writes the Standards?


The ESD Association and IEC standards are written by volunteer experts working within
Working Groups (WGs) at the ESDA and IEC Technical Committee 101 Electrostatics. The
WG experts are usually volunteers who may or may not be supported by their employers
or other organizations to work on ESD standards. In their professional life, they may be
engaged in electronics manufacturing companies, ESD-protective equipment manufactur-
ers, or suppliers, research, or consultancy work.
The ESDA has been developing test methods for ESD work since the early 1980s. Many of
the currently used standard test methods were developed by ESDA and then adopted, with
some revision or adaption, for the IEC system. They are then adopted as national standards
in various IEC participating countries around the world. In most cases, the requirements
of these standards and standard test methods specified in both systems are very similar.
The IEC was mandated by the World Trade Organization to draft electrotechnical
standards to facilitate trade worldwide. IEC TC101 is responsible for electrostatic-related
test methods and other documents. The IEC TC101 experts are nominated by their
national standards committees (NCs) to work on projects of interest to their countries. As
an example, in the United Kingdom, the BSI Committee GEL101 is the main NC following
and participating in TC101’s work. During development of a new document, drafts are
returned to GEL101 on a regular basis for comment, and these comments are considered
in the rewriting of a new draft. The process continues through various stages until a draft
International Standard is achieved by consensus of all the participating NCs. In 2017,
the participating members of IEC TC101 included 20 Participating members (who attend
meetings and return comments on draft documents) and 13 Observer members (Table 6.1).
While most IEC TC101 member countries adopt the 61340-5-1 standard as their national
standard, not all do. Notably, the United States maintains the ANSI/ESD S20.20 standard
as its American National Standards Institute (ANSI) standard.
When a new IEC standard is published, it is also submitted to the European standard-
ization organization CENELEC for possible adoption as a European Standard and to the
standardization bodies of the participating countries. This is then voted on by European
NCs participating in CENELEC. If adopted, it becomes a European norm (EN). In the
United Kingdom, European Norms are normally then adopted by member countries. In
the case of the United Kingdom, the UK standardization organization BSI, if adoption is
agreed, publishes the document as a BS EN documents. So, IEC 61340-5-1: 2016 became
EN 61340-5-1: 2016 in Europe and then published as BS EN 61340-5-1:2016 in the United
160 6 ESD Control Standards

Table 6.1 IEC TC101 Participating (P) and


Observer (O) members in 2017.

Argentina O-Member
Australia O-Member
Austria P-Member
Belgium P-Member
Bulgaria O-Member
China P-Member
Czech Republic P-Member
Denmark O-Member
Egypt O-Member
Finland P-Member
France P-Member
Germany P-Member
Hungary P-Member
Ireland P-Member
Israel O-Member
Italy P-Member
Japan P-Member
Korea, Republic of P-Member
Netherlands P-Member
Norway O-Member
Poland P-Member
Portugal O-Member
Romania O-Member
Russian Federation P-Member
Serbia O-Member
Spain P-Member
Sweden P-Member
Switzerland P-Member
Thailand O-Member
Turkey O-Member
Ukraine O-Member
United Kingdom P-Member
United States of America P-Member

Kingdom. All these steps and procedures take an amount of elapsed time. It is often several
years between commencement of a project at IEC and publication of an IEC standard and
then several further months before publication as BS EN in the UK or other national stan-
dard!
6.4 The IEC and ESDA Standards 161

6.4 The IEC and ESDA Standards


6.4.1 Standards Numbering
All standards are given a number according to a system. A list of ESDA and IEC test method
publications used in ESD work and current at the time of writing is given in Chapter 11. The
good news is that the average ESD practitioner in the electronics industry will not need to
have copies of all of these! Many of the modern ESD test methods are similar in their ESDA
and IEC forms. Beware, however, that there can also be some significant differences.

6.4.2 The Language of Standards


Standards use some words in a specific way. Many of the aspects described are “require-
ments” – which in standards terms means “you have to do this to comply with the standard.”
However, some aspects only have the status of “recommendations” – which means depart-
ing from their instructions does not represent noncompliance with the standard. It should,
however, be born in mind that there may be good reasons for acting according to the rec-
ommendations.
In the standards, the difference between “requirements” and “recommendations” is
often shown by the language used within the text. The word shall represents a “require-
ment,” but the word should represents a “recommendation.” For example, 61340-2-3:2016a
Section 8.2.1 states “The contact material surface shall have a volume resistance of less
than 103 Ω…” The shall means that for compliance the electrode manufacturer must select
the electrode material with this characteristic.
In places, some standards give “examples,” e.g. in 61340-2-1:2002 Figure 1 gives “Example
of an arrangement for measurement of dissipation of charge using corona charging.” Inter-
national Electrotechnical Commission. (2015a) Examples are for assistance of the reader.
Departing from the example is not necessarily a noncompliance, although it is often easiest
to follow the example when making a test setup.
In 61340-5-1:2016, Annex A is labeled as “Normative.” The word normative is another
way of indicating that the test methods in this annex are part of the “requirements” of the
standard. In contrast, a section marked as “Informative” means that the section can be
considered merely informational or “recommendations.”
IEC standards have “Normative references” sections that includes a list of standards that
are deemed essential to the application of the standard. Where a reference to a standard in
this section is dated (e.g. IEC 61340-4-1: 2003), this means that only this version must be
used – later versions of the standard may exist but do not apply and may not even make
sense in the context.
In contrast, an undated reference (e.g. IEC 61340-4-1 with no attached date) means that
the latest and current version of the standard should be referred to.
Some IEC documents also are designated “TS,” e.g. PD IEC/TS 61340-4-2 Electrostat-
ics – Part 4-2: Standard test methods for specific applications – Electrostatic properties of
garments (International Electrotechnical Commission 2013b). A Technical Specification
document is one for which the required support was not achieved for the publication of
an International Standard, the subject is still under technical development, or there is the
162 6 ESD Control Standards

future but not immediate possibility of agreement as an International Standard. It therefore


resembles, but does not have the strength and status of an International Standard. Technical
specifications normally reviewed within three years.
A third type of document is the Technical Report (TR), e.g. IEC TR 61340-5-2 Electrostat-
ics – Part 5-2: Protection of electronic devices from electrostatic phenomena – User guide.
These include data of a different kind from that normally published as an International
Standard. TR documents must be entirely informative in nature and must not contain mat-
ter implying that it is normative.
Some of the ESD Association standards are also ANSI. For example, the 20.20-2014 stan-
dard is ANSI/ESD S20.20-2014.
Many of the ESDA documents referred to here are Standards identified by the letter S
in the numbering system. These give “a precise statement of a set of requirements to be
satisfied by a material, product, system, or process that also specifies the procedures for
determining whether each of the requirements is satisfied” (DE Swenson 2017 private
communication). The letters STM indicate a Standard Test Method (e.g. ANSI/ESD
STM11.13-2015. ESD Association Standard Test Method for the Protection of Electrostatic
Discharge Susceptible Items – Two-Point Resistance Measurement). These documents give
“a definitive procedure for the identification, measurement, and evaluation of one or more
qualities, characteristics, or properties of a material, product, system, or process that yields
a reproducible test result.”
The letters SP identify a “Standard Practice document (e.g. ANSI/ESD SP15.1 Standard
Practice for the Protection of Electrostatic Discharge Susceptible Items – In-Use Resistance
Measurement of Gloves and Finger Cots). Standard Practice documents give “a procedure
for performing one or more operations or functions that may or may not yield a test result.”
Standard Practices may not give reproducible results.
Like the IEC system, the ESDA also has other types of documents that are not Standards.
Some of these are TRs and give “a collection of technical data or test results published as
an informational reference on a specific material, product, system or process.” These doc-
uments do not have technical requirements; they are informational in nature. Examples
are ESD TR20.20 and ESD TR53-01-06 (EOS/ESD Association Inc 2006b). The ESDA also
has ADV, or “Advisory,” documents (e.g. ESD ADV1.0-2017. ESD Association Advisory for
Electrostatic Discharge Terminology – Glossary). These are usually older documents that
would be published as SP or TR today.
There has been some effort to harmonize the ESDA and IEC series standards in various
ways. So, some of the documents in the series are to some extent similar or near equivalent
(Tables 6.2 and 6.3).

6.4.3 Definitions Used in Standards


The standards give specific definitions for some terms used within the standards. These
may differ from document to document and from version to version. For example, the
definitions of conductive, dissipative, and insulative often differ between standards in dif-
ferent industries and for different products such as footwear or packaging materials. There
may also occur changes with different versions of standards. For example, the resistance
range definitions of conductive and dissipative packaging materials have changed with
6.4 The IEC and ESDA Standards 163

Table 6.2 Near-equivalent ESD control program standards in the IEC 61340-5-1 and ANSI/ESD
S20.20 series of documents.

IEC 61340-5-1 series ANSI/ESD S20.20 series Content

IEC 61340-5-1 Protection ANSI/ESD S20.20 Development General requirements for an


of electronic devices from of an Electrostatic Discharge ESD control program
electrostatic Control Program for – Protection
phenomena – General of Electrical and Electronic Parts,
requirements Assemblies, and Equipment
IEC 61340-5-2 Protection ESD TR 20.20 Protection of User guides for implementing
of electronic devices from Electrical and Electronic Parts, an ESD control program
electrostatic Assemblies, and according to 61340-5-1 and
phenomena – User guide. Equipment – Handbook. ANSI/ESD S20.20
IEC 61340-5-3 Properties ESD Association. Packaging Properties and classifications
and requirements Materials for ESD Sensitive for packaging for
classifications for Items. ANSI/ESD S541 ESD-sensitive devices
packaging intended for ESD
sensitive devices.

Table 6.3 Near-equivalent test method documents in the IEC 61340-5-1 and ANSI/ESD S20.20
series.

IEC 61340-5-1 series ANSI/ESD S20.20 series Content

IEC 61340-5-4:2019 ESD TR53-01-06 Compliance Compliance verification


Verification of ESD Protective measurement techniques
Equipment and Materials.
IEC61340-2-3 Methods of ANSI/ESD STM11.11:2015b Surface General surface and
test for determining the resistance measurement of static volume resistance test
resistance and resistivity dissipative planar materials. methods
of solid planar materials ANSI/ESD STM11.12:2015c Volume
used to avoid electrostatic resistance measurement of static
charging dissipative planar materials.
ANSI/ESD STM11.13:2015d
Two-point resistance measurement.
IEC61340-4-3:2001 ANSI/ESD STM9.1:2014b Test methods for the
Footwear Footwear – Resistive resistance of
Characterization. footwear – not worn
IEC 61340-4-1 Electrical ANSI/ESD STM7.1:2013b Resistive Measurement of resistance
resistance of floor Characterization of of floor materials before
coverings and installed Materials – Floor Materials. and after installation
floors.
IEC 61340-4-5:2004 ANSI/ESD STM97.1:2015e Floor Measurement of resistance
Methods for Materials and Footwear – Resistance and body voltage of
characterizing the Measurement in Combination with person-footwear-floor
electrostatic protection of a Person. system
footwear and flooring in ANSI/ESD STM97.2:2016a Floor
combination with a Materials and Footwear – Voltage
person Measurement in Combination with
a Person.
164 6 ESD Control Standards

Table 6.3 (Continued)

IEC 61340-5-1 series ANSI/ESD S20.20 series Content

IEC61340-2-3 Methods of ANSI/ESD STM4.1:2006a Measurement of


test for determining the Worksurfaces – Resistance resistance to ground and
resistance and resistivity Measurements. to groundable point of
of solid planar materials work surfaces
used to avoid electrostatic
charging
IEC 61340-4-9:2016b ANSI/ESD STM 2.1 Garments. Resistance measurement
Garments of ESD control garments
IEC61340-2-3 Methods of ANSI/ESD STM12.1:2013a Measurement of resistance
test for determining the Seating – Resistive Measurement. to groundable point of
resistance and resistivity ESD control seating
of solid planar materials
used to avoid electrostatic
charging
IEC 61340-4-6:2015b ANSI/ESD S1.1:2013c Wrist Straps. Requirements and
Wrist straps. measurement of wrist
straps
IEC61340-2-3 Methods of ESD SP9.2:2003 Footwear – Foot Measurement of resistance
test for determining the Grounders Resistive of foot grounders
resistance and resistivity Characterization
of solid planar materials
used to avoid electrostatic
charging
IEC 61340-4-7:2017 ANSI/ESD STM3.1:2015a Measurement of
Ionization Ionization. performance of ionizers
IEC 61340-4-8:2014 ANSI/ESD STM11.31:2012b Bags. Measurement of
Discharge shielding. performance of ESD
Bags. shielding bags

different versions of 61340-5-1 and 61340-5-3 (International Electrotechnical Commission


2015c). When working according to a standard, the definitions used in that standard must
be applied. Incorrect use of these terms can result in inadvertent use of noncompliant equip-
ment or materials.

6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20


Standards
6.5.1 Background
The earlier editions of IEC 61340-5-1 and ANSI/ESD S20.20 were superseded by new
versions during 2007. The requirements of these new standards were broadly simi-
lar – previously 61340-5-1 had been significantly different in its approach from ANSI/ESD
S20.20.
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 165

IEC 61340-5-1 was adopted by CENELEC and replaced the first edition of EN61340-5-1:
2001 in Europe.
A new version of ANSI/ESD S20.20 was published in 2014. This was followed by a new
version 61340-5-1 in 2016. These versions were current at the time of writing, and the fol-
lowing sections summarize the requirements of these documents. The documents continue
to be updated in the light of technology and industry changes and the development of ESD
protection and control knowledge and understanding as time goes on. The reader should
of course check current versions of the standards for up to date information on them.

6.5.2 Documentation and Planning


The IEC 61340-5-1:2016 and ANSI/ESD S20.20-2014 standards require the organization to
write four ESD plans.
● ESD Control Program Plan
● ESD Training Plan
● (ESD Control) Product Qualification Plan
● Compliance Verification Plan
The standard does not dictate in detail what should be contained in these plans – it is
the responsibility of the ESD coordinator to specify this. The purpose and content of the
plans is further explained in the following sections. Equipment used in the ESD program
is normally expected to comply with the standard (where requirements are given) unless
a tailoring statement is given explaining the rationale and technical justification for the
departure is acceptable.
Although the requirements of 61340-5-1 are nearly identical, users seeking to comply
with these standards are advised to obtain copies and refer to them. There may be small
differences between the detail of the requirements and wording. Up-to-date copies should
always be obtained as the standards are updated every few years as knowledge and techno-
logical needs progress and as ideas of best practice change from time to time. The following
sections give an outline of the main requirements of the standards at the time of writing
(2019).

6.5.3 Technical Basis of the ESD Control Program


The ESD control program designed according to these standards is intended to provide
effective handling of ESD-sensitive (ESDS) devices with ESD withstand voltages down to
100 V HBM and 200 V charged device model (CDM). The scope of 61340-5-1 states that it
“applies to activities that: manufacture, process, assemble, install, package, label, service,
test, inspect, transport or otherwise handle electrical or electronic parts, assemblies, and
equipment.”
An ESD control program can be devised according to the standards for handling ESDS
devices with lower withstand voltages. In this case, the program may need additional con-
trol elements or adjusted requirement limits for the specified equipment.
It is important to understand that these standards are not intended to be used in non-
electronic applications and processes. Handling of, and processes involving, electrically
166 6 ESD Control Standards

initiated explosive devices, flammable liquids, gases, and powders are specifically excluded
by the scope. Some of these nonelectronic applications may be covered by specific national
regulations or other standards such as IEC 60079-32-1 International Electrotechnical Com-
mission. (2013a).

6.5.4 Personal Safety


The standards recognize that use of some ESD control equipment could conceivably expose
personnel to unsafe conditions under some circumstances. In other cases, safety or other
regulations may apply to materials, equipment, or processes in which ESDS devices are also
handled. In these cases, ESD control practices must be compliant with the relevant regu-
lations, laws, or codes of practice. ESD control practices never supersede correct personal
safety requirements.

6.5.5 ESD Coordinator


It is a requirement of the standards that a person must be appointed who has responsibility
for implementing the requirements of the standard including establishing, documenting,
maintaining, and verifying the compliance of the ESD program. Some organizations know
the ESD Coordinator by other titles (e.g. ESD program manager).

6.5.6 Tailoring the ESD Program


Parts of the standard may not apply in all cases. Where a part of the standard has been
assessed to be not applicable, requirements may be added, modified, or deleted as neces-
sary. These tailoring decisions must be documented including a rationale and technical
justification for the decision.
Tailoring statements are not required if the requirements of the ESD program are within
the requirements of the standard. For example, if the maximum resistance to ground
requirement of the ESD program for a floor is 10 MΩ and the requirement of the standard
is 1 GΩ, the requirement of the standard is not exceeded, and no tailoring statement is
required. On the other hand, if the maximum resistance to ground specified for chairs in
the ESD program were 10 GΩ, and the standard required a maximum of 1 GΩ, a tailoring
statement would be required for compliance.

6.5.7 The ESD Control Program Plan


IEC 61340-5-1:2016 states that “The ESD control program shall include all the adminis-
trative and technical requirements of this standard” and “The organization shall establish,
document, implement, maintain, and verify the compliance of the program in accordance
with the requirements of this standard.”
The ESD Control Program Plan must cover
● Training
● Product qualification
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 167

● Compliance verification
● Grounding and bonding systems
● Personnel grounding
● EPA requirements
● Packaging systems
● Marking

The ESD Control Program Plan is the main document for implementing and verifying
the ESD program and must be applied to all relevant areas. It should conform to internal
quality requirements.
IEC 61340-5-1 requires that the ESD control program documents the lowest ESD with-
stand voltage that can be handled by the program. As it can be difficult to find the ESD
withstand data on all components, it may be easiest to state the standard ESD withstand
voltages of 100 V HBM and 200 V CDM in the ESD Control Program Plan.

6.5.8 Training Plan


The Training Plan must define

● The personnel required to have ESD training


● The type and frequency of training
● A requirement for maintaining training records
● Where the records are stored
● Methods used to ensure trainee comprehension and training adequacy

Initial and recurrent training must be provided to all personnel who handle or come into
contact with ESDS devices. Initial training must be provided before personnel handle ESDS.

6.5.9 Product Qualification Plan


All ESD control items that are selected for use must be qualified. Test methods and require-
ments are given in the standards. Evidence for qualification may include

● Manufacturer data sheets


● Laboratory tests, including internal and third-party tests
● Ongoing compliance verification records for items that were installed before adoption of
the standard

6.5.10 Compliance Verification Plan


The Compliance Verification Plan is intended to check that the requirements of the ESD
program are met. This plan must document the following:

● The items to be measured and verified


● Pass criteria (measured parameter limits)
● Test methods used, including those not covered by the standard
● Frequency of measurements
168 6 ESD Control Standards

Test methods may be used that differ from those given in the standard, but evidence must
be provided in a tailoring statement that the results correlate with the reference standards.
Records must be kept providing evidence of compliance with the technical requirements
of the ESD Program.

6.5.11 Test Methods


Most test methods called up by IEC61340-5-1 and ANSI/ESD S20.20 are similar or nearly
identical. Compliance verification test methods are given in ESD TR53, and it is likely that
a similar document will be produced in the IEC system in due course
Two types of tests are required – product qualification tests and compliance verification
tests. Product qualification tests are to be used only when selecting equipment for use in the
ESD Program. They would not be intended for regular checking and testing of equipment
in use. Compliance verification tests are given for this purpose.
Compliance of equipment is evaluated based on measurements. The key measurements
used are
● Resistance to ground (Rg ) or to groundable point (Rgp ).
● Point-to-point resistance (Rp-p ).
● Packaging surface (Rs ) or volume resistance (Rv ) measurements. Packaging measure-
ments using a miniature two-pin probe is treated as surface resistance.
● Packaging ESD shielding (bags).
● Measurements of electric field and potentials.
● Body voltage measurements.
● Ionizer decay time and offset voltage.
A summary of the requirements for equipment used in the EPA is given in Tables 6.4–6.8.
ANSI/ESD S20.20 also gives some requirements for soldering iron-tip resistance to
ground, tip voltage, and tip leakage current (Table 6.9).

6.5.12 ESD Control Program Plan Technical Requirements


6.5.12.1 Safety
Technical limits given in these standards do not address lower resistance limits that may
be required for safety. Due consideration should be given to whether such limits may be
needed.

Table 6.4 Grounding requirements.

IEC 61340-5-1:2016 ANSI/ESD S20.20-2014


requirements requirements

Test Required Test Required


Grounding method method limit method limit

Protective earth National electrical National electrical ANSI/ESD <1 Ω


standard code S6.1:2014a
Functional ground National electrical National electrical ANSI/ESD S6.1 <1 Ω
standard code
Equipotential bonding See Tables 6.6 See Tables 6.6 ANSI/ESD S6.1 <1 GΩ
and 6.5 and 6.5
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 169

Table 6.5 Personal grounding requirements.

IEC 61340-5-1:2016 ANSI/ESD S20.20-2014


requirements and test method requirements and test method
Product Compliance Product Compliance
Measurement qualification verification qualification verification

Wrist band Interior ≤105 Ω Not specified Interior ≤105 Ω Wrist strap
(not worn) Exterior >107 Ω Exterior >107 Ω system test is
IEC 61340-4-6 ANSI/ESD S1.1 used
Wrist strap cord < 5 × 106 Ω or Wrist strap 0.8 × 105 ≤ R ≤
user defined system test is 1.2 × 106 Ω
IEC 61340-4-6 used ANSI/ESD S1.1
Wrist strap Not specified < 5 × 106 Ω Rg < 2 Ω Rg < 2 Ω
connection point
ANSI/ESD S6.1 ESD TR53
(not monitored)
grounding and
bonding systems
Wrist strap system Not specified < 3.5 × 107 Ω < 3.5 × 107 Ω < 3.5 × 107 Ω
test IEC 61340-4-6 ANSI/ESD ESD TR53 wrist
S1.1 Sec. 6.11 strap
Continuous Not specified Not specified User defined Manufacturer
monitors defined limit
ESD TR53
continuous
monitors
Footwear < 108 Ω Person-footwear < 109 Ω Rg < 109 Ω
IEC 61340-4-3 system test is ANSI/ESD STM ESD TR53
used 9.1 Footwear
ANSI/ESD STM section
9.2
Person, footwear Body voltage Rg < 109 Ω Body voltage Footwear
and flooring <100 V AND peak <100 V and Rg < 109 Ω
IEC 61340-4-5
system test Rg < 109 Ω floor Rg < 109 Ω
Periodic ESD TR53
IEC 61340-4-5 voltage test ESD STM97.1 Footwear
and 97.2 section
Flooring
Rg < 109 Ω
ESD TR53
Flooring section
Person footwear Not specified Rg < 109 Ω
system test
IEC 61340-5-1
Annex A
170 6 ESD Control Standards

Table 6.6 Requirements for benches, floors, and seats.

IEC 61340-5-1:2016 requirements ANSI/ESD S20.20-2014 requirements


and test method and test method

Product Compliance Product Compliance


Measurement qualification verification qualification verification

Bench surface, Rp-p < 109 Ω Rg < 109 Ω Rp-p < 109 Ω N/A
racks, trolleys
IEC 61340-2-3 IEC 61340-2-3 ANSI/ESD STM4.1 R < 109 Ω
etc. surface g
Rgp < 109 Ω Rgp < 109 Ω ESD TR53
IEC 61340-2-3 ANSI/ESD STM4.1 worksurface
ESD TR53 mobile
Rg < 109 Ω equipment
ANSI/ESD STM4.1
< 200 V
ANSI/ESD
STM4.2:2012a
Floor Rgp < 109 Ω Rg < 109 Ω Rg < 109 Ω Rg < 109 Ω
IEC 61340-4-1 IEC 61340-4-1 ANSI/ESD STM7.1 ESD TR53 Flooring
section
Rgp < 109 Ω
ANSI/ESD STM7.1
Rp-p < 109 Ω
ANSI/ESD STM7.1
Chair Rgp < 109 Ω Rgp < 109 Ω Rgp < 109 Ω Rg < 109 Ω
IEC 61340-2-3 ESD STM12.1 ESD TR53 seating

Table 6.7 Requirements for ESD control garments.

IEC 61340-5-1:2016 requirements ANSI/ESD S20.20-2014 requirements


and test method and test method
Product Compliance Product Compliance
Measurement qualification verification qualification verification

Static control Rp-p ≤ 1011 Ω or Rp-p ≤ 1011 Ω or Rp-p < 1011 Ω Rp-p < 1011 Ω
garment user defined user defined ANSI/ESD ESD TR53
IEC 61340-4-9 IEC 61340-4-9 or STM 2.1 garments
or user defined user defined
Groundable Rgp < 109 Ω Rgp < 109 Ω Rgp < 109 Ω Rgp < 109 Ω
static control
IEC 61340-4-9 IEC 61340-4-9 ANSI/ESD ESD TR53
garment
STM 2.1 garments
Groundable Not specified Not specified Rg < 3.5 × 107 Ω Rg < 3.5 × 107 Ω
static control
ANSI/ESD ESD TR53
garment
STM 2.1 garments
system
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 171

Table 6.8 Requirements for ionizers.

IEC 61340-5-1:2016 requirements ESD S20.20-2014 requirements


and test method and test method
Product Compliance Product Compliance
Measurement qualification verification qualification verification

Ionizer decay <20 s decay <20 s User-defined User-defined


neutralization (1000 V to (1000 V to ANSI/ESD ESD TR53
(decay) time 100 V) or user 100 V) or user STM 3.1 ionizer
defined defined
IEC 61340-4-7 IEC 61340-4-7
Ionizer offset ±35 V ±35 V ±35 V ±35 V
voltage IEC 61340-4-7 IEC 61340-4-7 ANSI/ESD ESD TR53
STM 3.1 ionizer

6.5.12.2 Grounding and Bonding Systems


All electrically conducting (noninsulative) items that might come into contact with ESDS
must be electrically connected together or to ground. Three methods are given for this.
● Grounding using protective earth. This is the preferred option in which all electrically
conducting items and personnel are connected to the electrical system protective earth.
● Grounding using functional ground. This covers the situation where electrically conduct-
ing items and personnel are connected to a functional ground such as a ground rod. It is
recommended that this functional earth is connected to the protective earth.
● Equipotential bonding. Where a ground is not available, the electrically conducting items
and personnel may simply be electrically bonded (connected together).
The terms ground and grounding are used whichever system is selected. Only one of these
systems should be used and one grounding system present in the EPA. If different grounds
were present, they could be at different voltages and cause ESD risk. A summary of the
grounding requirements is given in Table 6.4.

6.5.12.3 Personnel Grounding


All personnel must be grounded or equipotential bonded when handling ESDS. Seated per-
sonnel must be grounded via a wrist strap system. Standing personnel may be grounded
via a wrist strap or a footwear-flooring system. Where footwear-flooring is used (Table 6.5),
personnel must wear footwear on both feet, and the maximum body voltage generation
must be <100 V and the resistance from the person’s body through footwear and flooring
to ground must be <1 GΩ.
Earlier versions of IEC 61340-5-1 and ANSI/ESD S20.20 included an option for compli-
ance based on a resistance from body to ground <35 MΩ without body voltage measure-
ment. This was discontinued in these later versions, because it was found the resistance
limit alone did not necessarily guarantee achievement of sufficiently low body voltage .

6.5.12.4 ESD-Protected Areas (EPAs)


ESDS must be handled only outside ESD-protected packaging when they are in an EPA.
EPA boundaries must be clearly identified. Access to the EPA must be limited to personnel
who have completed ESD training or are escorted by trained personnel.
172 6 ESD Control Standards

6.5.12.5 Equipment Used in the EPA


Equipment commonly used in the EPA has measurable and verifiable technical require-
ments specified in the standards (Tables 6.6–6.8). These are measured using specified stan-
dard test methods (see Section 6.5.11 and Chapter 11).

6.5.12.6 Insulators
All nonessential insulators such as cups, packaging, and personal items must be removed
from any workstation where unprotected ESDS devices are handled. (In ANSI/ESD S20.20
they must be removed from the EPA.)
Where process essential insulators are present, the ESD threat must be evaluated.

● The electrostatic field must not exceed 5 kV m–1 at the position where ESDS devices are
handled.
● Where the surface potential of an insulator exceeds 2 kV, it must be kept at least 30 cm
from any ESDS device.
● A new requirement that where the surface potential of an insulator exceeds 125 V, it must
be kept at least 2.5 cm from ESDS devices.

Any ESD risk must then be mitigated by some means e.g. using ionizers.

6.5.12.7 Isolated Conductors


Requirements have been introduced covering isolated conductors. If a conductor that con-
tacts an ESDS cannot be grounded, the voltage between the conductor and the ESDS must
be reduced to <±35 V.

6.5.12.8 Hand Electrical Soldering and Desoldering Tools


ANSI/ESD S20.20 includes requirements for hand soldering and desoldering tools
(Table 6.9). These do not appear in 61340-5-1.

Table 6.9 Requirements for soldering and desoldering hand tools (ANSI/ESD S20.20 only).

IEC 61340-5-1:2016 requirements ANSI/ESD S20.20-2014 requirements


and test method and test method
Product Compliance Product Compliance
Measurement qualification verification qualification verification

Tip to ground Not specified Not specified <2.0 Ω <10 Ω


resistance ANSI/ESD S13.1 ESD TR53
Tip voltage Not specified Not specified <20 mV soldering iron or
ANSI/ESD
ANSI/ESD S13.1
S13.1 Sec. 6.1
Tip leakage Not specified Not specified < 10 mA
current ANSI/ESD S13.1
6.5 Requirements of IEC 61340-5-1 and ANSI/ESD S20.20 Standards 173

6.5.13 ESD Packaging


ESD-protective packaging and package marking “shall be in accordance with customer con-
tracts, purchase orders, drawing or other documentation.” When these documents do not
define ESD packaging, then packaging must be defined “for all material movement within
protected areas, between protected areas, between job sites, field service operations and to
the customer.” Packaging materials classified as dissipative, conductive, insulating, elec-
trostatic (electric) field shielding or ESD shielding, and shielding packaging are defined in
related standards 61340-5-3 and ANSI/ESD S541 (Tables 6.10 and 6.11).

Table 6.10 ESD Packaging resistance classifications.

IEC 61340-5-3:2015 requirements ANSI/ESD S541-2018 requirements


and test method and test method

Classification Surface (𝛀) Volume (𝛀) Surface (𝛀) Volume (𝛀)

Packaging – 104 Ω ≤ Rs < 1011 104 Ω ≤ Rv < 1011 104 Ω ≤ Rs < 1011 104 Ω ≤ Rv < 1011
Static IEC 61340-2-3 IEC 61340-2-3 STM11.11 STM11.12
dissipative
STM11.13
Packaging – Rs ≤ 104 Ω Rv < 104 Ω Rs < 104 Ω Rv < 104 Ω
conductive IEC 61340-2-3 IEC 61340-2-3 STM11.11 STM11.12
STM11.13
Packaging – Rs ≥ 1011 Ω Rv ≥ 1011 Ω Rs ≥ 1011 Ω Rs ≥ 1011 Ω
insulator IEC 61340-2-3 IEC 61340-2-3 STM11.11 STM11.12
STM11.13
“The user should determine
whether or not a specific packaging
configuration provides for a
Electrostatic Rs < 103 Ω Rv < 103 Ω
reduction of electric field strength at
(Electric) field IEC 61340-2-3 IEC 61340-2-3 the position in the package where
shielding
sensitive items are contained.”

Table 6.11 ESD packaging.

IEC 61340-5-3:2015 requirements ANSI/ESD S541-2018 requirements


Classification and test method and test method

ESD shielding (bags) < 50 nJ < 20 nJ


IEC 61340-4-8 ANSI/ESD S11.31
Low charging Not defined User defined.
(antistatic) Material having reduced amount
of charge accumulation compar-
ed to standard packaging
materials.
ESD ADV11.2
174 6 ESD Control Standards

In ANSI/ESD S20.20, where ESDS devices are placed on packaging and work performed
on them, the packaging is considered to be a work surface and work surface requirements
for resistance to ground apply.

6.5.14 Marking
Marking of ESDS devices, systems, or packaging must be according to customer contracts,
purchase orders, or other documentation. Where these are not defined, then the need for
marking should be considered in the ESD Control Program Plan. Where marking is needed,
it must be documented in the ESD Program Plan.

References

British Standards Institute (1984) BS 5783:1984. Code of practice for handling of electrostatic
sensitive devices. London, BSI.
British Standards Institute (1992) BS EN 100015-1:1992. Basic specification. Protection of
electrostatic sensitive devices. Harmonized system of quality assessment for electronic
components. Basic specification: protection of electrostatic sensitive devices. General
requirements. London, BSI.
British Standards Institute (1994a) BS EN 100015-2:1994. Basic specification. Protection of
electrostatic sensitive devices. Requirements for low humidity conditions. London, BSI.
British Standards Institute (1994b) BS EN 100015-3:1994. )Basic specification. Protection of
electrostatic sensitive devices. Requirements for clean room areas.London, BSI.
British Standards Institute (1994c) BS EN 100015-4:1994. Basic specification. Protection of
electrostatic sensitive devices. Requirements for high voltage environments. London, BSI.
British Standards Institute (2001) BS EN 61340-5-1:2001. Electrostatics. Protection of electronic
devices from electrostatic phenomena. General requirements. London, BSI.
British Standards Institute (2016) BS EN 61340-5-1:2016. Electrostatics. Protection of electronic
devices from electrostatic phenomena. General requirements. London, BSI.
Department of Defense. (1980a) MIL-STD-1686. Standard Practice. Electrostatic discharge
control program for protection of electrical and electronic parts, assemblies and equipment
(excluding electrically initiated devices). Washington, D.C., DoD.
Department of Defense. (1980b) MIL-HDBK-263. Military handbook. Electrostatic discharge
control handbook for protection of electrical and electronic parts, assemblies and equipment
(excluding electrically initiated devices. Washington, D.C., DoD.
EOS/ESD Association Inc. (1999) ANSI/ESD S20.20-1999. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2003) ESD SP9.2-2003. ESD Association Standard for the Protection
of Electrostatic Discharge Susceptible Items – Footwear – Foot Grounders Resistive
Characterization (not to include static control shoes). Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2006a) ANSI/ESD STM4.1-2006. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Worksurfaces - Resistance
Measurements. Rome, NY, EOS/ESD Association Inc.
References 175

EOS/ESD Association Inc. (2006b) ESD TR53-01-06. Technical Report for the protection of
electrostatic discharge susceptible items - Compliance Verification of ESD Protective Equipment
and Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2012a) ANSI/ESD STM4.2-2012. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – ESD Protective Worksurfaces – Charge
Dissipation Characteristics. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2012b) ANSI/ESD STM11.31-2012. ESD Association Standard Test
Method for Evaluating the Performance of Electrostatic Discharge Shielding Materials – Bags.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013a) ANSI/ESD STM12.1-2013. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items - Seating - Resistance
Measurement. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013b) ANSI/ESD STM7.1-2013. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Floor Materials - Resistive
Characterization of Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013c) ANSI/ESD S1.1-2013. Standard for protection of Electrostatic
Discharge Susceptible Items - Wrist Straps. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013d) ANSI/ESD STM2.1-2013. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Garments. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2014a) ANSI/ESD S6.1-2014. Standard for the Protection of
Electrostatic Discharge Susceptible Items – Grounding. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014b) ANSI/ESD STM9.1-2014. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Footwear - Resistive Characterization.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014c) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015a) ANSI/ESD STM3.1-2015. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Ionization. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2015b) ANSI/ESD STM11.11-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items – Surface Resistance Measurement of
Static Dissipative Planar Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015c) ANSI/ESD STM11.12-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items.Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015d) ANSI/ESD STM11.13-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Two-Point Resistance
Measurement. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015e) ANSI/ESD STM97.1-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Floor Materials and
Footwear – Resistance Measurement in Combination with a Person. Rome, NY, EOS/ESD
Association Inc.
176 6 ESD Control Standards

EOS/ESD Association Inc. (2016a) ANSI/ESD STM97.2-2016. Floor Materials and


Footwear – Voltage Measurement in Combination with a Person. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2016b) ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies, and Equipment. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2017) ESD ADV1.0-21017. ESD Association Advisory for
Electrostatic Discharge Terminology – Glossary. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2018) ANSI/ESD S541-2018. Packaging Materials for ESD Sensitive
Items. Rome, NY, EOS/ESD Association Inc.
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering component
level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf (Accessed: 10th May
2017).
International Electrotechnical Commission. (1998) IEC 61340-5-1:1998. Electrostatics - Part 5-1:
Protection of electronic devices from electrostatic phenomena - General requirements. Geneva,
IEC.
International Electrotechnical Commission. (2001) IEC 61340-4-3:2001. Electrostatics - Part 4-3:
Standard test methods for specific applications – Footwear. Geneva, IEC.
International Electrotechnical Commission. (2003) IEC 61340-4-1:2003+AMD1:2015 CSV.
Electrostatics - Part 4-1: Standard test methods for specific applications - Electrical resistance of
floor coverings and installed floors. Geneva, IEC.
International Electrotechnical Commission. (2004) IEC 61340-4-5:2004. Electrostatics - Part 4-5:
Standard test methods for specific applications - Methods for characterizing the electrostatic
protection of footwear and flooring in combination with a person. Geneva, IEC.
International Electrotechnical Commission. (2007) IEC TR 61340-5-2:2007. Electrostatics – Part
5-2: Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
International Electrotechnical Commission. (2013a) PD/IEC TS 60079-32-1. Explosive
atmospheres Part 32-1. Electrostatic hazards, guidance. Geneva, IEC.
International Electrotechnical Commission. (2013b) PD IEC/TS 61340-4-2:2013. Electrostatics -
Part 4-2: Standard test methods for specific applications - Electrostatic properties of garments.
Geneva, IEC.
International Electrotechnical Commission. (2014) IEC 61340-4-8:2014. Electrostatics - Part 4-8:
Standard test methods for specific applications - Electrostatic discharge shielding – Bags.
Geneva, IEC.
International Electrotechnical Commission. (2015a) IEC 61340-2-1:2015. Electrostatics - Part
2-1: Measurement methods - Ability of materials and products to dissipate static electric charge.
Geneva, IEC.
International Electrotechnical Commission. (2015b) IEC 61340-4-6:2015. Electrostatics - Part
4-6: Standard test methods for specific applications - Wrist straps. Geneva, IEC.
International Electrotechnical Commission. (2015c) IEC 61340-5-3:2015. Electrostatics.
Protection of electronic devices from electrostatic phenomena. Properties and requirements
classifications for packaging intended for electrostatic discharge sensitive devices. Geneva, IEC.
Further Reading 177

International Electrotechnical Commission. (2016a) IEC 61340-2-3:2016. Electrostatics. Part


2-3: Methods of test for determining the resistance and resistivity of solid materials used to avoid
electrostatic charging. Geneva, IEC.
International Electrotechnical Commission. (2016b) IEC 61340-4-9:2016. Electrostatics - Part
4-9: Standard test methods for specific applications – Garments. Geneva, IEC.
International Electrotechnical Commission. (2016c) IEC 61340-5-1:2016. Electrostatics - Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Electrotechnical Commission. (2017) IEC 61340-4-7:2017. Electrostatics - Part 4-7:
Standard test methods for specific applications – Ionization. Geneva, IEC.
International Electrotechnical Commission. (2018) IEC TR 61340-5-2. Electrostatics – Part 5-2:
Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
International Electrotechnical Commission. (2019) IEC TR 61340-5-4. Electrostatics – Part 5-4:
Protection of electronic devices from electrostatic phenomena - Compliance verification.
Geneva, IEC.

Further Reading

EOS/ESD Association Inc. (1995a) ESD ADV53.1-1995. Advisory for Protection of Electrostatic
Discharge Susceptible Items - ESD Protective Workstations. Rome, NY, EOS/ESD Association
Inc.
EOS/ESD Association Inc. (1995b) ESD ADV11.2-1995. Advisory for the Protection of
Electrostatic Discharge Susceptible Items - Triboelectric Charge Accumulation Testing. Rome,
NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (1999a) ESD TR13.0-01-99. Technical Report - EOS Safe Soldering
Iron Requirements. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (1999b) ESD TR15.0-01-99. Standard Technical Report for the
Protection of Electrostatic Discharge Susceptible Items-ESD Glove and Finger Cots. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (1999c) ESD TR50.0-01-99. Technical Report - Can Static Electricity
be Measured? Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (1999d) ESD TR50.0-02-99. Technical Report - High Resistance
Ohmmeter Measurements. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2000a) ESD TR2.0-01-00. Technical Report - Consideration For
Developing ESD Garment Specifications. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2000b) ESD TR2.0-02-00. Technical Report - Static Electricity
Hazards of Triboelectrically Charged Garments. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2001) ESD TR1.0-01-01. Technical Report - Survey of Constant
Monitors for Wrist Straps. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2002a) ESD TR3.0-01-02. Technical Report - Alternate Techniques for
Measuring Ionizer Offset Voltage and Discharge Time. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2002b) ESD TR4.0-01-02. Technical Report - Survey of Worksurfaces
and Grounding Mechanisms. Rome, NY, EOS/ESD Association Inc.
178 6 ESD Control Standards

EOS/ESD Association Inc. (2002c) ESD TR10.0-01-02. Technical Report - Measurement and ESD
Control Issues for Automated Equipment Handling of ESD Sensitive Devices Below 100 Volts.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2003) ESD TR50.0-03-03. Technical Report - Voltage and Energy
Susceptible Device Concepts, Including Latency Considerations. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2004) ESD TR55.0-01-04. Technical Report - Electrostatic Guidelines
and Considerations For Cleanrooms and Clean Manufacturing. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2005) ESD TR3.0-02-05. Technical Report - Selection and Acceptance
of Air Ionizers. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2011a) ANSI/ESD SP15.1-2011. Standard Practice for the Protection
of Electrostatic Discharge Susceptible Items – In-Use Resistance Measurement of Gloves and
Finger Cots. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2011b) ESD TR7.0-01-11. Technical Report for the Protection of
Electrostatic Discharge Susceptible Items – Static Protective Floor Materials. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2012) ANSI/ESD S11.4-2012. Standard for the Protection of
Electrostatic Discharge Susceptible Items - Static Control Bags. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2015a) ANSI/ESD S13.1-2015. Provides electrical
soldering/desoldering hand tool test methods for measuring current leakage, tip to ground
reference point resistance, and tip voltage. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015b) ESD TR17.0-01-15. Technical Report for ESD Process
Assessment Methodologies in Electronic Production Lines – Best Practices used in Industry.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015c) ESD TR53-01-15. Technical Report for the Protection of
Electrostatic Discharge Susceptible Items – Compliance Verification of ESD Protective
Equipment and Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016a) ANSI/ESD SP3.3-2016. Standard Practice for the Protection
of Electrostatic Discharge Susceptible Items – Periodic Verification of Air Ionizers. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016b) ANSI/ESD SP3.4-2016. Standard Practice for the Protection
of Electrostatic Discharge Susceptible Items – Periodic Verification of Air Ionizer Performance
Using a Small Test Fixture. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016c) ESD SP10.1-2016. Standard practice for protection of
Electrostatic Discharge Susceptible Items – Automated handling Equipment (AHE). Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2017a) ESD ADV1.0-2017. ESD Association Advisory for
Electrostatic Discharge Terminology – Glossary. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2017b) ANSI/ESD S8.1-2017. Draft Standard for the Protection of
Electrostatic Discharge Susceptible Items – Symbols – ESD Awareness. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2019) ESD Fundamentals Part 6: ESD Standards. Available from:
https://www.esda.org/index.php/about-esd/esd-fundamentals/part-6-esd-standards
(Accessed 26th January 2019).
179

Selection, Use, Care, and Maintenance of Equipment and


Materials for ESD Control

7.1 Introduction

Electrostatic discharge (ESD) control equipment can form a considerable part of the
investment cost of an ESD control program. It is important to be sure that the equipment
will do the task intended, for the intended life of the equipment. There may be many types
of similar equipment on the market, and the type chosen should be appropriate for use
in the facility and processes intended. Cleaning, care, and maintenance may be needed to
ensure that the equipment continues to work correctly over its life.
Unless equipment is single use and disposable, it is important that any failures of equip-
ment are detected, with failed equipment taken out of use for maintenance or replaced.
So, most equipment will require compliance verification testing at regular intervals.
Appropriate test methods and pass criteria must be defined. This is discussed further in
Chapter 9, and some test methods are discussed in Chapter 11.

7.1.1 Selection and Qualification of Equipment


The first task in selecting ESD control equipment is to identify its intended functions.
Aspects to be considered may include
● Physical or process task. What is the equipment there to do?
● ESD control functions and characteristics required.
● Safety considerations.
● Compliance with standards, if required.
● Tests or other methods of confirmation of functionality and ESD control characteristics.
● Intended life. Disposable or long-term use?
● Care and maintenance requirements.
● For long-term usage, compliance verification requirements and test methods. How will
failure be detected?
The second step is to decide how selected items may be qualified to confirm that they will
fulfill their intended function for their intended life. Product qualification can be done in
various ways and should address all the aspects that have been identified important.
In the simplest case, it may be sufficient to check data sheets to ensure that the required
functions and characteristics are addressed and that the product complies with relevant
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
180 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

standards. The standards may be to do with ESD control or any other aspect such as safety.
For example, footwear that is personal protective equipment (PPE) may, in Europe, need to
be compliant with a standard such as ISO 20345.
In some cases, it may be advisable to obtain one or more specimens of items for evalua-
tion or test. This may be done in-house or by a third party. As they are done only initially to
qualify the product before use, the qualification tests may be more in-depth than compli-
ance verification tests. They often are designed to check individual characteristics of parts
of the equipment in detail. For example, in qualifying an ESD control chair, it may be desir-
able to check that the seat back and arms as well as the seat are made of static dissipative
material and connected reliably together and to a groundable point, feet, or wheels. It may
be necessary to establish that more than one foot or wheel is made of noninsulating material
and capable of establishing a ground path to the floor.
In qualification tests, it is sometimes desirable to test the item isolated from its normal
working situation in an uninstalled condition. For example, a work surface point-to-point
surface resistance may be tested as well as the resistance from the surface to a groundable
point. The latter test does not require connection to ground. Often, it will be desirable to
qualify the item under “worst case” operating conditions. This usually means performing
tests under dry atmosphere <30% rh.
Conversely, in some qualification tests, it is desirable to test the operation of the equip-
ment as a system with other ESD control equipment. A good example of this is measurement
of body voltage generated on a subject wearing ESD control footwear and the flooring with
which it will be used. This test can be done only with representative samples of the footwear
and flooring under qualification (see Section 7.5.4).
Where compliance with a standard is required, testing should be done according
to the test methods required by the standard, and compliance with the standard pass
criteria should be established. Standards will often specify qualification testing under dry
atmosphere conditions. IEC 61340-5-1 and ANSI/ESD S20.20 usually require qualification
testing at 12 ± 3% rh (International Electrotechnical Commission 1998, 2016b; EOS/ESD
Association Inc 2014b).

7.1.2 Use
The intended use of equipment can be an important consideration, first in whether it is
needed at all and second in selecting its required characteristics. ESD control equipment
often acts as part of a system with other equipment. Careful thought should be given to
how the equipment is used and if it will help to give a convenient working environment
to the personnel in the area. Analysis of this can affect the combination of the ESD control
equipment selected. ESD control practices that are selected to be convenient to the operator
are more likely to be reliable in practice. This is discussed further in Chapter 10.

7.1.3 Cleaning, Care, and Maintenance of Equipment


It is important, when selecting ESD control equipment, to consider whether it may need
specific cleaning, care, and maintenance procedures and materials. These may differ from
similar equipment in unprotected areas. For example, ESD control floors and workstation
7.2 ESD Control Earth (Ground) 181

surfaces may need different cleaning procedures and materials than those in unprotected
areas, selected to preserve and enhance their ESD control properties.
Some equipment such as ionizers may need regular maintenance to keep their
performance in the long term.
Equipment that relies on grounding through feet or wheels often collects dirt on these
contact points, which can make the ground path fail or intermittent. Cleaning the wheels
or feet can bring the equipment back into compliance.

7.1.4 Compliance Verification


Equipment that is not single use only will need to have its properties tested and verified on
a regular basis to detect any failures or deviations from intended performance.
Compliance verification tests are normally simple efficient tests designed to test func-
tionality and performance in the installed situation. Where equipment works as part of a
system, it is often tested as a system. As these tests are repeated regularly and often many
times during an audit, it is important that they are quick and effective.
For example, a simple resistance to ground test is used to simultaneously test the condi-
tion of a work surface or floor and its ground connection. For a chair standing on an ESD
control floor, a simple resistance to ground measurement from the seat to the electrostatic
discharge protected area (EPA) ground simultaneously checks the chair and the connection
through its wheels and the floor to ground. Problems such as buildup of dirt on the wheels
are detected as an excessively high resistance reading.

7.2 ESD Control Earth (Ground)

7.2.1 What Does the ESD Control Earth Do?


A primary objective in ESD control is to bring all electrically conducting equipment, mate-
rials, and personnel to the same voltage so that they cannot become charged and become
an ESD source. This is done by connecting them to a common connection point called
ground. This connection of all conductors to one common point is called equipotential
bonding and establishes “zero volts” for the purposes of ESD control. Where conductors
are connected, electrostatic voltage differences are quickly equalized between them.
Elimination of voltage difference between eliminates the possibility of ESD occurring if
contact is made between them.

7.2.2 Choosing an ESD Control Earth


The common connection point is commonly called ground and is usually, but not always,
also connected electrically to the physical earth. One common and convenient way to do
this is to connect to the mains electricity supply protective safety earth. A second method
sometimes used is to connect to an earth rod buried in the ground for the purpose, known in
the ESD control standards as functional earth. If it is not possible to connect to a convenient
earth, for example in a vehicle or aircraft, then equipotential bonding alone will be sufficient
for ESD control.
182 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

It is undesirable to have more than one earth present in a facility. If there are multiple
earths and they are not connected together, they are likely to be at different voltages due
to earth currents or other phenomena. Often, there can be a variety of equipment in the
EPA that is already connected to mains protective “safety” earth. In this case, it is usually
convenient to also use mains protective earth as the ESD control earth. More information
on electrical power systems can be found in standards such as IEC 60364-1 and national
electrical codes of practice and regulations (International Electrotechnical Commission
2005).
If there is no mains protective earth available in the EPA, it can be convenient to use a
separate earth rod buried in the earth as the ESD control earth.
It is preferable to connect individual equipment (e.g. workstations) separately to ground
in “star” connection form (Figure 7.1) rather than “chain” or series connection (Figure 7.2).
This is because with the “chain” connection, failure of one connection can lead to several
ESD control items being disconnected from earth.
An exception to this is where an ESD control item (e.g. ESD floor) is designed to provide
a ground path for equipment (e.g. chairs or carts/trolleys) or personnel grounded through
it as a grounding system.

Figure 7.1 “Star” connection of ESD control benches.

Figure 7.2 “Chain” or series connection of ESD control benches.


7.3 The ESD Control Floor 183

7.2.3 Qualification of ESD Control Earth


Where the electrical protective earth is used as ESD ground, the qualification requirements
are likely to depend on national electrical code requirements. These should be tested by
suitably qualified personnel. It is extremely important to verify that the connection is made
to the earth and not live or neutral power connections.
Even where a temporary or semi-permanent connection to a mains power socket via an
earthing plug is used, it is wise to check that the socket is correctly wired before use.
The electrical connection between ESD ground and the earth connection should be
verified to be sufficiently robust and low resistance. ESD control standards 61 340-5-1 and
S20.20 also give some requirements that can be applied if national regulations are not
applicable.

7.2.4 Compliance Verification of ESD Control Earth


The electrical connection between ESD ground and the protective or functional earth con-
nection should be verified to be sufficiently low resistance. National regulations may require
this to be done by suitably qualified personnel.

7.2.5 Common Problems with Ground Connections


Mains sockets have occasionally been found to be incorrectly wired. Even where a tempo-
rary or semi-permanent connection to a mains power socket via an earthing plug is used, it
is wise to check that the socket is correctly wired before use.
Plug-in mains ground connectors are sometimes unplugged to allow plug-in of other
equipment and then inadvertently left disconnected. For example, a cleaner might remove
a ground plug to plug in a vacuum cleaner, or an engineer might do so to connect some
equipment.
Even hardwired ground connections can sometimes be disconnected by inadvertent and
undetected damage. For example, moving a bench might cause undetected disconnection
of a hardwired ground lead.

7.3 The ESD Control Floor


7.3.1 What Does an ESD Control Floor Do?
An ESD control floor, if required, can represent a large investment. It can also give signifi-
cant benefit and convenience in ESD control.
ESD control floors are used to give a convenient way of grounding anything that stands
on the floor, such as personnel, carts (trolleys), racks, chairs, and other free-standing equip-
ment (see Section 4.7.10). ESD control floor coverings typically provide an electrical path
between the floor surface and EPA ground via an underlying substrate material or ground-
ing structure. Floor finishes or treatments may reduce the tendency for charge generation
at the surface as well as providing an electrical path for charge to move to ground.
184 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

A floor also acts as a means of conjoining workstations or areas to make a single EPA
with no uncontrolled area between them. This can improve the convenience of handling of
electrostatic discharge–sensitive (ESDS) by eliminating the need for protective packaging
for transport between workstations.
The floor operates as a system with the items that it is intended to ground. So, in specifying
the characteristics of a floor, it is necessary to consider the items that will be grounded by
it. Typical systems using the floor include

● An operator’s body grounded through footwear and flooring


● A cart grounded from cart surface through the chassis, wheels, and floor
● A rack grounded from the shelf surface through the frame, feet, and floor
● A seat grounded from the seat surfaces through the frame, feet or wheels, and floor

Floors and floor covering materials of course have many non-ESD-related functions. They
must withstand the physical wear and tear expected for the lifetime required of the floor.
This may include considerations like use of forklift trucks and vehicles, or withstanding use
of chemicals. Floor materials for use in clean areas may need to be selected for low contri-
bution to particulate or other contaminants in the area. These factors must be included in
the overall selection of the floor material.

7.3.2 Permanent ESD Control Floor Material


Rubber and vinyl materials are among the most widely used floor coverings, supplied either
as sheet or as tiles. Some materials may be slippery when wet. Some may not withstand
heavy traffic. Some materials may contain carbon or may outgas, making them unsuitable
for clean room use.
Epoxy and polymeric coatings form a poured-on robust coating resistant to chemicals,
abrasion, and vehicle traffic wear. They are seamless, can be easy to maintain, and are suit-
able for clean room use.
High-pressure laminates are usually used as raised floors or floor mats. As they are often
sensitive to moisture, they may be unsuited for use where chemical or water spillage is
likely or on concrete substrates where high moisture levels could arise. They may change
considerably in resistance in response to changes in moisture levels.
Carpets may be attractive and absorb noise and can have lower maintenance requirement
than resilient materials. They are available in sheet or tile form. In tile form, individual worn
or contaminated tiles may be easily replaceable. Carpets may not, however, be suitable for
use in clean rooms or where heavy soiling or wear, water, chemical spills or hot solder, or
vehicle traffic may occur.

7.3.3 Semipermanent or Nonpermanent ESD Control Floor Materials


Semi- or nonpermanent ESD control floor materials include interlocking tiles, mats, floor
finishes, paints, coatings, and topical antistats. They can provide a quick and easy floor
covering solution, especially for small or temporary floor areas. They may have a relatively
short life and require regular retreatment or replacement.
7.3 The ESD Control Floor 185

Floor mats can be an easy and portable solution, especially for small areas, for example
around a single workstation. They can also provide a replaceable “sacrificial” surface
around an area where high contamination or solder damage is expected. Comfort floor
mats may be provided where operators are standing for long periods at workstations.
Some types of mat can tend to curl and ruck and form a trip hazard. They are often con-
nected to ground via a wire attached to a bonding point, and this may act as a trip hazard
or be prone to accidental disconnection.
Paints and coatings can be applied to existing floors such as concrete, giving easy applica-
tion and coverage of large areas. They have intermediate lifetime and tend to wear, requiring
reapplication. Some materials may not be suitable for clean areas due to their carbon load-
ing or tendency to abrade and shed particles.
Floor finishes such as topical antistats are treatments that are easy to apply and can be
applied to carpets. Their effect is usually temporary and requires regular refreshment. Their
effect often depends on moisture from the air, and they may become ineffective in low rel-
ative humidity atmospheric conditions. Some finishes may tend to be slippery and cause
slip hazard to personnel. Some may be susceptible to easily being washed off or worn away.
Careless application and maintenance may result in unreliable performance. Some finishes
may not be usable in clean room areas due to contamination issues.

7.3.4 Selection of Floor Materials


In selecting a floor material, the first considerations must be about how it will be used and
the processes that will occur there. Considerations may include
● Is it to be a permanent area or a temporary area?
● What are the activities and processes anticipated?
● Will it be a new installation or a cover to an existing floor?
● Will it be a clean environment in which particle shedding or other contamination would
be a concern?
● Are chemicals to be used in the area?
● Are there ergonomic considerations?
● What will be the maintenance and cleaning regime?
● Are there electrical considerations such as high voltage use?
● Are there possible safety considerations?
● Are there other safety or legislative requirements to consider?
● Will it need to support forklifts or vehicles or have other heavy usage considerations?
Where a floor may be used for handling solvents, explosives, or other flammable materials,
the floor may need to be specified for safety in this use. Local safety regulations may take
precedence.

7.3.5 Floor Material Qualification Test


Product qualification test of a floor should aim to test its performance as part of a system
with the equipment with which it is intended to be used, under the lowest atmospheric
humidity conditions likely in practice. Small specimens (e.g. 1 × 2 m) of coverings and coat-
ings or mats may be tested under low-humidity laboratory conditions. The tests might
include
186 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

● Resistance to a groundable point or simulated grounding structure, made at several points


on the surface. This is to get an indication of the likely installed resistance to ground.
● Point-to-point resistance, made at several positions and orientations on the surface.
● Resistance from body to ground (or groundable point) of a person wearing selected ESD
control footwear intended to be used in practice and standing on the flooring specimen.
● A walk test on the specimen by a person wearing the footwear intended to be used in
practice to test the body voltage generation of the selected footwear-flooring combina-
tion(s).
● If carts, seats, or other equipment are to be grounded through the floor, it is wise to test
that reliable grounding is achieved of the types of equipment intended to be used.
Suitable test methods are specified in standards such as IEC 61340-4-1, IEC 61340-4-5,
ANSI/ESD STM 97.1, and ANSI/ESD STM 97.2. (International Electrotechnical Commis-
sion 2003/2015, international Electrotechnical Commission 2005, EOS/ESD Association
Inc 2015, EOS/ESD Association Inc 2016)
In laboratory tests, the floor material is typically conditioned at the test humidity and
temperature required by the test standard (e.g. 12% rh 23 ± 3 ∘ C) for at least 24 hours before
testing under the same conditions. In some standards, the required conditioning time may
be as much as 48 or even 72 hours. The conditions may be determined by the test standard or
by the user’s requirements. One or more test specimens are mounted in a manner suitable
for test as determined by the test method or standard. This may involve adding groundable
points or a grounding method simulating the installed condition.
Resistance measurements are made using standard electrodes placed on the material
surface. The measurement is made at 10 V for resistance less than 1 MΩ, and 100 V for
resistance above 1 MΩ.
In a point-to-point measurement, two electrodes are placed on the surface 30 cm apart,
or some other distance defined by the standard. The resistance between the electrodes is
measured (see Section 11.8). This is repeated in several positions and orientations.
In the resistance to groundable point measurement, a single electrode is placed on the
specimen surface. The resistance between this and a groundable point or simulated ground
connection point is measured.
Where the floor is used for grounding personnel via ESD control footwear, qualification
tests of the footwear and flooring in combination will be required (Section 7.5.4). In measur-
ing the resistance from body to ground with a footwear-flooring combination, a test subject
wears the footwear under test. They stand on the flooring specimen, holding a handheld
electrode. The resistance between the handheld electrode and the floor specimen ground
point is measured. The maximum resistance is usually required to be below a required
value. Sometimes, the minimum resistance must also be above a required value, e.g. for
safety reasons.
With the body voltage walk test, the test subject wearing the footwear under test walks
on the flooring specimen under test. The floor specimen is grounded via the groundable
point. The walk style and footstep pattern to be used by the subject is determined in
some standards. The test subject holds a handheld electrode connected to an electrostatic
voltmeter and means of recording the voltage. The highest peaks in body voltage during
the test are recorded.
The final resistance to ground and performance of an installed floor material will
typically depend to some extend on the underlying substrate, grounding method, and other
7.3 The ESD Control Floor 187

installation conditions. For this reason, it can be advisable to test the proposed material
as a small installed area before deciding to install a large area of material. Some materials
(e.g. treatments and coatings) can by their nature be tested only in the installed condition.
There may also be concerns over whether a treatment could affect other aspects such
as the appearance of a treated material. In this case, it may be advisable to test a small
area before treating the full proposed application area. Qualification test on an installed
floor material will usually have to be done under ambient conditions. If possible, some
tests should be done under worst-case (e.g. low-humidity) conditions. Testing typically
includes
● Resistance to ground of the installed material, made at several points on the surface
● Resistance from body to ground (or groundable point) of a person wearing selected ESD
control footwear intended to be used in practice and standing on the surface
● A walk test on the specimen by a subject wearing the footwear intended to be used
in practice to test the body voltage generation of the selected footwear-flooring
combination(s)
ESD control standards may require these tests to be done as part of their ESD control
product qualification requirements. Where compliance with a standard is required, testing
should be done according to the standard, and compliance with the standard pass criteria
should be established.

7.3.6 Acceptance of a Floor Installation


After installation, a floor should be tested before use to make sure that it meets the required
performance. These tests are normally done under ambient conditions. If possible, some
tests should be done under worst-case (e.g. low-humidity) conditions.
The tests typically include resistance to ground test at several points on the surface. They
may also include test of resistance from body to ground and a walk test by a subject wearing
the footwear intended to be used in practice to confirm the body voltage generation of the
selected footwear-flooring combination(s).

7.3.7 Use of Floor Materials


The purpose of an ESD control floor is to provide a convenient means of grounding per-
sonnel, racks, seats, carts, or other items standing on the floor. To get the best from an
investment in an ESD control floor, it is important to make sure that full advantage of this
is taken when specifying equipment for use on the floor within the EPA.

7.3.8 Care and Maintenance of Floors


Dirt and contamination on a floor can considerably change both its charge generation and
its charge dissipation characteristics. Regular cleaning may be necessary to keep a floor
working at its optimum.
Use of ordinary floor cleaning products can change and impair the characteristics of a
floor. The cleaning processes and materials used should be specified according to the man-
ufacturer’s instructions to maintain good performance.
188 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

It is important to use a cleaning regime that will preserve and maintain the floor elec-
trostatic characteristics. Use of incorrect cleaning materials or polish can leave a coating
of wax or another contaminant material on it. This can seriously affect the surface charge
generation characteristics as well as the surface contact resistance of the floor. Guidance
should be obtained from the material manufacturer or supplier regarding suitable cleaning
methods and materials.
Where water or liquid cleaners have been used, leave to dry well before retest of the floor,
as the presence of residual moisture can dramatically affect the measured resistance value.

7.3.9 Compliance Verification Test


Floor compliance verification test is normally done using a simple resistance to ground
measurement (see Section 11.8). A standard electrode is placed on the floor surface.
A resistance meter is connected between this electrode and an ESD earth connection
point. Standards often require a minimum number of measurements to be made on a
floor or for a given floor area. It is good practice to also measure and note the atmospheric
humidity in the area tested. This, together with the measurement results, will over time
give an indication of variation of the floor resistance with humidity and show if the floor
resistance to ground goes out of specification under low-humidity conditions.
The frequency of testing should reflect the likelihood that the floor performance is
expected to change over time. For example, a permanent floor in a clean area, with a
history of no change over a long time, may need to be tested infrequently. In contrast, a
more temporary material subject to change, wear, or contamination may need testing on
a more frequent basis. An area subject to extreme wear or contamination may need to
be tested frequently. A mat subject to accidental disconnection may need frequent visual
checks in between electrical tests.
IEC 61340-5-2:2018 suggested a typical time interval of three months between compliance
verification tests.

7.3.10 Common Problems


Surface contamination such as dust, chemicals, or sprayed materials or use of incorrect
polishes or cleaning materials can seriously affect the surface charge generation character-
istics as well as the resistance to ground of the floor covering material. This can dramatically
increase the resistance to ground of equipment and personnel standing on the floor, and the
body voltage of walking personnel.
If a resistance to ground test has been failed, a first step is to make a visual check.
Any surface dirt or contamination should be cleaned off using a suitable cleaning
procedure. The test electrode itself should be checked, as this can also become contam-
inated. For a temporary installation such as a mat, the earth bonding system should be
checked.
Where water or liquid cleaners have been used, they should be left to dry before retest of
the floor. Moisture can dramatically affect the measured resistance value.
Electrodes can be initially cleaned of dust using a dry paper towel or clean cloth. If this is
insufficient, an alcohol wipe (if compatible with the electrode material) can give stronger
7.4 Earth Bonding 189

cleaning action while being quick in drying. The electrode should be allowed to dry com-
pletely before further measurements are made.
If areas of excessive resistance to ground are found on a floor, the area should be first
cleaned to check whether it may be due to contamination. If this does not restore the floor
performance, an evaluation should be made of the size of the area affected, whether the
performance has changed, whether low humidity may be a factor, and whether there may
be other areas in the process of deterioration. Depending on the results of this evaluation,
remedial action may need to be considered.
Small areas of high resistance can sometimes be due to a previously undetected problem
in a floor material or installation. If the area affected is very small and the effect on ESD con-
trol is minimal, then the position of the problem area may be noted, and the noncompliance
might be acceptable as insignificant. A key question here is whether the noncompli-
ant area prevents reliable grounding of an ESD control item, leading to ESD risk to
ESDS devices.

7.4 Earth Bonding


7.4.1 The Role of Earth Bonding Points
Earth bonding (grounding) points provide a way of connecting ESD control equipment
to EPA earth via earth bonding cords. They are available with, or without, built-in series
resistors (Figure 7.3). They may have a variety of connector types for connecting ESD
control equipment. The common connector types include 4 mm “banana” sockets and
10 mm circular studs.
Earth bonding points may be hardwired into a workstation or other equipment or may be
provided as a plug-in unit to connect to electrical mains sockets. Some types can be screwed
onto the underside of a bench and include a flying lead tail that can be hardwired to a
convenient earthing point.

Figure 7.3 A typical earth bonding (grounding) plugs that can be used to connect to mains
protective earth.
190 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

7.4.2 Selection of Earth Bonding Points


The main questions to be addressed in selecting earth bonding points are
● What will the bonding point connect to? Will it be hardwired in or make a temporary
connection?
● What format connection sockets are required by the equipment to be connected to the
bonding point?
● What is the maximum resistance to ground required through the bonding point?
● Is there a minimum resistance to ground required through the bonding point?
Where earth bonding connectors used for connecting measurement instruments to ground,
a connector without a series resistor included is preferable. Any internal resistance is added
to the measured resistance.

7.4.3 Qualification of Earth Bonding Points


Qualification should check that the connector form is as required and whether the correct
value series resistor is included (if required).

7.4.4 Use of Earth Bonding Points


For earth bonding points that plug into mains electrical sockets, it can be wise to do a visual
check that the plug is in place and connected before use. These items are often unplugged
by personnel wanting to use the mains socket for other equipment.
Before using a mains socket to provide ESD earth connection, check that the socket is
correctly wired. This can be done using a simple proprietary mains socket tester.

7.4.5 Compliance Verification of Earth Bonding Points


It is important to verify periodically that earth bonding points remain functional and
within specification. This can be done using a simple resistance to ground measurement
(see Section 11.8).
The frequency of testing chosen may depend on how reliable the bonding point is
expected to be. Plug-in connectors that could be easily disconnected, or points that should
be subject to damage, should be visually and functionally checked more often than rugged
hardwired points that are more durable.

7.5 Personal Grounding

7.5.1 What Is the Purpose of Personal Grounding?


The purpose of personal grounding is to maintain the body voltage of personnel who
handle ESDS at a low level so that they cannot be a significant source of ESD to ESDS
devices (see Section 4.7.9). In a standard ESD program handling ESDS devices of 100 V
human body model (HBM) withstand, the usual goal is to keep the body voltage below
7.5 Personal Grounding 191

100 V. To do this, a reliable and continuous electrical connection to ground must be


established and maintained. Where devices of lower HBM ESD withstand are handled, it
is wise to set a maximum body voltage as low, or lower than the device withstand voltage.
There are two main methods of grounding personnel – wrist straps and ESD control
footwear with ESD control flooring. Other methods such as grounding via garments or seat-
ing are sometimes used. It is not necessary to use both wrist straps and footwear-flooring
grounding, although providing both gives a useful redundancy that helps reliability of an
important ESD control measure.
While personal grounding controls the voltage developed on the body, it does not nec-
essarily eliminate voltages and fields from clothing and items worn or held by the person,
unless these are designed to be grounded via the body. EPA equipment such as garments
and tools are often designed to be grounded through the user’s body via personal grounding.
More unusually, systems have been designed to ground personnel via special ESD control
garments or seating although this is not currently common practice. It should not normally
be assumed that use of an ESD chair or garment ensures that personnel are grounded ade-
quately compliant with ESD control standards. Attempts to establish such as system should
be tested rigorously.
Personal grounding is one of the most essential ESD control measures for manual han-
dling of ESDS. Because of this, personal grounding is usually tested frequently, often every
working day before work is commenced. An alternative is to use a constant monitoring
system to monitor grounding.

7.5.2 Personal Grounding and Electrical Safety


Grounding of personnel handling ESDS devices is normally considered essential. In some
processes where high voltages are present, such as in-circuit or burn-in test, personal
grounding may conflict with safety considerations. Local electrical safety regulations
should always be considered, and personal grounding strategies consistent with their
requirements should be selected. Electrical safety is further discussed in Section 10.5.6.
The 61 340-5-1 and ESD S20.20 standards do not specify a minimum resistance to ground
for personal grounding. Nevertheless, if there is a possibility of contact with high volt-
ages, the user should consider specifying a minimum resistance to ground for the personal
grounding system. A level of resistance is normally included in items such as wrist band
ground cords, footwear, and flooring. The minimum resistance should be suitably specified
according to the voltage sources present in terms of resistance level and voltage withstand.
The level of resistance should limit the current possible due to possible contact with the
voltage source to a safe level. Some earlier standards made recommendations for this – the
IEC 61340-5-1:1998 recommended at least 750 kΩ per 250 Vac or 500 Vdc, to be increased
pro rata for increasing voltage. If there are local safety regulations, it is important that these
are observed and complied with. In some countries, a method such as ground fault circuit
interrupters (GFCIs) or circuit breakers must be used to open the circuit if personnel contact
high-voltage supplies.
Any protective resistors used must be able to withstand the voltages present. Failure
modes should also be considered, with failure to open circuit normally being essen-
tial. Failure to a short-circuit condition may expose the user to unacceptable electrical
shock risk.
192 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

If practical, the best option for both safety and ESD control is to prevent personnel from
touching the ESDS system while high voltages are present. Handling the system with no
voltages present is safe, providing there are no built-in energy sources such as high-voltage
batteries or charged capacitors.
When working in high-voltage systems, personnel may be required to wear high-voltage
protection gloves. These are an effective barrier to human body ESD at moderate levels
as well as essential protection for electrical safety, although ESD risk due to triboelectric
charging of the gloves may need evaluation.

7.5.3 Wrist Straps


7.5.3.1 Conventional Wrist Strap Systems
The purpose of the wrist strap system is to control body voltage by establishing a direct
electrical connection between the wearer’s skin and ground via an earth bonding point to
keep body voltage below a required level. The wrist strap system consists of a wrist band,
a ground cord, and an earth bonding point (Figure 7.4). The electrical and mechanical
properties of the wrist band and ground cord parts are specified in standards such as
IEC61340-4-6 and ANSI/ESD S1.1 (International Electrotechnical Commission 2015,
EOS/ESD Association Inc. 2013). These standards specify tests and limits for electrical
parameters such as wrist strap system resistance, wrist band inner, and outer surface
resistance. They also specify tests and limits for mechanical parameters such as band size,
breakaway force, ground cord extension, and bending life. The earth bonding point must
also be specified, established, and checked.

Figure 7.4 Typical wrist strap system.


7.5 Personal Grounding 193

The wrist band is a flexible band like a bracelet or watch strap that is worn in contact
with the wearer’s skin and must make good electrical contact with the skin. These bands
are normally worn on the wrist but are occasionally worn in contact with other parts of the
arms or legs. A ground cord makes electrical connection between the wrist band and an
earth bonding point. The earth bonding point provides a point where the ground cord can
be connected directly to EPA ground (earth).
Wrist bands come in various forms such as knitted or woven fabric, expanding metal
bracelet, or resin bracelet. Stick-on patches are also available. Wrist bands usually include
a hypoallergenic metal plate against the skin to promote good skin contact. The band inner
surface is also usually conductive to give electrical contact around nearly all the inner
surface of the band. The wrist band has a quick-release connection point for the ground
cord. This is designed to give a reliable connection strong enough to prevent accidental
disconnection but light enough for easy release. The disconnect force is normally in the
range 13–36 N.
The ground cord is an insulated wire with a connector on each end, used to connect the
wrist band groundable point to an earth bonding point. The ground cord includes various
features for safety and other considerations. The cord usually contains a current limiting
resistor at the wrist band end. Some types include a resistor at both ends. The cord wire must
be designed to be flexible and withstand constant flexing and dragging over equipment and
workstation edges. They are available in a range of length and colors and straight or coiled
wire construction. The earth bonding point connector is commonly a snap connector or
4 mm plug, although any suitable connector may be used.
The wrist strap system earth bonding point can be any suitable and reliable direct elec-
trical connection point to EPA earth. Dedicated earth bonding points are often provided on
workstations or equipment. It is not good practice to clip a wrist strap ground cord to a work-
station mat or other item that might introduce considerable addition resistance between
the ground cord and earth. Dedicated earth bonding points may include some resistance
providing the total wrist strap system resistance to ground does not exceed required upper
limits. Current standards IEC 61340-5-1 and ESD S20.20 specify an upper limit for the wrist
strap system resistance to ground of 35 MΩ although this may be changed according to the
user’s requirements.

7.5.3.2 Constant (Continuous) Monitor Wrist Strap Systems


A constant monitoring wrist strap system can be used to continuously test wrist strap
grounding during use. They are often used where high value or very sensitive ESDS
product are handled. The constant management system is designed to test the electrical
connections between the person and wrist band and the ground cord and connection to
the earth bonding point and set an alarm if any of these fail.
Constant monitoring systems are available in two types – single-wire systems and
two-wire systems. Single-wire wrist strap monitors can be used with ordinary wrist strap
systems, but two-wire wrist strap monitoring systems require a two-wire wrist strap
system to be used. Constant monitors use an electrical current to measure the wrist
strap system, and this can produce a low voltage on the wearer’s body during use. This
voltage may be unacceptable when handling the most sensitive (low ESD withstand
voltage) devices.
194 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Capacitance constant monitors use a single-wire wrist strap system. They detect the
person-earth capacitance of the person wearing the wrist strap using an alternating current
(AC) sensing signal applied to the system. Capacitance changes outside the set limits of the
monitoring system cause an alarm. Some disadvantages are that the monitor must be set
up for the individual using it and that false alarms may occur due to capacitance changes
rather than failure of the wrist strap system.
Impedance constant monitors are single-wire systems that use similar principles to the
capacitance monitor but detect circuit impedance changes rather than capacitance changes.
This eliminates the need for adjustment and can reduce false alarms.
Resistance constant monitors use a two-wire wrist strap system, measuring the resistance
of an out and return circuit through the wrist strap system and the wearer’s body. An alarm
sounds if the direct current (DC) resistance of this loop goes outside the accepted range.
The system may use a constant or pulsed DC voltage up to 16 V to make the measurement.
In most monitor designs of this type, the maximum voltage is evident only during an alarm
condition. The voltage is generally considerably less under normal conditions when the
person’s resistance is well under the set limit of the monitor. Some users may experience
skin irritation when a DC voltage is used.
Body voltage constant monitors use a two-wire system to measure the body voltage of
the user. This system may not detect an open circuit that results in apparently zero body
voltage. In some systems, an impedance or resistance monitor is incorporated to detect this
situation.
Each type of constant monitoring system requires an appropriate tester to verify the oper-
ation of the system. The single- and dual-wire wrist strap systems used with the monitors
can also be tested using resistance measurements.

7.5.3.3 Cordless Wrist Straps


I have never tested a cordless wrist strap that actually works in controlling body voltage to an
acceptable level. Grounding needs a continuous connection between the body and ground,
and a cordless wrist strap cannot provide this. Current standards require a maximum resis-
tance is maintained between the body and ground, and cordless wrist straps cannot provide
this connection.
Unconventional solutions like cordless wrist straps should always be carefully tested and
evaluated before adoption for use.

7.5.3.4 Wrist Strap System Selection


In a standard ESD program designed to protect ESDS devices down to 100 V HBM ESD with-
stand, an upper limit of 35 MΩ is normally specified for the resistance from the wearer’s
body to the groundable point. This should be reduced pro rata for handling lower ESD
withstand voltage devices. So, in an ESD program handling 50 V HBM devices, the max-
imum resistance from the wearer’s body to groundable point should be reduced to 17 MΩ.
In the manufacturing of magnetoresistive (MR) and giant magnetoresistance (GMR) heads
for disk drives, the upper limit for personnel grounding is often set at 10 MΩ. When han-
dling very sensitive devices, it would be wise to verify the maximum body voltages found
in practice during wrist strap qualification.
7.5 Personal Grounding 195

The intended reliability and durability of the wrist strap system should be considered.
The wrist band should be comfortable for the length of time the user is expected to use it.
Discomfort might encourage users to remove the strap and forget or omit to wear it while
handling ESDS.
Many wrist straps include a metal plate that contacts the skin under the ground cord
bonding plug, improving contact reliability. The rest of the inside of the band is normally
also conductive to give good all-round skin contact.
The length of ground cord, and whether it should be retractable, should be considered.
The type of earth bonding point connector should normally be standardized throughout the
facility so that there is no difficulty in using it in different areas as required.
If constant monitoring is to be used, it may be necessary to select a wrist strap system that
is designed specifically for use with the monitoring system.

7.5.3.5 Wrist Strap System Use


Wrist bands must be a good comfortable fit and make good direct contact with the wearer’s
skin. They should never be worn over clothing such as shirt or ESD garment sleeves.
The wrist strap is effective only while it is connected to a working earth bonding point.
Connection should be established before handling any ESDS and maintained constantly
while handling ESDS.
When seated, it is normally essential to use wrist strap grounding. Grounding via footwear
and flooring is unreliable in this situation as many seated personnel take their feet off the
floor from time to time, breaking the ground connection.

7.5.3.6 Wrist Strap Qualification


A wrist band ground cord can be qualified by simple resistance measurement of the ground
cord resistance. The types of connectors at each end of the cord should be confirmed to be
suitable for the intended use.
The complete wrist strap system should be confirmed to give, when worn by typical users,
system resistance acceptable within the ESD program requirements. Most organizations use
a proprietary wrist strap checker for compliance verification to check that the wrist straps
when worn pass this test.
A constant monitoring wrist strap system should be checked under typical operating con-
ditions to make sure it detects common types of wrist strap system failure but does not give
an unacceptable level of false alarms. The system should detect unworn or open circuit
wrist strap systems. When working with very sensitive ESDS, the voltage typically induced
on the wearer’s body should also be tested to make sure it does not risk damaging the ESDS.
Where compliance with a standard is required, testing should be done according to the
standard, and compliance with the standard pass criteria should be established.

7.5.3.7 Wrist Strap Cleaning and Maintenance


Wrist strap components (wrist band and cord) are normally replaced if they are found to
fail a compliance verification test.

7.5.3.8 Wrist Strap System Compliance Verification Test


As a key ESD control measure, it is important that wrist strap grounding works reliably. All
parts of the wrist strap system must be tested regularly to make sure they are working.
196 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Figure 7.5 Examples of proprietary portable wrist strap testers. Source (left): D E Swenson.

The wrist band and cord are normally tested as a system as worn, using a proprietary tester
(Figure 7.5). These are available as portable or wall-mounted testers, some of which can also
test footwear. The tester does a simple resistance measurement to verify the resistance from
the wearer’s body through to the cord groundable point is within acceptable maximum and
minimum limits. These limits are specified in the standards, although the user can select
different pass criteria for their specific circumstances. The resistance of the skin to band
contact is also tested in this measurement and can be highly variable. For a person with dry
skin, this resistance can be high and result in failure of the test.
In practice, these tests are usually done by personnel for themselves using proprietary
wrist strap checking instruments. These typically have built-in resistance checking, and
display pass and fail as colored lights. Some sophisticated systems may use this to control
access to the EPA via an entrance turnstile and automatically maintain compliance data.
In other facilities, the user may be required to record their wrist strap checking pass result
and relies on them to take remedial action if a failure is indicated.
With these instruments, the pass criteria are effectively provided by the instrument and
may or may not be adjustable. It is important to ensure that the pass criteria are aligned with
the ESD control program requirements. In current “off-the-shelf” testers, the upper limit is
often set at 35 MΩ with a lower limit of about 750 kΩ. These limits, and the possibility of
adjusting them if necessary, should be checked when qualifying a tester for use in an ESD
program.
It is also possible and acceptable to measure the wrist strap resistance from body to the
cord groundable point using a suitable resistance meter. The resistance meter must be
selected to be safe for use in this way, with limited output current to prevent electrical
shock risk. The measurement can be done with many types of instrument at 10–100 V test
voltage. The voltage used may depend on compliance with a standard or safety regulations.
Contact to the body should be made via a handheld electrode giving a reasonably large
skin contact area. Holding a small-point probe may not give adequate contact area for
reliable measurements.
7.5 Personal Grounding 197

It can be useful to keep failure data for the types of wrist straps in use, including for
example identification of skin contact, ground cord or connector failure, as well as manu-
facturer and type data. This allows the reliability of different wrist strap systems and their
components under operating conditions to be compared. The more reliable components
can be identified and selected for future purchase.
Testing the EPA wrist strap earth bonding points must also be done periodically to ensure
that any failure of these is discovered sufficiently quickly. This can be done using a sim-
ple resistance to ground test (see Section 11.8.3.5). A maximum resistance to ground for
earth bonding points, and a suitable test method, should be specified in the ESD program’s
Compliance Verification Plan.
The frequency of wrist strap testing can be determined according to factors such as the
level of use and established reliability of the wrist strap system. ESDS sensitivity and value
of the product handled, and possible consequences and cost of product failure, are also
relevant factors. If a wrist strap fails, any ESDS handled between failure and detection of
the failure might be damaged.
As personal grounding is of major importance in ESD control, it is common for the wrist
strap system to be tested at least each working day before the wearer commences handling
ESDS devices. In some ESD programs, wrist straps are tested every time the wearer enters
the EPA. When it is considered highly important to avoid risk of product damage and is
wished to immediately detect any wrist strap system failure, constant monitoring can be
specified.

7.5.3.9 Common Problems


Occasionally, personnel who have a dry skin do not make good low-resistance connection to
the wrist band. Proprietary skin moisturizer lotions are available that can be used to counter
this problem.
Some users attempt to ground the wrist strap by clipping to the edge of a bench mat or
other point not designated as an earth bonding point. This is not good practice as it may
introduce an unacceptable addition resistance of the mat to the wrist strap ground path.
Make sure the pass criteria of proprietary testers are aligned with the requirements of the
ESD program.

7.5.4 Footwear and Flooring Grounding


7.5.4.1 The Importance of Footwear and Flooring in ESD Control
ESD control footwear and flooring systems are intended to keep the body voltage on person-
nel below 100 V during walking and other activities. A well-chosen footwear and flooring
system can be used instead of wrist straps for reliable grounding of personnel who work
while standing. For seated personnel, use of a wrist strap is required as there is a risk that
the feet will be taken off the floor and contact with ground will be impaired or lost.
As has been previously explained, (see Section 4.7.9), the body voltage developed when
walking is dependent on the charge generated by the walking action and the resistance
through which it must pass in dissipation. Both charge generation and resistance to ground
of a footwear-flooring system are dependent on both the footwear and the flooring in
198 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

combination as they depend on the footwear-flooring material contact conditions. Both


are necessary for successful body voltage control.
Ordinary street or work shoes often have soles made from insulating materials that block
the flow of generated charge to ground. An ESD control floor gives the possibility of ground-
ing personnel (and equipment) via the floor, but ordinary footwear blocks the flow of charge
from the body to the floor. ESD control footwear provides the connection between the body
and the floor via the footwear. The final resistance to ground is given by the combination
of the contact resistance between the body and the footwear, the resistance of the footwear,
the contact resistance between footwear and floor, and the resistance through the floor to
ground.

7.5.4.2 Types of Footwear


There is a wide variety of ESD control footwear available to a modern ESD control program.
This allows footwear to be selected according to a variety of needs and job roles within the
ESD control program.
Heel and toe grounders are useful for fitting over street shoes or work shoes to give a
ground connection (Figure 7.6). They typically have a conductive strap that goes over the
shoe sole to contact the floor. This is attached to a conductive ribbon that can be tucked
into the shoe, contacting the wearer through the socks or in contact with the skin. Low-cost
one-use varieties are available that can be used by visitors.
Heel and toe grounders must be worn on both feet, as grounding depends on electrical
contact of the body through the grounder to the floor. If only one grounder were worn,
contact and grounding would be lost when that foot is raised from the floor. Toe grounders
are arguably better than those that ground via the heel alone. The heels are lifted from the
floor for more time during the walking action.
Booties and shoe covers are often used in clean areas for particle and contamination
control as well as ESD control (Figure 7.7). These simultaneously contain particulates and
provide grounding for the wearer.
ESD control shoes and boots are now available in a wide variety of styles (Figure 7.8).
They are also available with built-in safety features such as steel toe caps. Many ESD control
shoe types are hardly different in appearance from street shoes or work shoes. Many have a
visible label or symbol indicating their ESD control function to help in their identification.

Figure 7.6 Heel and toe grounders.


7.5 Personal Grounding 199

Figure 7.7 Booties or shoe covers.

Figure 7.8 A wide variety of shoe and work boot styles are available.

7.5.4.3 Footwear and Safety


In addition to its ESD control function, footwear can have a personal protection safety func-
tion. In Europe, PPE including footwear must comply with the requirements of the PPE
Directive. This means it must comply with the requirements of standards such as ISO 20345.
These requirements cover a range of physical properties including toe protection, penetra-
tion resistance, leak proofness, water penetration and absorption, and cut resistance.
200 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Electrical properties of footwear are classified under ISO 20345 as electrically “insu-
lating,” “antistatic,” or “conductive” depending on their electrical resistance. Antistatic
footwear has resistance between 100 kΩ and 1000 MΩ, and conductive footwear has
resistance less than 100 kΩ measured using ISO 20344 (International Organisation for
Standardisation 2011a,b). This test method is like, but slightly different from, test methods
used for ESD control footwear. The conditioning and test atmosphere conditions may also
be different. The result is that most types of antistatic and conductive footwear might be
suitable for EPA use, but they are not tested using the test methods and conditions required
for use in EPAs. PPE footwear that has not been qualified for EPA use should be tested and
qualified using appropriate tests.

7.5.4.4 Selection of Footwear


There is a wide variety of ESD control footwear available to a modern ESD control pro-
gram. This allows footwear to be selected according to a variety of needs and job roles
within the ESD control program. When selecting footwear for an individual, their gender
and personal preferences, job role requirements, and one-time or long-term usage can be
considered. Care should be taken where safety requirements may also be necessary, espe-
cially in Europe where there may be a need to comply with regulations on PPE (European
Union 2016). Guidance on selection, care, use and maintenance of PPE in Europe can be
found in CEN/TR 16832 (European Committee for Standardisation [CEN] 2015).
The types of footwear offered should be qualified in conjunction with all the types of floor-
ing with which they may be used to make sure they provide the ESD control performance
required.

7.5.4.5 Qualification of Footwear


Footwear selected for ESD control use must always be qualified for use with every type of
flooring with which it may be used. There are two basic types of qualification test.
The footwear must be tested to make sure that when worn by personnel, the resistance
from body to ground required by the ESD control program will be achieved (See Section
11.8.3.4).
The footwear must be tested for body voltage generation when worn by a user walking
on the floor types with which is it intended to be used. This is done using a walk test (see
Section 11.8.9). In most ESD control programs, the limit of body voltage is chosen to be
100 V, reflecting the goal of handling devices down to 100 V HBM ESD withstand. If lower
withstand voltage devices are handled, it may be necessary to use a lower body voltage limit.
Examples of body voltage waveforms obtained with different floors and the same footwear
are shown in Section 4.7.9.3 and Figure 4.8.
Initial qualification can be based on the resistance range of the footwear, tested unworn
according to a suitable standard such as IEC 61340-4-3 or ANSI/ESD STM9.1 (International
Electrotechnical Commission 2001; EOS/ESD Association Inc 2014a). Unfortunately, these
tests do not predict the performance of the footwear when worn by a person standing on
an ESD control floor. An important practical test is to determine the resistance from body
to ground achieved when a test subject wearing the footwear is standing on the floor with
which the footwear is to be used (IEC 61340-4-5:2004 or EOS/ESD Association Inc 2015)
Selected footwear should also be qualified in conjunction with the floor surface material
with which it will be used, to determine body voltage generated in a walk test (IEC 61340-4-5
7.5 Personal Grounding 201

or EOS/ESD Association Inc 2015). If the footwear is to be used with several types of floor
material, it should be qualified for use with all of them.
Qualification tests should, where possible, be tested with the flooring under the
worst-case conditions that can be envisaged during operation. This usually means testing
under low-humidity atmospheric conditions in a laboratory. Where a minimum resistance
for electrical safety is required, it may be necessary to establish performance under damp
conditions.
Where qualification with an existing floor is required, it is often not possible to test
the footwear-flooring combination in the laboratory. Qualification should be done under
the worst-case conditions practical under the circumstances. Confirmation of perfor-
mance under low-humidity atmospheric conditions when these arise is advisable. These
conditions often occur in the winter in colder climates.
The footwear should also be checked with the intended compliance verification test
equipment to make sure they reliably pass the intended compliance verification tests.
Where compliance with a standard is required, testing should be done according to the
standard and compliance with the standard pass criteria should be established.
Sampling and test of incoming ESD control footwear is advisable.

7.5.4.6 Use of Footwear


Although a wrist strap probably provides the most reliable personal grounding, personnel
who must stand or move around in their work can find wrist straps restricting or inconve-
nient. Using footwear-flooring grounding can provide an effective alternative for personnel
who stand or walk around during their work.
Grounding of personnel through footwear and flooring is effective only while the per-
son’s footwear is in contact with an ESD control floor. If the floor-footwear contact is broken,
grounding via footwear and flooring fails. This can happen, for example, if a person is seated
and takes their feet off the floor or if contact is only through the heels and the person stands
only on their toes. So, grounding through footwear-flooring is insufficient for seated person-
nel, and a wrist strap must be worn in this circumstance.
Heel or toe grounders must be worn on both feet. If a strap is worn only on one foot,
ground contact is lost when this foot is taken off the floor.
The resistance to ground and body voltage limiting performance of a footwear-flooring
combination can be changed by contamination on the footwear contact surface or the floor
surface. Contaminants such as asphalt, paints, varnishes, or chemicals can provide an insu-
lating barrier to current in the charge dissipation path. Water can have the opposite effect,
soaking into a sole material, reducing the footwear-floor contact resistance, lowering the
resistance to ground. This can be a problem if a level of resistance is required for electrical
safety reasons.
For these reasons, some organizations do not allow ESD control footwear to be worn out
of doors.

7.5.4.7 Compliance Verification of Footwear


Compliance verification of ESD control footwear is normally limited to measurement
of the resistance to ground (sole of footwear) when the foot is in contact with a metal
202 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

measurement plate. Proprietary testers are commonly available. These are often pro-
vided at the entrance to EPAs so that personnel can test their footwear before entry
to the EPA.
A disadvantage of many testers is that they only give a pass/fail indication rather than
a resistance value. Some automated data logging types also record the resistance value
measured. If necessary, a resistance measurement may be made using a suitable resistance
meter (see Section 11.8.3.3). The pass criteria of the footwear tester should be aligned with
the requirements of the ESD Control Program Plan documents.

7.5.4.8 Common Problems


Heel or toe straps can, unnoticed by the wearer, become dislodged and lose contact with the
floor during use. Some types of strap may have insufficiently long fixing straps to securely
hold the strap in place for personnel with large feet or shoes. It may be difficult to use straps
with some types of street footwear such as elevated heels or sandals.
Self-adhesive straps may not adhere well to the shoe if the wearer’s shoe soles were dusty
at the time of application.
ESD control shoes and boots often rely on moisture from the wearer’s feet to establish
contact from the skin through the socks to the shoe insole. If the wearer has dry skin or thick
man-made fiber socks, this contact can be impaired or take some time to be established. The
wearer may have to wear the shoes for several minutes before sufficient moisture is built
up, especially under cool atmospheric conditions. Atmospheric humidity conditions also
have an influence (Swenson et al. 1995).
Testers for wrist straps can sometimes be used to test footwear as well but may have wrong
resistance range. Current standards require a maximum resistance of wrist straps as worn
of 35 MΩ, whereas footwear as worn may have resistance up to 1000 MΩ.
In ISO 20345, safety footwear is classified as insulating, antistatic, or conductive.
Some users may wear insulating footwear for electrical shock protection, whereas others
may use antistatic or conductive safety footwear, for example for handling explosives
or flammable materials. Others may have footwear specified for ESD control rather
than personal protective use. Users may have poor awareness of these differences and
may just think they are wearing “safety footwear” or “ESD footwear.” This can result in
personnel inadvertently wearing the wrong type of footwear when they enter EPAs or do
particular jobs. This could be potentially unsafe if a person wore antistatic or conductive
footwear while working with high voltages. It could result in an ESD risk if a person
wore insulating safety footwear when within an EPA and handling ESDS devices. In the
EPA, this can be guarded against by requiring that personnel verify their footwear before
entry to the EPA.

7.5.5 Grounding via ESD Control Seating


A common misconception is that an ESD seat is designed to ground personnel sitting on
the seat. In most ESD programs and current ESD standards, this is not so. Nevertheless, in
some ESD control programs, seating has been successfully used to ground personnel. This
is discussed further in Section 7.9.8.
7.6 Work Surfaces 203

7.5.6 Personal Grounding via an ESD Garment


Groundable static control garments are available that can be used to ground the wearer
via the garment. These typically have a groundable point that can be connected to an earth
bonding point. They are also made from a relatively low-resistance material and must make
reliable contact with the body of the wearer. This is further discussed in Section 7.11.8.

7.5.6.1 Qualification of a Groundable Static Control Garment


Basic qualification of a garment fabric can be done on sample on the fabric. A point-to-point
resistance measurement is usually used. The resistance of some fabrics can vary with direc-
tion depending on how conductive fibers lie (warp and weft) and so measurements should
explore this possibility.
Qualification of a groundable garment should include tests to determine that all the pan-
els of the garment are made of material that has point-to-point resistance less than the
required upper limit and are electrically connected with the garment groundable point. The
garment must make direct reliable contact with the wearer’s body, and the wearer must be
reliably grounded through the system.
When grounding personnel through a garment, tests should be made to establish that
the body voltage is maintained below the level required by the ESD control program. Tests
should include the lowest-humidity conditions likely to be found in practice. ESD control
standards require that qualification is done at 12 ± 3% rh.

7.5.6.2 Compliance Verification of ESD Control Garments


ESD control garments used to ground personnel should be verified for grounding effective-
ness before use. Tests should verify the resistance of the material and the ground path. If the
garment is intended to ground the wearer’s body, then the entire ground path from body to
groundable point should be verified in the same way as for wrist straps. A proprietary tester
can be used for this providing the pass criteria are aligned with the garment requirements
in the ESD Control Program Plan.

7.6 Work Surfaces


7.6.1 What Does a Work Surface Do?
The purpose of providing a grounded static dissipative work surface is twofold (see
Section 4.7.10). First, the work surface material itself should not become charged and give
electrostatic fields that could lead to ESD risk. Second, the work surface provides a useful
way of draining charge from any noninsulative material or object that is placed on it. This
may include tools, assemblies, and components including ESDS.
The work surface provides an area in which the items being handled are brought to the
same voltage by being connected to EPA ground so that they present little ESD risk. Any
noninsulating item placed on the workstation surface can lose its charge to ground. The
characteristics of the work surface are key in determining the nature of the inevitable dis-
charge from any charged ESDS placed on the surface and ensuring that this discharge is
not damaging to the ESDS device.
204 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

The simplest EPAs for field work may consist of little more than a portable workstation
surface, personal grounding, and a connection to ground. In a simple case like this, the
boundary of the work surface may define the boundary of the EPA.

7.6.2 Types of Work Surfaces


There are a wide variety of work surface materials available and in use. Some are made from
a single material, whereas others have a multilayer structure.
Monolayer or homogenous materials have the same electrical properties through the bulk
of the material. These materials typically give a point-to-point resistance that varies with the
distance between the electrodes. The measured resistance to ground typically varies with
the distance between the measurement point and the groundable point.
Multilayer materials may have two or three layers. The top layer normally has resistance
between 10 kΩ and 1 GΩ. The underlying layer is normally low-resistance material. There
may be an additional bottom layer. Because of the underlying low-resistance layer, the
point-to-point resistance and resistance to groundable point measured on the mat surface
are relatively constant.
High-pressure laminates are typically single or multilayer rigid materials applied to a
substrate material. Some materials are affected by humidity. Their performance should
be qualified under low-humidity conditions. If a minimum resistance is also specified, it
should be tested under high-humidity conditions.
Mats and runners are usually flexible materials that can be used to cover surfaces that
were not designed for ESD control use. They may be single- or multilayer materials. They
may also be used as a sacrificial surface in areas of high contamination or to improve the
characteristics of high-pressure laminate surfaces.
Portable and field service work surfaces are typically light, flexible materials that can
conveniently be folded or rolled for inclusion in the tool kit of a field service engineer.

7.6.3 Selection of a Work Surface


In selecting a workstation surface, the first considerations must be about how the worksta-
tion is to be used and the processes that will occur there. Considerations may include
● Is it to be a permanent area or a temporary area?
● What are the activities and processes anticipated for the workstation, including the type
and form of ESDS handled?
● What physical properties are needed?
● Is there a need to resist chemicals?
● Will it be a clean environment in which particle shedding or other contamination would
be a concern?
● Are there electrical considerations such as use in the presence of high voltage, or handling
of printed circuit boards (PCBs) with on-board batteries?
● Are there possible safety considerations?
● Are there ergonomic considerations?
● What will be the maintenance and cleaning regime?
● Are there other safety or legislative requirements (e.g. fire retarding qualities) to consider?
7.6 Work Surfaces 205

The activities and processes and types of ESDS handled, along with whether it is a tempo-
rary or permanent facility, will probably largely determine the type of work surface selected.
For a field service workstation mat, compact storage and portability in a tool kit may be an
overriding consideration.
Durability considerations include the hardness, abrasion, and tear resistance of the
surface material. Aspects such as chemical (e.g. solvent), heat, and solder resistance may
also be a concern.
One significant point will be to decide whether a soft and cushioning surface is needed for
handling the ESDS device or a durable surface is needed for handling heavy or sharp items.
Even on a permanent workstation there may be a reason to use a temporary work surface.
Some ESD programs use a replaceable mat surface as a sacrificial cover where chemical
contamination or physical damage is a problem. Others use a mat to modify the surface on
some workstations, e.g. use of a cushioning mat on workstations that have a hard surface.
For mats, curling with age or due to storage in rolled or folded form may be a concern.
Where items such as cutting mats or trays are placed on the workstation surface, they
should comply with the requirements for worksurfaces.
If the workstation is in a clean room, this may restrict the work surface type to clean
room–compatible materials.
In workstations that use high voltages or handle ESDS devices containing batteries or
power sources, it may be necessary to use a work surface that has a specified minimum resis-
tance. The minimum resistance may be specified for safety, power drain, or short-circuit
elimination considerations. Where electrical safety is a concern, appropriate tests and speci-
fications must be selected that comply with local regulations. DC resistance measurements
as made for ESD control purposes may be inadequate for safety testing, especially where
high-voltage AC is present.
In a test workstation, proximity or contact with a conductive work surface could adversely
affect the operation of the ESDS under test.
A minimum resistance of the work surface may also be required to control a risk of
charged device ESD to ESDS devices placed on the surface (see Section 4.7.5). A metal
tray placed on a workstation surface can give charged device ESD risk if an ESDS device
is placed upon it.
Appearance can be a consideration in selecting a work surface. Colors can be used to
identify workstations for certain operations or for corporate identity identification.

7.6.4 Workstation Qualification Test


Workstation qualification test should address all the functionality, durability, chemical,
electrical, and other characteristics identified in the selection process.
Work surface qualification is typically done using standard point-to-point and resistance
to groundable point measurements (see Section 11.8). These should be done under con-
trolled atmospheric humidity representing the worst case likely to be experienced in prac-
tice. In most cases, this means testing under low-humidity conditions to make sure the
surface meets the upper resistance requirement under these conditions. If data sheet val-
ues are accepted as qualification, these should be obtained using the correct standards test
methods at 12 ± 3% humidity.
206 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

If a minimum surface resistance is specified, it should be verified under high-humidity


conditions.
Where compliance with a standard is required, testing should be done according to the
standard and compliance with the standard pass criteria should be established.

7.6.5 Acceptance of Work Surfaces


Work surfaces should be tested after installation and before first use. A resistance to ground
measurement can be used to verify that the work surface is grounded. A point-to-point resis-
tance measurement can be used to check that the work surface material surface resistance
characteristics are as expected from qualification data.

7.6.6 Cleaning and Maintenance of Work Surfaces


Work surfaces should be cleaned according to the manufacturer’s instructions. Ordinary
cleaning materials and procedures may not be suitable as they can leave a surface film of
wax, silicone, or other materials that can seriously affect the performance of the surface.
If a surface has been cleaned with water or a liquid cleaner, it should be left to dry before
testing as the residual moisture could affect the measurement results.

7.6.7 Compliance Verification Test of Work Surfaces


Current practice in compliance verification is to test the work surface resistance to EPA
ground. Current standards specify an upper limit for this, although an ESD program can
specify other pass limits to suit the expected operational performance of the selected items.
Visual checks can be a useful regular check of the ground connection, especially for
temporary connections, e.g. bonding cords between a mat and a mains electricity ground
connector.
The frequency of testing can depend on the expected reliability of the work surface mate-
rial and its connection, and the possible consequences of failure on ESD security. With a
permanently connected work surface in a clean area, with well-established history of mea-
surements showing little change, a relatively long interval between testing may be accept-
able. A bench mat placed on a surface for sacrificial use, known to become contaminated
by process chemicals and connected by a temporary connection to an earth bonding point,
may need to be tested very frequently.

7.6.8 Common Problems


Common failures can be due to failure or accidental disconnection of a ground connection
or contamination of the bench surface, for example with insulative process materials.
Any surfaces on which ESDS may be placed (including trays or packaging) should comply
with the requirements for worksurfaces.
Where other materials are placed on a worksurface, be careful that the ESD control
properties are not compromised. A common mistake is to use a cutting mat made of insu-
lating materials placed on an ESD bench, introducing a charged insulator risk. Another
7.7 Storage Racks and Shelves 207

common example is a metal tray placed on the bench surface can introduce charged device
ESD risk to ESDS devices placed on the tray.

7.7 Storage Racks and Shelves


7.7.1 Should It be an EPA Rack or Shelf?
Storage rack and shelf systems are used to store products and materials. Racks need no
special ESD control qualities and do not need to be in an EPA if they are used to store ESDS
device that are adequately protected in ESD-protective packaging for transport and storage
outside EPAs.
Racks and shelves used to store unprotected ESDS inside an EPA must have ESD control
qualities similar to workstation surfaces. They can also be used to store non-ESDS items
and products, providing any packaging used has suitable ESD control properties and the
proximity of such product does not introduce ESD risks. Non-ESDS items that are within
secondary (non-ESD control) packaging or in themselves might cause ESD risk (e.g. insu-
lating components such as enclosures) should be stored well away from exposed ESDS and
EPA shelves.
Workstations often have shelves above them for storage of test and IT equipment, tools,
materials, ESDS devices, and other items. If the shelf is used to store unprotected ESDS
devices, it must be treated as an EPA workstation surface. This means it must have the
electrical characteristics of a work surface and must be grounded. The usual ESD protective
requirements of an EPA workstation, such as control of insulating materials, must be
applied.
Workstation shelves that are not used to store unprotected ESDS can be treated as
unprotected areas. They may be constructed from non-ESD control materials, providing
they are sufficiently far from the position of any unprotected ESDS devices and no ESD
risks are produced by this. It requires training and discipline on the part of the operators
to ensure that unprotected ESDS are never placed on this shelf. The use of ESD control
shelves in some workstations and non-ESD control shelves in others should be avoided
if possible, as this can lead to confusion and placement of ESDS devices on non-ESD
shelves. If necessary, signs can be used to identify shelves that differ from the norm. If
unusual practices such as this are used, they should be carefully documented so that they
do not form a non-compliance with the documented ESD Control Program. Additional
training and compliance verification may be needed to ensure successful management of
the risks.
Single- or multilevel shelves and racks are often provided for storage of ESDS and
non-ESDS parts or product. The user will need to decide whether these should be specified
as EPA shelves or not. The racks should be specified as EPA racks if unprotected ESDS
devices are to be stored on them. If not EPA racks, then ESDS devices stored on them
will need to be protected within ESD-protective packaging for transport and storage in an
unprotected area.
EPA and non-EPA shelves should preferably not be mixed on the same rack as this
may lead to confusion. This could result in storage of unprotected ESDS devices on
208 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

unprotected area shelves or placing of insulative items on EPA shelves, causing ESD risks.
EPA racks and shelves should be maintained a suitable distance from any unprotected
area activities or storage. A distance of 0.5 m will normally reduce any electrostatic
fields from charged insulative materials to a sufficiently low level to minimize ESD
risks to all but extremely sensitive ESDS devices. A greater distance may give additional
confidence.
Factors indicating that racks may be better specified as EPA might include

● It would be convenient to store ESDS devices without packaging them in ESD protective
packaging.
● The items are to be taken to workstations in the same EPA after storage.
● The shelf can be sited for access without leaving an EPA. (If the floor between EPA work-
stations and the shelf is not an ESD control floor, transport to the shelf is effectively
through an unprotected area.)
● Non-ESDS and insulative items can be stored on a separate shelf unit.

The rack may be better specified as an unprotected area if

● Most items to be stored are not ESDS.


● It is desirable to store items in secondary packaging or insulative items on the same
shelves as ESDS.
● It is not inconvenient to package the ESDS within ESD-protective packaging for transport
and storage in an unprotected area.
● The items will go out of the EPA after storage.
● Access to the shelf is via an unprotected area (e.g. across a non-ESD control floor).

Drawer systems and carousels used to store unprotected ESDS can often be considered to
have the function of ESD control packaging and can be specified similarly.

7.7.2 Selection, Care, and Maintenance of Racks, and Shelves


When selecting EPA shelves and racks, similar considerations as for workstations may
apply, including the following:

● What will the physical requirements be (e.g. strength, size)?


● What is the type and form of ESDS handled?
● Will it be a clean environment in which particle shedding or other contamination would
be a concern?
● Are chemicals likely to be present?
● Are there electrical considerations such as high-voltage use or handling of PCBs with on
board batteries?
● Are there possible safety considerations?
● Are there ergonomic and convenience considerations?
● What will be the maintenance and cleaning regime?
● Are there other safety or legislative requirements (e.g. fire retarding qualities) to consider?

The care and maintenance aspects are like those for workstations.
7.7 Storage Racks and Shelves 209

7.7.3 Qualification Test of EPA Shelves and Racks


Workstation qualification test should address all the functionality, durability, electrical, and
other characteristics identified in the selection process. Qualification of the shelf surface
is like workstation surfaces, using standard point-to-point and resistance to groundable
point measurements (see Section 11.8). These should be done under controlled atmospheric
humidity representing the worst case likely to be experienced in practice. In most cases, this
means testing under low-humidity conditions to make sure the surface meets the upper
resistance requirement under these conditions. If a minimum surface resistance is speci-
fied, it should be verified under high-humidity conditions.
Where compliance with a standard is required, testing should be done according to the
standard, and compliance with the standard pass criteria should be established.

7.7.4 Acceptance of Shelves and Racks


Shelf and rack surfaces should be tested after installation and before first use. A resistance
to ground measurement should be used to verify that the work surface is grounded.
A point-to-point resistance measurement can be used to check that the work surface
material surface resistance characteristics are as expected from qualification tests.

7.7.5 Cleaning and Maintenance of Shelves and Racks


Shelf and rack surfaces should be cleaned according to the manufacturer’s instructions.
Ordinary cleaning materials and procedures may not be suitable as they can leave a surface
film of wax, silicone, or other materials that can seriously affect the performance of the
surface.
If a surface has been cleaned with water or a liquid cleaner, it should be left to dry before
testing as the residual moisture could affect the measurement results.

7.7.6 Compliance Verification Test of Shelves and Racks


Current practice in compliance verification is to test the shelf and rack surface resistance to
EPA ground. Current standards specify an upper limit for this as for work surfaces, although
an ESD program can specify other pass limits based on selected material characteristics.
Visual checks can be a useful regular check of the ground connection, especially for
temporary connections, e.g. bonding cords between a mat and a mains electricity ground
connector.
The frequency of testing can depend on the expected reliability of the shelf or rack surface
material and its connection and the possible consequences of failure on ESD security. With
a permanently wired-in shelf or rack surface in a clean area, with a well-established history
of measurements showing little change, a relatively long interval between testing may be
acceptable.

7.7.7 Common Problems


Common failures can be due to failure or accidental disconnection of a ground connection
or contamination of the shelf or rack surface, for example with insulative process materials.
210 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Some types of moveable shelf or rack systems can be difficult to distinguish visually from
similar systems not designed for EPA use. An example of this may be metal racks that have
plastic components joining the shelves to the frames. This can lead to unsuitable units being
inadvertently substituted for EPA shelf units during a work area rearrangement.

7.8 Trolleys, Carts, and Mobile Equipment


7.8.1 Types of Trolleys, Carts and Mobile Equipment
Trolleys, carts, and mobile equipment are used to store and move product, materials, or
equipment within an EPA or between EPAs. They may be simple mobile work surfaces or
shelves or specialist items such as PCB racks on wheels (Figures 7.9).
Mobile equipment and carts must not cause ESD risk to any ESDS device that they
come close to, may contact, or are used to store or transport. This usually means that the
cart must be made of noninsulating materials, and at the times and positions where ESD
product are loaded and unloaded, they must be grounded. At the times when the cart is
ungrounded, personnel should not be able to touch any unprotected ESDS on the cart,
even if the personnel are grounded. Personnel may safely handle ESDS on the cart only if
they are equipotential bonded to the cart.
Where carts have shelves on which unprotected ESDS devices may be placed, the shelves
should have a similar specification to EPA work surfaces. Where a mobile rack system
or similar equipment is effectively mobile ESD-protective packaging, requirements like
ESD-protective packaging can be applied.

Figure 7.9 Typical example of an


ESD control cart (trolley).
7.8 Trolleys, Carts, and Mobile Equipment 211

Figure 7.10 An ESD control seat.

Mobile equipment and carts can be grounded through noninsulative wheels, a drag chain,
or some other means to an ESD control floor. These can provide a convenient means of
grounding the cart if it is standing on the ESD control floor and has the advantage that it
needs no operator action. Qualification tests should demonstrate that grounding is effec-
tive when the equipment is standing on all the ESD control floors with which the cart will
be used.
Where there is no ESD control floor (e.g. at a stand-alone workstation), carts and mobile
equipment should be grounded via an earth bonding wire before and during loading and
unloading.

7.8.2 Selection, Care, and Maintenance of Trolleys, Carts, and Mobile


Equipment
In selecting a suitable mobile equipment or cart, considerations may include the following:
● Will the cart hold ESDS devices, or is it to be only used for tools, process equipment, or
materials?
● Will the cart work in one EPA or many EPAs?
212 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

● Are chemicals likely to be present?


● What will the physical requirements be (e.g. strength, size)?
● Will the cart remain in the EPA, or will it go through unprotected areas or even outside
between buildings? Will covers or additional protection be needed for this use?
● What is the type and form of ESDS devices handled?
● Are unprotected ESDS devices to be placed on the cart surface? If yes, does the ESDS
device have charged device ESD susceptibility?
● If used for storing or transporting ESDS, will they be within ESD protective packaging,
or will the cart provide the required level of ESD protection?
● When within EPAs, will an ESD control floor be available for grounding the cart? Will a
ground wire be required for grounding the cart at any stage?
● Can the cart ESD control characteristics be measured using standard electrodes, or will
special electrodes or techniques be required?
● Do the materials have permanent ESD control properties, or do they have coatings or
other control materials that might wear or degrade?
● Will it be a clean environment in which particle shedding or other contamination would
be a concern?
● Are there electrical considerations such as high-voltage use, or handling of PCBs with on
board batteries?
● Are there possible safety considerations?
● Are there ergonomic and convenience considerations?
● What will be the maintenance and cleaning regime?
● Are there other safety or legislative requirements to consider?
Where wheels are used as groundable points, it is advisable to make sure that two or more
noninsulative wheel groundable points are provided. This will provide a level of redundancy
to ensure that grounding is maintained even if a patch of contamination on a floor or wheel
prevents good contact. Some types of wheels, feet, or grounding chains do not reliably make
good contact with all types of ESD control floor.
Where cart wheels act as the groundable points, regular cleaning will usually be necessary
to keep them free of accumulated dirt contamination. Shelves may also need cleaning if
contamination is known to impair their properties.

7.8.3 Qualification of Trolleys, Carts, and Mobile Equipment


Carts that have shelves on which unprotected ESDS devices may be placed should be qual-
ified as for ESD control work surfaces. A point-to-point resistance measurement should be
used to check the surface properties. If charged device ESD is a concern for unprotected
ESDS devices placed on the surface, consider specifying a minimum point-to-point surface
resistance.
A resistance to groundable point measurement should be made to confirm the ground
path between all shelves and other parts and the wheels or groundable points. It is also
wise to do a resistance to ground test with the cart standing on the floor with which it will
be used to make sure the wheels make reliable contact through the floor to earth.
Resistance to ground tests should preferably be done with the cart standing on every floor
with which it will be used to check for reliable grounding.
7.8 Trolleys, Carts, and Mobile Equipment 213

For mobile PCB racks or other items that are specified like ESD control packaging, the
exposed surfaces should be tested to confirm that they are static dissipative or conductive.
Resistance to groundable point tests should be made to ensure that the surfaces are bonded
to the groundable points.
Any covers, doors, or sides to the cart or mobile equipment should be tested to make
sure they do not have exposed insulating surfaces that could charge up and give significant
electrostatic fields.
Where compliance with a standard is required, testing should be done according to the
standard, and compliance with the standard pass criteria should be established.

7.8.4 Compliance Verification of Trolleys, Carts and Mobile Equipment


Compliance verification of carts and carts that have shelves or work surfaces is typically
by testing the shelf surface resistance to EPA ground while standing on the EPA floor. Cur-
rent standards specify the same upper limit for this as for work surfaces, although an ESD
program can specify other pass limits.
Equipment such as mobile PCB racks may have surfaces specified as for ESD protective
packaging. Some of these, especially items such as slots or wells holding PCBs or other spe-
cific ESDS devices, may require testing of their surface resistance and resistance to ground.
This may be difficult to do using standard electrodes as the materials may be molded with
small and complex nonplanar surfaces. A suitable electrode system for testing these items
might have to be designed. Any special test methods or equipment used should be docu-
mented as part of the Compliance Verification Plan.
The frequency of testing can depend largely on the reliability of the ground connection.
With mobile equipment in a clean area, with a well-established history of measurements
showing little change, a relatively long interval between testing may be acceptable. Any
parts that have coatings that may be abraded or surfaces that may be damaged by normal
wear and tear should be regularly tested. Areas of conductive coating that have become
isolated through wear and may contact ESDS devices could be a serious source of ESD risk.

7.8.5 Common Problems


A common type of failure is excessively high resistance to ground due to contamination of
wheels with a buildup of accumulated dirt.
The wheels may make contact with the floor only over a small contact area. With some
types of floor material having sparsely distributed conductive elements, contact between
the wheels and conductive elements in the floor can be intermittent.
Contamination of shelf or rack surface, for example with insulative process materials,
can impair its ESD control properties.
Some types of mobile equipment can be difficult to distinguish visually from similar
systems not designed for EPA use. This can lead to unsuitable units being inadvertently sub-
stituted for ESD control shelf units during movement around a facility. Where this is a risk,
the equipment should be clearly marked in some way to differentiate the ESD-compliant
and noncompliant types.
214 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

7.9 Seats
7.9.1 What Is an ESD Control Seat for?
An ordinary chair not designed for ESD control usually has wheels or feet, structural parts,
and cover material made of insulating materials. The covering material and structural part
may become very highly charged in contact with the user’s outer clothing and by movement
of the chair over the floor. A person sitting on a highly charged seat might, if inadequately
grounded, rise to high induced voltage and be a source of ESD risk. The design of an ESD
control seat addresses these problems (see Section 4.7.10.5).
Wheeled seats can get charged by rolling action of the wheels on the floor. As well as
being a strong source of electrostatic fields, ordinary chairs can generate internal ESD that
can radiate electromagnetic interference (EMI) and interfere with electronic equipment
(Smith 1993, 1999).
An ESD control seat has a noninsulating covering material that can dissipate the charge
generated by contact with the user’s clothes. The structure of the chair is designed so that
an electrical connection is provided between the covering material and exposed chair parts
through to ground, usually via the feet or wheels or a grounding means such as a drag chain.
A common misconception is that an ESD seat is designed to ground personnel sitting on
the seat. In most ESD programs and current ESD standards, this is not so. One reason for
this is that it can be difficult to ensure reliable connection between the user’s body and the
seat through layers of intervening clothing. It would be difficult in many cases to ensure
that a sufficiently low-resistance ground connection and low body voltage is maintained.
Sitting personnel must normally be grounded via a wrist strap (see Section 7.4). Neverthe-
less, personnel have sometimes been successfully grounded via seating. This is discussed
further in Section 7.9.8.

7.9.2 Types of ESD Seating


A range of seating types are available including chairs and stools of many different heights
and style (Figure 7.10). Many are designed to be grounded through an ESD control floor
via conductive wheels or feet. Some may have additional grounding points such as bonding
points for ground cords.
The seat and back cover surfaces may be made of noninsulating fabric or vinyl. Chair
arms may be provided, made of noninsulating plastics. Metal footrests are often provided.
Fabrics used in upholstered chairs often have conductive fibers woven into the fabric.
Vinyl upholstery materials often have a thin conductive layer below the surface.

7.9.3 Selection of Seating


The seat style is primarily selected for its function and convenience to the user in the EPA
workstation that it is intended to be used. Safety considerations (e.g. stability during work-
ing activity) should be considered.
The seat should be constructed using noninsulating materials for all exposed surfaces
that might become charged by rubbing during use, especially those contacting the user’s
body. These surfaces should all be connected to the groundable point.
7.9 Seats 215

Where wheels or feet are used for grounding to an ESD control floor, it is advisable to
make sure that two or more noninsulative wheels or feet are provided. This will provide a
level of redundancy to ensure that grounding is maintained even if a patch of contamination
on a floor, wheel, or foot prevents good contact.
ESD control seats can be almost indistinguishable in appearance from ordinary seats. If
confusion with non-ESD control seating could be possible during use, consider identifying
the seats with appropriate markings.

7.9.4 Qualification Test of Seating


Qualification testing of seating should include testing to ensure that the exposed surfaces
that might become charged are not insulative and are electrically connected to the ground-
able points (wheels, feet, or earth bonding points). This should be done under dry humidity
conditions. Where compliance with a standard is required, testing should be done according
to the standard, and compliance with the standard pass criteria should be applied.

7.9.5 Cleaning and Maintenance of Seating


Build-up of dirt on the wheels or feet is a common cause of failure of compliance during
verification tests. Where cart wheels or feet act as the groundable points, regular cleaning
will usually be necessary to keep them free of accumulated dirt.

7.9.6 Compliance Verification Test of Seating


Compliance verification of seating is normally done using a simple resistance to ground
measurement from an electrode placed on the seat surface (see Section 11.8.1). This tests
the ground path from the seat surface through the grounding point (e.g. wheels and floor)
to the EPA earth.

7.9.7 Common Problems


For seats grounded through feet or wheels on an ESD control floor, a common cause of
grounding failure is the buildup of contamination and dirt on the wheels.
Wear and damage to the seat surfaces, especially those in contact with the user’s body,
can impair the ESD control properties of the materials. Repairs to ESD control seats should
never be made using insulative materials.

7.9.8 Personal Grounding via ESD Control Seating


A common misconception is that an ESD control seat is designed to ground personnel sit-
ting on the seat. In most ESD programs and current ESD standards, this is not so. One reason
for this is that it is difficult to ensure reliable connection between the user’s body and the
seat through layers of intervening clothing. Ground connection through wheels or feet can
also be unreliable as contamination can build up on the wheels/feet. It would be difficult
216 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

in many cases to ensure that a sufficiently low-resistance ground connection and low body
voltage is reliably maintained.
The resistance to ground through many types of seat is rather high for grounding person-
nel, with up to 109 Ω being allowed by current standards. In contrast, the limit required for
wrist strap grounding is set at 35 MΩ. Current standards require that sitting personnel must
normally be grounded via a wrist strap.
If seating is to be used for grounding personnel, these potential problems must be
addressed. The first difficulty is to ensure that reliable connection can be established
between the chair and the user. This may be easier in warm climates where fewer clothes
are worn because more skin may be exposed on the body where it will contact the chair. In
warmer temperatures, a greater level of body moisture helps make contact through clothing
fabric. In colder climates where several layers of clothes may be worn and low-humidity
conditions may arise, it can be difficult to establish reliable contact between the body and
the chair.
The second difficulty is to ensure that the chair provides reliable grounding to the ESD
control floor surface. A chair foot or wheel may contact a very small area of the floor and
is prone to buildup of dirt between the foot/wheel and the floor. This is particularly so in
areas where dust and contamination may be a concern.
As current standards do not use seats for grounding personnel, compliance would require
a tailoring statement giving the technical basis and rationale, backed by data, establishing
that the practice is acceptable. Qualification and compliance verification test methods and
pass criteria for suitable seating would need to be established as these are not provided by
standards in this case. For these reasons, a study would need to be done to establish the
reliability of grounding personnel through seating and establishing suitable tests and pass
criteria. The variation of resistance from the user’s body to ground and body voltage should
be studied over a range of temperature and humidity conditions. Cold and dry conditions
may provide a “worst case.” This and any other anticipated worst-case conditions should
be carefully studied. The results of the study should be documented in a report that can
be summarized and referenced in the ESD program documents and held on file for later
reference. An explanation of the practice and reference to the test report should be included
in the ESD program documents.
If seats are used for grounding personnel, then the body voltage must be maintained
less than the maximum required in the ESD program. This should be established using
body voltage measurements under simulated worst-case operating conditions including
low-humidity air conditions. One problem is that unless outer clothing is specified (e.g.
by use of ESD coats), the charge generation properties of outer clothing against the seat
cover material can be highly variable with clothing material changes.
If several types of seat are used, all these should be tested and qualified, and each seat type
to be used should be type tested to demonstrate that body voltage is adequately controlled
and establish the range for acceptable resistance to ground. Seats purchased for use should
then be restricted to the types tested and qualified.
Suitable test methods for compliance verification and pass criteria should be considered.
A simple resistance to ground test of the seat would not establish grounding of the user.
A resistance test from the body of the person sitting on the seat to ground would be needed to
verify adequate connection (Section 11.8.3). A body voltage measurement based on Section
11.8.9, made with the subject simulating normal working activity, would also be advisable.
7.10 Ionizers 217

7.10 Ionizers
7.10.1 What Does an Ionizer Do?
Ionizers are commonly used for reducing charge levels, voltages, and electrostatic fields
on essential insulators in use near ESDS devices (see Section 4.6.3). The primary means of
controlling voltages on conductors is to ensure that they are reliably grounded, but where
grounding cannot be used, it may be possible to use an ionizer to reduce voltages on conduc-
tors to an acceptable level. Reduction of electrostatic fields and voltages can be important for
reducing ESD risk and for reducing electrostatic attraction (ESA) of contaminant particles
to items in clean areas (Section 4.7.9).
An ionizer works by “spraying” positive and negative charges into the air as ions (see
Section 2.8). Opposite polarity ions drift under electrostatic fields to the surfaces that are the
sources of the fields. Arriving at the surface, they neutralize excess charges of the opposite
polarity or reduce the voltage by accumulating there.
The ions are produced from the air present in the environment. Producing air ions
in an ionizer does not produce any contamination and so ionization is suitable for
clean room use, although the erosion of emitter points in electrical ionizers should be
considered.
Use of ionization to neutralize charge and voltages suffers two main disadvantages. First,
it takes time to neutralize the charge or voltages, limited by the rate of arrival of ions at the
surface being neutralized. This can take many seconds. If charges are being generated at a
greater rate than can be neutralized by the arrival of air ions, the charge neutralization is
unsuccessful.
Second, depending on the type of ionizer used, the balance of positive and negative ions
produced by most types of ionizer is often not exactly balanced. This results in charging of
all ungrounded items or materials within the ionized region to a small offset voltage. The
user must establish that this offset voltage is insignificant for their purpose. The ionizer
imbalance typically changes with time and age.
Opposite polarity air ions attract each other due to their opposite charge. This means that
they tend to drift toward each other and recombine to form neutral molecules or particles.
The rate of arrival of ions, and hence charge or voltage reduction, can be highly dependent
on distance between the ionizer and the surface, ionizer orientation, local air flows, and
other factors. Effective neutralization occurs within a limited region around the output of
the ionizer and typically takes longer for greater distance from the ionizer. If an item is
within the ionized region for insufficient time, the surface voltages will not be completely
neutralized and may remain to cause ESD risk. In specifying an ionization system, it is
important to understand the size and position of the effective neutralization region, the
neutralization time, and the offset voltage characteristics.

7.10.2 Ion Sources


There are various ionizer technologies in use. Most use high voltages applied to
sharp-pointed needles within the ionizer to produce air ions by corona discharge.
The rate of ion emission from electrodes typically varies somewhat between electrodes
and for positive and negative polarity. Nuclear ionizers use nuclear radiation sources
218 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

to produce a well-balanced air ion source. X-ray ionizers use “soft” X-rays for the same
purpose.
Nuclear ionizers commonly use a polonium 210 nuclear source. This is packaged for
nuclear safety to prevent escape of the nuclear material but allows emitted alpha particles
to be released. These alpha particles (helium nuclei) collide with and split gas molecules
in the air to produce equal numbers of positive and negative ions and a well-balanced ion
source. Nuclear ionizers do not contain high voltages and so do not produce electrostatic
fields. They are unusual in that they produce a completely balanced ion stream and so have
zero offset voltage. One disadvantage is that the rate of ion generation is limited by the
rate of nuclear disintegration. The ionizer may be effective only for a short range although
airflow can be used to help distribution of the ions produced.
AC ionizers use AC high voltages applied to corona needle electrodes to produce positive
and negative ions alternately from the same electrodes. The high voltage AC is usually at
mains power frequency. As the ions of both polarities are produced at the same electrodes,
they can quickly recombine. A fan or other air movement system is often used to transport
the ions to the intended effective neutralization region.
DC systems have positive and negative DC high-voltage sources applied to separate
corona discharge needles. These needles are placed at some distance apart. Loss of ions
due to recombination is typically less than for AC systems. Lower air flow may then be
used to transport the ions through the intended region of effective neutralization. If the
ionization electrodes are far apart, the ion balance close to the electrodes may be poor,
resulting in high offset voltages close to the ionizer.
Pulsed DC ionizers used the same electrodes with both polarities of high voltage alter-
nately to generate both polarities of ions. The voltage is applied with a relatively low alter-
nating pulse rate, usually below 10 Hz. This results in a lower recombination rate than AC
systems, but the offset voltage may cycle negative and positive especially close to the ionizer
emitter electrodes. One advantage is that at the high voltage pulse rate, delivery of ions can
be adjusted to improve performance over the intended effective neutralization region.
Soft X-ray ionizers use X-ray sources to ionize the air, producing a well-balanced stream
of positive and negative ions. The ions are produced along the beam of the X-ray, which
can extend a meter or so from the ionizer. Personnel must be shielded from exposure to the
X-ray radiation. Advantages of the system are that well-balanced source of ions is produced
in a defined region without electric fields or air flow.

7.10.3 Types of Ionizer System


Ion sources can be built into various types of ionizer intended to be effective in regions of
various sizes or in different types of installation or situation. They can also be used in many
ways in processes or automated systems.
Room ionization systems are used when neutralization of charge sources is required over
a large production area or room, rather than localized to a workstation or process. Room
ionizer systems use multiple electrode systems as bars or grids, positioned to cover the
required area. They can use AC, DC, or pulsed DC ion sources.
Laminar flow bench ionizers are used to create a workstation in which contamination
is controlled, usually within a larger uncontrolled work area. Ionizers are used to control
7.10 Ionizers 219

electrostatic charging that can cause ESD and contamination (particle attraction) within
these workstations. AC, DC, pulsed DC, or nuclear ion sources may be used.
Work surface ionizers are used to control electrostatic charging in specific work surface
regions. AC, DC, pulsed DC, or nuclear ion sources may be used. Common types include
fans to take ions into the controlled region.
Where electrostatic control is needed in a small specific region within a process or
machine, point-of-use ionizers may be used. Compressed air nozzles may incorporate
ionizers to reduce electrostatic charge during blow-off particle removal. They may include
AC, DC, pulsed DC, X-ray, or nuclear ion sources.
Some systems use feedback to adjust ion balance and improve performance. These rely on
sensors to sense ion balance in the region around them. The sensor can assure balance only
in a limited region around it. Where balance is required over a large area, it may be better to
use several systems, each having their own sensor controlling smaller regions, rather than
a single system controlling a large area. Use of a feedback-controlled system can reduce
maintenance by automatically adjusting for changing emitter efficiency.

7.10.4 Selection of Ionizers


It may seem an obvious statement that an ionizer is normally required only if there is an
identified task for it to do. Nevertheless, ionizers have often been used in a “sticking plaster”
approach to try to solve or prevent real or imagined electrostatic control issues. They will
rarely give value for money if used in this way.
Perhaps the most common task for an ionizer system is to reduce charge and voltage
levels on process required insulators. Controlling voltages on ESDS devices or isolated con-
ductors may also be required. These insulators or conductors may be part of the product
being manufactured or handled, or they may be part of the process equipment. Understand-
ing the electrostatic issue that is to be addressed is an important first step in selection of a
solution.
A key first requirement for an ionizer system is that it must be able to deliver a sufficient
quantity of ions quickly enough to neutralize charge on the target object to sufficiently low
level in the timescale required in the process. In real situations, charge generated by the
process or handling as well as preexisting charge must be neutralized. If the charge is not
neutralized quickly enough, the ESDS devices may be put at unrecognized risk and the
benefit of ionization is compromised, or the process may have to be slowed (by introduc-
ing waits or reducing process speed) to allow time for effective neutralization. The latter is
often unacceptable. The necessary timescale for neutralization can vary from several tens
of seconds or longer (in a manual process or one organized to give built-in waiting times)
to a second or less in a fast automated process.
Selection of an ionizer needs consideration of a variety of environmental and process
factors, as well as the ability of the ionizer to reduce the charge and voltages to the required
levels within the required timescale. The size of the region to be controlled is an important
consideration. Often, an ionizer system must be engineered to suit the application. It must
always be evaluated and qualified within the actual operating area.
Ionizer performance is enhanced if airflow takes the ions into the working area to the
neutralization target. In regions where the airflow is limited or blocked, the effectiveness
220 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

of an ionizer can be considerably reduced. Some types of ionizers include fans to assist
air flow. The work area should be arranged to remove any obstruction to air flow to the
neutralization target.
Electrical ionizers use high voltages and sharp points. If insufficiently guarded, these can
give electrical shock or accidental injury risks to personnel. Electrical ionizers generate a
small amount of ozone that must not be allowed to become a threat to the health of per-
sonnel or to processes. Ionizers must meet local and national electrical, X-ray or nuclear
safety, ozone generation, or other regulations. Electrical ionizers also generate EMI that
may interfere with sensitive electronic equipment.
In summary, when selecting an ionizer, it may be necessary to consider the following:
● The application and task of the ionizer
● ESDS device withstand voltage
● Required ionizer offset voltage
● The size and position of the region to be controlled
● Whether the process can be modified to make the ionizer’s task easier or more reliable
● High voltages and electrostatic fields generated, and the regions and levels of ion imbal-
ance, especially near ionizer electrodes
● Environmental aspects such as airflow and blocks to ionized airflow
● Use of airflow to enhance neutralization
● Safety issues and regulations
● Maintenance requirements
● Installation, operation, and maintenance costs as well as equipment cost
● Need for associated facilities such as air supply
● Cleanliness and contamination, especially where air supplies and clean areas are involved
● Ozone and EMI effects (for electrical ionizers)
● For large systems, power distribution requirements
Nuclear ionizers should not be used in a chemical environment without consulting the
manufacturer.

7.10.5 Qualification Test of Ionizers


Once a task has been identified for an ionizer and an ionizer system has been selected for the
task, then it should be demonstrated that the system does indeed fulfill the role required.
This may involve a variety of standard and nonstandard tests. These should include spe-
cific tests to show that the neutralization task works in practice in the working environ-
ment. In-situ tests may be the only way to finally determine that an ionizer can do the task
intended. Once equipment is qualified, simpler acceptance tests may be required to check
equipment purchased works as selected.
Basic ionizer performance can be evaluated using standard tests using a charged plate
monitor (CPM) (see Section 11.8.8). Measurement of the charge decay time and offset volt-
age under repeatable fixed conditions form the basis of comparison of different ionizers.
Tests should be made of neutralization of charge of both polarities. Where compliance with
a standard is required, testing should be done according to the standard, and compliance
with the standard pass criteria should be established.
7.10 Ionizers 221

Tests should also be performed with the ionizer working in the process and environment
in which it will be used, testing performance at positions that represent the range of work-
ing positions of the ESDS devices or other items to be neutralized. Specific and nonstandard
tests should be devised to demonstrate that the ionizer reliably fulfills the task intended.
Tests might include, for example, investigation of the charge decay time and residual volt-
ages on an insulating item, an ESDS device, or an isolated conductor within the process
under conditions as close as possible to operating conditions. During in-situ qualification,
it may be useful to determine a suitable test arrangement or CPM location for compliance
verification testing. In some circumstances (e.g. within operating machines), it may be dif-
ficult or not viable to make measurements under working conditions.
It may be necessary to use nonstandard CPM instruments as well as other instruments
such as electrostatic voltmeters or field meters to quantify operation under real conditions.
As well as electrostatic tests, it may be necessary to test other aspects of ionizer performance
such as ozone buildup, particulate contamination, or EMI.
Where the ionizer is protecting against low-probability ESD events that may happen in
running production circumstances, the final proof of effectiveness may be demonstration
that it solves the identified problem and does not introduce another.

7.10.6 Cleaning and Maintenance of Ionizers


Electrical ionizers contain high voltages and needle electrodes and should be cleaned or
maintained only by competent personnel according to the manufacturer’s instructions.
Electrical ionizer emitter points need regular cleaning and maintenance or replacement.
The frequency required for this may depend on the emitter materials and ionizer design as
well as the operating environment. Variation of rates of deterioration of the emitter perfor-
mance affects both the decay time and the ionizer balance and offset voltage.
Radioactive ionizers experience gradual decay of their strength due to decay of the
radioactive isotope. They require replacement once the strength has dropped to an
unacceptable level. The radioactive source is normally returned to the manufacturer for
replacement and disposal. Radioactive ionizers may also require leak test for safety.
X-ray ionizers may need periodic return to the manufacturer for replacement of the X-ray
source. Functionality can be tested by measurement of decay time and offset voltage as for
other ionizer types.

7.10.7 Compliance Verification Test of Ionizers


Standard test methods are available for using a CPM to measure the decay time and off-
set voltage of ionizers for compliance verification (see Section 11.8.8). A repeatable test
arrangement or locations should be used, preferably determined during ionizer in-situ qual-
ification, to ensure that verification tests give results representative of the performance
required at the position of the item to be neutralized.
The test frequency should be chosen appropriately according to the circumstances and
experience. In a situation where high-value and high-sensitivity components are handled,
it may be desirable to do functional tests and quantitative tests relatively frequently. Visual
checks can show correct orientation of movable fan ionizers. Less frequent checks may be
222 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Figure 7.11 Typical ceiling mounted pulsed DC ionizer charging of an object at tabletop height.
Source: D E Swenson.

used in a situation where the performance is less critical or where verification history has
shown reliable and consistent performance.

7.10.8 Common Problems


Inadequate maintenance of electrical ionizers can result in excessive offset voltage devel-
oping due to changes in emitter efficiency. The result is that items in the vicinity become
charged to the offset voltage and can cause undetected ESD risk. Isolated conductors that
might contact the ESDS device can provide the most serious risk. I have seen ionizers that
charge items to nearly 1000 V due to neglect. Pulsed dc ionizers can also induce time varying
offset voltages (Figure 7.11).
Chemical contamination or excessive humidity can affect the high-voltage sources of
electrical ionizers, causing internal leakage currents that can reduce the ion output and
lengthen neutralization times and increase offset voltage.

7.11 ESD Control Garments


7.11.1 What Does an ESD Control Garment Do?
A well-designed and correctly worn ESD control garment can protect ESDS devices han-
dled by the wearer from electrostatic fields and ESD from highly charged outer clothing.
Where coats or coveralls are required for cleanliness or other reasons, use of a garment
specified for ESD control prevents the garment from being a source of electrostatic fields or
discharges (see Section 4.7.10.8) and prevents electrostatic fields and discharges originating
from clothing under the garment from affecting the ESDS device (Paasi et al. 2005a,b).
Use of an ESD control garment does not remove the need to ground personnel via a wrist
strap or footwear and flooring.
7.11 ESD Control Garments 223

7.11.2 Types of ESD Control Garments


Current ESD standards IEC 61340-5-1 and ANSI/ESD S20.20 define three classes of
garment.
● Static control garments that provide ESD control without necessarily grounding the gar-
ment. These have point-to-point resistance specified in the standards.
● Groundable static control garments that have a defined groundable point. They have a
point-to-point resistance and resistance to groundable point specified in the standards.
● A groundable static control garment system that can be used to ground the wearer via
the garment. These have a point-to-point resistance and resistance to groundable point
specified in the standards. They include a means of reliably connecting the wearer to
ground via the garment such as an integrated wrist strap or conductive cuff electrically
connected to the groundable point.
IEC 61340-4-9 describes the garment types and characteristics (Table 7.1) in more detail,
together with the test methods used to qualify and verify the garment characteristics (Inter-
national Electrotechnical Commission 2016a).

Table 7.1 Types and characteristics of ESD control garments according to IEC 61340-4-9:2016.

Description and Resistance


use of garment Garment type Tests used range (𝛀)

Garment has some Static control Resistance point to point <1011


electrostatic field garment
suppression properties
Garments with Groundable static Resistance point to point <109
designated grounding control garment Resistance to groundable point
point system (garment in
combination with a
person)
Garments in continuous Groundable static Resistance point to point <109
electrical path with the control garment Resistance to groundable point
wearer but not used as
the primary ground path
Grounded with dual paths Groundable static Resistance point to point <109
to ground via continuous control garment Resistance to groundable point
monitoring equipment system (garment in
that requires two separate combination with a Integrated wrist strap in <3.5 × 107
paths to ground person) accordance with IEC 61340-4-6

Grounded via a single wire Groundable static Resistance point to point <109
continuous monitoring control garment Resistance to groundable point
equipment system (garment in
combination with a Integrated wrist strap in <3.5 × 107
person) accordance with IEC 61340-4-6

Garment used as primary Groundable static Resistance point to point <109


ground path for personnel control garment Resistance to groundable point
system (garment in
combination with a Integrated wrist strap in <3.5 × 107
person) accordance with IEC 61340-4-6
224 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Disposable garments, as the name implies, are intended to be used a small number of
times. They are typically made from nonwoven materials. ESD control properties are often
provided by finishes or treatments that have a limited life. Their properties may be depen-
dent on atmospheric humidity and be impaired under low-humidity conditions. They may
be useful in processes where contamination gives a short garment life and a need for fre-
quent replacement.
Reusable topically treated garments may require retreatment after each cleaning.
Treatments typically depend on atmospheric humidity and become inadequate under
low-humidity conditions. Some materials have conductive fibers but also require topical
treatment. For these materials, the properties may be less dependent on humidity.
Garments with permanent ESD control properties should maintain their ESD control
properties for their intended life. They usually have a grid or stripes of conductive fibers
woven into the fabric (Figures 7.12 and 7.13). If not grounded, exposed conductive fibers
could represent an ESD source.
Static control garments can be made in various materials. Some are disposable, whereas
others are reusable. They may be made from homogenous untreated fabrics or coated fab-
rics or may have conductive threads aligned in one direction only or in two directions as
a grid.
Groundable static control garments include a means of grounding the conductive
fibers in the fabric through one or more groundable points. The panels of these garments
should be electrically connected. Fabrics that have a grid pattern may produce the most
reliable connection. Connection between panels may be improved by including additional
non-insulative (electrically conducting) material in the seams. The garment can be
grounded via the wearer’s body through conductive wrist cuffs or other areas of direct

Conductive threads

Textile Textile

Figure 7.12 Stripe (left) and grid (right) conductive fiber patterns (Paasi et al. 2005a,b).

Conducting material Insulating material

Conducting Core conducting Surface Hybrid


fiber fiber conducting conducting
fiber fiber

Figure 7.13 Examples of ESD control garment fiber types (Paasi et al. 2005a,b).
7.11 ESD Control Garments 225

contact with the wearer’s skin, providing the wearer is grounded through a wrist strap or
footwear-flooring system. Alternatively, grounding may be achieved via a ground cord.
Fabrics that have a stripe pattern rather than grid may not have a good connection
between the conductive fibers. These may need topical treatments or conductive material
in seams to improve connection between the fibers.
Fabrics that depend on topical treatments can degrade unevenly and leave areas of fabric
isolated from the ground connection.
Some experts have been dissatisfied with measurement of garment materials only on
the basis of material resistance. Buried conducting fiber materials are usually rejected by
resistance measurements as the measurement electrodes do not make contact with the con-
ducting fiber. These simple tests do not characterize many of the properties of garment
materials (Paasi et al. 2004, Baumgartner 2000). Potentially relevant factors that are ignored
by the resistance test method include

● The triboelectric charging propensity of the fabric


● The effect of grounding of the fibers on the protective performance of the fabric
● ESD risk from unearthed conducting fibers in the material
● ESD risk from charged insulating areas of nonhomogenous material
● Effectiveness of protection against electrostatic fields from underlying charged fabrics
● The effect of the grounded wearer on the protective performance of the garment

In the early 2000s the ESTAT garments European Project researched the importance of
these factors and produced recommendations for use and test of ESD-protective garments,
considering surface and core conducting fiber and stainless steel thread materials. Their
recommendations have not been widely adopted in ESD standardization, but their final
report makes interesting reading (Paasi et al. 2005b). Their results broadly confirmed the
value of simple resistance measurements in evaluating groundable garment materials.
They found that the ESD-protective fabric and garment characteristics were very depen-
dent on the fabric and garment design, atmospheric humidity, and whether the garment
and the wearer were grounded. The main functions of an ESD control garment are

● Reduction or elimination of electrostatic fields produced by clothing worn beneath the


garment
● Prevention of ESD originating from the clothing worn beneath the garment
● The ESD control garment should not itself charge up and cause external electrostatic
fields or be a potential source of damaging ESD

Charging of ESDS devices by accidental rubbing against the garment fabric could not be
easily averted by choice of fabric but could be reduced by minimizing the risk of rubbing, e.g.
by avoiding loosely fitting sleeves. Loose garments have lower protective value than close
fitting garments. ESD could occur from garment materials with ungrounded conducting
threads or sufficiently large insulating areas (over 20 × 20 mm). Most effective ESD control
was achieved for materials having surface resistance in the range 100 kΩ–100 GΩ.
Paasi et al. (2005b) classified garments as Class A (using electrically continuous, low
charging, static dissipative, or conductive materials and grounded) or Class B (low charging
but need not have measurable electrical conductivity, and grounding not required). These
226 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

approximately correspond with the “groundable static control garment” and “static control
garment” classifications of IEC 61340-4-9 (Table 7.1).
Grounding the conducting parts of the groundable garment (e.g. to the grounded wearer)
improves the performance of the garment both in terms of external electrostatic fields and
as a potential source of ESD.

7.11.3 Selection of ESD Control Garments


ESD control garments may be required for protection of ESDS devices against the possible
effect of ESD. Where flammable materials are handled, for example solvents, the garments
may be classed as PPE and in some countries subject to regulation. Guidance on selection,
care, use, and maintenance of PPE in Europe can be found in CEN/TR 16832.
The first question in the process of selecting an ESD control garment is, is it necessary
to specify them? There is no simple answer to this. When developing an ESD program, the
ESD coordinator will need to evaluate the need based on many considerations.
In other cases, the ESD coordinator must evaluate whether ESD risks associated with
ordinary everyday work clothes are likely to be significant in the context of the ESDS devices
handled and other considerations. Various factors can be considered (Table 7.2).
While finding that ESD control garments are not always necessary, Paasi et al. (2005b)
recommended that they should be seriously considered if the ESD withstand voltage of
ESDS was 500 V charged device model (CDM) or 1 kV HBM or less. They should also be
specified if over garments are required for contamination control or where indicated by
high cost and consequences of ESD failures and high reliability required of product. For
cleanroom applications, the chargeability of the material is the main consideration.

Table 7.2 Factors that might be considered in decision to issue ESD control garments.

Factors that might indicate ESD control Factors that might indicate ESD control
garments are desirable garments are unnecessary

Low CDM ESD withstand voltage Components handled have moderate or high
components. ESD withstand voltages and low or moderate
High cost and low withstand voltage ESDS. cost.
Consequence of ESD failure is highly Consequence of occasional ESD failures is
undesirable. acceptable to some extent.
Clean area where ESA is undesirable. ESA is not a significant issue.
Likelihood that high charge generating Likelihood that single layer low-charging
clothing that is not close fitting may be worn clothing will be worn, e.g. warm moist climate.
e.g. cool dry climate.
High reliability product or market where low Consumer market where moderate failure
ppm failure rate is required. rates are accepted.
Use of ESD control garments may help
establish ESD control culture or help
demonstrate best practice and care (e.g. to
visitors).
7.11 ESD Control Garments 227

The main ESD control risk from ordinary garments is that they might charge highly and
give an electrostatic field near ESDS devices. This electrostatic field could induce high volt-
ages on devices, leading to charged device or charged board ESD. ESA can also be a concern
in clean areas, where electrostatic fields can encourage attraction of contamination to the
items handled. In most cases, the areas of the body most likely to come close to ESDS are
the front of the torso and arms. It is therefore particularly important to shield ESDS from
electrostatic fields from clothing in these areas. The clothing of the lower body is often less
important as it is less likely to come near to unprotected ESDS devices.
Woven materials can also release fibers that can give unacceptable contamination in
clean areas. Coveralls used in clean areas can be made entirely from ESD control materials
designed for low fiber release to reduce ESA of particulate contaminants.
Once the decision to use ESD garments is made, a suitable type must be selected. In some
cases (e.g. clean room), this will be a relatively simple matter of specifying an ESD control
version of a garment that is required for a specific purpose (e.g. contamination control or
personal protection against hazards). Ordinary personal protective clothing may be made
from insulative materials that can be a strong source of electrostatic fields.
An ESD control garment should cover the underlying garments completely over the key
areas of the body that need to be covered. As a guide, where the clothes are likely to come
within 30 cm of ESDS devices, they should be covered by an ESD control garment. The
material should be designed to reduce the electrostatic fields generated by underlying gar-
ments. It should not itself be a source of electrostatic fields or ESD. Materials used should
not contain low resistance fibers or metal parts that could charge and become a source of
ESD or could be an electrical safety concern.
According to Paasi et al. (2005b), a dense grid of conductive fibers (e.g. 5 × 5 mm) gives
improved electrostatic field control. For materials with conductive stripes, the stripes
should be less than 10 mm apart. For a grid fabric, the grid squares should be 20 × 20 mm
or less. They found that core conductive fiber fabrics cannot give good grounding of the
fibers under dry air conditions, but this does not necessarily preclude their value in ESD
protective garments. They did not recommend use of a garment as the ground path of the
wearer.
In some countries (e.g. in Europe), garments that have a PPE function are subject to
regulations that require they comply with specific safety-related regulations. Individual
organizations may also have safety rules that must be followed when selecting garments
for use within that organization.

7.11.4 Qualification Test of ESD Control Garments


Where compliance with a standard is required, testing should be done according to
the standard, and compliance with the standard pass criteria should be established.
Current standard test methods rely on surface resistance of the material to be deter-
mined using point-to-point resistance measurements (see Section 11.8.2). For garments
that have a groundable point, a resistance to groundable point measurement should
also be made. Resistance should be measured between panels as well as across panels.
Where the garment is intended to make contact with the wearer’s body, this should also
be tested.
228 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Many garment materials are affected to some extend by atmospheric humidity. Cotton is
inherently hygroscopic, but man-made fibers such as PET and nylon are usually less so. Top-
ical treatments are often highly dependent on humidity for their performance. So, resistance
measurements should be made under controlled low-humidity conditions. Testing over
the full range of temperature and humidity conditions likely in practice may be advisable.
Garments that use fabrics that have core conductive fibers cannot be successfully mea-
sured using surface resistance measurements, as the test electrodes do not make good con-
tact with the buried core conductive fibers. No standard test methods are currently defined
for measurement of these materials for use in electronics environment (Swenson 2011). The
user would need to define suitable test methods and pass criteria for qualification of these
materials.
Paasi et al. recommended that as well as surface resistance tests for qualification of gar-
ments, EN1149-3 Method 1 could be used to evaluate tribocharging, and EN1149-3 Method
2 evaluate induction charging of all types of garment materials (British Standards Institute
2004). Holdstock et al. (2003) have developed a “capacitance loading” test method that can
be used to evaluate charging of garment materials.
Garments made from fabrics that claim permanent ESD control properties should have
tests that compare measurements on new fabrics with measurements on fabrics after
repeated cleaning simulating long-term use. IEC 61340-5-2 recommends typically 50
cleaning cycles should be used.

7.11.5 Use of ESD Control Garments


The garment size should be selected to provide a reasonable fit to the wearer while wearing
their usual clothing. For normal ESD control use, the garment should cover the material of
the clothing beneath the garment at least on the arms and torso. Clothing should be worn
fastened and should not be worn loose and open, exposing the clothing beneath.
Groundable static control garments should be connected to ground before the wearer
handles ESDS. They should remain grounded during all handling of ESDS devices by the
wearer.

7.11.6 Cleaning and Maintenance of ESD Control Garments


Garments that have become damaged should be replaced or repaired according to the man-
ufacturer’s recommendations. A repaired garment should be shown to pass compliance
verification tests before use.
Laundry should be according to the manufacturer’s instructions. Thorough rinsing
should be included to ensure that cleaning chemicals are removed. An auditable tracking
system should be used to monitor cleaning. Compliance verification tests should be used
to verify the ESD control properties at least on a sample basis after cleaning.

7.11.7 Compliance Verification of ESD Control Garments


Garments should be tested for compliance verification on a regular basis and especially
after cleaning. Garment materials that rely on treatments or additives will usually need to
7.12 Hand Tools 229

be tested more frequently that those that include a conductive fiber matrix. For garment
materials that have measurable surface resistance, standard point-to-point resistance test
methods can be used (see Section 11.8.2). For groundable garments, the resistance to the
groundable point should also be tested.
Resistance test methods are not suitable for verification of garment materials that use
buried conductor fibers. If these materials are selected for use, suitable verification test
methods and pass criteria will need to be defined.
The frequency of compliance verification tests should reflect the expected permanence
of the garment fabric technology.

7.11.8 Personal Grounding via an ESD Garment


Groundable static control garments are available that can be used to ground the wearer
via the garment. These typically have a groundable point that can be connected to an earth
bonding point. They are also made from a relatively low-resistance material and must make
reliable contact with the body of the wearer.
Basic qualification of a garment fabric can be done on sample on the fabric.
A point-to-point resistance measurement is usually used. The resistance of some fab-
rics can vary with direction depending on how conductive fibers lie (warp and weft) and
so measurements should explore this possibility.
Qualification of a groundable garment should include tests to determine that all the pan-
els of the garment are made of material that has point-to-point resistance less than the
required upper limit and are electrically connected to the garment groundable point. The
garment must make direct reliable contact with the wearer’s body, and the wearer must be
reliably grounded through the system.
ESD control garments used to ground personnel should be verified for grounding effec-
tiveness before use. Tests should verify the resistance of the material and the ground path.
If the garment is intended to ground the wearer’s body, then the entire ground path from
body to ground should be verified before use in a similar manner to wrist strap test.

7.12 Hand Tools

7.12.1 Why Have ESD Hand Tools?


Hand tools include any tools that are held in the hand such as screwdrivers, pliers, lead
cutters, tweezers, vacuum pick-up tools, etc. Tools that are not designed for ESD control
often have metal parts that are electrically isolated by, for example, an insulating handle
(see Section 4.7.10.6). These isolated metal parts can become charged, and if they contact
an ESDS, then a potentially damaging ESD could occur. Most ESD tools are designed to
ensure that all parts of the tool are grounded, usually through contact with the user’s hand.

7.12.2 Types of Hand Tool


The types of hand tool of concern are mainly those that are likely to come into direct contact
with ESDS during use. These may include tools such as pliers, cutters, and desoldering tools.
Screwdrivers and adjustment tools may also be of concern.
230 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

Hand tools are normally made ESD safe by ensuring that metal parts have a grounding
path through to the user’s hand. This usually means that insulative parts of the tool have
been replaced by noninsulative parts.
In some cases, when handling ESDS devices that have low CDM ESD withstand, it may
be necessary to ensure that metal parts (even though grounded) do not contact the ESDS
devices. The materials that contact the ESDS devices may need to be resistive and a min-
imum resistance be specified to reduce the strength of any discharge that may occur on
contact with the ESDS devices.

7.12.3 Qualification Test of Hand Tools


Where compliance with a standard is required, testing should be done according to the stan-
dard (if specified), and compliance with the standard pass criteria should be established.
Unfortunately, the current versions of 61340-5-1 and ESD S20.20 do not specify require-
ments or test methods for hand tools. So, the user will have to develop test methods and
qualification criteria.
Many hand tools (e.g. hand cutters or pliers) consist of a metal part that in a non-ESD
control tool would be insulated from the user’s hand by a handle made of insulating mate-
rial. In an ordinary tool, this can represent an isolated metal part that can become charged
and then contact and discharge to the ESDS device. To prevent this, the insulating handle
is replaced with a noninsulative material. The resistance between the tool metal part and
the user’s hand should be below a level that will allow any charge on the metal to dissipate
in a short time. In most cases, a dissipation time of up to a second or so would be accept-
able. If the capacitance C of the metal part can be measured or estimated, an upper limit of
the resistance from the metal part to ground R can be specified as RC < 1. As an example,
if the tool metal part has a capacitance of 10 pF, R can be up to 1011 Ω and be acceptable.
In practice, it is difficult to reliably measure high resistances. It may be better to specify
a pragmatic lower resistance limit determined to be easily achievable with products avail-
able on the market. Many ESD control tools in practice have resistance <1 GΩ. If a CPM
is available, charge decay measurements are often easier and more reliable than resistance
measurements on high resistance tools.
An early version of 61340-5-1 (IEC 61340-5-1:1998) specified resistance and charge decay
tests and pass criteria that could be adopted or modified for use.
Some types of hand tools, for example handheld instruments, have significant areas of
exposed housing. If this is made of insulating material, it could become charged and give
an electrostatic field that could cause ESD risk if brought close to an ESDS device. In this
case, a charging test of the material under dry air conditions may be advisable to evaluate
if unacceptable high voltage is generated and held on the material.

7.12.4 Use of Hand Tools


Hand tools are usually grounded by contact with the user’s hand. If the user must also wear
gloves, use of insulating glove materials would prevent grounding of the tools. These could
then become charged conductors that present an ESD risk.
So, gloves used with hand tools must be selected to present sufficiently low resistance
through the glove to provide acceptable grounding of the tools.
7.13 Soldering or Desoldering Irons 231

7.12.5 Compliance Verification Test of Hand Tools


Where compliance with a standard is required, testing should be done according to the stan-
dard (if specified), and compliance with the standard pass criteria should be established.
Unfortunately, the current versions of 61340-5-1 and ESD S20.20 do not specify require-
ments or test methods for hand tools.
A simple test can be used to verify the resistance from a tool tip via the user’s hand
to ground (see Section 11.9.5.2). Alternatively, a CPM can be used in a voltage decay test
(Section 11.9.8.1). If the tool is usually held in a gloved hand, the system of tool-glove-hand
to ground can be verified in either of these ways in one measurement.

7.12.6 Common Problems with ESD Control Hand Tools


Possibly the most common problem experienced with ESD control hand tools is that these
are confused with similar non-ESD control versions. This can result in non-ESD control
tools being present and used in the EPA. This can be avoided if tools having suitable ESD
control identification markings are used.
Use of insulating glove materials prevents grounding of the tools. These can then become
charged conductors that present an ESD risk.

7.13 Soldering or Desoldering Irons

7.13.1 ESD Control Issues with Soldering or Desoldering Irons


Many of the issues of concern for soldering irons are to do with potential Electrical Over-
Stress (EOS) damage due to injection of potentially damaging leakage currents and voltages
rather than ESDs (EOS/ESD Association Inc.1999). Energy-sensitive devices are likely to be
more susceptible to damage from these currents and voltages than voltage-sensitive devices,
providing the soldering iron bit is grounded.
For a soldering iron to be safe, the tip to ground open circuit tip voltage must not exceed
20 mV, the tip to ground short circuit current must not exceed 10 mA, and the tip resistance
to ground must not exceed 2 Ω.

7.13.2 Qualification of Soldering Irons


ESD S20.20 requires that for soldering or desoldering irons three parameters are measured.
These are the tip resistance to ground (or groundable point), tip voltage, and tip leakage
current. The IEC 61340-5-1 standard lists no requirements for soldering and desoldering
tools.

7.13.3 Compliance Verification of Soldering Irons


ESD S20.20 requires that for soldering or desoldering irons only the tip resistance to
ground (or groundable point) is measured. This is usually easily done with a multimeter
(Section 11.9.6).
232 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

7.14 Gloves and Finger Cots


7.14.1 Why Have Gloves and Finger Cots?
There are two main reasons why items such as gloves or finger cots are used in processes
(see Section 4.7.10.7). First, the product being handled may need to be protected from the
oils, salts, microbes, or other contaminants present on the operator’s skin. Gloves may also
be used to improve grip when handling product, tools or other items.
Second, in some processes the operator may need to protect their hands from heat, cold,
chemicals, sharp edges, high voltages, or other threats present in the process. In this case,
the gloves may be classified in some countries as PPE and may be subject to regional,
national, or organization PPE regulations and requirements or other safety regulations.
Local safety requirements must always be observed.
Three possible ESD-related threats are introduced when gloves or finger cots are used.
First, a potentially insulative material is introduced between any item held in the hand
and the wearer’s body. This is a concern in any circumstance where it is desirable for
a handheld item to be grounded via the operator’s hand, for example in use of ESD
hand tools.
Second, for some types of gloves, the materials of the gloves could charge and give elec-
trostatic fields that could lead to ESD risks to ESDS product handled.
Third, contact between the material of a glove or finger cots is likely to give some level
of charging of the ESDS or other items handled. This can lead to charged device, charged
PCB, or other ESD risks.

7.14.2 Types of Gloves and Finger Cots


There are many types of gloves and finger cots made from a variety of materials. PPE gloves
are required for safety reasons in some countries and organizations for some processes. The
local regulations and requirements for these must always be observed.
Gloves may be disposable or reusable. The condition of reusable gloves should be mon-
itored and periodically verified to make sure they remain fit for use. Some types of gloves
and finger cots may use topical treatments or antistatic additives. In some processes, these
could cause unacceptable contamination of the items handled.
Latex, vinyl, and nitrile gloves or finger cots are typically disposable. They may use topical
treatments or antistatic additives.
Fabric gloves may contain surface conductive or core conductive fibers. In activities
where it is important to provide a grounding path between items held in the hand and
the operator’s hand, surface conductive fibers should be used. Core conductive fibers are
unlikely to maintain this conductive path unless topical treatments are also used.

7.14.3 Selection of Gloves or Finger Cots for ESD Control


PPE gloves are required for safety reasons in some countries and organizations. The local
regulations and requirements for these are beyond the scope of this book but must always
7.14 Gloves and Finger Cots 233

be observed. Guidance on selection, care, use, and maintenance of PPE in Europe can be
found in CEN/TR 16832, and the gloves may be subject to requirements given in EN16350
(European Committee for Standardisation [CEN] 2014).
Selection of gloves and finger cots should first identify and address the main reasons
for needing them, e.g. personal protection or avoidance of product contamination. These
may well dictate the basic form of glove to be selected. Other considerations may then
include

● Disposable or multiuse
● Physical and safety protection considerations
● Need for grounding of handheld items
● Possible charging of items handled
● Cost and sensitivity of ESDS handled
● Ease and convenience of use
● Suitability for use in clean areas
● Possible consequences of contamination due to antistats

Some types of glove, e.g. thin latex and vinyl gloves can have surprisingly good properties
for ESD control use even though they were not intended for this purpose. If tests confirm
this is the case, qualified types may be usable in the ESD Control Program.

7.14.4 Qualification Test of Gloves and Finger Cots


Where compliance with a standard is required, testing should be done according to the
standard, and compliance with the standard pass criteria should be established. Where they
are not specified by standards, tests and pass criteria will need to be established as part
of the ESD control program. Most current standards at the time of writing do not specify
requirements for ESD control gloves and finger cots.
As well as qualification based on physical or safety functions, qualification based
on ESD control functions is required. Physical and PPE properties will often be listed
on data sheets. There are two types of test often specified for ESD control gloves and
finger cots. First, resistance tests can be used to establish whether a grounding path
exists between the wearer’s hand and a conductive item held in the hand. These can
be done directly as a resistance measurement or indirectly as a charge decay mea-
surement. Gloves that comply with EN16350 have resistance <100 MΩ through the
material. Second, where ESDS devices are directly handled, electrostatic charging of
items handled can be a concern. Some test methods that can be considered are given
in Sections 11.9.7, 11.9.8.2, 11.9.8.3 and 11.9.9.
Theoretical upper limits for resistance through the glove and charge decay can be cal-
culated in the first instance as for tools (see Section 7.12.3). In practice, it is difficult to
reliably measure high resistances, so it may be better to specify a pragmatic lower resistance
limit easily achievable with product available on the market. If a CPM is available, charge
decay measurements are often easier and more reliable than resistance measurements of
high-resistance glove materials.
234 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

7.14.5 Cleaning and Maintenance of Gloves


Reusable textile gloves may be cleaned according to the manufacturer’s instructions. Topical
treatments may need to be renewed.

7.14.6 Compliance Verification Test of Gloves and Finger Cots


Where disposable gloves are used, a compliance verification test might not be needed pro-
vided a suitable qualification procedure is adopted. For reusable gloves, compliance verifi-
cation tests should be used to check that basic ESD control properties established during
qualification have not been lost. Suitable test methods and pass criteria should be defined.
The frequency of testing should consider aspects such as the following:
● The possibility of contamination
● Established experience of reliability and life
● Cost and sensitivity of the ESDS handled
● The likely consequences of changes to glove electrostatic properties
Perhaps the easiest test to verify a glove or finger cot is a resistance to ground test, done
with the item worn by a grounded person (see Section 11.9.7)

7.14.7 Common Problems with Gloves and Finger Cots


Gloves and finger cots may often be supplied in ordinary polythene packaging. If taken into
an EPA, this can be a source of noncompliant secondary packaging materials that lead to
ESD risks. To prevent this, secondary packaging should be removed before the gloves are
taken into an EPA.

7.15 Marking of ESD Control Equipment


ESD control equipment may be marked as a way of distinguishing them from similar
non-ESD control items. Only items qualified for ESD control should be marked in this
way. A marking symbol recommended by 61340-5-2 and ANSI/ESD S8.1 for identifica-
tion of ESD control equipment is shown in Figure 7.14 (International Electrotechnical
Commission 2018; EOS/ESD Association Inc 2017).

Figure 7.14 Symbol recommended for marking ESD


control equipment. Source: Reproduced by permission of
the EOS/ESD Association Inc.
References 235

References

Baumgartner G. (2000) ESD TR2.0-01-00. ESD Association Technical Report – Consideration For
Developing ESD Garment Specifications. Rome, NY, EOS/ESD Association Inc.
British Standards Institute. (2004) BS EN 1149-3:2004. Protective clothing – Electrostatic
properties – Part 3: Test methods for measurement of charge decay. ISBN 0 580 43736 1
EOS/ESD Association Inc. (1999) ESD TR13.0-01-99. ESD Association Technical Report EOS
Safe Soldering Irons requirements. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013) ANSI/ESD S1.1-2013. Standard for protection of Electrostatic
Discharge Susceptible Items - Wrist Straps. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014a) ANSI/ESD STM9.1-2014. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Footwear - Resistive Characterization.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014b) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015) ANSI/ESD STM97.1-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Floor Materials and
Footwear – Resistance Measurement in Combination with a Person. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2016) ANSI/ESD STM97.2-2016. Floor Materials and
Footwear – Voltage Measurement in Combination with a Person. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2017) ANSI/ESD S8.1-2017 Draft Standard for the Protection of
Electrostatic Discharge Susceptible Items – Symbols – ESD Awareness. Rome, NY, EOS/ESD
Association Inc.
European Committee for Standardisation (CEN) (2014) EN16350-2014 Protective gloves -
Electrostatic properties. Brussels, CEN.
European Committee for Standardisation (CEN) (2015) PD CEN/TR 16832:2015. Selection, use,
care and maintenance of personal protective equipment for preventing electrostatic risks in
hazardous areas (explosion risks) Brussels, CEN.
European Union (2016) Regulation (EU) 2016/425 of the European Parliament and of the
Council of 9 March 2016 on personal protective equipment and repealing Council Directive
89/686/EEC. Available from: https://eur-lex.europa.eu/legal-content/EN/TXT/?qid=
1484921753526&uri=CELEX:32016R0425 [Accessed 17th Aug. 2019]
Holdstock, P., Dyer, M.J.D., and Chubb, J.N. (2003). Test procedures for predicting surface
voltages on inhabited garments. In: Proc. of the EOS/ESD Symp. EOS-25, 300–305. Rome, NY:
EOS/ESD Association Inc.
International Electrotechnical Commission. (1998) IEC 61340-5-1:1998. Electrostatics - Part 5-1:
Protection of electronic devices from electrostatic phenomena - General requirements. Geneva,
IEC.
International Electrotechnical Commission. (2001) IEC 61340-4-3:2001. Electrostatics - Part 4-3:
Standard test methods for specific applications – Footwear. Geneva, IEC.
236 7 Selection, Use, Care, and Maintenance of Equipment and Materials for ESD Control

International Electrotechnical Commission. (2003/2015) IEC 61340-4-1:2003+AMD1:2015


CSV. Electrostatics - Part 4-1: Standard test methods for specific applications - Electrical
resistance of floor coverings and installed floors. Geneva, IEC.
International Electrotechnical Commission. (2004) IEC 61340-4-5:2004. Electrostatics - Part 4-5:
Standard test methods for specific applications - Methods for characterizing the electrostatic
protection of footwear and flooring in combination with a person. Geneva, IEC.
International Electrotechnical Commission. (2005) IEC 60364-1:2005. Low-voltage electrical
installations - Part 1: Fundamental principles, assessment of general characteristics,
definitions. Geneva, IEC.
International Electrotechnical Commission. (2015) IEC 61340-4-6:2015. Electrostatics - Part 4-6:
Standard test methods for specific applications - Wrist straps. Geneva, IEC.
International Electrotechnical Commission. (2016a) IEC 61340-4-9:2016. Electrostatics - Part
4-9: Standard test methods for specific applications – Garments. Geneva, IEC.
International Electrotechnical Commission. (2016b) IEC 61340-5-1:2016. Electrostatics - Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC
International Electrotechnical Commission. (2018) IEC TR 61340-5-2. Electrostatics – Part 5-2:
Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
International Organisation for Standardisation (ISO) (2011a) ISO 20344:2011. Personal
protective equipment – Test methods for footwear. Geneva, ISO.
International Organisation for Standardisation (ISO) (2011b) ISO 20345:2011. Personal
protective equipment -- Safety footwear. Geneva, ISO.
Paasi, J., Nurmi, S., Kalliohaka, T. et al. (2004). Electrostatic testing of ESD-protective clothing
for electronics industry. In: Proc. Electrostatics 2003 Conference, Edinburgh, 23–27 March
2003. Inst. Phys. Conf. Ser. No. 178, 239–246.
Paasi, J., Nurmi, S., Kalliohaka, T. et al. (2005a). Electrostatic testing of ESD-protective clothing
for electronics industry. J. Electrostat. 63 (6–10): 603–608.
Paasi J, Fast L, Lemaire P, Vogel C, Coletti G, Peltoniemi T, Reina G, Smallwood J, Börjesson A.
(2005b) Recommendations for the use and test of ESD protective garments in electronics
industry. Estat Garments Project VTT Research Report No. BTUO45-051338. Available from:
http://estat.vtt.fi/publications/vtt_btuo45-051338.pdf [Accessed 11th Oct 2017]
Smith, D.C. (1993). A new type of furniture ESD and its implications. In: Proc. of the EOS/ESD
Symposium. EOS-15, 3–7. Rome, NY: EOS/ESD Association Inc.
Smith, D.C. (1999). Unusual forms of ESD and their effects. In: Proc. of the EOS/ESD Symp.
EOS-21, 329–333. Rome, NY: EOS/ESD Association Inc.
Swenson D E. (2011) Understanding core conductor fabrics. In: Proc. Electrostatics 2011. J.
Phys. Conf. Se. 301012051 doi: https://doi.org/10.1088/1742-6596/301/1/012051
Swenson, D.E., Weidendorf, J.P., Parkin, D.R., and Gillard, E.C. (1995). Paper 3.6. Resistance to
ground and tribocharging of personnel, as influenced by relative humidity. In: Proc. of the
EOS/ESD Symp. EOS-17, 141–153. Rome, NY: EOS/ESD Association Inc.
Further Reading 237

Further Reading

EOS/ESD Association Inc. (2016) ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
Paasi J, Kalliohaka T, Luoma T, Soininen M, Salmela H, Nurmi S, Coletti G, Guastavino F, Fast
L, Nilsson A, Lemaire P, Laperre J, Vogel C, Haase J, Peltoniemi T, Viheriäkoski T, Reina G,
Smallwood J, and Börjesson A. (2004) Evaluation of existing test methods for ESD garments
VTT Research Report BTUO45-041224 Available from: http://estat.vtt.fi/publications/vtt_
btuo45-041224.pdf [Accessed 11th Oct 2017]
239

ESD Control Packaging

8.1 Why Is Packaging Important in ESD Control?

Ordinary packaging materials are made from a wide variety of materials (Figure 8.1).
Many of these are made of insulating materials such as plastics that charge up easily giving
electrostatic fields and electrostatic discharge (ESD) risks. Others such as papers and
cardboard may have characteristics that vary widely with type and moisture content. Those
that are not insulative at moderate humidity often become insulative at low humidity
below about 30% rh.
Ordinary (non-ESD control) papers and cardboard have highly variable properties that
depend on moisture content and atmospheric humidity as well as the grade of material.
Figure 8.2 shows the variation of surface resistance of some ordinary papers found in the
author’s office, with atmospheric humidity. Many types of cardboard or papers have resis-
tance that varies by several orders of magnitude and may become insulative under dry air
conditions. Some types may be insulative even under moderate humidity conditions. The
unpredictability of the characteristics of these materials makes them undesirable in an area
where static electricity is to be controlled. For this reason, ordinary papers and cardboard
are usually excluded from areas where electrostatic control is required. The moisture con-
tent of papers and cardboard can also be affected by processes. For example, papers passed
through a photocopier may emerge in a dried and charged condition.
Cardboard can be formulated for ESD protective packaging use and is often used to make
boxes. Papers formulated for use in ESD protected areas (EPAs) are also available.
ESD protective packaging is required for two ESD control purposes. First, there is a need
for materials that have known and defined characteristics and can be relied on to provide
little in the way of ESD risk when used in an EPA. Second, materials and packaging are
required that can protect ESD–sensitive (ESDS) devices from ESD risks when outside the
EPA during storage and transport.
ESDS devices, printed circuit boards (PCBs), and assemblies are protected within an
EPA by reducing the ESD threats and sources to a level where ESD damage is insignificant.
When the ESDS devices are to be taken outside the EPA, protective packaging is used to
provide protection from these ESD threats and sources. The ESD protection function may
be combined with other functions such as organizing the ESDS devices into batches or
quantities, types or other groupings, and protection against physical damage or humidity.
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
240 8 ESD Control Packaging

Figure 8.1 Secondary packaging materials.

1.E+13

Office
1.E+12
News
Surface resistance (Ω)

Computer
1.E+11
Viking outer
Viking inner
1.E+10
Mylar
Jiffy bag
1.E+09
Cardboard

1.E+08

1.E+07
20 30 40 50 60 70 80
Relative humidity % rh

Figure 8.2 Variation of resistance of some papers with atmospheric humidity.

ESD protective packaging is now available in an extraordinary variety of types and forms
(Figure 8.3) to suit different types of ESDS devices.
ESD packaging is so important that the International Electrotechnical Commission
(IEC) and ESD Association (ESDA) standards systems have specific standards dealing
with ESD packaging. In the IEC system, ESD packaging requirements are given by IEC
61340-5-3 (International Electrotechnical Commission 2015) and American National
8.2 Packaging Functions 241

Figure 8.3 Some examples of the extraordinary variety of packaging types in use having ESD
control functions.

Standards (ANSI)/ESD S541 in the ANSI/ESDA system. Nevertheless, the topic of ESD
packaging remains one of the most poorly understood areas of ESD control.

8.2 Packaging Functions


ESD control packaging has one or more of the following major ESD protective functions:
● Protection against electrostatic fields.
● Protection against direct ESDs.
● Prevention of buildup of electrostatic charges due to tribocharging.
● The surfaces in normal contact with ESDS devices have sufficiently high resistance to
avoid risk of discharge of on-board batteries or power sources, or charged device ESD
damage.
A variety of ESD packaging solutions are available that have one or more of these prop-
erties. Packaging materials taken into an EPA workstation where exposed ESDS devices
are handled should always be specified as ESD protective packaging. Ordinary (non-ESD
protective) packaging often can easily become charged and cause ESD risks.
Packaging usually also has other functions that have nothing to do with ESD control but
are nevertheless important, for example:
● Moisture barrier properties
● Transparency, so that the contents may be seen and identified or counted
242 8 ESD Control Packaging

● Physical protection
● Holding product ready for use in processes such as automated handling and assembly
● Bearing identification markings or other information
When selecting packaging, all the properties required at all stages of use must be considered.

8.3 ESD Control Packaging Terminology

In ESD control packaging, various terms have been defined by standards or are in common
usage. Unfortunately, they are not always consistent even where standard definitions exist.

8.3.1 Terminology in General Usage


Intimate packaging is the ESD packaging material on the inside of a package that may
contact any ESDS devices within the package. Proximity packaging is any ESD packaging
material that surrounds the intimate packaging. Secondary packaging is any non-ESD con-
trol packaging that may be added around the ESD packaging for physical protection or other
purposes.
Figure 8.4 shows an example of an integrated circuit (IC) protected within an ESD pack-
aging box. The intimate packaging is the foam that holds the IC within the box. The box
itself is proximity packaging. The two together give “ESD shielding” protection. Once sealed
within the box, the device is fully protected against ESD risks and can be taken out of the
EPA or put within secondary (non-ESD protective) packaging for transport or storage.
The electrostatic properties of the materials also have commonly used terms. These cover
the resistive properties of the materials, their ability to attenuate electrostatic fields or act
as a barrier to ESD, and their charge generating properties.
The electrical resistance of materials used in ESD packaging are often classified as con-
ductive, static dissipative, or insulative, or variations on these terms. A conductive material
has comparatively low electrical resistance. A static dissipative material has intermediate
resistance. An insulative material has high electrical resistance and would not normally be

Figure 8.4 Intimate and proximity packaging.


8.4 ESD Packaging Properties 243

used in exposed surfaces of ESD packaging as it may be prone to charge generation and
retention. The exact resistance levels used for these classifications may vary with different
ESD standards. The definitions used in IEC 61340-5-1 (International Electrotechnical Com-
mission 2016b) and ESD S20.20 standards systems (EOS/ESD Association Inc. (2014, 2016))
at the time of writing are given in Section 8.8.
A material may be classified as “electrostatic (or electric) field shielding” if it attenu-
ates electrostatic fields. A material that acts as a barrier to ESD may be classified as “ESD
shielding.” Different ESD packaging standards may vary in the detailed wording of these
definitions and terminology.
The term antistatic has unfortunately been widely used to mean many different things.
Many people refer to ESD packaging as “antistatic” packaging. In some ESD packaging
standards, the term is not used at all because of the risk of confusion. In others, it may be
used with specific meanings.
The property of low electrostatic charge generation is desirable in ESD control packaging
materials. Materials that have lower charge generation properties than ordinary materials
may be classified as “low-charging” materials. They may also be called antistatic materials
by some.
These terms may have specific definitions in ESD packaging standards (Sec. 8.8 and
Tables 8.). The resistance properties of the materials are defined in terms of the surface
resistance Rs or volume resistance Rv measured using defined test methods. Electrostatic
(electric) field shielding is also defined in these standards in terms of the material resis-
tance. In contrast, ESD shielding is defined in two different ways. For bags, it is defined
as the energy measured appearing within the package in response to a standard ESD
waveform applied to the outside of the package in a standard test method. For packaging
other than bags, ESD shielding is defined as a combination of criteria.

8.4 ESD Packaging Properties


ESD packaging must have defined properties, most of which can be tested by the user using
standardized test methods. These test methods are given in Chapter 11. The ESD control
properties of common concern are considered in the following sections.

8.4.1 Triboelectric Charging


The propensity of a material to generate and accumulate electrostatic charge is important
as it is related to the likelihood that the material may cause electrostatic fields and voltage
close to the ESDS. Another aspect is the propensity for a component to become charged by
contact with the packaging material. These are separate but related issues.
Triboelectric charging happens whenever two materials come into contact. It happens
whether they are the same material or different materials. The charging often becomes
evident only when the charged materials separate. The amount of charge generated does,
however, depend on the materials involved, their surface condition, and many other factors
such as surface contaminants including moisture from atmospheric humidity. It follows
that triboelectric charge generation is not a fixed property of the material but depends on the
244 8 ESD Control Packaging

materials that contact it and the circumstances. Materials can be classified or ranked accord-
ing to their relative charging polarity in the triboelectric series (see Section 2.2). Triboelec-
trification is, however, a highly variable effect and notorious for its poor reproducibility.
Materials may be ranked differently by different experimenters in their triboelectric series
(Cross 1987).
Some materials have been designed to be “low charging” or “antistatic” in that their
propensity to generate electrostatic charge is less than other materials. The definitions used
in IEC 61340-5-1 and ESD S20.20 standards systems at the time of writing are given in
Section 8.8. It has been found difficult to standardize tests of this property due to the irre-
producibility of results obtained. No agreement has been possible on pass criteria for a
low-charging property, and it remains a comparative parameter.
Triboelectric charging is often confused with the ability of a material to hold an electro-
static charge and surface or volume resistance properties. There is no correlation between
resistance and triboelectric charge generation properties. In tribocharging, two materials
are involved in contact and charging. If one of these is an insulator, it will become charged
even if the other contacting material is not an insulator. It is quite possible for triboelectric
charging to be stronger when a material contacts a static dissipative or conductive material
than it is when it contacts an insulating material.

8.4.2 Surface Resistance


The surface resistance of a material is related to its ability to dissipate electrostatic charge
and prevent charge buildup. As the term suggests, it is a measure of the resistance across the
surface of a material. This may be measured using a concentric ring electrode, miniature
two-pin probe, or a point-to-point electrode system (see Section 11.8.4).
It is important to understand that the surface resistance of a material is unrelated to
the charge generation properties. The fact that lower voltage buildup may be found with
a lower resistance material is normally due to more rapid charge dissipation. The triboelec-
tric charge generated by contact with the material may be the same or even greater than with
a higher resistance material that shows high voltage buildup. This can be important where
an insulating material such as a device package becomes charged in contact with a resistive
packaging material. The charging of the insulative device package is due to the charge gen-
eration properties of the two materials in combination, not the packaging material resistive
properties.

8.4.3 Volume Resistance


Volume resistance is a measure of the resistance through a material and hence the possibil-
ity that charge or current can flow through the material.
Many modern ESD packaging materials are made of nonhomogenous material, e.g. mul-
tilayered film or surface-coated cardboard. Surface resistance measurements only give the
surface properties of the material. For a nonhomogenous material, each surface may have
a material with different properties, and they may be separated by one or more other mate-
rials with other properties. The electrical resistance between inner and outer surfaces of
a package may be of interest. It may be undesirable to have the inner and outer surfaces
8.4 ESD Packaging Properties 245

electrically isolated from each other. On the other hand, it may be desirable to have a high
resistance or even insulating barrier layer that prevents significant conduction of direct ESD
current between the inner and outer surfaces (see Sec. 8.4.5).

8.4.4 Electrostatic Field Shielding


Electrostatic field shielding refers to the ability of a material or system of materials to act
as a barrier to electrostatic fields. It is important in situations where a sensitive component
within a package must be shielded from the effects of electrostatic fields occurring outside a
package. Electrostatic field shielding is not the same property as electromagnetic shielding
in electromagnetic compatibility work.
The electrostatic field shielding property is related to the conductivity of the material,
because it is related to the ease with which the charge can rearrange in response to the
field, to equalize the potential across the material and hence nullify the field within the
package. A package made from purely insulating material does not give any electrostatic
field shielding effect because the charges cannot rearrange quickly in response to the field.
A good conductor such as metal forms an excellent electrostatic field shield as the charges
within it rearrange almost instantaneously. The field within a container made of an excel-
lent conductor such as metal is zero because the voltage is quickly equalized around the
conductor. This type of container is known as a Faraday cage (see Chapter 2).
For materials of intermediate resistance, the time taken for charges to redistribute around
a container made of the material depends on the effective resistance and capacitive (charge
storage) effects. The higher the resistance, the longer it takes for the charges to redistribute
in response to a change in external field. Various consequences follow from this.
After a long enough time, the field inside a container made of resistive material is nulli-
fied just as it would be with a container made of a good conductor. If, however, the external
electrostatic field is changed rapidly, a transient electrostatic field can arise within the con-
tainer while the charges redistribute. The strength and duration of the field depend on the
characteristic decay time of the charge redistribution. An example of this is shown in the
case of the pink polythene bag response of Figure 8.5b.
If the external field is constantly varying, the effect within the container depends on the
relative speed of variation compared to the characteristic decay times of the charge redis-
tribution within the material of the container. If the external field varies slowly compared
to the container characteristic decay times, good electrostatic field shielding may be seen.
If the external field varies rapidly compared to the characteristic decay times, significant
varying internal fields may arise. Figure 8.5 shows the internal field sensed within an ordi-
nary polythene bag, a pink polythene bag, and a metallized ESD shielding bag during a step
change in external electrostatic field of 5 kV m−1 .
In the case of the polythene bag, the material is a good insulator, and charge moves very
slowly on its surface. The response shows the bag is effectively transparent to the field in the
short term, although charge migration over a period of seconds starts to reduce the internal
field. There is also an initial field due to charge on the bag surface before the external field
is applied.
A package made of material such as pink polythene (see Section 8.7.1.1) has an effec-
tive resistance that depends highly on humidity can also have highly variable electrostatic
246 8 ESD Control Packaging

(a) 6.0
Applied field (kV/m)
5.0
4.0
Electrostatic field (kV/m)

3.0 Field in bag


2.0
1.0
0.0
–0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
–1.0
–2.0
–3.0
–4.0 Time (seconds)

(b) 6.0
Applied field (kV/m)
5.0
4.0
Electrostatic field (kV/m)

3.0 Field in bag


2.0
1.0
0.0
–0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
–1.0
–2.0
–3.0
–4.0
–5.0 Time (seconds)
(c) 6.0
Applied field (kV/m)
5.0
Electrostatic field (kV/m)

4.0 Field in bag

3.0

2.0

1.0

0.0
–0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
–1.0 Time (seconds)

Figure 8.5 (a) Transient electrostatic field arising within an insulating package (polythene bag)
due to external varying field. (b) Transient electrostatic field arising within an intermediate
resistance package (pink polythene at 50% rh) due to external varying field. (c) Transient
electrostatic field arising within a metalized ESD shielding bag at 50% rh due to external varying
field.
8.5 Use of ESD Protective Packaging 247

field shielding effect. Under dry conditions, pink polythene can show little electrostatic field
shielding. Under humid conditions, the same material can show considerable electrostatic
field shielding although transient fields may arise within the package (Figure 8.5b). The
moisture adsorbed by surface antistats allows slow charge migration to reduce the inter-
nal field. When the external field is removed, an internal transient field is produced as the
charges again migrate to a new equilibrium.
In the case of an ESD shielding bag, there is a lower resistance metallized shielding layer
that allows rapid movement of charge in response to the external field. So, the internal field
transient is small (Figure 8.5c).
Until recently, ESD packaging standards, electrostatic field shielding materials have been
classified according by their surface or volume resistance. The definitions and requirements
given in IEC 61340-5-1 and ESD S20.20 standards systems at the time of writing are given in
Section 8.8.

8.4.5 ESD Shielding


ESD shielding refers to the ability of a material (or system of materials) to act as a barrier
to ESDs occurring to the material. It is important where an ESDS component must be pro-
tected against the effects of ESD occurring to the package. ESD shielding is not the same
property as electromagnetic shielding in electromagnetic compatibility work.
An important element of ESD shielding packaging is that there must be a barrier prevent-
ing ESD that impinges against the outside of the package from penetrating and causing ESD
risk to the ESDS devices on the inside of the package. The definitions and requirements
given in IEC 61340-5-1 and ESD S20.20 standards systems at the time of writing are given
in Section 8.8.

8.5 Use of ESD Protective Packaging

8.5.1 The Importance of ESD Packaging Properties


8.5.1.1 Charge Generation and Retention
The charge generation and retention aspect of ESD control packaging is important in pre-
venting ESD risk within an EPA and to any ESDS within the package.
Within the EPA, electrostatic fields must be reduced or eliminated to prevent induced
voltages on ESDS devices leading to ESD occurring when they contact other conductors.
In most cases, it is important that the exposed surfaces of ESD control packaging material
cannot generate and hold electrostatic charge so that it cannot be the source of electrostatic
fields. The tendency to generate and hold charge depends on both the charge generation
properties and the resistance (charge dissipation properties) of the material. In general,
packaging brought into the vicinity of ESDS devices should not be insulating as it would
charge up and cause ESD risk.
It is desirable to minimize the charging of the ESDS device by contact with ESD control
packaging, as it could lead to charged device ESD risk when the ESDS device is removed
from the packaging and contacts a conductive surface. Many ESDS devices have exposed
248 8 ESD Control Packaging

insulating surfaces as well as conductive parts. These parts charge up by contact with the
intimate packaging surface. Charging of the ESDS is dependent on the mutual charging
properties of the ESDS materials and the intimate packaging material and is not related
to the resistance of the intimate packaging material. Charge removal from the conducting
parts of the ESDS device depends on the ESD packaging material resistance, but charge is
not removed from the insulating parts of the ESDS device in this way. In some cases, it can
be most important to minimize electrostatic charging of the ESDS device in this way, and
that is not necessarily done by selecting a non-insulating material.
Where a charged device contacts a low-resistance material, the discharge that occurs can
give charged device ESD risk. This can be minimized by avoiding contact between the ESDS
device and low resistance materials. It is usually desirable for ESD control packaging to have
surface resistance greater than a minimum value (at least 10 kΩ) for this reason.
For very small components, charge retention on the component or on the ESD control
packaging can also result in uncontrolled ejection of the components from the package by
electrostatic attraction or repulsion forces.

8.5.1.2 Electrostatic Field Shielding


Electrostatic fields could form a risk in situations where
● The ESDS is inherently susceptible to electrostatic fields.
● Induced voltages on conductors on the ESDS device could lead to excess voltage and break
down across a part of the device.
● The field could give voltage differences and movement or contact between the ESDS
device and other ESDS devices or conductors could lead to ESD occurring.
In practice, there are few components that are known to be susceptible to damage directly
from electrostatic fields. An example is the photomask reticle, in which induced voltages
between adjacent conductors can lead to ESD between them (See Section 9.5.1.2). Depend-
ing on the circumstances and susceptibility of the ESDS devices to damage from electro-
static fields, electrostatic field shielding may or may not be necessary for packaging protect-
ing ESDS devices outside the EPA.
Effective electrostatic field shielding depends on whether the charge redistribution
needed to balance field changes can happen as quickly as the field is changing. This
depends on various factors including the resistance of the material. For response to rapidly
changing fields, a low material resistance is required, although for slow-changing fields, a
high resistance material can be effective.

8.5.1.3 Electrostatic Discharge Shielding


Most ESDS devices are susceptible in some way to damage from direct ESD to the ESDS
device. This is usually the greatest risk to an unprotected ESDS devices. Most ESDS devices
require protection against direct ESD when outside the EPA. ESD protective packaging for
protecting these ESDS devices when outside the EPA must have ESD shielding properties.

8.5.2 Packaging Used Within the EPA


Within an EPA, ESD risk is normally eliminated by the design of the EPA and the ESD
control equipment used within it. The primary requirement of packaging used within the
8.5 Use of ESD Protective Packaging 249

EPA is that it should not introduce ESD risk to ESDS handled there. For this reason, pack-
aging used in the EPA must not have exposed insulating materials that could charge up
and cause ESD risks. Exposed packaging surfaces must be low-charging, static dissipative
of conductive. This applies equally to intimate and proximity packaging materials.
Clearly, secondary (non-ESD-protective) packaging usually contains exposed insulating
materials or at least materials that have not been characterized for ESD purposes. These
must be kept well away from ESDS devices due to the possibility they could cause ESD risk.
It is therefore normal practice to keep these materials away from workstations where ESDS
may be handled. Many ESD programs find it preferable to keep these materials out of the
EPA altogether.
Occasionally ESD protective packaging could be required within the EPA to protect ESDS
devices against specific threats that are not well controlled by standard ESD control mea-
sures. It is more usual that additional ESD protective properties will be required where ESD
packaging will be used to protect ESDS devices taken out of the EPA.
Some types of ESDS may place additional requirements on ESD packaging materials that
they may contact or be stored within. For example, for a PCB containing a battery, a mini-
mum surface resistance may be specified for intimate packaging due to a concern that the
battery charge may be drained by contact with the material.

8.5.3 Packaging Used to Protect ESDS Outside the EPA


The ESD packaging required to protect ESDS outside the EPA could vary according to the
type and form of ESDS and specific way in which ESD could occur to it. Other requirements
unrelated to ESD control may have a strong influence on the type of packaging selected. The
minimum requirements for ESD packaging used in an EPA normally also apply, because the
ESDS device must be placed into the package or taken out of the package within an EPA.
In general, an ESDS device such as a PCB or semiconductor component will need to be
within ESD shielding packaging for transport or storage outside the EPA. This requires a
barrier of high-resistance or insulating material to be built into the package in a way that
does not cause ESD risk when within the EPA.
Sometimes the ESD risk is limited to specific threats that may be countered by simple
protective packaging solutions. For example, a module may be housed within an enclosure
that provides an effective barrier against ESD, but a residual ESD risk of contact with con-
nector terminals or flying leads may be present. It may be sufficient to provide protective
covers that prevent possible contact with connector terminals in this way.

8.5.4 Packaging Used for Non-ESD Susceptible Items


While a need for ESD packaging for the protection of ESDS devices is clear, it is less obvious
why it should be used with nonsusceptible items. The key point is that any material brought
into an EPA must not cause ESD risk to ESDS devices in the vicinity. Packaging made from
insulating materials could cause ESD risk, so ordinary secondary packaging materials are
normally excluded from the EPA. Any suitable type of ESD packaging could be used to
package nonsusceptible components within the EPA.
250 8 ESD Control Packaging

8.5.5 Avoiding Charged Cables and Modules


ESDS devices that are contained within equipment and not exposed to the outside world
can normally be assumed to be not susceptible to damage by ESD. A possible exception to
this is where ESDS internal parts are connected to connectors and they may be subject to
ESD to the connector, for example from connection of charged cables.
Cables are often supplied in polythene packages that can charge highly, especially under
dry atmosphere conditions. When the cable is taken out of its packaging, it can acquire
a high voltage of several kilovolts by triboelectrification. Highly charged packaging close
to a cable can also induce high voltages on the conductive cores. This represents a highly
energetic ESD source that can discharge on connection to equipment connectors. Cables
can also be charged by handling (See Section 2.6.6).
A similar problem can occur when a module or equipment housed in an insulating enclo-
sure is packaged in a polythene wrapper or other highly insulating packaging. High levels
of charge can arise on the equipment housing or nearby packaging that can induce high
voltages on internal PCBs. These may then discharge on connection to a cable or other
equipment.
This problem can be reduced or avoided by packaging the cable or module in a
low-charging material. Where low atmospheric humidity is not expected, pink polythene
can be a suitable material. This material, however, loses its low charging properties below
about 30% rh. Below this, humidity level another static dissipative material should be used.
Pink polythene should always be used with caution – this material is further discussed in
Section 8.6.2.
Where triboelectric or induced voltages may occur on the internal parts of a module,
connecting or touching of potentially sensitive flying leads or connector pins to wiring or
other conductive items should be avoided. It may be necessary to cover the connectors or
leads to prevent accidental contact.

8.6 Materials and Processes Used in ESD Protective


Packaging
8.6.1 Introduction
The basic materials used in ESD protective packaging are often polymers of some type.
These may have some additive material or treatment or a surface coating to give them the
desired ESD control properties.

8.6.2 Antistats, Pink Polythene, and Low-Charging Materials


Antistats are long surfactant type molecules that have a hydrophilic (water attracting) end
and a hydrophobic (water repelling) end. The hydrophobic end is attracted to the poly-
mer material, and the hydrophilic end attracts water from the air. Antistats used in pink
polythene typically are ethoxylated fatty acid amines or amides – chemicals that give them
their low-charging properties in conjunction with atmospheric moisture (Havens 1989).
“Amine-free” materials typically use amides rather than amines.
8.6 Materials and Processes Used in ESD Protective Packaging 251

Pink polythene is polythene that has been loaded or surface coated with a chemical
antistat. The pink color is a colorant that was originally added to differentiate static
control materials from ordinary secondary packaging. Nevertheless, a pink color does not
guarantee that a material has low-charging characteristics. Users of these materials should
be careful to ensure they know their characteristics before use.
In some types, a form of amine was used that caused oxidation of some metals and stress
cracking in some plastics. For this reason, some manufacturers switched to amide-based
antistats. An effect on device solderability has also been a concern.
The materials are made by compounding the antistat additive with the polymer base
material, usually low-density polyethylene. Other polymers may be used to change proper-
ties such as stiffness and sealability. A material such as crushed silica or calcium carbonate
is added to counter the tendency for the finished film to stick to itself. The finished mate-
rial may have a greasy feel if the proportions of additive, polymer, and antiblock are out of
balance. The additive may contaminate any surface that it contacts. The final material can
be made into film, sheets, or foams or can be injection molded into trays or other forms.
For effective performance, first the antistat must diffuse to the material surface. Second,
atmospheric moisture must be adsorbed to the surface. The amount of antistat at the sur-
face depends on the material formulation but also can vary and reduce with age of the
material, due to evaporation of the antistat. The amount of antistat present at the surface
depends on the relative rates of evaporation and diffusion to the surface from the bulk mate-
rial. The amount of atmospheric moisture available to be adsorbed to the antistat varies
with atmospheric relative humidity. These factors mean that the material properties can
change considerably with atmospheric humidity and age. At low humidity and extended
age, the material may become ineffective. This means that many pink polythene materials
have short shelf life.
Pink polythene materials have also been implicated in solderability concerns and crack-
ing of polycarbonate materials. The latter is due to the additive dissolving into the polycar-
bonate material. Properties such as label adhesion and printability may be reduced.

8.6.3 Static Dissipative and Conductive Polymers


Static dissipative and conductive polymers are polymer materials that have been loaded
with conductive particles such as carbon black, carbon fibers, metal fibers or flakes,
metal-coated fibers, or metal powders (Drake 1996). More recently carbon nanotubes or
fibers have also been used. Resistance levels from 100 GΩ to as low as 1 Ω can be obtained.
The resistance of the material can be relatively free from dependence on atmospheric
humidity. The base materials are commonly nylon (carpets or equipment parts), polyethy-
lene (flexible packaging), polypropylene (tote boxes), acrylonitrile butadiene styrene (ABS,
engineering plastic parts), polyvinylchloride (PVC) (flooring), and others such as rubber,
polyimides, polyester, and phenolics.
The conductivity of the material is produced by contact between conductive particles in
the insulating polymer matrix establishing continuous conductive paths through the mate-
rial. The resistance of the material is not a linear relationship with the percentage loading
of additive (Blythe and Bloor 2008). At low additive loading, the effect on resistance is small
because there are few contacts between conductive particles and hence few conductive
252 8 ESD Control Packaging

Log material resistivity (Ωm)


1014 Ωm

In this region small


changes in
composition give large
changes in resistivity

0 Volume fraction of conductive filler 1


Percolation
threshold 10–2 Ωm

Figure 8.6 Typical variation of polymer resistance with conductive particle loading. Source: Blythe
and Bloor (2008).

paths through the material. When the additive loading increases beyond a level, an abrupt
reduction in resistance occurs (Figure 8.6). This level is known as the critical loading or
percolation threshold. The level is influenced by the nature of the resin substrate, size dis-
tribution, and shape of additive particles and the quality of contact between the particles.
Different fillers are added at different loading to produce a given resistance range. The per-
colation threshold occurs around 10–20% loading for spherical conductive particles but at
lower concentrations for fibrous conductive materials.
The material resistance drops quickly with increasing conductive particle loading above
a threshold loading level. At the same time, particle loading variation will inevitably be
present due to variability of mixing. This means that there can be considerable variation of
material resistance over short distances. Processes producing final packaging products can
also affect the local additive loading and material conductivity and result in high resistance
or insulating regions in the material.
The aspect ratio of the particles is the ratio of particle length to diameter. High-aspect
ratio particles improve contact between particles and hence higher conductivity
(lower resistance).
Adding the conductive particulate material influences the physical properties of the
material, such as flexural modulus, tensile strength, hardness, viscosity, and heat distortion
temperatures. Sloughing or particle shedding can be a problem and can prevent use in
clean areas.
These materials are commonly used to make packaging products such as tote boxes, bags,
tubes, bins, and trays by vacuum forming, injection molding, or film extrusion.
8.6 Materials and Processes Used in ESD Protective Packaging 253

Carbon black is the most common additive used to produce conductive or static dis-
sipative materials as they are cost effective. Carbon-loaded materials are normally black
in color. Carbon black in low loading is used in materials not intended for ESD control,
giving black color and other properties. At these low loadings, the materials remain highly
insulating. So, it is important to understand that a black color does not necessarily imply
a material has static dissipative or conductive properties or is suitable for ESD control use!

8.6.4 Intrinsically Conductive or Dissipative Polymers


Intrinsically conductive and dissipative polymers (ICPs and IDPs) are polymer materials
that have intrinsic conducting properties such as polyaniline or poly pyrrole. These can be
used as coatings or additives to other polymers. They can be used to make conductive fibers
for use in fabrics.
Blending these materials with conventional polymers can result in stable materials in
the static dissipative and conductive resistance ranges. They have been mixed commer-
cially with ABS, polycarbonate, polyester, nylon, and polyurethane and can be translucent,
nonsloughing, and available in many colors.

8.6.5 Metallized Film


Thin metallized films on a polymer substrate (e.g. polyester) are usually laminated in mul-
tilayer structures with other materials to give electrostatic field and ESD shielding and
moisture barrier properties.
The metallization is typically aluminum. The metallization used in ESD shielding bags
is typically thin to be transparent. Thicker metallization used for moisture barrier bags is
typically opaque.

8.6.6 Anodized Aluminum


Aluminum is a highly conductive material, but anodization forms a resistive surface layer of
alumina. Anodized aluminum can be used to make “boats” (holders) or machine parts for
automated handling and tubes for components. The aluminum base can be grounded. The
anodization layer is thin and can be damaged and may also have defects such as pinholes.
The breakdown voltage of the layer can be as low as a few hundred volts and so a charged
component at higher voltage could break down and discharge through the layer on coming
into contact (Bellmore 2001; Smallwood and Millar 2010).

8.6.7 Vacuum Forming of Filled Polymers


Filled conductive and dissipative polymer sheets can heated and be vacuum formed into
complex shapes over a solid mold. The process can result in changes to the conductivity
of the material, especially where it is deeply drawn. Pockets can become isolated from the
surrounding material.
254 8 ESD Control Packaging

8.6.8 Injection Molding


Filled conductive and dissipative polymer materials are often injection molded to form rigid
shapes. The molding process often modifies the conductivity of the material, even produc-
ing nonconductive regions.

8.6.9 Embossing
Embossing can be used to change the surface profile of a material. This can in some cases
be used to modify charge generation properties by reducing surface contact area.

8.6.10 Vapor Deposition


Vapor deposition is used to form a metallized coating on polymer surfaces, for example in
ESD shielding bag or moisture barrier bag manufacture.

8.6.11 Surface Coating


Materials may be surface coated with other materials to give the desired surface properties.
Examples include coating of a cardboard box with a conductive surface layer. With flexible
film materials, coating can be done in moving web form.

8.6.12 Lamination
Several layers of materials with different properties may often be laminated together to form
a package with the desired properties. Examples are ESD shielding bags and moisture bar-
rier bags. Some laminated materials may be vacuum formed, but the forming process may
alter the properties of the laminated layers.

8.7 Types and Forms of ESD Protective Packaging


Modern ESD protective packaging products take many forms. Packaging materials and
products are used individually or as part of packaging systems. These systems are com-
binations of packaging materials that are used to give a required set of physical protections,
ESD control, and other properties.

8.7.1 Bags
There are various types of ESD packaging bags on the market that have different com-
binations of physical and ESD control properties. They may be heat sealable or have a
zip or self-adhesive seal. They are available in low-charging, static dissipative, or conduc-
tive or metallized ESD shielding materials. Low-charging bubble wrap bags are also avail-
able. Some materials are transparent, whereas others are opaque. Some examples of bags
designed for ESD control or protection are shown in Figure 8.7. The bag properties specified
in ANSI/ESD S11.4 (EOS/ESD Association Inc. 2012a) and MIL-PRF-81705-D standards are
given in Table 8.1.
8.7 Types and Forms of ESD Protective Packaging 255

Figure 8.7 ESD packaging bags include pink polythene, black polythene, and metalized ESD
shielding bags among other types.

Table 8.1 ESD control bags standard classification and properties specified.

EMI
Standard Type Electrostatic property shielding
Surface Electrostatic
Low resistance Charge field ESD Moisture
charging or resistivity decay shielding shielding barrier

MIL-PRF- Type I Yes Yes Yes Yes Yes Yes


81705D
Type II Yes Yes Yes
and Ea)
Type III Yes Yes Yes Yes Yes
ESD S11.4 Level 1 Yes Yes Yes Yes Yes
Level 2 Yes Yes Yes Yes Yes
Level 3 Yes Yes Yes Yes
Level 4 Yes Yes
Level 5 Yes Yes

a) Type II has been dropped from MIL-PRF-81705E


256 8 ESD Control Packaging

8.7.1.1 Pink Polythene Bags


“Pink” polythene is a low-cost low charging material that relies on moisture from the air to
give it low-charging properties. It can give reduced charging of materials with which it is in
contact but can leave a contaminating layer of antistat on the surface.
Originally pink polythene was given its pink color to distinguish for ESD control purposes
it from ordinary polythene. Modern versions may be colorless or have various colors. It is a
transparent material and so the bag contents can be easily seen.
The measured surface resistance can vary considerably with atmospheric humidity.
Above about 30% rh, it is usually in the static dissipative range. Below about 30% rh, the
measured surface resistance may rise above 1011 Ω (100 GΩ), and at worst the material
behaves not very different from ordinary (non-ESD) polythene.
This type of bag does not have ESD shielding properties. At high atmospheric humidity,
it may have a variable level of electrostatic field shielding, but this is not usually sufficient
or reliable.
Pink polythene bags are mainly useful for replacing polythene secondary packaging for
use in an EPA in applications where the humidity is not expected to drop below 30% rh.
They may be used to enclose documents or components that are not susceptible to ESD
damage and eliminate electrostatic charge generators from the EPA.
Pink polythene bags are an example of “TYPE II” from the U.S. military stan-
dard MIL-PRF-81705-D or Level 5 in ANSI/ESD S11.4 and “antistatic” in ANSI/ESD
S541 (EOS/ESD Association Inc. 2018). The type was subsequently dropped from
MIL-PRF-81705E.

8.7.1.2 Conductive (Black Polythene) Bags


Black polythene bags are made from carbon-loaded material and have relatively low surface
and volume resistance around 103 –104 Ω (1–10 kΩ). The material is black and opaque, and
the contents cannot easily be seen without opening the package.
The material gives excellent electrostatic field shielding and charge dissipation. It does
not necessarily give low-charge generation to materials in contact with it. As the material
has low-volume resistance, a discharge to the outside of the bags can be conducted through
the material to items within the bag. The material does not in itself have ESD shielding
capability.
The charge may be transferred through the volume of the material to the device instead
of around the material to ground.
Conductive bags are classified as Type 4 in ESD S11.4.

8.7.1.3 Metalized ESD Shielding Bags


ESD shielding bag materials have a multilayer structure that provides a barrier layer pre-
venting direct ESD from being conducted through to the inside of the bag. The structure also
contains a metallization layer that provides an electrostatic field shield and a path for ESD
current to flow around the bag material. The material is generally sufficiently transparent
to view the contents of the package to some extent, without opening the package.
There are two common types of this structure. In the “metal out” structure (Figure 8.8), a
dissipative polyethylene inner intimate layer is bonded to a buried insulative polyester layer.
This insulative layer forms a barrier to ESD current flow through the material. Outside this
8.7 Types and Forms of ESD Protective Packaging 257

Figure 8.8 Typical “metal-out” ESD shielding bags material structure.

Figure 8.9 Typical “metal-in” ESD shielding bag material structure.

layer is a metallization layer. The metallization layer is protected by an outer dissipative


coating.
In the “metal in” structure, the inner dissipative polyethylene layer has outside it the
metallization layer, which is now buried (Figure 8.9). Outside this is a polyester barrier
layer, and this is coated by a dissipative outer layer.
ESD shielding bags can be used to provide ESD shielding protection for ESDS compo-
nents, boards, and assemblies in unprotected areas. They are classified as TYPE III under
MIL-PRF-81705 and Level 3 in ESD S11.4 (see Table 8.1). The ESD shielding performance of
bags is measured using an energy attenuation test (see Section 11.8.6) in which a standard
human body model ESD stress is applied to the outside of the bag. The energy transmitted
to a sensor in the inside of the bag is measured. Figure 8.10 shows results of this test for a
selection of 16 pink, black, and ESD shielding bag types from the market.

8.7.1.4 Moisture Barrier Bags


Moisture barrier bags typically have static dissipative outer and inner surfaces with a
low-charging treatment or coating. They typically provide electrostatic field and ESD
shielding and include moisture vapor barrier protection (see Section 8.8.2). This protects
moisture sensitive items and improves long-term storage.
Moisture barrier bags are typically similar in structure to metallized ESD shielding bags
but are physically stronger than an ESD shielding bag (Figure 8.11). They may have one
or more metallization layer that may be thicker than that of an ESD shielding bag. This
metallization layer is often opaque rather than transparent. The moisture barrier bag is
used when moisture barrier protection or maximum ESD shielding protection is needed.
Moisture barrier bags are classified as Type I in MIL-PRF-81705 and Level 1 or Level 2 in
ESD S11.4 (see Table 8.1).
258 8 ESD Control Packaging

Figure 8.10 Comparison of ESD shielding capability of 2 pink polythene, 3 black polythene, and
11 ESD shielding bag specimens. Source: Smallwood and Robertson (1998).

Figure 8.11 Typical moisture barrier material structure

8.7.2 Bubble Wrap


Bubble wrap material includes sealed gas bubbles for physically cushioning the package
contents. It is available in low-charging material such as pink polythene blow molded
from low-density polyethylene with antistatic additives. It is also available in colors other
than pink.

8.7.3 Foam
Foams are available in insulative, low-charging, dissipative, and conductive materials. They
may be used to cushion items within boxes or as a means of holding leaded package devices
securely (Figure 8.12). A high resistance foam insert within a box or other package system
can help give a barrier to ESD for ESD shielding.

8.7.4 Boxes, Trays, and PCB Racks


Boxes, trays, and PCB racks are commonly made from conductive or static dissipative
carbon-loaded polymers (Figure 8.13). They may have covers or dividers and may be shaped
8.7 Types and Forms of ESD Protective Packaging 259

Figure 8.12 Foam is often used as intimate packaging or cushioning in packaging systems (left
and mid) black carbon loaded static dissipative foam (right) pink low charging polythene foam.

Figure 8.13 ESD packaging boxes.

specifically to suit the product they are to contain. They may be designed for use with
components or assemblies of finished products. They may be made from injection-molded
materials such as polypropylene or vacuum formed from materials such as high-density
polyethylene.
Shipping and storage boxes may also be made from cardboard, typically faced with con-
ductive or static dissipative coatings and containing a foam insert.
Packaging systems based on boxes can include conductive layers or coatings to give elec-
trostatic field shielding and air gaps, high-resistance materials, or even buried insulating
layers as a barrier to ESD for ESD shielding.
Joint Electron Device Engineering Council (JEDEC) and waffle trays (Figure 8.14) are
stackable trays containing multiple pockets in a matrix, designed to hold component pack-
age size and form made to dimensions defined by JEDEC standards. Trays may also be made
to custom sizes.
260 8 ESD Control Packaging

Figure 8.14 JEDEC waffle tray.

8.7.5 Tape and Reel


Tape and reel packaging (Figure 8.15) is typically used to hold surface-mount semicon-
ductor or passive components for automated assembly. The components are held in small
pockets within the tape. These are then covered by a light cover tape held in place with

Figure 8.15 Tape and reel packaging.


8.7 Types and Forms of ESD Protective Packaging 261

adhesive. The tape is then wound onto the reel for transport, storage, and dispensing of
components.
Tape and reel packaging is available in paper or polymer insulative materials for passive
component handling, as well as in dissipative or conductive ESD control materials for use
with ESDS. The reel itself can be made of insulative, dissipative, or conductive material.
When used to contain ESDS devices, the tapes and reel should be made from static dissipa-
tive of conductive materials. The intimate surfaces of the tapes should be static dissipative
or conductive and low charging.
The small feature sizes of the tapes make them impossible to test with current standard
test methods that use large electrodes. The tape forming method can affect the resistivity of
the material and can result in isolation of conductive or dissipative regions within the tape
pockets.

8.7.6 Sticks (Tubes)


Stick magazines (Figure 8.16) are used for transport and storage of ESDS components or to
feed devices to automatic placement machines for surface and through-hole board mount-
ing. Several stick magazines loaded with components are often placed in boxes or bags to
give defined component quantities. The devices are held within the stick using end pins or
plugs, which should also be made of static dissipative material.

Figure 8.16 Component sticks.


262 8 ESD Control Packaging

Stick magazines or tubes are usually constructed of extruded rigid, clear, or translucent
PVC. Metal sticks have also been used, and carbon-loaded static dissipative or conductive
plastic tubes are available. The extrusion shape is tailored to component package types or
shapes. The extruded magazines are surface treated, usually by dipping the tubes in a liquid
solution. The pins and plugs are also treated.

8.7.7 Self-Adhesive Tapes and Labels


Conventional tapes and labels can generate high levels of electrostatic charge. Self-adhesive
tapes and labels have been developed for applications where use in EPAs is required.
ESD tapes and labels typically have surface resistance in the static dissipative range up to
1011 Ω. A low-charging adhesive is used that also allows the charge generated by separating
the label from its backing material to be dissipated quickly.
Tapes supplied on paper cores may be unsuited for use in clean rooms due to particulate
contamination risk. Some are available on static dissipative plastic cores suitable for clean
room use.
Beware that some self-adhesive “ESD tapes” are designed for sealing and marking pack-
aging rather than for use in the EPA – these may be made of insulating materials unsuitable
for use in an EPA.

8.8 Packaging Standards

8.8.1 ESD Control and Protection Packaging Standards


The ESD standard systems IEC 61340-5-1 and ANSI/ESD S20.20 have specific standards
that deal with ESD protective packaging classification and properties. These are IEC
61340-5-3 and ANSI/ESD S541, respectively. These standards give specific classification of
ESD packaging materials in terms of their electrostatic charging properties (Table 8.2) and
resistance range (Table 8.3) and define electrostatic field shielding and ESD shielding prop-
erties (Table 8.4). For product qualification of packaging materials, they are preconditioned
and tested under prescribed environmental conditions. For example, in IEC 61340-5-3,
conditioning for ≥48 and test at 23 ± 2 ∘ C and 12% ± 3% relative humidity is required. These
conditions are set because dry air conditions typically represent “worst-case” conditions
under which the material resistance is likely to be at its highest.

Table 8.2 Classification of packaging materials electrostatic charging properties in IEC


61340-5-3:2015 and ANSI/ESD S541-2019.

Definition
Classification IEC61340-5-3:2015 ESD S541-2019

Low charging Not defined User-defined level that will ensure that ESDS items will
not be charged excessively (produces unacceptable risk
of discharge)
8.8 Packaging Standards 263

Table 8.3 Classification of packaging materials in IEC 61340-5-3:2015 and ANSI/ESD S541-2019
based on their surface and volume resistance range.

Definition
Classification or term IEC61340-5-3:2015 ESD 541-2019

Dissipative Rs ≥ 104 Ω and < 1011 Ω Provides an electrical path for charge to
Rv ≥ 104 Ω and < 1011 Ω dissipate from the package
Rs ≥ 104 Ω and < 1011 Ω
Rv ≥ 104 Ω and < 1011 Ω
Surface conductive Rs < 104 Ω Provides an electrical path for charge to
dissipate from the package
Rs < 104 Ω
Volume conductive Rv < 104 Ω Provides an electrical path for charge to
dissipate from the package
Rv < 104 Ω
11
Insulative Rs ≥ 10 Ω Rs ≥ 1011 Ω
11
Rv ≥ 10 Ω Rv ≥ 1011 Ω

Table 8.4 Classification of packaging materials as electrostatic field or ESD shielding in IEC
61340-5-3:2015 and ANSI/ESD S541-2019.

Definition
Classification or term IEC 61340-5-3:2015 ESD 541-2019

Electrostatic field Capable of attenuating an See “electric” field shielding


shielding electrostatic field
Rs or Rv < 103 Ω
Electric field shielding See “electrostatic” field shielding Attenuates electrical fields
Electrostatic discharge Capable of attenuating an Protects packaged items from the
shielding (bag) electrostatic discharge. effects of static discharge that are
Calculated energy within a bag is external to the package and limits
<50 nJ when tested to 61340-4-8 current flow through package
(International Electrotechnical Calculated energy within a bag is
Commission 2014). <20 nJ when tested to ANSI/ESD
(Edition 2: …or an equivalent STM11.31 (EOS/ESD Association
test method modified to Inc. 2012b)
accommodate the product.)
Electrostatic discharge Intimate packaging shall be User defined
shielding (other conductive or dissipative.
packaging) A barrier layer or a defined air
gap attenuating ESD energy shall
be included.
No component of the packaging
system shall cause ESD risk
when taken within EPA.
264 8 ESD Control Packaging

These ESD packaging standards also give examples of markings that are recommended
for use to identify packaging used in ESD control and protection. These are used to identify
the packaging material or warn the user that ESDS devices may be contained within the
package. This is discussed further in Section 8.10.

8.8.2 Moisture Barrier Packaging Standards


8.8.2.1 Handling, Packing, Transport, and Use of Moisture-Sensitive Devices
Automated assembly processes and the use of surface-mount devices (SMDs) led to dam-
age such as cracks and delamination occurring due to from the solder reflow process. These
problems arise because atmospheric moisture is absorbed into permeable component pack-
age materials. During solder reflow, SMDs are exposed to temperatures over 200 ∘ C. Rapid
moisture expansion and materials mismatch can result in cracking and/or delamination of
material interfaces within the device.
IPC/JEDEC J-STD-033D describes the standardized levels of exposure for moisture/
reflow-sensitive SMDs and the requirements for handling, packing, and shipping to
avoid moisture/reflow-related failures (IPC, JEDEC 2018). The standard applies to all
devices subjected to bulk solder reflow processes during PCB assembly, including plastic
encapsulated packages, process sensitive devices, and other moisture-sensitive devices
made with moisture-permeable materials (epoxies, silicones, etc.) that are exposed to the
ambient air.
The standard calls for drying of moisture-sensitive devices followed by packing within
moisture barrier packaging. IPC/JEDEC J-STD-033D calls for use of moisture barrier bags
compliant with MIL-PRF-81705 Type I. The materials of these packaging materials are
designed to limit moisture passage to ≤0.0310 g m−2 per day through the material. Moisture
barrier packaging is marked with a symbol specified in MIL-PRF-18705D and ESD S11.4
(Figure 8.17). The materials are heat sealed after packing, with desiccant and a humidity
indicator card.

8.8.2.2 MIL-PRF-81705
MIL-PRF-81705D 2004 gave specifications for heat-sealable electrostatic protective flexible
barrier materials used for military packaging of ESDS for use by the US Department of

Figure 8.17 Example of moisture sensitive device packaging label specified in MIL-PRF-18705D
and ESD S11.4.
8.8 Packaging Standards 265

Table 8.5 Applications of static control moisture barrier material classifications defined by
MIL-PRF-81705D.

Type Properties Application

I Water vapor proof, electrostatic protective, Water vapor proof, electrostatic, and
electrostatic, and electromagnetic shielding electromagnetic protection of microcircuits,
and semiconductor devices, such as diodes,
field effect transistors, and sensitive resistors
II Transparent, waterproof, electrostatic Use where transparency and static
protective, static dissipative dissipation is required and contact with oil or
grease is not contemplated
III Transparent, waterproof, electrostatic Use where a transparent, waterproof,
protective, electrostatic shielding electrostatic-protective, and electrostatic
field protective barrier is required

Table 8.6 Static control moisture barrier material requirements of MIL-PRF-81705D.

Electrostatic property requirements


Surface
resistivity Electrostatic ESD Electromagnetic
Type 𝝆s (𝛀 sq−1 ) Static decay shielding shielding attenuation

I Inner: <2 s max 30 V peak 10 nJ >25 dB


105 ≤ 𝜌s < 1012
Outer:
𝜌s < 1012
II Inner: <2 s
105 ≤ 𝜌s < 1012
Outer:
𝜌s < 1012
III Inner: <2 s Max 30 V peak 10 nJ >10 db
105 ≤ 𝜌s < 1012
Outer:
𝜌s < 1012
Test method ASTM D257 FED STD 101 EIA 541 2 probe ANSI/ESD Specified in
method 4046A method 1000 V S11.31 standard
(superseded by
MIL STD 3010
METHOD 4046)

Defense. It classified materials as Type I, Type II, or Type III (Tables 8.5 and 8.6). Each
type has two subclasses: Class 1 for unlimited use, and Class 2 for use on automated
bag making machines only. The standard specifies physical characteristics such as seam
strength, water vapor transmission rate, thickness, and transparency as well as electrostatic
properties.
266 8 ESD Control Packaging

Type I – Water vapor proof, electrostatic protective, electrostatic, and electromagnetic


shielding.
Type II – Transparent, waterproof, electrostatic protective, static dissipative.
Type III – Transparent, waterproof, electrostatic protective, electrostatic shielding.
The main properties and applications of the materials classified by 80715D are summarized
in Table 8.5. The ESD-related requirements of these materials are summarized in Table 8.6.
A new version, MIL-PRF-81705E:2009, dropped the Class II classification.
MIL-PRF-81705D specifies some different properties than those specified in ESDA and
the IEC standard. First, the surface properties are specified in terms of surface resistivity
measured according to ASTM D257 (ASTM International 2014), not surface resistance.
Second, “electrostatic shielding” is specified in terms of voltage measured using the EIA
541 (Electronic Industries Assoc. 1988) two-probe method of ESD S11.31. Third, “static
decay” is specified, measured according to FED STD 101 Method 4046A (superseded by
MIL STD 3010 METHOD 4046 (Department of Defense 2002)). These differences are fur-
ther discussed in Section 8.8.3.

8.8.2.3 ESD Association ANSI/ESD S11.4


The ESDA ESD S11.4 defines five levels of bags for static control purposes (Table 8.7), listed
here:
Level 1 bags are moisture barrier bags for device packaging intended to protect items that
will be subjected to reflow soldering. They provide moisture barrier properties prevent-
ing excess moisture absorption that can lead to device body cracking. They include ESD
and electrostatic field shielding properties. The inner surfaces are low charging to avoid
charge accumulation on or in the bag. The inner and outer surface resistance may differ
for each surface and allows charge to dissipate from the inside or outside surface of the
bag when the surface is grounded. They are typically a multilayer structure of metal foil
and plastic layers containing or coated with an antistat.
Level 2 bags are moisture barrier bags for general packaging to protect items that will not
be subjected to reflow soldering. They have ESD and electrostatic field shielding prop-
erties to protect electronic items from ESD. They have low-charging properties to avoid
charge accumulation on or in the bag. Their inside and outside surface resistance may be
different for each surface and allows charge to dissipate when the surface is grounded.
They typically have a multilayer structure of metalized plastic and contain or are coated
with an antistat.
Level 3 bags provide ESD and electrostatic field shielding properties intended to protect
ESDS electronic items. Their inner and outer surface resistance may differ and allows
charge to dissipate when the surface is grounded. They are typically a multilayer struc-
ture of transparent metalized plastic and plastic layers containing or coated with an
antistat.
Level 4 bags are conductive and intended for protection of ESDS items. Their surface and
volume resistance allow charge to dissipate from the bag when in contact with ground.
They have electrostatic field shielding properties but do not provide ESD shielding. They
are typically constructed from extruded plastic containing conductive materials and have
the same surface resistance on inner and outer surfaces.
Table 8.7 Static control moisture barrier bags levels defined by ESD S11.4.

Property
Low Electrostatic ESD
Level Application Structure charging field shielding shielding Surface resistance (𝛀) Moisture barrier

1 Device packaging: Multilayer Yes Yes Yes Yes. Inner surface may be different Yes
Items subject to metalized plastic <20 nJ from outer surface Inner: ≤0.002 g/100 in.2 /d
reflow soldering 104 ≤ Rs < 1011
Outer:
Rs < 1011
2 General packaging: Multilayer plastic Yes Yes Yes Yes. Inner surface may be different Yes
Items not subject to <20 nJ from outer surface Inner: ≤0.02 g 100 in.2 /d
reflow soldering 104 ≤ Rs < 1011
Outer:
Rs < 1011
3 “Static shielding” to Multilayer plastic Yes Yes Yes Yes. Inner surface may be different No
protect ESDS <20 nJ from outer surface Inner:
104 ≤ Rs < 1011
Outer:
Rs < 1011
4 Conductive to Conductive Yes Yes No Yes. No
protect ESDS extruded plastic Rs < 1011
5 Static dissipative to Extruded plastic Yes No No Yes. Inner and outer surface No
protect ESDS with antistat typically similar Inner:
104 ≤ Rs < 1011
Outer:
104 ≤ Rs < 1011

Test method STM11.31 STM11.11 (EOS/ESD Association ASTM F1249 (ASTM


Inc. 2015a) International 2013)
268 8 ESD Control Packaging

Level 5 bags are static dissipative bags for protection of ESDS items. They have low-charging
properties intended to avoid charge accumulation on or in the bag that could be damag-
ing to ESDS items. Their inner and outer surface resistance is typically similar and is
intended to allow charge to dissipate when the surface is grounded. They are typically
made from extruded plastic containing or coated with an antistat.

8.8.3 ESD Control Packaging Measurements


ESD-related standard measurements of packaging materials and products fall broadly into
four types, listed here:
● Resistance measurements to characterize the resistance across a surface or through a
material
● Charge decay measurement
● Tribocharging measurement
● ESD shielding
Packaging materials may, of course, be also tested for many other non-ESD-related proper-
ties such as physical protection and moisture barrier properties.
Various standard measurement methods have been produced over the years, used by dif-
ferent ESD packaging standards. Test methods specified for use within the 61340-5-3 and
ESD S541 standards are given in Section 11.5.
It is important to understand that in general, different test methods are likely to give dif-
ferent results due to variations in test electrode systems, procedures, and test voltages. It
is important to make sure that when specifying and selecting materials for use within an
ESD control program that the characteristics are specified and tested using appropriate test
methods. Standards normally specify test methods that are to be used to make measure-
ments to evaluate properties for compliance with the limits set in the standard.
The test methods required by 61340-5-3 and S541 are nearly identical in most cases.
Table 8.8 gives the approximate equivalence of IEC and ESDA test standards. There may
be some differences between the test methods given in the standards in some cases.
One apparent major difference is seen in the specification of material surface resistivity
(measured according to ASTM D257) in MIL-PRF-81705 and surface resistance (measured
according to IEC 61340-2-3 and ANSI/ESD S11.11) in the ESDA and IEC 61340 standards.

Table 8.8 Approximate equivalence of ESD packaging test standards in the


IEC 61340-5-x and ESDA standards systems.

Standards system test


method standards

Measurement IEC 61340-5-3 ANSI/ESD S541

Surface resistance 61340-2-3 ESD S11.11


Volume resistance 61340-2-3 ESD S11.12
Point to point surface resistance 61340-2-3 ESD S11.13
ESD shielding 61340-4-8 ESD S11.31
8.9 How to Select an Appropriate Packaging System 269

Surface resistivity is defined as a material surface parameter that is corrected for the mea-
surement electrode form (see Section 1.7.2). In contrast, surface resistance is the result
of a measurement made on the surface, with no correction for electrode form. The elec-
trodes specified in IEC 61340-2-3 (International Electrotechnical Commission 2016a) and
ANSI/ESD S11.11 give a factor of 10 reduction in the resistance measured, compared to the
surface resistivity of the material. So, a material with surface resistivity of 1012 Ω according
to MIL-PRF-18705 would also be at the surface resistance limit of 1011 Ω as specified and
measured according to IEC 61340-5-3 or ANSI/ESD S541.

8.9 How to Select an Appropriate Packaging System

8.9.1 Introduction
In selecting suitable packaging for an ESDS, many factors may need to be considered. These
include the type of ESDS and likely ESD threats, the environment through which it will
pass, physical and environmental protection requirements, and customer requirements.
The overall cost of the packaging is likely to be a significant consideration.

8.9.2 Customer Requirements


The first consideration is, does the customer place mandatory requirements on the packag-
ing that have been specified in contracts? If so, there may be only one option – to comply
with these requirements.
Even where no contractual customer requirements exist, it may be necessary to consider
likely customer and marketing needs. An example might be packaging a computer PCB for
retail sale. The package may need to be transparent and show the product attractively in
a retail display stand. It may be necessary to design the package according to the type of
display stand on which it will appear. The package may need to display product features
and description or marketing or technical information to help the buyer select it.
A package may need to be tamper proof or designed to help prevent counterfeiting. Alter-
natively, it may need to be designed to allow inspection of the contents in some way.

8.9.3 What Is the Form of the ESDS Device?


One of the first considerations is to examine the form of the ESDS device as this may elim-
inate some packaging possibilities or suggest others. ESDS devices may be components or
devices, PCBs, modules or assemblies. The size, weight, and value of the ESDS could be
important considerations. While small components may be packaged one or more in a bag,
a large component or module may need a sizable or even custom-built box to contain it.

8.9.4 ESD Threats and ESD Susceptibility


ESD can occur when two conductors at different voltages touch (or come sufficiently close)
so that a discharge can occur between them, and the ESD current that flows is sufficient
270 8 ESD Control Packaging

to do damage. ESD can be prevented by ensuring that contact between the ESDS device
and other conductors cannot happen, by ensuring that significant voltage differences do
not exist between the ESDS device and conductors that it may make contact, or that when
ESD occurs, its strength (peak current) is limited and discharge energy is absorbed by high
resistance of the contacting material.
There are two types of ESD threats that should be considered. The main one is protection
of the ESDS devices against direct ESD that may occur from an external ESD source or from
a charged ESDS on contact with an external conductor.
Electrostatic fields are a second threat. While most ESDS devices do not suffer damage
directly from electrostatic fields, they can set up the conditions under which ESD can occur
by inducing high voltages on conductors within the ESDS device (see Section 2.4.4).
Where the ESD threats to the ESDS are not well understood, then it is best to protect the
ESDS device against direct ESD to or from the ESDS device. Protection against electrostatic
fields should also be considered. It may often be most cost effective to take this approach
rather than analyze and evaluate the specific ESD threats to the ESDS device.
If the specific ESD threats to the ESDS device are well defined, it may be possible to choose
ESD packaging that directly prevents those threats. An example is an ESDS electronic mod-
ule built into a metal shielding housing, in which the only threat is the possibility of ESD
to connector pins. In this case, it may be sufficient to protect the connectors against direct
ESD, for example by use of static dissipative or conductive connector cover caps. In some
cases, even insulative caps may be acceptable if these do not present ESD risk within EPAs.
There is a risk of ESD wherever a conductive part of an ESDS device can contact another
conductor (person, equipment, or object) in the environment. The ESD threat, of course,
depends on the susceptibility of the exposed ESDS part to the ESD source with which it
makes contact. Possible sources of ESD include
● ESD from charged personnel
● ESD from charged packaging, objects, or equipment
● The possibility that the ESDS may become charged and discharge to an item that they
touch
The number of ways in which ESD can enter an ESDS device can vary tremendously with
the type and form of ESDS device. A PCB or multipin component can have many ESD entry
points that could touch an external item or person resulting in ESD. Some types of modules
may have few contact points limited to exposed flying leads or connector pins. In some
cases, connector pins may be recessed in a connector housing to the extent that contact
with them is highly unlikely. In other cases, a connector may extend in a way that it is
extremely likely that it will be the first point of contact with another item.
The ESD susceptibility of an ESDS device, and ways in which ESD can enter the ESDS
device, is likely to change with the build state. Ultimately, many ESDS products are pro-
tected against ESD by being built into a housing or potted, preventing ESD from occurring
in all but a small number of well-defined ways. The final product may be considered insen-
sitive to ESD. In some markets, e.g. Europe, the product may need to pass ESD susceptibility
tests before it can be placed on the market. Some customers may have similar requirements.
A change in build state can also introduce a new ESD risk. Potting or building a PCB into
a plastic housing can prevent ESD into all parts except connector pins. At the same time, it
8.9 How to Select an Appropriate Packaging System 271

introduces the risk that the housing can become highly charged and induce high voltages
on the internal PCB, giving a risk of damaging “charged module” ESD on connection to a
wiring loom or other external item.

8.9.5 The Intended Packaging Tasks


The packaging task may require consideration of various aspects such as the quantity of
ESDS device housed, physical protection needs, package life, and recyclability.
In some cases, only one ESDS device is packaged, and in other cases multiple ESDS
devices must be packaged. An example might be a tray designed to accept multiple PCBs
for storage and transport.
Packaging may need to protect the contents against physical impact, crushing, and other
physical threats. This requirement is likely to be greater for packaging that is intended to
protect ESDS devices for transport outside the EPA and between sites, especially using com-
mercial transportation networks.
The packaging may be used in transport or storage, or both. Storage may include tasks,
such as display in a retail setting, that may place special requirements on the package. Pack-
ages may need to be designed for stacking in specific storage areas or using specific stacking
and storage methods.
The packaging may be single use or multiple use. Packaging products meant for single
use may have lower initial cost. Some types of packaging (e.g. bags) may be less physically
robust than others (e.g. boxes and rigid or semirigid containers). The initial cost of these
more robust packages may be relatively expensive in initial cost but may be the least expen-
sive choice when all lifetime costs are considered. For reusable packaging, the cost of the
material can be amortized over the anticipated number of times it can be reused. A means
of testing for detecting failed or damaged packages may be required. Where multiple use
is a possibility, a cost/benefit analysis should be made to evaluate the packaging material
for the intended purpose over its lifetime. The cost of returning packaging for reuse may
need to be considered with other lifetime costs. These costs may include collecting, sort-
ing, cleaning, and preparing materials for the return shipment and reuse, testing costs, and
transport costs.
Where packaging is to be reused, it may be necessary to test and verify the properties of
each element of the package during its lifetime. Packaging surface and volume resistance
properties can be evaluated by the user using simple test methods. Other types of packaging
properties (e.g. ESD shielding) are difficult to evaluate without sophisticated equipment.
For this reason, metallized ESD shielding bags may be considered one-use packaging. The
cost of testing should be considered as part of the lifetime cost of reusing the packaging.
Some materials may have a limited shelf life that needs to be considered if items or the
packaging material are to be stored for long periods.
Recycling and recovery of discarded or failed packaging materials should be considered.

8.9.6 Evaluate the Operational Environment for the Packaging


If the packaging is to be used only within an EPA, then the main ESD protective task may
be to ensure that the packaging does not charge up and provide ESD risk to ESDS items.
272 8 ESD Control Packaging

In many cases, the packaging will need to protect the ESDS device for transport and stor-
age outside the EPA. In this case, it must protect against the anticipated ESD threats as well
as provide no ESD risks to ESDS devices while within the EPA.
Within an EPA, the package may be handled only by trained personnel, which may reduce
the requirement, for example for physical protection. Outside the EPA the package may
be handled by untrained personnel and must withstand the physical threats likely to be
encountered when handled in the same way as any other package.
Within an EPA, it is often necessary to use ESD packaging for non-ESDS items. Secondary
packaging must be kept well away from ESDS devices. Many ESD programs keep secondary
packaging out of the EPA completely. Any suitable ESD packaging can be used to package
non-ESDS within, and for entry into an EPA. For example, a conductive or static dissipative
tote box can equally be used to contain non-ESDS items. Conductive or static dissipative
bags can be used for non-ESDS items instead of ordinary polythene bags. Low-charging
(pink polythene) can also be used for non-ESDS device inside or outside the EPA, providing
the atmospheric humidity within the EPA is not expected to go below about 30% rh. Pink
polythene should also be checked regularly to make sure it has not lost its properties due to
loss of antistat.
If the package is required to enter a clean area, it may be necessary to select materials that
will not give a risk of contamination. This may prevent use of some common ESD control
materials such as carbon-loaded materials or pink polythene. Intrinsically dissipative or
conductive polymer materials may be more acceptable in this case.
The physical and environmental threats are likely to be more severe where packages go
out of the factory for transport between sites or to the customer.
Use of third-party courier or postage services may require a greater level of protection
than use of company transport services. Packages may be subject to transport by air, ship,
rail, or road during a journey.
Where the package crosses national borders, there may be a need to inspect the contents
by personnel who are untrained in ESD control procedures.
The temperature and humidity likely to be encountered during transport and storage may
be subject to considerable variation. The packaging may need to protect against wide tem-
perature and humidity extremes during vehicle or air transport. Even within an EPA, the
humidity may vary dramatically with season, weather, and the presence of heat sources
such as ovens.
Packaging used within an EPA should always be selected to maintain its properties to
the lowest humidity level that may be encountered in practice. Modern packaging stan-
dards usually require materials to be characterized at low humidity (e.g. 12% rh). Some
materials (e.g. pink polythene) rely on antistats and atmospheric moisture for their prop-
erties. These materials be used with caution where the humidity is likely to drop below
about 30% rh. Below this humidity, their properties may become little better than ordinary
secondary packaging materials.
If the package is to take the items to a customer, then customer requirements may be
a strong consideration. Some customers may only accept items packaged as specified in
contracts or standards.
The way in which the package is used, stored, or handled at the customer’s site may give
specific requirements for the package and package marking. For example, a retail customer
8.9 How to Select an Appropriate Packaging System 273

may require packaging that looks attractive for retail display of the product. The package
may need to carry information on the contents to help the consumer select a product for
purchase.
In some cases, it may be sufficient to protect the package contents from ingress of rain,
for example during postage and transport. In other cases, the need for moisture protec-
tion may include maintaining low-humidity conditions within the package to avoid process
problems at a later stage. High humidity can cause problems to electronic parts including
corrosion and difficulties in soldering. It may be necessary to enclose parts susceptible to
such problems within moisture barrier packaging.
The expected temperature range during transit and storage could be an important con-
sideration. In a sealed package, the relative humidity level doubles with a 10 ∘ C fall in
temperature. If the temperature falls below the dew point (see Section 1.11) condensation
may occur on surfaces within the package. A desiccant may be used to reduce the air humid-
ity to a minimum, or the air can be evacuated during sealing to prevent this. The package
may need to have a moisture barrier to prevent vapor ingress.
Removing the water in the internal space of the package with a desiccant can result in
the “worst-case” low-humidity conditions where electrostatic charging may be a concern.
Even without the presence of desiccant, low-humidity conditions can arise due to temper-
ature changes. When ambient temperature increases, the internal relative humidity drops,
halving with a 10 ∘ C rise in temperature.

8.9.7 Selecting the ESD Packaging Type and ESD Protective Functions
8.9.7.1 Intimate Packaging
Intimate packaging is the material in contact with the ESDS device within the package.
It must not charge up appreciably to avoid electrostatic risk to the ESDS device and other
items in an EPA.
The material must not cause significant charging of the ESDS device that is in contact
with it. At first sight, a packaging material that does not retain charge might be expected to
also minimize charging of the ESDS device. However, charge accumulation on any material
is the product of two characteristics – charge generation and the dissipation of charge from
the material. Typically, the ESDS device has some insulating surfaces as well as conducting
surfaces. Any charge generated by contact of these insulating surfaces with the packaging
can be retained on the ESDS device. In contrast, the packaging material typically should be
a noninsulative material and dissipate the charge produced on this material. So, the charge
retained on the packaging is not correlated to the charge retained on the ESDS device. For
low charging of the ESDS device, a material that will give low charge generation character-
istics with the ESDS device should be selected.
It is impossible to avoid some charge generation by contact of the ESDS device with
the intimate packaging material. For dissipation of charge generated by contact with the
ESDS device, the intimate packaging material should be static dissipative or conductive
(see Section 8.3.1 and Table 8.3). Any making or breaking of contact with the ESDS device
will result in ESD at some level. If a high-resistance intimate packaging material is used,
the ESD that occurs has its peak current limited by the high resistance of the material. The
274 8 ESD Control Packaging

energy in the discharge is mainly absorbed in the intimate packaging material rather than
the ESDS device.
It may also be necessary to set a lower limit to the surface resistance of the intimate pack-
aging. For example, a PCB with on-board battery may require an intimate packaging of
sufficiently high resistance to avoid discharge of the battery by contact with the packaging.
The physical design of the intimate packaging can also give benefits. For example, a pur-
pose made rigid material can be shaped to hold the ESDS device firmly while introducing an
air gap to minimize contact with the surfaces of the ESDS device (and hence charging) and
at the same time act as a barrier to direct ESD. As another example, foam inserts within
a box can be used to hold ESDS devices of a variety of size or form firmly with minimal
movement and can virtually eliminate making and breaking of contact with ESDS leads.

8.9.7.2 Proximity Packaging


The proximity packaging design and materials will depend very much on the circumstances
of usage. Table 8.9 shows the characteristics required of ESD packaging when used inside
the EPA and outside the EPA. As ESD packaging must go into an EPA, it should not charge
appreciably to avoid risk to ESDS items within the EPA. The surfaces that may become
exposed within the EPA should be low charging and static dissipative or conductive.
Proximity packaging will often also be designed to give many of the physical and other
protective functions required of the package for transport or storage of the ESDS, within or
outside the EPA.
Where the package will go out of the EPA, the need for electrostatic field or ESD shielding
should be considered (see Section 8.4.4 and 8.4.5). The most important of these is usually
ESD shielding against direct ESD from ESD sources in the environment. In an unprotected
area, these could include ESD from personnel, carts, and other substantial metal items.
ESD shielding is provided by a barrier preventing or attenuating ESD current flowing
through and from the outside of the package to the ESDS device. The objective is to reduce to
a low level the ESD current that can flow through to the ESDS device. This can be provided
by an air gap, a layer of high resistance material, or even (as in ESD shielding bags) a buried
layer of insulative material. It is important that any insulating material cannot be exposed
to give ESD risk when in an EPA.

Table 8.9 Electrostatic characteristics of ESD packaging required inside and


outside the EPA.

Packaging property Inside EPA Outside EPA

Low charging Required Required


Surface resistance Static dissipative Static dissipative or conductive
(inner and outer) or conductive
Electrostatic field Not required Optional, depending on ESDS
shielding failure modes
ESD shielding Not required Usually required, depending on
ESDS failure modes
8.10 Marking of ESD Protective Packaging 275

Electrostatic field shielding is given by inclusion of a layer of low-resistivity material in


the package. This should enclose the whole intimate package forming a “Faraday cage” (see
Section 2.4.3).

8.9.7.3 Packaging Systems


The functions of ESD packaging can be provided by one carefully designed packaging item
(e.g. ESD shielding bag) or can be provided by a system of packaging materials that in com-
bination provide the functions required. Figure 8.4 shows an example of a box packaging
system. The foam intimate packaging that gives some degree of attenuation of ESD cur-
rent that can flow to the ESDS gives physical cushioning to the ESDS and holds the device
securely. The conductive box proximity packaging acts as a Faraday cage giving electrostatic
field shielding and significantly reduces voltage differences within the box. It also encour-
ages ESD current to flow around, rather than through the package. The combination gives
both ESD shielding and electrostatic field shielding.
There may be many aspects of the packaging that require consideration for reasons unre-
lated to ESD control and protection. These may include physical protection against crush-
ing, shock and vibration, strength, cleanliness, chemical contamination, plastic compati-
bility, marking and bar code reading, transparency, and testability.
The complete packaging system for transport and storage outside an EPA is likely to
include secondary outer packaging to give some of the protection unrelated to ESD control.

8.9.8 Testing the Packaging System


In some cases, it may be necessary to test the packaging system to ensure that the con-
tents will not become damaged in transit or storage. The tests may include physical tests
such as vibration or drop tests as well as environmental tests such as rain, humidity, and
temperature cycling.
Electrostatic and ESD testing and pass criteria should be selected to reflect the anticipated
ESD threats and sensitivity of the ESDS. They may include aspects such as application of
ESD to the package or electrostatic charging evaluation of the ESDS enclosed.
Packaging should be checked and tested before reuse. The first test that should be done is
a visual check for damage during usage. ESD protective packaging should never be reused
if there is visible damage that might affect its performance. Bags can easily be damaged by
puncture by leads or sharp edges on components or PCBs within. Coatings can be torn off
by removal of documents or labels attached to the outside of a package.

8.10 Marking of ESD Protective Packaging


There are various types of ESD packaging markings encountered in current practice. These
inform the user that the package is likely to contain ESDS or give details of the type of
packaging material. At the time of writing the IEC 61340-5-1 and ESD S20.20 standards
require that marking is done in accordance with customer contract requirements if they
exist. If no customer requirements are in place, then these standards require that the need
for marking requirements must be determined and if necessary suitable markings defined.
276 8 ESD Control Packaging

Figure 8.18 Example of packaging marking to identify ESD control packaging or materials
according to IEC 61340-5-3, ANSI/ESD S541 and ANSI/ESD S11.4. Source: Reproduced by
permission of the EOS/ESD Association Inc.

The ESD packaging standards IEC 61340-5-3 and ANSI/ESD S541 give examples of ESD
packaging marking that can be used with ESD control and protective packaging. The outer
surface of an ESD protective package should be marked with a symbol and/or wording
identifying it as a package that is expected to contain ESDS devices (Figure 8.18). This is
so that personnel handling the package can easily recognize that the package may contain
ESDS devices and should not be opened further outside an EPA. Where a packaging system
has several materials in layers, each material layer should be identified as an ESD control
material. The “hand” symbol should also normally have a letter code beneath, indicating
the main packaging function codes as follows: S for electrostatic discharge shielding,
F for electrostatic field shielding, C for electrostatic conductive, and D for electrostatic
dissipative.
Unfortunately, in practice packaging manufacturers are rather variable in how they
mark ESD control packaging materials and products. Markings from older and superseded
standards are often still seen. The letter codes showing the packaging function are often
missing.

References

ASTM International. (2013) ASTM F1249 13 Standard Test Method for Water Vapor
Transmission Rate Through Plastic Film and Sheeting Using a Modulated Infrared Sensor.
West Conshohocken, PA, ASTM.
ASTM International. (2014) ASTM D257-14 Standard Test Methods for DC Resistance or
Conductance of Insulating Materials. West Conshohocken, PA, ASTM.
Bellmore, D. (2001). Anodized aluminium alloys – insulators or not? In: Proc EOS/ESD Symp.
EOS-23, 141–148. Rome, NY: EOS/ESD Association Inc.
Blythe, T. and Bloor, D. (2008). Electrical Properties of Polymers, 2e. Cambridge University
Press. ISBN: 978-0-521-55838-9.
Cross, J.A. (1987). Electrostatics Principles, Problems and Applications. Adam Hilger. ISBN:
0-85274-589-3.
Department of Defense (2002) MIL-STD-3010). Test Method Standard. Testing Procedures for
Packaging Materials. Test Method 4046 Electrostatic Properties, 33–41. Washington, D.C.:
DoD.
Department of Defense (2004) MIL-PRF-81705D:2004). Military Specification. Barrier
Materials, Flexible, Electrostatic Protective, Heat Sealable. Washington, D.C.: DoD.
References 277

Department of Defense (2009) MIL-PRF-81705E:2009). Military Specification. Barrier Materials,


Flexible, Electrostatic Protective, Heat Sealable. Washington, D.C.: DoD.
Drake N. (1996) Polymeric materials for electrostatic applications. RAPRA Report ISBN
1-85957-076-3
Electronic Industries Assoc. (1988) ANSI/EIA-541-1988. Packaging material standards for ESD
sensitive items. Washington D.C., USA, Electronic Industries Association.
EOS/ESD Association Inc. (2012a) ANSI/ESD S11.4-2012. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items - Static Control Bags. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2012b). ANSI/ESD STM11.31-201. ESD Association Standard Test
Method for Evaluating the Performance of Electrostatic Discharge Shielding Materials – Bags. 2
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014). ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015a) ANSI/ESD STM11.11-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items – Surface Resistance Measurement of
Static Dissipative Planar Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015b) ANSI/ESD STM11.12-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items–Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015c) ANSI/ESD STM11.13-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Two-Point Resistance
Measurement. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016). ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2018) ANSI/ESD S541-2018. Packaging Materials for ESD Sensitive
Items. Rome, NY, EOS/ESD Association Inc.
Havens, M.R. (1989). Understanding pink poly. In: Proc. EOS/ESD Symp. EOS-11, 95–101.
Rome, NY: EOS/ESD Association Inc.
International Electrotechnical Commission. (2014). IEC 61340-4-8:2014. Electrostatics - Part
4-8: Standard test methods for specific applications - Electrostatic discharge shielding – Bags.
Geneva, IEC.
International Electrotechnical Commission. (2015) IEC 61340-5-3:2015. Electrostatics - Part 5-3:
Protection of electronic devices from electrostatic phenomena - Properties and requirements
classification for packaging intended for electrostatic discharge sensitive devices. Geneva, IEC.
International Electrotechnical Commission. (2016a). IEC 61340-2-3:2016. Electrostatics. Part
2-3: Methods of test for determining the resistance and resistivity of solid materials used to avoid
electrostatic charging Geneva, IEC.
International Electrotechnical Commission (2016b) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
IPC, JEDEC. (2018) Handling, Packing, Shipping and Use of Moisture/Reflow Sensitive Surface
Mount Devices. J-STD-033D. ISBN# 978-1-61193-348-2
278 8 ESD Control Packaging

Smallwood, J.M. and Millar, S. (2010) Paper 3B4). Comparison of methods of evaluation of
charge dissipation from AHE soak boats. In: Proc. EOS/ESD Symp, 233–238. Rome, NY:
EOS/ESD Association Inc.
Smallwood J M, Robertson C J. (1998) Evaluation of Shielding Packaging for Prevention of
Electrostatic Damage to Sensitive Electronic Components. ERA Report 97-1079R, ERA
Technology Ltd., Cleeve Rd, Leatherhead, Surrey, KT22 7SA

Further Reading

EOS/ESD Association Inc. (1995). ESD ADV11.2-1995. Advisory for the Protection of
Electrostatic Discharge Susceptible Items - Triboelectric Charge Accumulation Testing. Rome,
NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2017) ANSI/ESD S8.1-2017. Draft Standard for the Protection of
Electrostatic Discharge Susceptible Items – Symbols – ESD Awareness Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2017). ESD ADV1.0-2017. ESD Association Advisory for Electrostatic
Discharge Terminology – Glossary. Rome, NY, EOS/ESD Association Inc.
Fowler S. (2000) ESD protective packaging. Available from: http://www.esdjournal.com/
techpapr/twenty1/intro.doc [Accessed 10th Nov. 2017]
Gale S F. (2006) Zero tolerance for ESD. Solid State Technology http://electroiq.com/blog/
2006/09/zero-tolerance-for-esd [Accessed 29th Nov. 2017]
Huntsman J. R., Yenni D. M., Mueller G. E. (1980) Fundamental requirements for static
protective containers. Nepcon West VI pp. 624–635
Huntsman, J.R. and Yenni, D.M. (1982). Test methods for static control products. In: Proc. of
EOS/ESD Symp. EOS-4, 94–109. Rome, NY: EOS/ESD Association Inc.
Huntsman, J.R. (1984). Triboelectric charge: its ESD ability and a measurement method for its
propensity on packaging materials. In: Proc. EOS/ESD Symp. EOS-6, 64–77. Rome, NY:
EOS/ESD Association Inc.
International Electrotechnical Commission. (2018) IEC TR 61340-5-2:2018. Electrostatics – Part
5-2: Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
Koyler, J.M. and Anderson, W.E. (1981). Selection of packaging materials for electrostatic
discharge (ESDS) items. In: Proc. EOS/ESD Symp. EOS-3, 75–84. Rome, NY: EOS/ESD
Association Inc.
Matisoff, B. (1997). Handbook of Electronics Manufacturing Engineering, 3e. Springer.
Swenson, D.E. and Lieske, N.P. (1987). Triboelectric charge-discharge damage susceptibility of
large scale IC’s. In: Proc. EOS/ESD Symp. EOS-9, 274–279. Rome, NY: EOS/ESD Association
Inc.
Swenson, D.E. and Gibson, R. (1992). Triboelectric testing of packaging materials: practical
considerations – what is important? What does it mean? In: Proc. EOS/ESD Symp. EOS-14,
209–217. Rome, NY: EOS/ESD Association Inc.
Texas Instruments. (2002) Electrostatic Discharge (ESD) Protective Semiconductor Packing
Materials and Configurations. Application Report SZZA027A - April 2002
Further Reading 279

Vermillion R. (2014) The Silent Killer: Suspect/Counterfeit Items and Packaging. In


Compliance. Available from: https://incompliancemag.com/article/the-silent-killer-
suspectcounterfeit-items-and-packaging [Accessed 29th Nov. 2017]
Vermillion R. (2013) Pin Holes & Staples Lead to Diminished Performance in Metallized Static
Shielding Bags. InCompliance. Available from: https://incompliancemag.com/article/pin-
holes-a-staples-lead-to-diminished-performance-in-metallized-static-shielding-bags
[Accessed 29th Nov. 2017]
Vermillion R. (2016) Have Suspect Counterfeit ESD Packaging & Materials Infiltrated the
Aerospace & Defense Supply Chain? Interference Technology. Available from:
https://interferencetechnology.com/suspect-counterfeit-esd-packaging-materials-
infiltrated-aerospace-defense-supply-chain [Accessed 29th Nov. 2017]
Vermillion R. J., Fromm L.. (n.d.) A Study of ESD Corrugated. Available from: http://talkpkg
.com/Papers-Presentations/Presentation/AHPStudy.pdf [Accessed 29th Nov. 2017]
281

How to Evaluate an ESD Control Program

9.1 Introduction
There are always several ways of looking at the “fitness for purpose” of an electrostatic
discharge (ESD) program. As this phrase implies, it is necessary first to understand the
various objectives that the ESD control program may be addressing. The primary purpose
is presumably to control ESD risks and reduce ESD damage to an acceptable level.
A second purpose is often to help satisfy customers that the organization takes sufficient
care in the manufacture or handling of their product commensurate with the market and
reliability requirements. In this way, the ESD control program may form a positive contri-
bution to the marketing of the product handled. Customers may audit the facility from time
to time as part of their supplier quality assurance procedures. Some customers may require
that their suppliers comply with an ESD control standard, and some of these may insist on
compliance with their own preferred ESD control standard.
Another important aspect of evaluation is of the cost effectiveness of the ESD control pro-
gram. It is usually desirable to attempt to get the maximum benefit from the least investment
in resources (time and money). Evaluating the likely benefit from the ESD control program
may be difficult but can be helpful in selecting the objectives of the program as well as
deciding the level of resources to invest in it.

9.2 Evaluation of ESD Risks


9.2.1 Sources of ESD Risk
There are two overall types of ESD risk that are normally controlled in an electrostatic dis-
charge protected area (EPA).
● Direct ESD to or from the device
● Electrostatic fields that could lead to ESD to or from the device or in some cases to damage
to electrostatic discharge–sensitive (ESDS) devices

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
282 9 How to Evaluate an ESD Control Program

Identification and evaluation of these is discussed further in Chapter 9. The usual sources
of ESD risk include
● Charged personnel
● Charged metal or conductive objects or materials
● Charged devices
Other ESD risks may also occur, such as
● Charged board ESD
● Charged cable ESD
● Charged modules or assemblies
Understanding the ESD risks is an important step toward determining effective ESD con-
trol measures.

9.2.2 Evaluation of ESD Susceptibility of Components and Assemblies


The process of evaluating ESD risks must start with an evaluation of the ESD susceptibility
of the ESDS component or assembly. In some cases, component human body model (HBM)
and charged device model (CDM) ESD withstand voltage data can be obtained from device
data sheets. Unfortunately, not all manufacturers’ data sheets give information on this. Fur-
thermore, the ESD susceptibility of printed circuit boards (PCBs), modules, and assemblies
is not usually tested. So, the evaluation must often proceed based on some assumptions.
When handling ESDS devices of unknown ESD withstand voltage, it is often easiest to
implement an ESD program with standard ESD control measures such as those given in
IEC 61340-5-1 or ANSI/ESD S20.20. These are designed to control the most common ESD
risks for devices of ESD withstand voltage as low as 100 V HBM and 200 V CDM.
If possible, any components that might have particularly low ESD withstand voltage or
unusual ESD susceptibility should be identified. It is particularly important, if possible, to
identify any components with ESD withstand voltage less than 500 V HBM or 250 V CDM,
or those sometimes called “Class 0” devices. These should be handled with particular care.
Components with ESD withstand voltage less than 100 V HBM or 200 V CDM may require
special or adapted ESD control requirements.
The ESD susceptibility of the items handled usually changes through a production pro-
cess. An ESDS component is typically assembled into a PCB assembly. This may then be
assembled into a higher-level assembly or module. At some stage, this assembly is built into
a final working product. The ESD risks and susceptibilities are likely to be different at each
build stage. In many cases, it is convenient to implement standard ESD control measures
for handling these items of unknown ESD susceptibility.
The final working system product may well be not considered to be ESD susceptible. So,
at some build state the product becomes treated as not susceptible to ESD. The build state
at which the product can be considered no longer susceptible to ESD is often debatable as
it is rarely determined by design. Where the product has been subjected to, and passed, an
ESD test as part of electromagnetic compatibility (EMC) immunity testing then this gives
an assurance that it is also relatively robust to ESD damage. Nevertheless, there could be a
remaining possibility of some level of susceptibility to ESD to untested parts. An example
might be susceptibility to cable discharge to connector pins.
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 283

A final product that is not susceptible to ESD damage often relies on a housing or covers
for protection of the ESDS internal parts. If the housing or covers are removed and the ESDS
parts exposed, it may be necessary to again consider the item to be susceptible to ESD.
An assembly containing ESDS devices should be considered also to be ESDS devices
unless there is good reason or evidence that it is adequately protected against ESD at that
build level. The following possible ESD sources should be considered:
● Direct ESD from charged personnel to the ESDS parts of the product, especially via hand-
held metal tools
● ESD occurring on contact with charged metal parts, chassis, or machinery
● The possibility that the internal parts could reach high voltage by induction or tribocharg-
ing leading to ESD on contact with another conductive item
● ESD occurring due to exposed terminations of a charged product contacting external con-
ductive items
● Connection of charged cables or wiring looms to the assembly
Where only specific ESD threats occur in a process step, it may be sufficient to use spe-
cific ESD control measures rather than general ESD control measures. General ESD control
measures may not adequately protect against specific ESD risks.

9.3 Evaluating Process Capability Based on HBM, MM,


and CDM Data
9.3.1 Process Capability Evaluation
In recent times it has become a goal to be able to evaluate a process and determine its capa-
bility of handling devices in terms of HBM, machine model (MM), or CDM ESD withstand
voltage levels. This has become more important as more devices are handled that have low
ESD withstand voltages (ESD TR17.0-01 2015; Lin et al. 2014; Halperin et al. 2008). This is
also linked to a trend toward reducing device on-chip ESD protection ESD withstand volt-
ages toward levels that are not excessively above those easily handled by processes (Industry
Council 2010, 2011; ESDA Association 2016b).

9.3.1.1 A Structured Approach to Process Evaluation


Evaluation of process risks in manual or automated processes is best done using a structured
approach in which the ESDS path is followed from beginning to end through the processes
(e.g. Gaertner 2007; Halperin et al. 2008; Jacob et al. 2012; ESD TR17.0-01 2015; ESD SP10.1
2016a). In every process step, the ESD risks are evaluated. This may include contact with
personnel, contact with potentially charged metal objects, and the risk of the ESDS devices
contacting a low-resistance conductor when in a charged state. Halperin et al. (2008) has
called this transitional analysis.

9.3.1.2 Evaluate the Critical ESDS Path Through the Process


From the point of view of ESD risk, the positions that unprotected ESDS devices can take in
the process define a critical path. The ESD risk occurs, and ESD threat must be evaluated,
only in a region around this critical path (ESD SP10.1 2016a). ESD sources outside this
region, a safe distance from the ESDS device, do not provide an ESD risk.
284 9 How to Evaluate an ESD Control Program

The responses to risks are usually specified in accordance with IEC 61340-5-1 and ESD
S20.20 or the guidance of ESD SP10.1. For example, ESD SP10.1 recommends that in
a region 15 cm around the ESDS critical path, all conductive machine parts should be
grounded and insulative parts made static safe. S20.20 and 61340-5-1 require that the
electrostatic field at the position of the ESDS device is <5 kVm−1 , and items having surface
voltage greater than 2 kV must be kept >30 cm from the ESDS device. The responses to
risks may also need special measures not prescribed by these documents. These standards
also require that
● Personnel handling unprotected ESDS devices must be grounded according to the stan-
dards.
● Metal objects contacting the ESDS devices must be grounded. If they cannot be grounded,
the voltage difference between the ESDS devices and metal object must be reduced to
below a hazard threshold level (±35 V in 61340-5-1:2016 and 20.20-2014).
● To reduce charged device ESD risks, low-resistance conductors that contact the ESDS
device may be replaced with intermediate- or high-resistance conductors.
● To address field-induced charged device ESD risk, nonessential insulators that might
become charged are removed from the vicinity of the ESDS devices. Essential insulators
are treated according to ESD risk.
● To address charged device ESD risks caused by contact charging or rubbing of ESDS
devices, process steps that cause tribocharging of the ESDS devices are minimized.
Notice that all the ESD risks listed involve contact between the ESDS devices and a person
or material. For most ESDS devices, where there is no contact, there is no ESD risk. The
contact that causes ESD may, however, occur during a following step rather than the step
at which charging occurs.
Electrostatic fields usually provide an ESD risk only where there is contact with the ESDS
device in the presence of the field. There are, however, a small number of ESDS device types
(e.g. reticles) that may be damaged by electrostatic field alone due to induced voltage differ-
ences causing breakdown of internal low-voltage insulating layers or air gaps. Smallwood
(2019) has shown that damage could be caused to some voltage-sensitive devices such as
metal-oxide-semiconductor field-effect transistors (MOSFETs) with high-impedance termi-
nals, by charge injection from fast-changing electrostatic fields.
When evaluating processes, it’s likely that the ESD specialist will need assistance from
process specialists to ensure that ESDS movements are correctly evaluated, risk points
identified, relevant measurements made, and suitable solutions found. Examples are given
by Gaertner (2007), Halperin et al. (2008), Lin et al. (2014), and ESD TR17.1-01-15. The
TR17.1-01 document draws on several earlier works.
TR17.1-01 following Jacob et al. identifies an evaluation procedure steps as follows:
● Assess potential risk
● Identify measurement points
● Make measurements
● Assess measurement results
● Define and implement corrective actions
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 285

The first step involves demonstration of the process under real conditions, with the goal of
identifying potential risks and measurement points. Making measurement and evaluating
the results presents its own set of challenges. Some of these are discussed by Lin et al. (2014).

9.3.1.3 Use of ESD Withstand Data in Process Evaluation


Unfortunately, it is not easy to correlate process capability with component HBM, MM, or
CDM ESD withstand data. Some general guidance can be offered. In practice, the analysis
and assumptions given in this section are approximate. Unfortunately, accurate methods of
evaluating ESD risk in real production situations are currently unavailable. In some cases,
the estimates given are likely to significantly overestimate the ESD risks to the device, but
in other cases the risk may be overestimated or underestimated.
In principle, the ESD risk could be understood by knowing the susceptibility of each type
of ESDS devices to each ESD source. The risk could then be controlled by maintaining the
parameters of ESD from each source below thresholds at which ESD damage are known to
occur. To this end, HBM, MM, and CDM ESD susceptibility tests have been developed to
allow reproducible measurement of the ESD susceptibility of components (see Chapter 3).
The component ESD susceptibility is quoted as an ESD withstand voltage, which is the
highest test voltage in the ESD source that did not result in damage to the component.
At first sight then, a strategy for risk evaluation and management might seem clear – if
the real-world voltages in ESD sources can be kept below the withstand voltage levels, no
ESD damage should be possible. Specification of ESD control measures would then be a
matter of defining controls that can maintain the source voltages below the risk threshold
defined by component ESD withstand voltages (Steinman 2010, 2012).
● The body voltage on personnel should be kept below the HBM ESD withstand voltage.
● The voltage on metal objects that contact ESDS devices should be kept below the MM
ESD withstand voltage.
● The voltage on devices should be kept below the CDM ESD withstand voltage.
While this simple view and strategy is useful and partially valid, it represents a significant
oversimplification. Nevertheless, this may be the best starting point that we have currently,
representing an estimated “worst-case” capability for many cases. Steinman (2010) sug-
gested that the limits used should be 50 percent, rather than 100 percent of the respective
ESD withstand voltages.
In the case of discharges between the ESDS devices and metal objects, the MM and CDM
withstand voltages probably give a conservative limit to work to (Steinman 2010; Tamminen
and Viheriäkoski 2007). As an example, if the ESDS device makes contact with a conductor
of resistance >10 kΩ rather than a metal object, voltages on the ESDS devices considerably
greater than the CDM ESD withstand voltage do not cause damage. In the case of human
body ESD, the HBM withstand voltage probably gives a conservative limit in many cases
but not necessarily all, depending on the particular type of sensitivity of the ESDS devices.
Component susceptibilities are not in fact normally to the ESD source voltage, but to
other parameters such as the peak discharge current or charge power or energy transferred
to the device in the discharge. These parameters are related to the source voltage by circuit
286 9 How to Evaluate an ESD Control Program

parameters such as inductance, resistance, and source capacitance. Real-world sources are
unlikely to have the same capacitance, resistance, and inductance as the HBM, MM, and
CDM models. This means that HBM, MM, and CDM discharges do not happen in the real
world – they happen only in the component susceptibility test equipment. Real-world ESDs
are simply discharges from the human body, a metal part, a device, or another source. Each
source has very different ESD waveform characteristics from the HBM, MM, and CDM mod-
els and would yield a different withstand voltage (TR17.0-01, Tamminen and Viheriäkoski
2007; Gaertner and Stadler 2012). So, ESDS devices might be more or less susceptible to
damage from real sources than it is to HBM, MM, or CDM. A simple action such as changing
the position of an ESDS devices can change a relevant parameter such as its capacitance as
well as the voltage and hence the ESD damage susceptibility. Tamminen and Viheriäkoski
(2007) found that the real charged device ESD risk in a process can be significantly less than
suggested by the CDM withstand voltage. As components approach a flat grounded surface,
the device voltage could in some cases reduce by 95 percent compared to the initial value
away from the surface. This drop varied with component package, being least for tall com-
ponents such as dual in-line (DIL) packages. Similar results were found with a component
approaching a floating PCB. They were able in some cases to get a better estimate of ESD
risk by taking account of the initial voltage and charge on the device, especially where the
geometry of the device and environment were taken into account. They commented that
in placing of a component, the resistive characteristics of solder paste may affect discharge
characteristics and ESD risk.
Even in the well-controlled world of the device ESD susceptibility tester, there can be
significant differences in parameters due to parasitic components in the tester. These can
lead to variation in withstand voltages measured for the same device with different testers.
Changing a device package type can also lead to changes in measured withstand voltage.
Some ESDS devices such as PCBs or assemblies are not normally tested for susceptibil-
ity using standard HBM, MM, or CDM ESD. Even if they were, the susceptibility would be
likely to change for every possible contact or ESD entry point on the ESDS devices. A com-
plex ESDS device such as a PCB thus could have a large number of possible contact points.
Susceptibility to ESD could vary with each new component added to an assembly and with
each ESDS position in an assembly process.
Some workers have suggested that it would make more sense to specify parameters such
as peak current (Bellmore 2004; Smallwood and Paasi 2003), charge, power, or energy
instead of the source voltage as withstand data parameters. This added complexity would
at least allow use of relevant parameters in evaluation of ESD risk in real-world situations.
Even this would not necessarily give the complete answer. It can be difficult to measure real
ESD parameters in a process, especially in automated equipment. Inserting measurement
equipment and sensors usually modifies the parameters that govern the ESD waveforms
that are being measured. In fast-moving automated equipment, it can be difficult to add
measurement equipment safely and without modification of the process.
A further complication is that it is not always easy to obtain ESD withstand data for all
components. In practice, this may not matter too much as it is likely that the process eval-
uation will focus on the most sensitive components. As long as these are identified and the
ESD control designed for handling them, then the less-sensitive components are probably
adequately protected.
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 287

9.3.2 Human Body ESD and Manual Handling Processes


ESD risk from charged personnel is notionally related to the HBM withstand voltage of
components. As a first approximation, it is assumed that the voltage developed on personnel
handling a device should be limited to less than the HBM withstand voltage of the most
sensitive device handled. So, in an ESD control program handling 100 V HBM devices, the
body voltage of personnel handling ESDS devices should not exceed 100 V. If 50 V HBM
devices are handled, the body voltage should be limited to less than 50 V.
If lower withstand voltage devices are handled, the allowed body voltage should be
reduced accordingly. For example, for handling 50 V HBM devices, the maximum body
voltage should be 50 V. The maximum wrist strap resistance to ground of 35 MΩ is deter-
mined to limit body voltage to <100 V. For 50 V maximum body voltage, this maximum
resistance should be reduced pro rata to 17 MΩ.
When working with low HBM ESD withstand voltage devices, it is wise to directly evalu-
ate body voltage produced during usual process and handling activities to ensure that body
voltage does not exceed required maximum levels. This is because resistance to ground is
only one factor related to body voltage. Charge generation is also highly important and is
neglected if only resistance to ground is measured. Charge generation depends the materi-
als and equipment used by the operator and can change dramatically if the type of material
or equipment, or surface conditions (including contamination of surfaces) or atmospheric
humidity, are changed. Evaluation of body voltage generation, for example with a change of
footwear or flooring type or condition, can be especially important for safe manual handling
of low HBM ESD withstand voltage devices.

9.3.3 ESD Risk Due to Isolated Conductors


The MM ESD test simulates a “two-pin” ESD risk that occurs if a charged external metal
object can discharge causing current flow through the ESDS device and through a second
pin to another conductor. (HBM and MM tests are “two-pin” ESD. Charged device ESD is
a “one pin” because the device, once charged, discharges through a single pin.) ESD risk
could occur, for example, if a charged external object discharges through the ESDS device
to ground or in charged board ESD if the ESD current flows across the PCB through the
ESDS device.
In principle, if the ESDS device is susceptible to the energy in the discharge, all the stored
energy in the MM test capacitor is delivered to the ESDS device. This energy, Emm , can be
calculated from the MM test capacitor capacitance C (200 pF) and test voltage V by
Emm = 0.5 C V 2
The energy Emm can then be assumed to be the maximum that the device will survive in
discharge from another metal object. The maximum voltage V s allowed on any other source
of capacitance Cs can then be calculated from
Emm = 0.5 Cs Vs2
As many real sources are much less than 200 pF, a considerably higher voltage than the
MM ESD withstand voltage can often be allowed. In practice, the capacitance of the source
is likely to be variable and can be difficult to measure.
288 9 How to Evaluate an ESD Control Program

This analysis depends on the ESDS device being susceptible to the energy in the discharge.
If it is susceptible to the voltage or charged transferred in the discharge, then a different
analysis would need to be applied (Paasi et al. 2003). For simplicity, it is tempting to specify
that the maximum voltage between any external metal ESD source and the ESDS device
should not exceed the MM ESD withstand voltage. This would, however, greatly overesti-
mate the ESD risk, and it could be difficult to maintain these low voltages. In practice, the
MM ESD susceptibility test is being discontinued by IC manufacturers.
As with charged device ESD, if the contacting material has significant resistance, this
absorbs some of the discharge energy and to some extent protects an energy-susceptible
ESDS device. If the ESDS device is voltage or charge susceptible, resistance in the contacting
material may not have any protective effect.

9.3.3.1 What Is an Isolated Conductor?


A primary ESD control measure is that all conductors that may make contact with ESDS
devices are grounded if at all possible. In unusual cases, it may be found that some con-
ductors that make contact with ESDS devices cannot easily be grounded and have to be left
electrically isolated. In this case, the ESD risk may need to be evaluated and controlled. The
ESDS device itself is not usually grounded.
Although the standards give some requirements for dealing with isolated (ungrounded)
conductors that touch the ESDS device (see Section 6.5.12.7), they do not actually give
any definition of an isolated conductor, or criteria by which they may be identified.
Future standards may address this issue, but in the meantime, if in practice the resistance
from a conductor to ground is less than 1 GΩ (109 Ω), it can usually be considered to
be grounded. If the resistance from the conductor to ground is greater than 100 GΩ
(1011 Ω), it should be considered isolated. Between 1 and 100 GΩ, some evaluation may be
required. In this case, if the voltage produced on the conductor during operation remains
sufficiently low during normal operation, then it can probably be considered adequately
grounded. The question is, what voltage is considered “sufficiently low”? The answer
may depend on the sensitivity of the ESDS device or may be specified by ESD control
standards.

9.3.3.2 How to Deal with Isolated Conductors


Figure 9.1 gives a simple flowchart for detecting and dealing with isolated conductors. Con-
ductors that cannot make contact with ESDS devices can be disregarded as they do not
provide an ESD risk.
The root of the ESD risk is that the voltage difference between the conductor and the
ESDS device may be sufficient to give damaging ESD. This can be a type of charged device
(one-pin) ESD, or more like ESD from a charged metal object (two-pin). If the conductor
is a low-resistance material, it could represent a significant ESD risk. The voltage between
the conductor and the ESDS device must be maintained at a sufficiently low level to control
this risk.
The current standards IEC 61340-5-1 and ANSI/ESD S20.20 give some requirements for
ungrounded conductors. IEC 61340-5-1 requires that the voltage difference between the
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 289

Identify conductor

Yes Is conductor Yes


Can the conductor No risk
touch ESDS? grounded?

No
No

Yes Ground the


No risk Can conductor be
grounded? conductor

No

Evaluate and
ameliorate risk

Document
evaluation and
ESD control
measures

Figure 9.1 Dealing with isolated (ungrounded) conductors.

ESDS device and the conductor must be less than ±35 V. Measuring this is easier said than
done. It requires two measurements.
● The voltage range and polarity on the ESDS devices must be measured under working
conditions.
● The voltage range and polarity on the conductor must be measured under working con-
ditions.
● The range in difference between the two voltages must be calculated.
Measurement of voltage on the ESDS device and conductor under working conditions
may be far from easy in practice, especially under working conditions in a fast-moving
automated process. If the ESDS device or the conductor is small, contact or noncontact
electrostatic voltmeters capable of measurement on small objects must be used (Steinman
2010, 2012).
One way to bring the voltage difference between a conductor and an ESDS device to a
low level is to place them in the ion stream from an effective ionizer. Given sufficient time
and low charge generation processes, each will come to the same voltage - the offset voltage
of the ionizer. This may not work well, however, in a fast-moving process where sufficient
time may not be available or if charge generation processes are greater than the rate of
neutralization that the ionizer can achieve in the installation.
290 9 How to Evaluate an ESD Control Program

If the ESD concern is restricted to charged device ESD (i.e. an otherwise unconnected
ESDS device makes contact with the conductor) and if the conductor has surface resistance
>10 kΩ, the ESD risk is likely to be negligible (see Section 9.3.4).

9.3.4 Charged Device ESD Risks


Charged device ESD risks are often the main risks occurring in automated processes and in
any situation where a charged device can contact a highly conductive material (e.g. metal).
Charged device ESD is a “one-pin” ESD risk that can occur by the device pin contacting
a low-resistance conductor. Contacting a highly resistive conductor gives a reduced risk of
damage because the ESD current is reduced by the resistance of the material at the point
of contact. For a simple analysis, an assumption can be made that if the voltage difference
between the device and the conductor is less than the CDM withstand voltage, ESD damage
is unlikely to occur.
In this case, a visual inspection, observing the path taken by ESDS devices through
the system, can be undertaken to detect the locations where contact with metal or other
low-resistance conductors may occur. At these contact points, charged device ESD risk is
apparent if the voltage difference between the ESDS devices and the contacted metal could
exceed the ESDS devices CDM withstand voltage.
Charged device ESD could also be a risk in a manual process if the ESDS device contacts
a low-resistance material. This is the main reason why intimate ESD packaging and bench
mats should usually be made from materials having surface resistance greater than about
10 kΩ. When ESDS devices can contact materials having surface resistance <10 kΩ, charged
device ESD risks should be evaluated.
It is often not easy to determine the voltage difference between the device and contacting
material, particularly in an automated process during operating conditions. If one of the
conductors is grounded, then at least the voltage of this is known to be zero. If the conductor
is not grounded, grounding the conductor does not prevent charged device ESD as it is
the charged device itself that is the ESD source. Grounding the contacting conductor does
simplify the analysis of risk and ensures that the conductor cannot become charged and an
additional source of ESD risk (see Section 9.3.3).
In practice, it is the peak current in the discharge that normally causes charged device
damage, and the risk of this can be reduced by increasing the resistance of the material con-
tacting the ESDS devices. So, charged device ESD risk can be reduced either by maintaining
the voltage difference less than the CDM withstand value or by increasing the contacting
material resistance, or preferably both. The charged device risk is greatly reduced by con-
tact with a sufficiently resistive material. The resistance of the contacting material reduces
the ESD peak current and absorbs some of the energy of the discharge current. ESD S20.20
and IEC 61340-5-1 recommend a minimum contacting material surface point-to-point resis-
tance of 10 kΩ for avoiding charged device ESD risk.
Adding resistance into the ground path of a low resistance conductor, for example by
adding a discrete resistor, does not reduce charged device ESD risk as the discharge is into
local capacitance of the conductor (Wallash 2007). Resistance in the ground path does not
act to limit the ESD current into this capacitance (Figure 9.2) The device capacitance can
be discharged with the peak current limited only by the impedance of the discharge (e.g.
spark) and the conductive material.
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 291

Device
Capacitance
between
device and
ESD
surface

Minimal
Metal surface surface
resistance

Resistance in
ground path

(a)

Device
Capacitance
between device
ESD and surface

Surface
Resistive resistance
surface

(b)

Figure 9.2 Charged device ESD to metal (above) or resistive (below) surface. Adding resistance to
a ground path from a contacting conductor does not limit the charged device ESD current as using
a resistive surface does.

Evaluation of the voltage likely to be produced on a device in a production process is


not easy (see Section 11.9.4). Voltages can be produced on the device by triboelectrification
during handling or by induction due to nearby electrostatic fields. If it is to be measured,
it must be done in a way that minimizes the effect of the measurement equipment on the
device voltage. The measurement method must also distinguish the voltage on the device
from any other electrostatic fields and voltages in the region. This requires use of a low
capacitance very high input impedance contacting or noncontacting electrostatic voltmeter
capable of accurate measurement of voltage on small items.
If a measurement of the device voltage is successfully made, it remains a challenge to
ensure that this represents the maximum likely during the full range of possible process
conditions including low atmospheric relative humidity.
If contact with an isolated conductor occurs, then the voltage arising on the conductor
must be evaluated in the same way as for the device. The range of possible voltage difference
between the device and the conductor must then be calculated.
292 9 How to Evaluate an ESD Control Program

If reduction of the voltage produced on the device by triboelectrification is needed, then


the charging effect due to contact with materials leading to that process point must be evalu-
ated. Careful change of the materials that contact the device may be one way of reducing tri-
boelectrification. Against this, triboelectrification is a variable phenomenon, and attempts
to use it in ESD control are often unreliable.
Evaluating and reducing voltages induced on the device may be simpler. The electrostatic
fields occurring due to voltages on materials close to the device must be evaluated. The
voltages induced on the device by these fields cannot exceed the original voltage of the
field source. Therefore, if the voltages produced are less than the device CDM withstand
voltage, they can be considered insignificant. As with measurement of the device voltage, it
remains a challenge to ensure that this represents the maximum likely during the full range
of possible process conditions including low atmospheric relative humidity.
If the voltage is greater than the CDM ESD withstand voltage, then its effect can be
reduced or eliminated by measures such as

● For a charged insulator, it may be possible to replace the insulator with a grounded con-
ductor.
● Increasing the separation distance between the voltage source and the device path.
● Adding an electrostatic field shield between the voltage source and the device.

9.3.5 Damage to Voltage-Sensitive Structures Such as a Capacitor or a


MOSFET Gate
Some components such as capacitors and MOSFET gates can be damaged by injection of
sufficient charge to raise the voltage to a breakdown voltage value. (see Section 3.4.3.) ESD
risk to these components can be evaluated in terms of the charge that can be transferred to
the ESDS devices by triboelectrification or in an ESD event (Paasi et al. 2003, 2006; Small-
wood and Paasi 2003; International Rectifier AN-986 2004). This can happen by various
means including the following:

● Charging by contact with another material


● Contact with a charged person or conductor
● Field-induced voltage differences
● Charging by ion influx from an unbalanced ionizer output

The parts of an ESDS device most at risk from this are likely to be high-impedance lines
that have low capacitance and low-voltage breakdown. This result in sensitivity to damage
from electrostatic fields even where no contact is made with the ESDS device.
Paasi et al. (2006) showed that for a MOSFET, the ESD risk threshold was adequately
described by a charge threshold, whereas the HBM voltage value was inadequate as an indi-
cator of ESD risk. The damage risk occurred when the induced charge could elevate the volt-
age to the gate breakdown value. When the MOSFET was mounted on a PCB, the gate break-
down voltage could replace the charge threshold as a risk indicator. The real ESD risk is of
course modified by the circumstances and the probability of making contact to the MOSFET
in a way that could cause ESD damage. ESD risk did not arise directly from the electrostatic
field strength but rather from charge induced on the MOSFET or PCB. The field threshold
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 293

for damage to a device on a PCB was lower than for an unmounted component because the
PCB had greater area exposed to the field and greater induced charge for the same field.

9.3.6 Evaluating ESD Risk from Electrostatic Fields


ESD risk is in many cases a result of the electrostatic field from charged materials such as
essential insulators but can also be from equipment such as old types of cathode ray tube
(CRT) display equipment screen or high-voltage cables. Nonessential field sources should
be dealt with by removing them from the vicinity of the ESDS device.
The risk often arises because voltage differences arise between the ESDS device and
another conductor in an electrostatic field. If the ESDS device comes close enough to
or touches another conductor at a different voltage, ESD will occur between them. The
conductor may or may not be grounded – ESD will occur in either case.
The classification of insulating items as essential or nonessential is a matter of opinion.
The same item may be considered essential in one process and nonessential in another. For
example, it may be easy to remove papers from one workstation process, but in another it
may be difficult to proceed without them if they must be updated or signed on completion
of process steps.
In most cases, the risk from insulators can be assessed by a simple process of evaluation
such as the procedure shown in Figure 9.3.
Most ESDS devices are not inherently sensitive to damage directly from electrostatic
fields. The ESD risk is usually significant only if there are significant electrostatic fields
and there is the possibility of contact between the ESDS device and another conductor
within the field. If there is no contact with the ESDS device within the field, it may not be
necessary to control the field. If the insulator is sufficiently far from the ESDS device, the
electrostatic field arising at the position of the ESDS device may be negligible. (Take care
that the ESDS device and risk of contact are not likely to be moved into a position where
the field may be significant.) If the insulator is not likely to be handled or moved or become
charged, the risk of a field arising may be negligible.
Electrostatic charging of the item should preferably be measured under worst-case
conditions, which usually means under low humidity (<30% rh) ambient atmosphere. In
practice, measurements may have to be done under ambient atmospheric conditions, as
a humidity-controlled facility is usually not available. Nevertheless, an initial evaluation
done under ambient condition at higher humidity may give a useful first evaluation and
should then be repeated when the weather conditions give lower humidity.
The question then arises – what level of charging can be considered negligible? Unfortu-
nately, this may not be easy to answer and depends on the withstand voltages of the ESDS
device being handled and other factors. For example, if it is charged device damage to the
ESDS device that is of concern, if the voltages produced on the device are lower than the
CDM withstand of the ESDS device, they can be considered negligible. The voltage induced
on a conductor can never exceed the voltage of the electrostatic field source.
In practice, higher voltages than the ESD withstand voltage may also be negligible, but
evaluation of this is more difficult. Charged device ESD damage is caused by peak ESD
current and not the source voltage. For ESD between an ESDS device and a conductor, the
peak ESD current is determined by the total impedance of the discharge circuit through
294 9 How to Evaluate an ESD Control Program

Measure electrostatic
charging under worst-
case conditions

Is charging
negligible under Yes
worst-case
conditions?

No

Could ESDS make


Yes
contact with another
conductor within the
field?

No

Is the ESDS Yes Plan and


device to implement
electrostatic appropriate ESD
field? control measure
Document
No
evaluation and
ESD control
measures

Figure 9.3 A simple electrostatic field risk evaluation.

any spark and the impedance of the contacting material at the point of contact as well as
voltage difference between the ESDS device and conductor. The impedance of this circuit
typically changes with material characteristics and ESDS device position and orientation.
Standards may also give requirements that can be used to evaluate fields from charged
insulators. For example, the IEC 61340-5-1:2016 standard gives requirements that the elec-
trostatic field at the position of the ESDS device must be <5 kVm−1 . In IEC 61340-5-1:2016
and ESD S20.20-2014, insulators charged to >125 V must be kept at least 2.5 cm from the
ESDS device, and if charged to >2 kV must be kept >30 cm from the ESDS device. If these
conditions are fulfilled, the electrostatic fields and voltages can be considered negligible for
the purposes of these standards. Nevertheless, these limits may not be sufficiently low or
well specified for safe handling of some types of ESDS devices (Stadler et al. 2018). This is
a subject of research and discussion among experts at the time of writing.

9.3.6.1 Risk to ESDS Devices Due to Electrostatic Fields


Swenson investigated the susceptibility of a voltage-sensitive component to damage from
an electrostatic field and helped establish field limits for use in ESD control (Swenson
9.3 Evaluating Process Capability Based on HBM, MM, and CDM Data 295

2012). He used a 14-pin DIL package positive metal-oxide semiconductor (PMOS) device
(Siliconix SM110CJ with 200 V HBM and 125 V CDM ESD withstand voltage) and a
discrete MOSFET (Motorola 3 N157 with 200 V HBM and 150 V CDM ESD withstand
voltage) as his test devices. He mounted these in an electrostatic field created by charged
plates of various sizes charged by momentary contact with a preset high voltage. Devices
were placed on a glass plate or a grounded metal plate. A grounded metal-wire probe
was touched to a device pin while within the field. The discharge current waveform
was recorded. Swenson then produced charts showing the damage rates experienced by
devices at different applied voltage (electrostatic field) and separation distance from the
ESDS devices.
With the SM110C device, Swenson found that the field level required for damage was
greater when the device was placed on a ground plane compared to when it was on a glass
plate. With the device on the glass plate, it is likely that the field coupled to the device is
intensified. With the ground plane present, some of the field is coupled to the plane, the field
at the device is reduced, and voltage required for damage is increased. Differences were also
observed with the 3N157 device. Swenson noted that damage actually occurred when the
devices were touched by the ground lead, creating a field-induced charged device ESD.
Swenson went on to perform experiments with a uniformly charged insulating plastic
plate. He found that with the SM110C device on a ground plane, direct contact between
the ESDS device and the charged plastic plate did not damage the device until the charged
insulator voltage was 18 kV. When the device pin was grounded in the presence of the field,
damage occurred at 10 kV. (In practice, the surface voltage and surrounding field will be
altered by the presence of conductors such as the ground plane and ESDS device. The insu-
lator voltage was measured using a field meter at 2.5 cm distance. (The measured insulator
surface voltage would not have been the same as at the time of damage occurring.)
In experiments with the 3N157, again direct contact between the ESDS device and the
charged plastic plate did not damage the device until the charged insulator voltage was
18 kV. Touching a device pin with a grounded probe resulted in damage with the insulator
charged to only 2 kV.
Swenson’s experiments showed that an ESDS device, if grounded via a metal contact in
the presence of an electrostatic field, experiences a potentially damaging ESD. The device
becomes charged in the process and, if the field is removed, can subsequently experience
another potentially damaging ESD. The risk to the ESDS device depends on the size of the
electrostatic field source and its distance from the ESDS device, as well as the charged sur-
face voltage. Low voltages close to the device can be of concern where very sensitive devices
are handled. It is the act of grounding, rather than the field, that caused the damaging ESD.
Direct contact with an insulator did not result in damage up to high charge levels. Swen-
son’s results show that proximity or contact between a charged insulator and ESDS device is
not a significant ESD risk in most circumstances, providing the ESDS device does not make
contact with a conductor in the presence of the field. Small charged objects provide lower
risk than large objects – Swenson found damage occurred to both devices when they were
grounded in the presence of plates >4 cm × 4 cm size charged to 500–1000 V. The presence
of a ground plane nearby increased the field required for damage.
Swenson’s results supported the strategy of removing nonessential insulators from the
vicinity of the ESDS device to reduce ESD risks. They also confirm that field strength is an
296 9 How to Evaluate an ESD Control Program

indicator of these risks where contact between the ESDS device and a grounded conductor
occurs in the presence of the field.
It is arguably the CDM ESD withstand voltage of the devices that is of most relevance to
the ESD risk in Swenson’s experiments. The voltages required for damage in the experiment
(>500 V) were considerably greater than the CDM ESD withstand voltage of the devices (125
and 150 V).

9.3.7 Troubleshooting
Where failures have been experienced and ESD damage is suspected, the first action should
be to check that normal ESD control processes are in place and are operating correctly. Rec-
tify any deviations from the ESD control program and evaluate whether they could cause
the damage.
Check that personnel handle the ESDS device using the required procedures and personal
grounding equipment. Human nature sometimes causes personnel to invent more quick
and convenient procedures that may not be so effective for ESD control, especially when
under pressure to work fast.
If possible, obtain ESD withstand data for the damaged components to help identify pos-
sible sensitivity to likely ESD sources. If possible, failure analyze the damaged components
to establish, or rule out, ESD as a possible cause. Semiconductor manufacturers will some-
times failure analyze components on behalf of their customers.
If atmospheric humidity data is available, check whether failure rates correlate with low
atmospheric humidity. If this correlation exists, it can be a strong indicator that ESD is a
likely cause of damage and ESD control is insufficient. Low humidity can occur in areas
where strong heat sources are present, such as near ovens or near the cooling fan exhausts
of equipment or computers. Local low humidity conditions can occur near heat sources
even in air-conditioned rooms where humidity is controlled.
If there is no obvious deviation from the ESD program that could be responsible for the
failures, more detailed evaluation of the process may be required, including following the
path of the ESDS device through the process. Try to narrow down the problem to part of the
process. For example, if failed products pass tests at one stage and then fail at a later stage,
it is likely that the problem has occurred in the process steps between the two test points.
Critically examine the process for stages at which the ESDS device contacts other con-
ductors. In a manual process, this should include handling by personnel. These contact
points are points at which ESD is likely to occur. Processes in which ESDS devices may
contact metal or highly conductive items should be examined carefully for ESD risks. This
is discussed further in Section 9.4.
One area that is often overlooked is where an ESDS device is placed on a test jig that
makes contact prior to test. If the ESDS device becomes charged before contact or contact is
made in the presence of an electrostatic field, ESD can occur. Test jigs often contain essential
insulating parts that could charge and give rise to an electrostatic field during use. I have
also often found jig covers made from high-charging materials – these could induce voltages
on ESDS devices immediately prior to contact with test jig contacts. The possibility that the
test jig contacts and wiring might be charged or energized should also be checked.
Another often overlooked area is connection of cables to an ESDS device. ESD can occur
if the cable is charged, if the ESDS device is charged, or if the connection is made in the
9.4 Evaluating ESD Protection Needs 297

presence of an electrostatic field. Cables can be highly charged as a result of being supplied
packaged within a polythene bag or taken off a reel. The presence of charged packaging can
also induce high voltage on cables, modules, or assemblies nearby.
It is often assumed that a potted module is immune to ESD. A module can have its exterior
highly charged by contact with, or storage within, insulating packaging. The surface charge
on the module can then induce high voltage on internal conductors that can lead to ESD
when connected to cables or other conductors.
One area that can be problematic is handling or testing ESDS devices where high voltages
are present, e.g. test process. Safety considerations often dictate that personal grounding or
other standard ESD control measures cannot be used. Personal protective equipment such
as high-voltage protection gloves or footwear may be demanded.

9.4 Evaluating ESD Protection Needs


9.4.1 Standard ESD Control Precautions Do Not Necessarily Address all ESD
Risks
Standard ESD control precautions applying ESD control equipment by rote address most
well-known ESD risks, but they do not necessarily address all ESD risks. Every step of
processes where ESDS devices are handled should be scrutinized for possible ESD risks
(Gaertner 2007). This includes any automated processes. Any possible risks found should
be evaluated. Risks in general occur where the ESDS device is brought into contact with
other conductors. Wherever this happens, the following questions should be asked:

● Is the conductor grounded? If not, what would be the ESD risk if it became charged?
● Could the ESDS device become charged before touching the conductor? Could the voltage
difference between the conductor and the ESDS device result in ESD risk?
● What is the resistance of the conductor? Is it high enough to prevent ESD risk?

ESDS devices or isolated conductors can attain high voltage by two means. First, they
can become charged by contact with other materials (triboelectrification). Second, they can
attain high voltage under the influence of an electrostatic field.
Insulating materials are often the main source of electrostatic fields when they become
charged by contact with other materials. Nonessential insulators are therefore removed
from the vicinity where unprotected ESDS devices are handled, or even from the EPA. Many
insulators are, however, an essential part of the product or process. These essential insula-
tors cannot be removed from the vicinity of the unprotected ESDS devices. The ESD risk
associated with the presence of the insulator must be evaluated.
If an ESDS devices is exposed to an electrostatic field but does not contact another con-
ductor, there is unlikely to be a risk of ESD damage (Gaertner 2007). Few ESDS devices
are susceptible to damage from electrostatic fields alone. An exception may be where ESDS
devices contain voltage-sensitive devices such as MOSFETs with high-impedance circuit
nodes (Smallwood 2019). So, the presence of an electrostatic field from a charged insula-
tor does not necessarily amount to an ESD risk. An ESDS device can enter a field and exit
without damage. If, however, contact is made with another conductor while within the
298 9 How to Evaluate an ESD Control Program

electrostatic field, ESD will occur. An ESDS device that enters the field uncharged will exit
the field uncharged unless charge has been lost or gained by ESD, by conduction through
contact with other materials or via unbalanced ionization.
Triboelectric charging of ESDS devices can happen whenever the ESDS device contacts
other materials. The amount of charge generated depend on the material that contacts the
ESDS device and many other factors and conditions (e.g. humidity and rubbing action). It is
often thought that triboelectrification occurs with contact only by insulating materials. This
is not correct as any material contacts give triboelectrification, although if a conductor is
grounded, the separated charge escapes from the conductor without any voltage developing.
Insulating parts of an ESDS device such as PCB substrate, coatings, or potting compound
can be charged by contact with conductors, and the charge developed on the insulator is not
removed by contact with the conductor. In some cases, a greater level of charging has been
found for contact with static dissipative materials than with insulators or lower-resistance
conductors (Viheriäkoski et al. 2012). The voltage developed on the charged item depends
on the amount of charge and its capacitance. Capacitance is in turn dependent on the prox-
imity of other materials and conductors.
The risk of damage occurring when an ESDS device at high voltage contacts a conductor
depends on the part of the ESDS device that is subjected to ESD, the magnitude (energy,
current, or other parameter related to possible damage) of the ESD, and the sensitivity of
the ESDS device to the form of ESD that occurs. Unfortunately, it is difficult to predict the
susceptibility of ESDS devices to a real ESD event even if HBM, MM, or CDM ESD withstand
voltage data is available.
Common ways in which an ESDS devices could attain high voltage include
● Triboelectrification during handling by operators wearing gloves
● Triboelectrification during transport by suction grip
● Triboelectrification during contact with a support wheel or conveyor
● Induced voltages due to proximity to charged insulators, machine parts, computer
screens, clothing, gloves, or another electrostatic field source
● Triboelectrification during removal of a label, masking material, or adhesive item
● Triboelectrification during brushing to remove debris
● Induced voltages due to assembly into a charged insulative housing
● Triboelectrification during rubbing or removal adhesive tape or packaging from a plastic
housing or potted component
Common ways in which conductors often contact ESDS devices include
● Touching a metal tool or machine part
● Placing of a component on a PCB track
● A PCB track coming against an end stop in a machine
● A PCB track contacting “bed of nails” pogo pin or support
● Contact of a cable with a terminal on the ESDS device
If an ESDS device is found to attain high voltage and contact another conductor while in
this state, an ESD risk is clear. Similarly, a risk is clear if an isolated (ungrounded) conductor
is found to attain high voltage and contacts an ESDS device while in this state. To avoid these
risks, a means must be found of reducing voltage difference between the conductor and the
9.4 Evaluating ESD Protection Needs 299

ESDS device before contact. A possible alternative is to replace the conductor with a static
dissipative material to reduce the ESD current occurring on contact.
It is usual, if possible, to ground any conductor that contacts the ESDS device. This elim-
inates the risk that the conductor will attain high voltage due to triboelectric or induction
charging. It does not, however, eliminate the possibility that the ESDS device may attain
high voltage due to triboelectrification or induction. If this occurs, charged device ESD can
occur on contact.
Understanding the process and evaluating the associated risks can be particularly difficult
for automated processes. A working automated process is usually fast and inaccessible. See-
ing the steps can be difficult and measuring ESD-related parameters during live operations
is often impossible. Some common practices used in automated equipment are discussed
in Chapter 5.

9.4.2 Evaluating Return on Investment for ESD Protection Measures


It can be useful to estimate return on investment (ROI) for individual ESD control measures
or the whole ESD control program (see Section 9.5.4). This can be done to help decide which
particular ESD control measures should be prioritized or may be omitted.

9.4.3 What Is the Maximum Acceptable Resistance to Ground?


In Section 4.5 we looked at a simple model of electrostatic charge build-up and definitions
of insulator. This model also allows us to evaluate the upper limit of resistance to ground
Rg in many circumstances. This can be looked at in two ways.
● In a quasicontinuous charging process, for a given charging current, what is the allowable
voltage build-up?
● How long would it take a stored charge to dissipate?
In practice, we do not have to do this type of evaluation if the circumstance is covered by
ESD control standards. There may, however, be circumstances not covered by standards or
in which the requirements of standards need to be questioned for some practical reason.

9.4.3.1 Charging Current in a Quasicontinuous Process


In a quasicontinuous charging process, the charging current I can sometimes be measured
using a picoammeter or electrometer. This might be useful, for example, in an automated
process where grounding of some moving machine parts may be difficult.
Typically, measurement of the charging current would be attempted in worst-case (usu-
ally dry atmospheric humidity) conditions where possible. If a maximum voltage build-up
V max can be specified, then Ohm’s Law can be used to specify a maximum resistance to
ground Rgmax .
Vmax = IRgmax
If, for example, a maximum charging current on a part is found to be 10 nA (10−9 A) and
the maximum voltage acceptable is specified as 10 V, the maximum resistance to ground is
10/10−8 = 109 Ω (1 GΩ).
300 9 How to Evaluate an ESD Control Program

9.4.3.2 Maximum Decay Time


In some cases, it may be that a conductor with measurable capacitance C has been identified
as a possible ESD risk and must be grounded to prevent charging. Often, as a general guide,
it can be assumed that if there are no continuous charging processes present, a charge decay
time 𝜏 less than a second or so would be adequate to ensure this. In a fast-moving automated
process, a faster decay time might be desirable. The maximum resistance to ground Rgmax
can then be specified as

Rgmax C = 𝜏

An example might be the resistance through the handle of a handheld tool. If the capac-
itance from bit to hand through the handle is measured to be 20 pF and a decay time of
one second is deemed to be adequate, the maximum acceptable resistance through the han-
dle is 1/10−11 = 1011 Ω (100 GΩ).
In practice, it might be convenient to set a lower resistance that is easier and more con-
venient to measure (see Section 9.4.5).

9.4.4 Should There Be a Minimum Resistance to Ground?


There are usually possible two reasons that a minimum resistance to ground might be
specified.

● Protection of personnel against electric shocks in a fault condition with high voltages
present (see Section 4.7.6)
● Incorporation of inherent resistance in a material that contacts an ESDS device, for reduc-
tion of charged device ESD risk (see Section 4.7.5)

If neither of these is necessary, minimum resistance to ground need not be specified.

9.4.5 ESD from Charged Tools


A common risk in manual assembly is that of ESD due to charged tools. Many ordinary
hand tools have a metal part (e.g. cutter, plier, or screwdriver bit) that is mounted in an
insulating handle. This is an isolated conductor that could attain high voltage by induc-
tion or triboelectrification. If this metal part contacts an unprotected ESDS device during a
process, a potentially damaging ESD could occur.
To avoid this risk, the insulating tool handle can be replaced by a conducting material,
usually having resistance in the range 1 MΩ–100 GΩ. The metal bit is then connected to the
grounded operator’s hand via the conducting handle material.
In some cases, where charged device ESD risk is a concern, the tool bit may be replaced
with a high-resistance material to reduce charged device peak ESD current levels below the
likely damage threshold.
Grounding of the tool is achieved unless the operator is wearing insulating gloves. If they
do so, the tool is isolated from the grounded hand by the glove, and contact with the glove
is likely to charge the tool to some voltage. So, where ESD control tools are used, any gloves
used must also be conducting through the grip area (see Section 9.4.6).
9.4 Evaluating ESD Protection Needs 301

In practice, it might be convenient to set a lower resistance that is easier and more conve-
nient to measure. For example, setting an upper limit of 20 MΩ might make the tool more
easily verifiable using a lower specification resistance meter or even a wrist strap checker.

9.4.6 Use of Gloves or Finger Cots


Some processes require the operator to wear gloves to protect their hands or the product
being handled. Contact with the glove material handled causes triboelectrification of the
items handled.
If the glove is a conducting material, the charge retained on the conductors of the ESDS
devices can be reduced via grounding via the conducting glove and operator’s body. The
resistance through the glove is usually chosen in the range 1 MΩ–1 GΩ, although up to
100 GΩ may be acceptable in some cases. Any charge retained on insulating parts of the
ESDS device is not removed, but it is usually the voltage on the conductors of the ESDS
device that is the critical issue in charged device ESD control.

9.4.7 Charged Cable ESD


Long cables are often supplied on reels, and short cables are often supplied in insulating
plastic packaging. The insulating sheaths of the cables and the core conductors can easily
become charged by triboelectrification between the core and sheath and between the sheath
and packaging or other sheath areas. An additional voltage is often induced on a core due
to proximity with the charged insulating materials. If a cable is removed from packaging,
its core voltage can rise to several kilovolts (see Section 2.6.6).
ESD can occur when the cable is connected to an ESDS devices. Damage could occur
if the circuitry connected to the ESDS device connector has insufficient immunity to ESD
damage.

9.4.8 Charged Board ESD


PCBs usually have some accumulated voltage during manual or automated production pro-
cesses due to triboelectrification or induced by nearby electrostatic fields. ESD occurs if a
conductor of the PCB contacts another conductor during handling. This can happen by
contact with a machine part such as end stop, metal PCB support, or “bed of nails” pogo
pin contact.
It can be difficult to prevent this type of ESD, although charge on a PCB can be neu-
tralized using an ionizer. The ESD peak current, and potentially damaging effect, can be
limited by ensuring that materials that contact the ESDS device have high resistance (e.g.
100 kΩ–100 GΩ) rather than being low resistance or metallic.

9.4.9 Charged Module or Assembly ESD


Some types of system components have an ESDS PCB contained within a potted block or
plastic housing. The component electronics may have only a flying lead or connector for
connection to the external wiring looms or system components. In other cases, an ESDS
302 9 How to Evaluate an ESD Control Program

PCB may be contained within a subassembly isolated from the main assembly. In each of
these cases, the ESDS PCB can achieve high voltage by triboelectrification or induction due
to nearby electrostatic fields. ESD can occur when the ESDS PCB is connected to another
system component. Damage can occur if the connection and ESD occurs to a susceptible
part of the circuit.
It can often be difficult to prevent this type of problem unless the system components are
designed to avoid it. The risk of damage can often be reduced by connecting first to a 0 V,
ground, or power connection before I/O connections are made.
When packaging for transport modules that have flying leads, the ESD protective packag-
ing should be specified to prevent accidental contact between the flying leads and external
items.

9.5 Evaluation of Cost Effectiveness of the ESD Control


Program
9.5.1 The Cost of an Inadequate ESD Control Program
In an interview with Halperin, Brandt (2003) reported that independent consultants and
corporate studies have found that ESD losses can be as high as 10% of annual revenues
with an estimated average negative impact of 6.5% of revenues. Based on 1997–2001 num-
bers, this represented estimated losses of $84 billion per annum to international electronics
industry. While this is difficult to verify it represents significant losses, these are not just
due to material costs, which are often the smallest part of the ESD impact, but include costs
such as rework, warranty, field service, and customer service. Customers may also experi-
ence costs such as lost productivity due to failures during operation. The total cost of even
a small ESDS device loss may be significant.
Ever since the recognition of the need for ESD control, engineers suspecting ESD damage
have asked two questions (Halperin 1986).
● How do I know static-related problems are affecting our operations, and to what degree?
● How do I define static impact in a way that attracts management attention and support?
Most company managers have to identify priorities for their resources. They respond by
releasing resources and action to clearly defined problem areas that have specific causes and
measureable value impact, with a probable return on investment. Lack of quantified infor-
mation on cost or impact, and value of investment, can lead to inadequate investment in the
issue. Halperin went on to discuss ways of estimating the cost of ESD to the organization
and recommend how to present this to management to encourage their support.
It might seem at first sight that the cost of the lack of adequate ESD control should be
easily determined, as the cost of product failed due to ESD within the organization and at
the customer’s site. Unfortunately, this is not so simple for various reasons.
One problem is that few organizations evaluate failures to the level at which ESD failures
could be identified. This failure analysis can be expensive and time-consuming, often taking
several day’s work. The outcome may be inconclusive. ESD failures often are difficult to
distinguish from electrical overstress (EOS) failures, and EOS failures could sometimes be
caused by earlier ESD.
9.5 Evaluation of Cost Effectiveness of the ESD Control Program 303

The costs of an inadequate ESD control program can include (Smallwood et al. 2014)
● Repair and replacement of ESDS devices product failures
● Dealing with product unreliability or drift in characteristics
● Failure analysis
● Purchase of materials that have incorrect properties for ESD control
● Production delays
● Need to overstock commonly failing components
● Expenditure on unnecessary or ineffective control materials or equipment
● Dealing with customers failures occurring in the field
● Dealing with disputes with customers about their perception of adequacy of ESD control
in the facility
● Effects on product and company reputation, and sales
One way to estimate component losses is by throughput evaluation (Halperin 1986). This
includes the following steps:
1. Identify the ESDS devices and determine the discrepancy between the volume purchased
and the volume used in production.
2. Analyze ESDS usage, including average inventory levels and locations, requisitioning
departments, purchase volume, and unit cost.
3. Define burden costs associate with ESDS devices and assemblies.
4. Evaluate the overall impact of ESD damage.
One approach is to identify the finished products produced during a particular period that
contain ESDS components or assemblies, especially those that contain very sensitive (low
ESD withstand voltage) devices. The number of product produced is usually easily found
from production statistics, and the number of devices or assemblies used in the product
can be calculated. Sometimes a difference between planned and realized product numbers
is evident, which can itself indicate that something was going on that might be worth inves-
tigation, such as excessive rework, shortage of parts, or field problems.
If the ESD withstand data for the ESDS devices can be obtained, then the lowest ESD
withstand devices are the ones most likely to be worth further investigation. Unfortunately,
this data is often difficult to obtain as it is not always published on device data sheets.
For the devices selected for further investigation, inventory and purchasing records can
be obtained. The actual number of ESDS items purchased in the period can be obtained,
including the starting and final inventory and the cost of the ESDS devices. It can also be
useful to find details of the locations or departments at which the ESDS devices are used.
Analysis of these data and the production figures may reveal a discrepancy between the
number of devices leaving in the product and the number purchased and remaining in stor-
age. An excessive difference may indicate, for example, rework or field service consumption
due to failures. Of course, this would not in itself confirm ESD as the failure cause in all
cases, but identification of any failure mode is usually a useful outcome. The analysis may
give useful indication of where to focus further resources.
Knowledge of the number of failed devices and their cost allows calculation of failed
component cost, but this represents only a part of the cost of the failures. If the number
of reworked items or field or other failures can be identified, the associated costs can
304 9 How to Evaluate an ESD Control Program

be estimated. These may include labor, facility, power, and other expenditure. Average
costs may be easier to estimate than real costs per individual failed item. Halperin (1986)
described preparation of the data in the form of a spreadsheet table, listing the sensitive
parts, and for each part the ESD withstand voltage, difference between numbers sourced
and those used in final in product, cost of each item, estimated “burden” of associated
costs, and total lost cost of the components and associated burden. The magnitude of total
cost per item give a means of ranking for further work, and the overall total cost for all
items gives a view of the possible cost of failures.
The cost of the failure is likely to vary with the production stage at which the failure is
found, including field failures. The cost of failure typically increases as the product advances
through the production process. Field failures are usually the costliest of all. With some
markets and product types (e.g. satellites and aerospace), the cost of a field failure, in eco-
nomic and other terms, is extremely high and could include equipment downtime, loss of
the product, and even threat to life or property. Some product may be irreplaceable or unser-
viceable. This consideration in itself may justify considerable care in ESD control during
product manufacture, storage, transport, and handling.
Halperin (1986) provided an example of this type of analysis. An interesting case study
was also later provided by Helling (1996) (see Section 9.5.4).
While it may be desirable to confirm ESD damage with failure analysis of the failed com-
ponents, few organizations do this in practice. One reason may be that failure analysis of
components can be a time-consuming process requiring several person-days specialist effort
and equipment. The outcome is not always conclusive in identification of ESD failures. In
particular, ESD damage and electrical overstress damage often show similarities. The situ-
ation can be worsened if adequate ESD control is not maintained at all stages of handling
including between discovery of failure and into failure analysis, as ESD could still damage
an already failed component!
Dangelmayer (1999) reported several interesting case studies that formed a convincing
demonstration of the economic benefit of ESD control. These used several different
approaches to establish the source or presence of ESD failures and benefit of ESD controls,
including the following:
● Use of failure analysis to identify ESD failures
● Correlation of ESD control deviations and ESD losses
● Comparison of failure rates of batches of product assembled with, and without, ESD con-
trol measures
● Reproduction of failure of components using simulated ESD
● Testing of devices before and after operations to determine the source of ESD damage

9.5.2 The Benefit Arising from of the ESD Control Program


It can be difficult to know with certainty the cost of ESD failures to the business. It can be
much easier to establish at least a standard ESD control program addressing the usual ESD
risks. Nevertheless, the cost of ESD failures can be considerable and has been evaluated in
some studies (e.g. Helling 1996; Halperin 1986). One problem is that failure analysis to a
level that can positively identify ESD failures can be time-consuming and therefore expen-
sive. It can be difficult to distinguish between failures due to EOS and ESD (Lin et al. 2014).
9.5 Evaluation of Cost Effectiveness of the ESD Control Program 305

One common reason for my clients to first contact me for help has been to evaluate their
ESD control after a customer audit has declared it inadequate. The customer is not always
right in terms of their perception of control of real ESD risks. They do, however, sometimes
opt to take their business elsewhere due to perceived inadequate of ESD control in the orga-
nization’s facility. A dispute with a customer over this costs time and resources and is not
good for customer relations, even if their evaluation is of ESD control is incorrect.
Implementation of a good ESD control program, with good documentation of ESD con-
trol processes and compliance with a standard, can do much to convince a customer that
adequate care is taken and ESD control is effective. For some customers, compliance with
an ESD control standard can be a prerequisite to doing business with a supplier organiza-
tion. In this situation, the organization’s positive attitude to ESD control can even become
a useful marketing benefit.

9.5.3 Evaluation of the Cost of an ESD Control Program


The costs of implementing an ESD control program includes aspects such as
● Identification and analysis of ESD related failures during production
● Identification of process stages at which failures are found and possible ESD sources
● Estimation of cost of failure and rework at each stage
● Analysis of failures occurring at the customer site
● Estimation of cost of ESD control
● Setup and maintenance of EPAs and acquisition of ESD control equipment
● Documentation of the ESD control process
● Compliance verification
● ESD training
● Maintenance and replacement of ESD control equipment
Some of these are easier to estimate than others. Some costs are likely to vary considerably
with product type and market characteristics.
One question that is usually worth considering is, what would be the cost of an ESD
failure occurring at the customer’s site? The answer to this will help put into context the
importance of ESD control for the organization. A throwaway product in a consumer mar-
ket might indicate that minimal ESD control sufficient to maintain acceptable failure rates
is adequate. In a high-reliability high-value product market, the cost of a single failure,
and associated costs like downtime or safety issues, can justify considerable investment.
Some products lines require failure levels in the parts per million range. In a market such
as satellite manufacture, a single failure may be intolerable.

9.5.4 ROI in ESD Control


Effective investment in ESD control should lead to a worthwhile return. According to
Halperin (Brandt 2003), a properly implemented ESD control program can have a return
on investment exceeding five to one within six months. Unfortunately, there are few
published research studies on this aspect of ESD control program evaluation. An early
example is Downing (1983). An ROI estimate can also be made for an individual ESD
306 9 How to Evaluate an ESD Control Program

control measure. Gumkowski and Levit (2013) examined the use of air ionization and
compared different types of ionizers in a semiconductor manufacturing process.
Helling (1996) found that internal studies had shown that a failure to observe an ESD con-
trol measure in their facility typically resulted in about 1% of ESDS devices being stressed
by ESD. About 10% of stressed devices resulted in a defect or failure. He therefore assumed
an ESD failure rate of 0.1% per ESD control fault. He calculated on this basis failure rates
for five production lines. He added the cost of repair of ESD failures. He found that 60% of
failures were found at PCB test, and 30% at system test. Ten percent of failures were found
at the customers site. His calculations were compared over two business years. The repair
costs at each stage was calculated as follows:

Stage repair cost = number of product × %failures × ESD failure rate (0.1%)
× repair cost per item

An example summary of some of Helling’s results for a manual process handling 80 000
products per year is given in Table 9.1. Two interesting things can be seen from this example.
First, the cost of failures increases as the failure is discovered at later stages in production.
Second, the most expensive cost is that of failures at the customers site. These characteristics
are likely to be typical of most cost of product failure profiles.
Helling estimated the repair costs for other processes in the same way and was then able
to add these to estimate the cost of ESD damage related repairs to the business. He then
estimated the cost of ESD control measures required for an example facility in terms of
packaging, ESD control equipment, training, audit, and other items for each business year
under consideration. The ROI could then be calculated as the cost/benefit ratio of expendi-
ture on ESD control compared to anticipated saving on ESD failures. For the two business
years considered he found ROI values of 3:1 and 11:1, respectively. He commented that ESD
protection faults led to the product being more expensive than necessary, and this alone,
aside from the loss of customer reputation that results from defects and failures, cost the
business more than ESD protection measures.
Other workers have reported high return on investment figures. Danglemayer (1999)
reported in some of his case studies ROI as high as 185% and even 950%. In another case
study, an investment of $1000 resulted in a saving of $6 000 000. He concluded that using
ESD controls can lower operating costs with a ROI of up to 1000% while simultaneously
enhancing product quality and reliability.

Table 9.1 Costs of repair of ESD damage in a manual process (Helling 1996).

Failures found ESD failure Stage item Total stage


Stage at this stage (%) rate repair cost (DM) repair cost (DM)

PCB test 60 0.1% 100 4800


Final test 30 400 9600
Customer’s site 10 2000 16 000

Total cost to business 30 400


9.5 Evaluation of Cost Effectiveness of the ESD Control Program 307

9.5.5 Optimizing an ESD Control Program


Establishing and maintaining a good ESD control program requires investment in time
and resources. A key to achieving a good return on investment is acting through knowl-
edge and understanding. Each action should be taken and ESD control measures should be
applied for a reason addressing a need. The need may be to address an ESD risk or improve
the effectiveness or efficiency of the ESD control process or improve value (e.g. in cus-
tomer perceptions or compliance with a standard). Improving cost effectiveness of an ESD
control program is likely to require balancing trade-offs between costs, for example in equip-
ment purchases, documentation time, training, and compliance verification (Smallwood
et al. 2014).
ESD controls applied without knowledge and understanding can often be unnecessarily
expensive or make little real contribution to ESD control. ESD control equipment often
works as part of a system with other equipment (see Section 4.7.8.2). If the system is not
well understood, another part may be omitted or incorrectly specified. The result may be
that the system may be compromised or even rendered ineffective.
Investment is often thought of in terms of ESD control equipment and materials, but it
also needs resources such as the following:
● Planning, development, and implementation time
● Documentation time
● Selection and qualification of ESD control equipment, packaging, and materials
● Purchase, installation, and commissioning of EPA equipment and materials
● Development and regular provision of an ESD training program
● Development and regular execution of a compliance verification program
● Product failure detection, tracking, and analysis
All of these costs are really investments that should together provide a return on the
investment. ESD control applied without knowledge and understanding is akin to financial
investment applied without knowledge and understanding. Either can result in expendi-
ture with little return. Furthermore, the investment in ESD control should be appropriate
to the value of the product, market needs, and consequences of an ESD-related failure. A
low-cost throwaway consumer product with unimportant consequence of failure may merit
minimal ESD control investment. A high-cost high-reliability product with a high cost of
failure would merit larger investment.
Many of the costs are inter-related. For example, a complex ESD control program using
a variety of control techniques and equipment and different implementation in different
areas is also likely to have more complex training and compliance verification needing
extensive documentation. Each variation on ESD control equipment or procedure needs
documentation, training, and compliance verification.
Implementation of different ESD control measures in different areas can sometimes lead
to equipment savings. It may also lead to confusion in workers who enter the different areas,
and problems with maintaining compliance. It can also lead to conflict with customers who
audit the areas, especially if unconventional ESD control measures are used or common
standard ESD controls are omitted, as they may believe the implemented control measures
are inadequate or incorrect.
308 9 How to Evaluate an ESD Control Program

In contrast, a simple standardized ESD control program implementing the same con-
trol measures in different areas may lead to reduced training and compliance verification
needs and costs and help prevent operator confusion. It may also reduce the risk and cost
of conflict with customers or auditors.
Against this, equipping EPAs with the same equipment, whether it is necessary for ESD
risk control or not, is likely to increase equipment and compliance verification costs. Any
additional equipment must of course be regularly tested and verified.
When optimizing an ESD control program, it is essential to analyze and understand the
ESD risks that are present in the processes and facilities. Without this, ESD controls may
be implemented that are not necessary or risks may be present that are not addressed. ESD
risks are often not obvious or well understood. Adequate skills and experience are required
for successful analysis of these. Implementation of a standard will address the standard
well-known ESD risks but may not cover more unusual ESD risks (Gaertner 2007).
With a little knowledge and understanding, the cost of ESD control can be high and
the effectiveness low (Smallwood et al. 2014). This is perhaps most likely to occur when
the low ESD withstand voltage devices are handled. With a high level of knowledge and
understanding, the cost of ESD control can be reduced and the effectiveness maximized.
Investment in high-level training in the principles and practice of ESD control for the ESD
coordinator and other personnel working on development of the ESD control program can
be worthwhile if not essential. Strategies for optimizing the ESD control program are further
discussed in Section 10.7.

9.6 Evaluation of Compliance of an ESD Control Program


with a Standard

9.6.1 Two Steps to Compliance Evaluation


There are arguably two steps to evaluating compliance of an ESD control program with a
standard. First, the ESD program documentation must be evaluated for compliance with
the standard. The second step is to evaluate the ESD program in practice for compliance
with the ESD program documents. If both steps find compliance, then compliance with the
standard is achieved.

9.6.2 Using Checklists to Evaluate Compliance of Documentation with a


Standard
Compliance of an ESD control program documentation with a standard can be evaluated
with the aid of a check list of the requirements of the standard. The check list is compiled
by going through the standard in detail, noting every requirement.
The following tables are an example of a check list compiled in this way, with reference
to the IEC 61340-5-1:2016 standard. For compliance, all the items in the tables should be
addressed by the ESD program plans. The compliance evaluation response can be in terms
such as “satisfactory,” “major non-compliance,” “minor non-compliance,” or “not applica-
ble” to suit the user.
9.6 Evaluation of Compliance of an ESD Control Program with a Standard 309

The IEC 61340-5-1 states under the heading “ESD Coordinator” that “A person shall be
assigned by the organization with the responsibility for implementing the requirements of
this standard including establishing, documenting, maintaining, and verifying the compli-
ance of the program.” This translates into the checklist of Table 9.2.
Table 9.3 gives a checklist of the broad requirements of the ESD Control Program Plan.
The detail of compliance of the Compliance Verification Plan is covered in Table 9.7. As an
example, the Compliance Verification Plan might be counted as “defined” if some elements
of compliance verification are covered in some way, e.g. testing of wrist straps. That does not
mean it is necessarily considered adequate – further requirements are tested in Table 9.7.
Table 9.4 covers the requirements for “tailoring” given by the standard. This does not mean
that some tailoring must be specified – if none is needed, none needs be documented.

Table 9.2 ESD Coordinator duties.

Compliance
Requirement evaluation Notes

The organization must assign an ESD


coordinator

Stated responsibilities of ESD coordinator include


Establishing the program
Implementing the requirements of the
standard
Documenting the program
Maintaining the program
Verifying the program

Table 9.3 ESD control program plan – topics covered.

Compliance
Requirement evaluation Notes

ESD training plan is defined


Product qualification plan is defined
Compliance Verification plan is
defined
Grounding/bonding systems are
defined
Personnel grounding is defined
EPA requirements are defined
Packaging systems are defined
Marking requirements are defined
The ESD Control Program Plan is
applied to all relevant aspects of the
organization’s work
310 9 How to Evaluate an ESD Control Program

Table 9.4 Tailoring.

Compliance
Requirement evaluation Notes

Evaluation of applicability of each


requirement is adequate.
Documentation of tailoring decisions
is adequate.
Tailoring statements cover situations
where limits of the standard are
exceeded.

Table 9.5 Content of the ESD training plan.

Compliance
Requirement evaluation Notes

Define all personnel that are required


to have training.
Initial awareness and prevention
training to be provided before
personnel handle ESDS devices.
Recurrent training is required.
The type of training is defined for all
relevant personnel.
Frequency of training is defined for all
relevant personnel.
There is a requirement for maintaining
training records.
The location where records are kept is
adequately documented.
The training methods used are
adequately documented.
The methods used to ensure
comprehension and training adequacy
are documented.

Table 9.5 covers the required content of the ESD Training Plan in detail, and Table 9.6
covers the detailed content of the ESD Control Product Qualification Plan. Table 9.7 covers
in more detail the required content of the Compliance Verification Plan.
Table 9.8 concerns the definition of grounding and bonding systems. Suitable grounding
methods must be defined, but not all the methods in this table must be used in all facili-
ties. Usually only one will be sufficient. For example, many facilities will use the electrical
protective earth as ESD earth (ground).
9.6 Evaluation of Compliance of an ESD Control Program with a Standard 311

Table 9.6 ESD control product qualification plan.

Compliance
Requirement evaluation Notes

All ESD control items selected for use


are qualified in some specified way.
The technical requirements to be
verified are adequately defined.
The pass criteria are adequately
defined.
Any test methods used are adequately
defined and documented.
ESD control items not listed in the
standard but considered to be part of
the ESD control program are qualified

Table 9.7 Compliance verification plan.

Compliance
Requirement evaluation Notes

The technical requirements to be


verified are adequately defined.
The pass criteria are adequately
defined.
The frequency of verifications is
adequately defined.
All test methods used are adequately
defined and documented.
If test methods are nonstandard,
correlation with standard
measurements is adequately
documented.
Test methods used to verify items not
covered in the standard are
documented with corresponding test
limits.
Compliance verification records are
required to be established.
Compliance verification records are
required to be maintained.
Test equipment is capable of making
the required measurements
312 9 How to Evaluate an ESD Control Program

Table 9.8 Grounding and bonding systems.

Compliance
Requirement evaluation Notes

It is required that all conductive and


dissipative equipment are connected to
ground or each other.
At least one of the following are defined
as “ground.”
Grounding via
electrical protective
earth
Functional ground
Equipotential
bonding

Table 9.9 Personal grounding.

Compliance
Requirement evaluation Notes

All personnel are required to be


grounded.
When personnel are seated, they are
required to wear a wrist strap.
Personnel using footwear-flooring
grounding must wear footwear on both
feet.
Resistance to ground and body voltage
criteria are both addressed.

Table 9.9 concerns personal grounding. The detailed pass criteria given in current stan-
dards 61340-5-1:2016 and ESD S20.20-2014 are given in Chapter 6, in this case in Table 6.5.
Table 9.10 gives general requirements for EPAs. Requirements for EPA equipment such as
floors, bench mats, and chairs are given in in Chapter 6, in this case in Table 6.6.
Table 9.11 covers the requirements for ESD protective packaging. The classification of
packaging materials and requirements of IEC 61340-5-3:2015 are given in in Chapter 6, in
this case in Tables 6.10 and 6.11. Table 9.12 covers the requirements for marking for ESD
control purposes.

9.6.3 Evaluation of Compliance of a Facility with the ESD Control Program


The task of evaluation of compliance of a facility with its ESD Control Program Plan should
be specified in the Compliance Verification Plan part of the documentation. Check lists can
also be used to aid this process. Every requirement of the ESD Control Program Plan should
9.6 Evaluation of Compliance of an ESD Control Program with a Standard 313

Table 9.10 General requirements for EPAs.

Compliance
Requirement evaluation Notes

Handling of unprotected ESDS devices


must always be within an EPA.
The boundaries of the EPA must be
clearly identified.
Access to the EPA is limited to trained
personnel or personnel escorted by
trained personnel.
All nonessential insulators must be
removed from locations where
unprotected ESDS devices are handled.
ESD threat evaluated such that field
<5 kV m−1 .
ESD threat evaluated such that 2 kV
potentials are kept >30 cm from ESDS
devices.
ESD threat evaluated such that 125 V
potentials are kept >2.5 cm from ESDS
devices.
Where fields or potentials exceed
limits, method of mitigating the ESD
risk is defined.
Where conductors come into contact
with ESDS devices and cannot be
grounded, the voltage difference
between the conductor and ESDS
devices is reduced below 35 V.

Table 9.11 Packaging.

Compliance
Requirement evaluation Notes

Use of packaging in accordance with


customer contracts, purchase orders,
drawing, or other documentation is
required where it is defined.
(61340-5-1) Packaging requirements
must be defined for ESDS devices not
covered by customer contracts or
documentation, in accordance with
the packaging standards.
(61340-5-1) Packaging where required
is defined for all material movement
within EPAs, between EPAs, between
job sites, field service operations, and
to the customer.
314 9 How to Evaluate an ESD Control Program

Table 9.12 Marking for ESD control purposes.

Compliance
Requirement evaluation Notes

Marking according to customer


contracts, purchase orders, drawing, or
other documentation is required where
it is defined.
Evaluation of need for marking is
required where marking is not defined
by customer contract, purchase orders,
drawing, or other documentation.
If marking is required, marking
requirements must be documented in
ESD Control Program Plan.

be tested in the Compliance Verification Plan and hence in evaluation of compliance of the
facility with the plan.

9.6.4 Common Problems


External auditors often audit and comment on an ESD Program based on what is practiced
in their own facility. This is a mistake – an organization’s facility and practices should always
be audited against its own documented ESD Control Program Plans. So, if an organization
has a well-documented ESD Control Program Plan that is compliant with a standard and
complies with the plans in practice, this can go a long way in preventing adverse comments
from customers who are used to doing things in a different way.
Tailoring allows the organization to do things in a different way to that required by the
standards. But, if the tailored practice is not adequately documented, it represents a non-
compliance. Good documentation of a tailored ESD control measured, with documenta-
tion of the technical rationale and any tests supporting its implementation, is necessary to
defend the user against accusations of non-compliance. The detailed documentation may
be in a separate report cross referenced by the ESD Control Program Plan.

References

Bellmore D. G. (2004) Paper 4A.6. Characterizing Automated Handling Equipment Using


Discharge Current Measurements. In: Proc EOS/ESD Symp. EOS-26. Rome, NY, EOS/ESD
Association Inc.
Brandt M T. (2003) What does ESD really cost? Available from: http://circuitsassembly.com/
cms/images/stories/pdf/0306/0306esd.pdf [Accessed 6th March 2019]
Dangelmayer, T. (1999). ESD Program Management, 2e. Clewer. ISBN: 0-412-13671-6.
Downing, M.H. (1983). ESD Control Implementation and Cost Avoidance Analysis. In: Proc
EOS/ESD Symposium EOS-5, 6–11. Rome, NY: EOS/ESD Association Inc.
References 315

EOS/ESD Association Inc. (2014). ANSI/ESD S20.20–2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical
and Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive
Devices). Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015). ESD TR17.0–01-15. Technical Report for ESD Process
Assessment Methodologies in Electronic Production Lines – Best Practices used in Industry.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016a) ESD SP10.1–2016. Standard practice for protection of
Electrostatic Discharge Susceptible Items–Automated handling Equipment (AHE). Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016b) ESD Association Electrostatic Discharge (ESD) Technology
roadmap – revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf [Accessed: 10th May 2017]
Gaertner R. (2007) Do We Expect ESD-failures in an EPA Designed According to International
Standards? The Need for a Process Related Risk Analysis. Proc. EOS/ESD Symposium
EOS-29 Paper 3B.1 pp. 192–197
Gaertner R., Stadler W. (2012) Paper 3B.5. Is there a Correlation Between ESD Qualification
Values and the Voltages Measured in the Field? In: Proc. EOS/ESD Symposium EOS-34.
Rome, NY, EOS/ESD Association Inc.
Gumkowski G., Levit L. (2013) EOS-35 Paper 7B4. A New Look at the Financial Impact of Air
Ionization. In: Proc. EOS/ESD Symposium. Rome, NY, EOS/ESD Association Inc.
Halperin, S.A. (1986). Estimating ESD losses in the complex organisation. In: Proc. EOS/ESD
Symp. EOS-8, 12–18. Rome, NY: EOS/ESD Association Inc.
Halperin S. A., Gibson R, Kinnear J. (2008) EOS-30 2B-2. Process Capability & Transitional
Analysis. In: Proc. EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Helling, K. (1996). ESD protection measures – Return on investment calculation and case
study. In: Proc. EOS/ESD Symp. EOS-18, 130–144. Rome, NY: EOS/ESD Association Inc.
Industry Council on ESD Target Levels (2010) White paper 2: A case for lowering component
level CDM ESD specifications and requirements. Rev. 2.0. http://www.esdindustrycouncil
.org/ic/en/documents/6-white-paper-2-a-case-for-lowering-component-level-cdm-esd-
specifications-and-requirements [Accessed: 10th May 2017]
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering component
level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf [Accessed: 10th May
2017]
International Electrotechnical Commission. (2015) IEC 61340–5-3:2015. Electrostatics.
Protection of electronic devices from electrostatic phenomena. Properties and requirements
classifications for packaging intended for electrostatic discharge sensitive devices. Geneva, IEC.
International Electrotechnical Commission (2016) IEC 61340–5-1: 2016. Electrostatics – Part
5–1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Rectifier. (2004). ESD Testing of MOS Gated Power Transistors. AN-986. http://
www.infineon.com/dgdl/an-986.pdf?fileId=5546d462533600a40153559f9f3a1243 [Accessed:
10th May 2017]
316 9 How to Evaluate an ESD Control Program

Jacob P., Gärtner R., Gieser H., Helling K., Pfeifle R., Thiemann U., Wulfert F., Rothkirch W.
(2012) EOS-34. Paper 3B.8. ESD risk evaluation of automated semiconductor process
equipment – A new guideline of the German ESD Forum e.V. In: Proc. EOS/ESD Symp.
Rome, NY, EOS/ESD Association Inc.
Lin N., Liang Y., Wang P. (2014) DOI: 10.1109/ICEPT.2014.6922951. Evolution of ESD process
capability in future electronics industry. In: Proc. 15th Int. Conf. Electronic Packaging Tech.
Bristol, England, IOP Publishing.
Paasi, J., Smallwood, J., and Salmela, H. (2003). EOS-25 Paper 2B4. New Methods for the
Assessment of ESD Threats to Electronic Components. In: Proc. EOS/ESD Symp, 151–160.
Rome, NY: EOS/ESD Association Inc.
Paasi, J., Salmela, H., and Smallwood, J.M. (2006). Electrostatic field limits and charge
threshold for field-induced damage to voltage susceptible devices. J. Electrostat. 64: 128–136.
Smallwood J M. (2019) Can ElectroStatic Discharge Sensitive electronic devices be damaged by
electrostatic fields? In: Proc. Electrostatics 2019. J. Phys. Conf. Se. Vol. 1322 01 2015 Available
from: https://iopscience.iop.org/article/10.1088/1742-6596/1322/1/012015/pdf [Accessed
Oct. 2019]
Smallwood J., Paasi J., (2003) Assessment of ESD threats to electronic devices, VTT Research
Report No BTUO45-031160
Smallwood J., Taminnen P., Viheriaekoski T. (2014) EOS-36. Paper 1B.1. Optimizing
investment in ESD Control. In: Proc. EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Stadler W., Niemesheim J., Seidl S., Gaertner R., Viheriaekoski T. (2018) EOS-40 Paper
1B.4.The Risks of Electric Fields for ESD Sensitive Devices. In: Proc. EOS/ESD Symp. Rome,
NY, EOS/ESD Association Inc.
Steinman A. (2010) EOS-32 Paper 3B3. Measurements to Establish Process ESD Compatibility.
In: Proc. EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Steinman A. (2012) EOS-34 Paper 2B.4. Process ESD Capability Measurements. In: Proc.
EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Swenson D. E. (2012) EOS-34 paper 3B.6. Electrical fields: What to worry about? In: Proc.
EOS/ESD Symp. Rome, NY, EOS/ESD Association Inc.
Tamminen, P. and Viheriäkoski, T. (2007). EOS-29 Paper 3B3. Characterization of ESD Risks in
an Assembly Process by Using Component-Level CDM Withstand Voltage. In: Proc.
EOS/ESD Symp, 202–211. Rome, NY: EOS/ESD Association Inc.
Viheriäkoski T, Ristikangas P, Hillberg J, Svanström H, Peltoniemi T. (2012). Paper 3B.2.
Triboelectrification of static dissipative materials. In: Proc. EOS/ESD Symp. Rome, NY,
EOS/ESD Association Inc.
Wallash, A. (2007). EOS-29 Paper 2B8. A Study of “Soft Grounding” of Tools for ESD/EOS/EMI
Control. In: Proc. EOS/ESD Symp, 152–157. Rome, NY: EOS/ESD Association Inc.
317

10

How to Develop an ESD Control Program

10.1 What Do We Need for a Successful ESD Control


Program?

10.1.1 The ESD Control Strategy


An electrostatic discharge–sensitive (ESDS) device needs protection from potentially
damaging electrostatic discharge (ESD) at all times. This can be done by adopting a simple
overall ESD control strategy.
● Unprotected ESDS devices are handled only in an area where ESD risks are controlled
to an acceptable level. This area is usually called an electrostatic discharge protected area
(EPA).
● Outside EPAs, in unprotected areas (UPAs), the ESDS device is always protected within
ESD protective packaging.
If this strategy is implemented and the ESD packaging and control measures used within
the EPA are sufficient and appropriate, it’s likely that the ESD program will be effective.
Notice that this strategy immediately implies that an ESDS device is not removed from its
ESD protective packaging unless it is within an EPA.
This chapter looks at how the ESD control program can be developed and documented
according to the International Electrotechnical Commission (IEC) 61340-5-1: 2016 stan-
dard and determine the control measures that will be used in the EPA. ESD control equip-
ment is discussed in more detail in Chapter 7. ESD protective packaging is discussed in
Chapter 8. ESD measurements are covered in Chapter 11 and ESD training in Chapter 12.

10.1.2 How to Develop an ESD Control Program


Dangelmeyer (1990), based on his experience at AT&T, commented that developing, imple-
menting, and managing a successful ESD program requires a system approach from product
design to customer acceptance. A well-managed program can be much more effective than
one well stocked with expensive supplies. He found that 12 critical factors form the basis of
a successful ESD control program.
1. An effective implementation plan
2. Management commitment
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
318 10 How to Develop an ESD Control Program

3.
A full-time ESD coordinator to serve as a consultant and oversee the plan
4.
An active committee to help implement the ESD program
5.
Realistic requirements
6.
Training for measurable goals
7.
Auditing using scientific measures
8.
ESD test facilities to qualify and test ESD control equipment
9.
A communication program to keep people aware of the ESD issue and demonstrate
progress
10. Systematic planning
11. Human factor engineering to take care of employee needs and render human error
unlikely
12. Continuous improvement

Effective implementation of these factors will produce an effective and cost-effective


ESD control program, and continuous improvement sustains the success. Dangelmeyer
states that failure to continuously improve the process will translate into complacency and
deterioration.
Effective ESD protection requires the following elements:

● Identification of ESDS device.


● Understanding in which processes and facilities they must be exposed and handled in an
unprotected state. These areas will be EPAs. All other areas will be regarded as UPAs.
● Ensuring that within EPAs, ESD risks are adequately controlled. This is done by speci-
fication of appropriate ESD control equipment (e.g. wrist straps, bench mats) as well as
removal of items and equipment that do not comply with ESD control requirements or
could cause ESD risk (e.g. ordinary packaging materials).
● Ensuring that when outside EPAs, ESDS devices are protected from ESD damage
(e.g. by enclosure within equipment or ESD protective packaging).
● Training of personnel in ESD-related matters.
● Regular checks and tests (Compliance Verification) to ensure that equipment is working
and procedures are being observed correctly.
● Documentation of ESD control, training, and compliance verification procedures and
requirements.

To specify and use appropriate and effective ESD controls, it is important to understand
ESD control technology and the role of different items of ESD control equipment. Speci-
fication of equipment without this understanding can lead to an overly expensive and yet
poorly effective ESD control program (Smallwood et al. 2014). The equipment specified may
be unnecessary or may fail to function as a system in the absence of other items.

10.1.3 Safety and ESD Control


Safety considerations and local safety regulations must always be given top priority over
ESD control.
One area in which apparent contention between safety and ESD control requirements
often occurs is in handling ESDS devices that are in a powered state with high voltage
10.1 What Do We Need for a Successful ESD Control Program? 319

present or that contain batteries that could give electric shock risk to personnel. Another
area is where processes require the operator to use personal protective equipment against
the effects of chemicals, high temperatures, or other risks.
Often, by careful evaluation of the risks and process steps, ways of working can be found
that both maximize safety and minimize ESD risks. As an example, handling an unpro-
tected ESDS device may be required while setting up for live testing with high voltages
present. Using personal grounding with live voltages present could give electric shock risks.
If it can be arranged that the handling of the ESDS device can be done before the high
voltages are present, the safety risk is removed. Not touching an ESDS device when live
voltages are present at the same time prevents ESD risk from charged personnel and pre-
vents electric shock to the personnel! If it is necessary to work on the ESDS device in
a live state, precautions such as use of insulating rubber gloves may be required to pro-
tect the person against electric shock risks. The same gloves may give some protection
to the ESDS device against human body ESD at body voltages up to the breakdown volt-
age of the gloves. (If the person’s electrostatic body voltage could exceed the breakdown
voltage of the glove, an ESD risk could result, but more importantly the protection pro-
vided by the glove against shock risk is compromised.) When using rubber gloves in this
way, an ESD risk could arise from electrostatic fields due to charging of the glove mate-
rial. This is usually a lower risk but should be evaluated and if necessary countered in
some way.
In practice, the effect of electric shock depends on the electrical current flowing through
the body, rather than the source voltage present. The effects were summarized by Dalziel
(1972) and become more hazardous as current flow level and duration of the shock
increases. The sensitivity of the body to shock also depends on the body parts that are
subjected to the current flow, and whether the source is direct current (DC) or alternating
current (AC), and the frequency of AC current. Women are typically susceptible at lower
current levels than men, and individuals vary greatly in their sensitivity. Other conditions
such as presence of moisture can be influential. As current is increased, the effects start
with perception of the current (tingling or warmth) and can include at higher levels painful
shocks, stopping of breathing, inability to let go, burns, ventricular fibrillation (cessation
of heart action), and immediate or delayed death.
Where safety guidelines or regulations do not cover a situation, the guidelines in
(Table 10.1) are often given in electrical textbooks, e.g. Nave and Nave 1985. Safety can be
evaluated by comparing possible electrical currents with these thresholds.

Table 10.1 Electrical current physiological effects for AC 60 Hz current.

Current level (mA) Physiological effect

1 Threshold of perception
5 Maximum harmless current
10–20 Sustained muscular contraction “let go current”
50 Pain, possible fainting, exhaustion, heart, or respiratory effects
>100 Ventricular fibrillation, death possible
320 10 How to Develop an ESD Control Program

10.2 The EPA


10.2.1 Where Do I Need an EPA?
A major contribution to ESD protection may be made by minimizing handling of ESDS
devices in an unprotected state. Where handling is necessary, some form of ESD control is
required – meaning the ESD protected areas.
Any area in which unprotected ESDS devices may be handled in an unprotected state
needs to be an EPA in some form. That does not necessarily mean it needs to have all the pos-
sible range of ESD control measures and equipment. To be effective, the EPA must have the
following features:
● A known boundary in which ESD control measures are applied
● ESD control measures addressing the significant ESD risks to an acceptable level
The equipment and practices are discussed in more detail in the following sections.
The first step in determining ESD hazards is to identify where ESDS components, printed
circuit boards (PCBs), or other items (ESDS device) are handled in an unprotected state in
the facility. Many semiconductor devices and some passive devices are ESD susceptible.
In general PCBs, modules and similar assemblies should be considered ESD susceptible if
they contain ESDS devices and are not protected from ESD by being fully contained within
an enclosure or some other factor. Even PCBs or modules that are fully contained within
an enclosure may have some susceptibility to ESD that may occur via a connector or flying
leads.

10.2.2 Boundaries and Signage


A clear boundary, obvious to all personnel approaching the entrance to the EPA, is neces-
sary. If this is not present, personnel will not know whether they are inside or outside the
EPA and where the ESD control measures are to be applied (a simple control measure is to
keep untrained personnel out of the EPA and reduce handling of unprotected ESDS devices
to a minimum).
Modern ESD standards require that the EPA is marked by signage visible to personnel
before they enter the EPA.
EPA boundaries should be carefully chosen to include all areas where ESD control is
required (because unprotected ESDS devices are to be handled) but preferably exclude areas
or processes where ESD control is not required or not possible. Including inappropriate
processes within an EPA can lead to difficult noncompliance issues and confusion over
acceptable ESD practices. Nevertheless, EPAs and UPAs should be designed for operator
convenience, as convenient practices will assist compliance and inconvenience can encour-
age noncompliance.

10.3 What Are the Sources of ESD Risk in the EPA?


The next step is to identify the potential ESD sources in each process. The most common of
these are
10.4 How to Determine Appropriate ESD Measures 321

● A charged person touching the ESDS device


● A charged metal or conductive object, tool, or another item touching the ESDS device
● The ESDS device becoming charged and touching a conductive item (e.g. metal part or
equipment)
Electrostatic fields are not usually in themselves damaging (with a few exceptions).
However, electrostatic fields help set up the conditions in which ESD can occur because
any isolated conductor (e.g. metal parts or the device itself) within the electrostatic field
will attain a voltage. If two conductors touch (or become sufficiently close to each other)
within an electrostatic field and at least one is isolated (i.e. not grounded), then ESD will
occur between them due to voltage induced on the isolated conductor.
A particularly damaging form of ESD can occur when an ESDS device contacts a high-
conductivity (low-resistance, e.g. metal) item. An ESD event occurring in this circumstance
can have a short-duration high discharge current due to the low resistance and inductance
of the discharge circuit between them. The susceptibility of the ESDS device to this type of
event is characterized by its charged device model (CDM) withstand voltage. This type of
damage can often be avoided by ensuring that, where possible, the device instead contacts
a higher resistance (>104 Ω) material.

10.4 How to Determine Appropriate ESD Measures


10.4.1 ESD Control Principles
The first step in determining the appropriate ESD control measures is to define the bound-
ary of each EPA. These must then be marked and signed so that personnel can easily identify
which areas are EPA and which are uncontrolled areas. Within the EPAs, the ESD control
measures can then be determined.
● Personnel handling ESDS devices must be grounded so that they cannot be at a high
enough voltage to damage the ESDS device that they touch. This is normally accepted to
mean that the body voltage on personnel must not exceed the human body model (HBM)
ESD withstand voltage of the most sensitive ESDS device. For an ESD program handling
100 V HBM devices, the maximum voltage allowed on a person handling ESDS devices
must be less than 100 V.
● Any metal or conductive items that contact ESDS devices must be grounded where pos-
sible to ensure that they are not charged.
● Sources of high electrostatic fields (e.g. charged insulators or equipment that generates
external electrostatic fields) must be kept far enough away from ESDS devices not to risk
inducing high voltages on the device (or on isolated conductors that may contact the
ESDS device).
An ESDS device will often be in an isolated (nongrounded) state and must make contact
with metal objects or other components during the process. ESDS devices are not normally
grounded as an ESD protection measure. Because of this, it will normally have some level of
residual voltage on it. If the ESDS device is susceptible to charged device ESD and the volt-
age difference is sufficient, a risk of damage could occur, so the act of grounding an ESDS
322 10 How to Develop an ESD Control Program

device to a low-resistance conductor often carries a charged device ESD risk. This risk can
be minimized by contacting the ESDS device through a resistive material (see Section 10.3)
rather than a low-resistance material such as a metal. Where ESDS devices must contact
conductive items (including other devices, tools, or PCBs), the ESD risks should be evalu-
ated, and ESD prevention measures may be required e.g. to reduce the voltages possible on
the ESDS device and conductors or increase the resistivity of the conductor.
The ESD control measures required within the EPA will be selected to address the ESD
risks found. In a typical EPA with manual processes, these will include
● Grounding of personnel handling ESDS devices either by wrist strap or by ESD control
footwear and flooring. (Occasionally other means e.g. personal grounding garments may
be used.) Suitable ground points for wrist straps must be provided.
● Where personnel are to be seated, ESD control seating is used, and personal grounding
via wrist straps is used.
● Specification of grounded ESD control materials for surfaces on which unprotected ESDS
devices could be placed, e.g. workstations and carts.
● Insulating materials are evaluated to decide whether they are essential to the process
or not.
⚬ Nonessential insulators should be kept well away from unprotected ESDS devices.
The ESD program may specify that they are excluded from the EPA.
⚬ Insulators that are essential to the process must be evaluated to determine whether
any ESD risk arises in their use. If an ESD risk is found, some means of ameliorating
it must be devised. Often this may be done by using an ionizer.
● Ionizers may be specified for control of voltages on essential insulators or nongroundable
isolated conductors
● ESD control packaging may be selected to protect ESDS devices when outside the EPA or
to prevent ESD risks arising from packaging materials used inside the EPA.
● ESD control garments may be specified to reduce the risk of electrostatic fields arising
from clothing worn by operators.
All the equipment and processes in the EPA should be designed with these require-
ments in mind. Well-designed common ESD control equipment normally fulfills these
requirements.
Other control items such as ESD control tools may also be specified. These control mea-
sures are further discussed in Section 10.5.13.4.
Sometimes ESD risks are identified that are not addressed by the usual ESD control mea-
sures and equipment. In this case, special ESD control measures and precautions must be
devised.
Operator safety should always be considered when developing an ESD program
(see Section 10.1.3).

10.4.2 Select Convenient Ways of Working


If ESD control procedures are inconvenient, it’s likely that they may not be properly and
consistently followed, especially when in a hurry or under pressure. If, on the other hand,
they provide the easiest and most convenient way of working, it is likely they will be
consistently followed.
10.5 Documentation of ESD Procedures 323

As an example, in a Stores area it may be sufficient to provide a simple single workstation


for occasional inspecting or counting incoming ESDS products. If the operator does not
need to sit and is highly mobile in their work, it may be most convenient to provide them
with ESD control footwear and a local ESD control floor mat around the workstation for
personal grounding. With no ESD control seat in use and footwear and flooring selected to
give adequate grounding, no wrist strap is required. Merely walking onto the ESD control
mat establishes the connection to ground.
If instead the operator were required to connect a wrist strap for personal grounding
as they enter and leave the workstation, they may occasionally forget in the heat of busy
activity.
In an otherwise similar facility, it could be that the operator needs to sit for their work for
longer periods at the workstation. In this case, an ESD control seat should be provided. The
ESD control floor must now withstand the repeated rolling action of the seat, and the seat
can be selected to be conveniently grounded through its wheels and the floor. The ESD floor
area must be large enough to ensure that the operator and the seat remain on the mat during
use. A wrist strap must now be provided for effective grounding of the person during seated
work, as grounding through footwear and flooring is not normally relied up for seated
operations.

10.5 Documentation of ESD Procedures


10.5.1 What Should the Documentation Cover?
For long-term effective ESD control, four aspects should be documented.

● The ESD Control Program Plan, specifying the equipment and other ESD control mea-
sures to be used.
● An ESD Training Plan, specifying necessary ESD-related training requirements.
● An ESD Control Product Qualification Plan, specifying the criteria and requirements by
which ESD control equipment and products will be selected for use in the ESD control
program.
● A Compliance Verification Plan, specifying the checks and test program to be used.

When writing the ESD control program, bear in mind that there are usually at least two
audiences who must clearly understand the documents. First, the personnel who must
work with the documents, using them to maintain and implement the ESD control pro-
gram, must understand them.
Second, auditors (first, second, or third party) may need to audit the ESD control program
and check that what is happening in practice reflects the documented ESD control require-
ments. These auditors may need also to check that the documents are compliant with the
selected standard. It can save a lot of time and effort for auditors if the documentation is
laid out in a way that quickly shows that these requirements are met. If parts of the ESD
control program are documented in separate documents or procedures or are provided in a
different form (e.g. paper or intranet-based documents), suitable cross-referencing can be
used to make this clear.
324 10 How to Develop an ESD Control Program

10.5.2 Writing an ESD Control Program Plan That Is Compliant


with a Standard
Often it is advantageous to develop an ESD control program that is compliant with an
ESD control standard such as IEC 61340-5-1 or American National Standards (ANSI)/ESD
S20.20 (EOS/ESD Association Inc 2014). While compliance with an ESD control standard
is not a necessary part of effective ESD control, it can be an effective way of demonstrating
to customers that ESD control is taken seriously and has reached a certain level of bench-
marked professionalism. A summary of the main requirements of the 61340-5-1 and S20.20
standards (which are almost identical) is given in Chapter 6. New versions of these stan-
dards will be published from time to time. These may have some differences from the cited
versions but are expected to remain broadly similar. A copy of the chosen standard should
be kept for reference.
When working to a standard, it is usually worth selecting the standard that is most likely
to be recognized by the market or geographical area in which the products will be marketed.
It is worth noting that it is easily possible to write an ESD control program that complies
with both 61340-5-1 and 20.20 standards.
One way to commence documentation that is to be compliant with a standard is to first
compile a checklist of requirements of the standard.
Using the same or similar section headings as given in the standard can give a compliant
document structure that is also easily seen by an auditor to address the requirements of
the standard. Table 10.2 gives an example of a three-level heading structure based on the
headings of IEC 61340-5-1: 2016. The headings can be modified as wished to add clarity for
the intended readers.

Table 10.2 An example of a document heading structure


derived from the headings of IEC 61340-5-1: 2016.

Introduction
Scope
Terms and definitions
Personal safety
ESD control program
● ESD control program requirements

● ESD coordinator

● Tailoring ESD control requirements

ESD training plan


ESD control product qualification
Compliance verification
ESD control program technical requirements
● ESD ground

● Personal grounding

● ESD protected areas (EPAs)

⚬ Handling ESDS device and access to the EPA


⚬ Insulators
⚬ Isolated conductors
⚬ ESD control equipment
● ESD protective packaging

● Marking of ESD related items

References

This was used to produce the example of Appendix A.


10.5 Documentation of ESD Procedures 325

Once the heading structure is in place, the text under each heading can be added accord-
ing to the requirements for ESD control in the facility and processes, in compliance with
the requirements of the standard. In some cases, it can be appropriate to include text based
on that of the standard being followed to ensure compliance with the standard. The follow-
ing sections give an example of how the basis of an ESD program based on IEC 61340-5-1:
2016 can be drafted from the heading structure. An outline of the content of each heading is
given. The requirements of 61340-5-1: 2016 are discussed in more depth in Chapter 6. Once
the document content is drafted, the headings can, if necessary, be reorganized to better
reflect the content and for clarity.
The ESD program documents should be written to conform to the organization’s internal
quality system procedures.

10.5.3 Introduction Section


An introduction section can include anything that will be useful to the reader in introducing
the document and ESD control program. A brief description of typical ESDS handled in the
facility, and the main ESD risks, can be useful.

10.5.4 Scope
The section summarizes the scope of the ESD program. This should specify the areas and
activities in which ESDS devices are handled and will be covered by the ESD program. It
should cover all applicable parts of the organization’s work. The standard lists “activities
that: manufacture, process, assemble, install, package, label, service, test, inspect, transport,
or otherwise handle” ESDS devices.
The scope should also specify the range of ESD susceptibility of the components covered
by the ESD program. The default range from the standard gives “electrical or electronic
parts, assemblies, and equipment with withstand voltages greater than or equal to 100 V
HBM, 200 V CDM, and 35 V for isolated conductors.” If the devices handled are limited
to an ESD withstand voltage range higher than this default, the default can normally be
accepted. If devices of lower ESD withstand voltage are handled, a withstand voltage range
should be specified that includes the lowest withstand voltage device to be handled. For
example, if 60 V HBM or 180 V CDM devices are handled, the ESD withstand voltage range
should be specified to include these, e.g. greater than or equal to 50 V HBM and 150 V CDM.

10.5.5 Terms and Definitions


The terms and definitions used in the adopted standard should be used. Using different
definitions for terms specified in the standards can lead to great confusion. It can also lead
to, for example, purchase of incorrectly specified equipment if the term is used in equipment
specification. Where terms are not defined in the standards, they should be clearly and
carefully defined in this section.

10.5.6 Personal Safety


It is always worth stating clearly that compliance with local safety laws and regulations is
essential. These requirements take precedence over standard ESD requirements. Specific
safety issues could be covered here.
326 10 How to Develop an ESD Control Program

10.5.7 ESD Control Program


10.5.7.1 ESD Coordinator
61340-5-1 and S20.20 require that an ESD coordinator is assigned with responsibility for the
ESD control program. 61340-5-1 lists these responsibilities as “implementing the require-
ments of this standard including establishing, documenting, maintaining, and verifying the
compliance of the program.” It could be easiest just to quote these requirements. Having an
ESD coordinator is a good way of ensuring that there is a recognized point of contact for all
personnel who need help in compliance with ESD control matters.
It’s not necessary to name the person here – doing so could just make it necessary to
update the document if they leave the role. Just make it necessary to appoint one and define
their role. The ESD Coordinator will normally have to specify ESD control equipment, pro-
cedures, and training and so should have a good understanding of ESD control practice.
An allowance should be made for specialist training and update for the role as well as time
and resources necessary for the role.

10.5.7.2 Tailoring ESD Control Requirements


An ESD program may, or may not, need to tailor the requirements of the ESD program so
that they lie outside the requirements of the standard. Either way, it is worth specifying in
the ESD Control Program Plan that the process is allowed and how it is to be done and
documented.
Tailoring of ESD control equipment requirements is also discussed in Section 10.5.13.4.

10.5.8 ESD Control Program Plan


In 61340-5-1:2016 this heading gives the requirement that an ESD control program is writ-
ten and that it must address ESD training, product qualification, compliance verification,
grounding/bonding systems, personnel grounding, EPA requirements, packaging systems,
and marking. It also notes that the plan is the principal document for implementing and ver-
ifying the program, the program must conform to internal quality system requirements and
apply to all applicable facets of the organization’s work. Providing all these are addressed
in the ESD Control Program Plan, this heading is not needed within it.

10.5.9 ESD Training Plan


The ESD Training Plan should define the personnel who need ESD training, the type of
the training they need, and when it is to be given. It should define the need for training
records, and where they are to be kept. Tests used to evaluate the effectiveness of training
should also be covered.
One way of doing this is to develop a training matrix that lists the types of personnel that
require training in the left column of a matrix table and the types of training in columns
across the page. The training appropriate to each type of trainee can then be indicated in
each cell of the table (see Table 12.1).
ESD training is discussed in more detail in Chapter 12.
10.5 Documentation of ESD Procedures 327

10.5.10 ESD Control Product Qualification


The ESD control products that are to be used within the ESD Control program must be
qualified to demonstrate that they will do the job intended, for the duration of their required
working life. The way that this is done should be documented under this heading. The
standards give requirements for some basic tests for the standard ESD control equipment.
The tests and pass criteria required will vary with the type of product being qualified
(see Chapter 7). In the simplest cases, the tests and pass criteria may replicate the measure-
ments used in compliance verification. For example, qualification of a wrist band or ground
cord would be based on a simple end-to-end resistance measurement. The same measure-
ment could be made in compliance verification, although in practice a combined system
test is used of the wrist band and cord as it is worn by the operator.
In other cases, additional tests are made in qualification that are not repeated in compli-
ance verification. For example, a workstation bench mat would be qualified by measure-
ment of the resistance from the surface to a groundable point (earth bonding point). This
measurement is related to the compliance verification system test of bench surface resis-
tance to ground. But we often also want to know what is the surface resistance characteristic
of the bench mat material, as too low resistance could give a charged device ESD risk. So, in
selecting and qualifying a bench mat, we do a point-to-point resistance measurement. We
do not normally repeat this in compliance verification tests, as material resistance changes
are adequately tested in the resistance to ground system test.
Qualification of ESD control products can be evaluated based on data sheet values (if you
trust them), in-house tests, or third-party evaluations. Taking the trouble to test things
in-house can be useful in eliminating products that do not behave as the data sheets specify
and in understanding how well (or not!) ESD control products work (Smallwood et al. 2014).

10.5.11 Compliance Verification Plan


This heading must define the routine tests that will be made to check compliance of the
ESD control products and equipment used in the ESD control program. Everything that is
specified must be tested. The test methods used, the pass criteria, and how often the tests
will be made must be documented. If there is ESD control equipment not specified by the
standard, compliance verification test methods, pass criteria, and frequency of testing must
also be defined for these. One way to conveniently specify all these things is in the form of
a table, such as the examples of Tables 10.3 and 10.4.
The standards give reference to test method standards that must usually be applied for
testing ESD control equipment. Normally these will be documented by the user as simple
test procedures that can easily be followed in compliance verification. Chapter 11 gives
examples of these test methods. Often the test procedures are in other documents – these
should be referenced in the Compliance Verification Plan.
If the user wants to use test methods that are different from those specified by the stan-
dard, it must be shown that the results of the test method used correlate with the standard
test methods. A reference should be included here to the test reports that demonstrate
correlation.
The need to keep compliance verification records must be defined.
328 10 How to Develop an ESD Control Program

Table 10.3 An example of specification of requirements for personal grounding equipment with
test method and test frequency.

ESD control item Test method Pass criteria Test frequency

Wrist strap, while worn, TM1 <35 MΩ and On each entry to EPA
resistance from body to >750 kΩ indicated by
groundable point green light on tester
ESD control footwear, while TM2 <10 MΩ and On each entry to EPA
worn, hand to metal foot plate >750 kΩ indicated by
green light on tester
ESD control footwear worn while TM3 <1 GΩ 6 monthly
standing on ESD control floor,
hand to ground.
Body voltage generated while TM4 100 V peak body 6 monthly
wearing ESD control footwear voltage
and walking on ESD control floor

Table 10.4 An example of EPA equipment pass requirements, test methods and test frequency
summarized as a table.

ESD control item Test method Pass criterion Test frequency

Workstation or rack surface TM5 Rg < 1 GΩ 6 monthly


Floor TM5 Rg < 1 GΩ 6 monthly
Seat TM5 Rg < 1 GΩ 6 monthly
Wrist strap earth bonding point (temporary) TM6 Rg < 5 MΩ Daily before use
Wrist strap earth bonding point (permanent) TM6 Rg < 5 MΩ Monthly
ESD garment TM7 Rp-p < 100 GΩ 6 monthly
ESD tool TM8 Rg < 1 GΩ 6 monthly
Ionizer decay time TM9 <5 s 6 monthly
Ionizer offset voltage TM8 ±35 V 6 monthly

10.5.12 ESD Program Technical Requirements


10.5.12.1 Documenting Technical Requirements
In the ESD program document, these requirements could be specified in any one of several
places (e.g. in this section or under product qualification and compliance verification). It
is probably advisable not to repeat the requirements in more than one place. If they are
later changed, one of the places they are specified could be missed, and this could lead to
conflicting or ambiguous requirements being specified in the document.

10.5.12.2 ESD Ground


The ESD control grounding method that is to be used (i.e. mains electricity protective
earth, functional earth, or equipotential bonding point) must be specified so that it is easily
10.5 Documentation of ESD Procedures 329

recognized by the user. All ESD control items will be connected to a common ground point
using this method. Grounding should comply with requirements of the National Electrical
Code. Any requirements of this that apply to the ESD ground can be specified or referenced
under this heading, if appropriate.
It is worth specifying that all conducting items and materials that contact ESDS devices
must be connected to this common ground point if possible. Any conductor that contacts
ESDS devices but for some reason cannot be electrically connected to ground must be
treated as an isolated conductor.
It is not good practice to have two different ground points in an EPA that are not connected
together as they could be at different voltages. It follows that if mains electricity protective
earth is present in the EPA, this should be used as, or connected to, ESD ground if possible.

10.5.12.3 Personal Grounding


Personal Grounding Equipment
Personal grounding is one of the key control measures of an ESD control program.
61340-5-1:2016 gives some basic requirements that are worth stating in the ESD program
document.
● All personnel must be grounded according to the requirements of the ESD program when
handling ESDS devices.
● Seated personnel must be grounded via a wrist strap.
● Standing personnel can be grounded via a wrist strap or via footwear and flooring.
● When grounding via footwear and flooring is used, ESD control footwear must be worn
on both feet. The resistance from the person’s body to ground through the system of per-
son, footwear, and flooring must be less than 1 GΩ and the maximum voltage generated
on the body (measured using a walk test) must be less than 100 V (for an ESD program
protecting devices down to 100 V HBM).
The main requirement is that personnel handling ESDS devices should be grounded.
Whether wrist straps or footwear-flooring grounding, or both, are used is optional and
should be defined by the ESD control program. An exception is the case of seated personnel
handling ESDS device, where current ESD standards require that a wrist strap must be used.
Note that ESD seating is not commonly specified in such a way that reliable grounding of
personnel can be achieved and should not normally be considered a method of personnel
grounding (See Section 7.9.8).
Other methods of grounding have been used in some ESD control programs, for example
grounding the person via specially designed groundable ESD control garments or seats.
Use of nonstandard grounding methods such as this under a current standard ESD control
program would require tailoring, with documentation of the rationale for the methods and
evidence that it achieves the required personal grounding performance.
It is necessary to specify the limits of the measurable parameters for grounding equipment
used in the ESD program. 61340-5-1:2016 gives a table (table 2 in the standard) that specifies
these for product qualification and compliance verification.
It can be helpful to specify the requirements, test methods, and frequency of testing in
a single table such as the example of Table 10.3. The test methods TM1–TM4 must be
written up and referenced. The pass criteria and test frequency should be specified after
330 10 How to Develop an ESD Control Program

consideration of the standard requirements and appropriateness of the facility and process.
In this example, the resistance of footwear, while worn by personnel, measured between
the hand and a metal plate under the foot, is specified to be <10 MΩ, which is much less
than the requirements of the standard. As it is within the requirements of the standard, no
tailoring explanation or justification is required.
As a key ESD control measure, it is wise to test personal grounding (wrist straps or
footwear) before use on each day of use. Some organizations require personnel to test these
at the beginning of each work shift or on each entry into an EPA. It is good practice to keep
records, such as a simple log (e.g. signed list) of this test. A failed wrist strap should always
be removed from use and repaired or replaced.
Where wrist straps are used, wrist strap grounding points should be regularly checked
for connection to ground. Where these are of a temporary nature (e.g. via connection to a
ground plug in a mains socket), tests should be more frequent due to the risk that accidental
disconnection of the ground path may occur. In the standard, testing of wrist strap bonding
points is addressed under “EPA Equipment.”

Occasional Verification of ESD Control Footwear Worn on ESD Control Flooring


The resistance from the operator’s body to ground through footwear and flooring and the
body voltage generated while walking are dependent on the type of footwear and flooring
used. These parameters are primarily confirmed during product qualification. Thereafter,
only the combinations of footwear and flooring checked in product qualification should be
used in practice. If a new combination is used, it should be qualified before use.
Nevertheless, these parameters can be affected by floor or footwear contamination,
humidity, and other factors. So, it is wise to check them occasionally to make sure nothing
has changed to compromise their effectiveness.

Personal Grounding Testers


Perhaps an obvious point, although often missed, is that personal grounding (wrist strap or
footwear) testers must be specified to test according to the parameter limits specified in the
ESD Control Program Plan. In practice, it may be easiest to select suitable personal ground-
ing testers and then specify the parameter limits that they use, in the ESD Control Program
Plan. Many personal grounding testers use upper parameter limits specified in the stan-
dards. Most also specify minimum limits, but some may not do so. Minimum limits are
not specified in the standards but are user defined, usually based on a minimum resistance
from body to ground for reduction of electrical shock risk if a live conductor is accidentally
touched. Minimum values selected are often around 750 kΩ for up to 250 VAC and 500 VDC
power systems.

10.5.13 ESD Protected Areas


10.5.13.1 Handling ESDS Devices and Access to the EPA
First, make it clear that handling unprotected ESDS devices must be done only in an EPA.
This applies equally to ESDS devices that are protected by ESD protective packaging or
those that would normally be enclosed e.g. within equipment. It is worth stating explicitly
that an ESDS device must not be taken out of ESD protective packaging unless it is within
an EPA.
10.5 Documentation of ESD Procedures 331

The boundaries of the EPA must be clearly identified and noticeable to personnel entering
or leaving the EPA. The way that the organization does this should be defined and described
here. Examples of typical signage should be shown.
The 61340-5-1: 2016 requires that “Access to the EPA shall be limited to personnel who
have completed ESD training. Untrained individuals shall be escorted by trained personnel
while in an EPA.” A clear statement like this should be included here, or at some other
appropriate point in the ESD Control Program Plan.

10.5.13.2 Insulators
This section should be used to explain and define identification and treatment of essential
and nonessential insulators used in the EPA. A similar approach can be taken to dealing
with other known electrostatic field sources.
Essential insulators are those that are required to be present as part of the product
or process, without which the process cannot be completed. Nonessential insulators
are those without which the process can be completed. The most common source of
high electrostatics fields in the workplace include insulators used in plastic (non-ESD
protective) packaging, documentation (file dividers and covers), equipment housings and
exposed construction, and other components used in the product. While electrostatic fields
do not in general cause direct damage to ESDS devices, they do lead to induced voltages on
ungrounded conductors and lead to the conditions in which ESD can arise if the conductors
are, or touch, ESDS devices. Contact between the ESDS device and a metal item while
within an electrostatic field can give a charged device ESD. It may be worth explaining this
risk in this section or somewhere else in the ESD control program such as the introduction.
There are essentially three ways of dealing with nonessential insulators in an EPA. The
organization should decide which approach they are going to take in each case and write
the ESD control program accordingly.
The first approach is to replace the insulator with a noninsulating (conducting) material
and ground it.
The second approach is to remove or eliminate the nonessential insulator from the EPA.
This approach is simple and clear and often easy to manage and train personnel to do. In
many cases, ESD control materials (static dissipative or conductive) can replace insulators
for use in the EPA. In some facilities and processes, however, it could give great inconve-
nience for various reasons. Inconvenience usually leads to difficulty in compliance.
A third approach is to keep nonessential insulators a minimum distance from worksta-
tions where ESDS devices are handled. This approach can sometimes be used in convenient
practices. (An example might be to have sign-off papers or computer equipment at a work-
station next to, but sufficiently far from, the workstation area at which work is done on the
ESDS device.) The disadvantage can be that effective procedures can be less clear to EPA
personnel and may require greater training for reliable implementation. Insulators or ESDS
can easily be transported from a position where they are allowed to a position where they
might cause ESD risk by personnel who are unwary or lack sufficient understanding of the
ESD control procedures.
The requirements for electrostatic field or voltage limits should be stated in the ESD
Control Program Plan. These limits can be taken from the standard, or other limits can be
defined. If the electrostatic field and voltage limits defined in the ESD Control Program Plan
332 10 How to Develop an ESD Control Program

are outside the limits given in the standard, this requires justification and documentation
as tailoring. This section should also define the actions to be taken if these limits are found
to be exceeded.
Papers and cardboard are highly variable materials that can vary by orders of magni-
tude in their electrical characteristics between grades and with humidity changes (and the
weather!). So, it is considered good practice to keep these materials away from ESDS devices.
61340-5-1: 2016 requires that “All non-essential insulators and items (plastics and paper),
such as coffee cups, food wrappers, and personal items shall be removed from the worksta-
tion or any operation where unprotected ESDS are handled.” A clear statement such as this
should be included in this section of the ESD Control Program Plan.
Pink polythene material used for ESD control relies on the presence of humidity for its
low-charging properties. It is often used within an EPA as a replacement for polythene
packaging e.g. for enclosing non-ESDS device items or documents. This material becomes
ineffective and can act like an insulator at low ambient humidity (<30% rh), charging highly.
There are other common items that could occur in practice and must be controlled
because they generate electrostatic fields, such as cathode ray tube (CRT) displays. These
should be identified, and how they will be treated should be defined.
Essential insulators cannot, by definition, be removed from the process in which they are
used. A strategy for dealing with these is shown in Figure 10.1. The first task is to determine
whether the insulator provides a significant ESD risk to the ESDS device in the process. This
can be evaluated by determining whether it can charge up and give a significant electrostatic
field at the possible position of the ESDS device. If not, it is necessary only to document the
evaluation.

Identify essential insulator

No
Could it give electrostatic fields Document in the ESD
at the site of the ESDS device? Program Plan
Controlled Risk

No
Yes Can it be replaced by a
grounded conductor?

No
Yes Can it be moved away from the
ESDS device?
Use an ioniser
or other means of risk
Replace with reduction
grounded conductor Yes
Controlled Risk
Move it away
Document in the ESD
Program Plan
Controlled Risk

Figure 10.1 A strategy for dealing with essential insulators.


10.5 Documentation of ESD Procedures 333

If the essential insulator can provide significant electrostatic fields at the position of the
ESDS device, it remains to decide what to do with it. Sometimes they can be replaced with
static dissipative materials and grounded. An example might be part of a test or assembly
jig that does not need to be made of insulating material. Sometimes, although the insulator
must remain nearby, it can be moved sufficiently far away from the position where ESDS
devices are handled to reduce the field at the position of the ESDS devices to a low level.
If neither of these techniques can be used, an ionizer or some other means of reducing the
electrostatic field must be found. Whichever technique is used should be documented in
the ESD Control Program Plan.

10.5.13.3 Isolated Conductors


The 61340-5-1:2016 standards states that “All items that come into contact with ESDS device
and are capable of conducting electricity shall be connected to ground or electrically bonded
in order to eliminate differences in potential” (Section 5.3.2 of the standard).
The standard also says that “if a conductor that comes into contact with an ESDS
device item cannot be grounded or equipotential bonded together, then the process shall
ensure that the difference in potential between the conductor and the contact of the ESDS
device item shall be less than 35 V.” It is worth including this statement, or an equivalent,
in the ESD control program.
In practice, it is easy to say and difficult to do this. If at all possible, all conductors that
contact the ESDS device should be grounded per Section 5.3.2 of the standard. One way
of bringing the ESDS device and an isolated (ungrounded) conductor to nearly the same
voltage is to use an ionizer to neutralize the charge on them to the same level. It is, however,
difficult to measure and verify that this control is working as intended, especially for small
items and automated processes.
The standard is not clear as to what should be considered an isolated conductor. Perhaps
a good, although deliberately vague, definition of an isolated conductor is “a conductor
on which electrostatic charge can build up, during normal operations, to a level where it
could cause ESD risk on contact with ESDS.” This definition implies that some evalua-
tion is required to know whether the charge can build up on the conductor during normal
operations or not.
Perhaps the easiest way to specify this would be to specify a maximum resistance from
the conductor to ground. If the resistance from the conductor to ground is greater than
the limit, the conductor is considered isolated. While the standard does not specify such
a limit, from other specifications in the standard, resistance to ground below 1 GΩ would
usually be considered adequately grounded. For small items and where charge generation
or voltage changes due to capacitance changes are not an issue, up to 100 GΩ might be
acceptable resistance to ground. Above 100 GΩ, resistance to ground will almost certainly
be considered isolated unless it can be shown that the voltage on the item is reliable kept at
zero during operation.
If an isolated conductor is found necessary, as it cannot easily be grounded and charge is
found to build up on it during normal operation, it is likely to be necessary to specify control
measures on a case-by-case basis. A strategy for dealing with ungrounded conductors is
given in Figure 10.2. If the ungrounded conductor cannot touch an ESDS device, then there
is probably no ESD risk. In this case, it is enough to document this evaluation.
334 10 How to Develop an ESD Control Program

Identify isolated (ungrounded)


conductor

Can it be grounded?
No

Yes Could it touch an ESDS?


No

Ground it
Controlled risk
Yes
Use an ioniser

Controlled risk Controlled risk

Document in the ESD


Program Plan

Figure 10.2 A strategy for dealing with isolated (ungrounded) conductors.

If the isolated conductor could touch an ESDS device, one way to reduce the voltage
on it is to use an ionizer. If an ESDS device and isolated conductor spend sufficient time
close together in the ion stream from an ionizer, they are both likely to reach a similar
voltage near the offset voltage of the ionizer. The time taken to achieve this will depend
on the ionizer and process circumstances. To show that this has been successful, it would
be necessary to measure the voltages on the ESDS device and the conductor with an elec-
trostatic voltmeter and compare them. Measurement of these voltages would require an
electrostatic voltmeter capable of measuring voltage on small items. This is not an easy
task, especially in a fast-moving automated process.

10.5.13.4 ESD Control Equipment


How to Specify ESD Control Equipment
The ESD control equipment and materials used in EPAs must be specified both for qualifi-
cation and compliance verification. This means that test methods and pass criteria must
be specified. The 61340-5-1:2016 standard specifies basic limits for work surfaces, stor-
age racks and carts, wrist strap bonding points, flooring, ionizers, seats, and static control
garments (see Chapter 6). It is not essential that all the items listed in the standard are used
in an EPA, but any equipment used should be specified to comply with the requirements
of the standards. For example, an ESD control program may not specify ESD control gar-
ments. Or, if personnel handling ESDS devices are always grounded via wrist straps, it may
not be necessary to specify ESD control footwear. However, if these items are specified, they
should be specified according to the standard used or covered by tailoring of the standard
requirements.
10.5 Documentation of ESD Procedures 335

Requirements for other items not specified in the standard may also need to be specified,
e.g. gloves, finger cots, tools, or soldering irons. Test methods and pass criteria will also need
to be specified for these items.
Sometimes it is desirable to specify a limit that is different from that given in the standard.
If this limit is within the range given by the standard, it does not represent tailoring of the
ESD program. For example, the organization may have selected a floor material that when
installed has a resistance to ground below 10 MΩ. In this case, 10 MΩ could be specified as
the upper limit for compliance verification to pick up changes in material performance or
due to surface contamination.
In contrast, if the proposed limit is outside the range given in the standard, this should be
considered a tailored requirement. As an example, many organizations that established an
ESD control program under earlier versions of 61340-5-1 would have accepted seats with
resistance to ground up to the previously specified limit of 10 GΩ. In the 2016 version, the
limit for seats was dropped to 1 GΩ. For compliance with 61340-5-1:2016, the organiza-
tion would have the choice of replacing all the seats according to the new requirements or
including a tailored limit of 10 GΩ in their ESD control program. If they choose the tailored
option, the decision should be supported by some tests demonstrating that no additional
risk is provided by this specification.
Don’t use words like conductive, antistatic, or dissipative to specify equipment unless they
are defined for that product in a standard that is referenced as part of the specification.
Contrary to popular usage, these words do not have generally standardized meanings for
many products. Even worse, they can mean different things for the same product in different
industry areas and for different products. So, unsuitable products may be purchased if using
these words as the specification.
Instead, all products are best specified using verifiable parameters (e.g. resistance), mea-
sured using standard test methods. It is important to specify the test method used, because
different test methods may give different measurement results. Common measurable
parameters include resistance to ground or to a groundable point for many types of ESD
control equipment, resistance point to point on a surface, material or garment, surface or
volume resistance for packaging, and decay time and offset voltage for ionizers.
An example of ESD control equipment summarized with test method, pass criteria, and
test frequency is given in Table 10.4. The pass criteria and test frequency should be specified
after consideration of the standard requirements and the needs of the facility and processes
present. The test methods TM5–TM9 must be written up and referenced. These should be
based on those given in IEC 61340-5-4 (International Electrotechnical Commission 2019)
or ESD TR53-01-15 (EOS/ESD Association Inc 2015; see Chapter 11).

Bench Mats and Other Surfaces on Which ESDS Devices Are Placed
Any surface on which unprotected ESDS device could be placed, for example storage racks
and cart shelves, should be subject to the same requirements as work surfaces. If ESD con-
trol packaging materials are used in this way, they should also be specified as per work
surfaces.
Compliance verification of these surfaces is normally specified and measured as resis-
tance between the surface and ESD ground (Rg ) (see Section 11.8.1).
336 10 How to Develop an ESD Control Program

ESD Control Flooring


ESD control flooring can be a useful way of establishing an EPA containing multiple work-
stations. Without the EPA floor, each workstation is effectively a stand-alone EPA with
UPAs in between. ESD risks must be considered when transporting ESDS devices through
these UPAs in between workstations, and ESD packaging is likely to be required. Using ESD
flooring, the necessity of this packaging can be eliminated.
The function of ESD flooring is to provide convenient grounding of personnel wearing
ESD footwear or using seats, carts, racks, and any other items standing on the floor.
The resistance of ESD control flooring should be chosen according to its purpose. The
standards specify a maximum resistance to ground of 1000 MΩ. Organizations often have,
or source, a floor with much lower resistance than this limit. A lower-resistance floor is
more likely to give reliable low body voltage on personnel using suitable footwear or may
be required to help give low system resistance to ground for some items.
If a floor has a considerably lower resistance to ground than required by the standard, it
may be desirable to set the upper limit for flooring resistance to ground in the ESD Con-
trol Program Plan at a level just above the maximum expected in practice for the installed
floor. Compliance verification measurements will then flag up rising floor resistance due
to contamination or other factors.
Compliance verification of floors is normally specified and measured as resistance
between the floor surface and ESD ground (Rg ) (see Section 11.8.1).

Seating
If seating is needed at a workstation where ESDS devices are handled, then it must be
specified for ESD control according to the standard requirements. Seats are normally most
conveniently grounded through an EPA floor, and so an ESD control floor must usually also
be present. In the absence of an ESD control floor, it is sometimes possible to ground the
seat through a ground cord, but this can be inconvenient and may easily become discon-
nected. The requirement for seating should be thoroughly investigated before adding to the
ESD Control Program Plan.
Compliance verification of seats is normally specified and measured as resistance
between the seat surface and ESD ground (Rg ) (see Section 11.8.1).

Tools
When handling very sensitive ESDS devices, any tools that could contact ESDS devices
(e.g. cutters or pliers) could provide a risk of ESD damage. This can be addressed by spec-
ifying ESD tools in which the parts that contact ESDS devices are grounded through the
user’s body.
Compliance verification of tools can be specified and measured in various ways. There
is no standard test specified in current standards. One method is to specify the resistance
between the tool contact tip and ESD ground (Rg ) when the tool is held by the operator
(see Section 11.9.5).
Another way is to specify the maximum charge decay time measured using a charged
plate monitor (CPM) (see Section 11.9.8.1).
10.5 Documentation of ESD Procedures 337

Gloves and Finger Cots


Gloves and finger cots may need to be specified for ESD control, especially when used to
handle ESDS devices or ESD control tools. Compliance verification of gloves and finger
cots can be specified and measured in various ways. There is no standard test for these
items specified in current standards, although annex A of S20.20-2014 points to ANSI/ESD
SP15.1 for testing gloves and finger cots.
One method is to specify the resistance between the glove contact area and ESD ground
(Rg ) when the glove is worn by the operator (see Section 11.9.7). Another way is to specify
the maximum charge decay time measured using a CPM (see Section 11.9.8.2). The latter
method can also indicate whether the CPM plate is significantly charged by contact with
the glove material.

Ionizers
Ionizers may be needed if there are essential insulators or isolated conductors present that
evaluation has shown can charge up and cause ESD risk (see Section 4.6).
Ionizers are specified for compliance verification in terms of charge decay time and offset
voltage, measured using a CPM (see Section 11.8.8).

ESD Control Garments (Coats)


It is not always necessary to use ESD control garments in every ESD control program. Those
that choose to use ESD garments are often those that have heightened concern over ESD
damage, for example those involved in handling highly sensitive ESDS devics or producing
items for a high reliability market. Some ESD coordinators choose to use ESD control gar-
ments to help give a favorable impression to visitors regarding the care taken in the facility,
and wearing a special garment in the EPA can help reinforce EPA discipline.
If an ESD control garment is required by the ESD program, it should be selected for com-
pliance with the standard. These are usually specified in terms of point-to-point resistance
(Rp-p ) of the garment material (see Section 11.8.2.2) or the resistance to a groundable point
(Rgp ) for a groundable static control garment (see Chapter 6). User-defined test methods
and pass criteria can also be used.
Some types of ESD control garments are available that do not easily pass the standard tests
because they are not made from resistive materials. If these are to be used in a standard (IEC
61340-5-1 or ANSI/ESD S20.20) ESD control program, then tailoring will need to be used
to give the rationale and justification for their use (see Section 6.5.6). Suitable test methods
and pass criterial will need to be defined to qualify and verify these garments.

10.5.14 ESD Protective Packaging


The 61340-5-1:2016 standard states first and foremost that “ESD protective packaging and
package marking shall be in accordance with customer contracts, purchase orders, drawing,
or other documentation.” For organizations that use packaging agreed with customers, a
statement like this should be included in the ESD control program.
Where customers or other documentation does not define ESD protective packaging, ESD
protective packaging requirements must be defined in the ESD Control Program Plan for
338 10 How to Develop an ESD Control Program

wherever packaging is needed, according to IEC 61340-5-3. In situations where packaging


is not needed, it need not be specified.
For otherwise unprotected ESDS devices taken outside the EPA, including packaging
used for sending ESDS devices to the customer, ESD protective packaging must be defined
that gives adequate protection against the ESD risks likely to be met in an unprotected area.
Packaging used within the EPA must also be specified according to IEC 61340-5-3.
ESD protective packaging is normally specified in terms of the surface resistance (Rs ; see
Section 11.8.4), the volume resistance (Rv ; see Section 11.8.5), and for ESD shielding bags,
the energy determined in an ESD shielding test. The latter is not normally a test that many
users will do themselves due to its complexity. For packaging systems other than bags, ESD
shielding capability is evaluated as the presence of a barrier against ESD current being con-
ducted to the inside of the package (see Section 11.8.6).

10.5.15 Marking of ESD-Related Items


The 61340-5-1:2016 standard states that “ESDS, system or packaging marking shall be in
accordance with customer contracts, purchase orders, drawing, or other documentation.”
For organizations that have marking of ESD control–related items agreed with customers,
a statement like this should be included in the ESD control program.
Organizations that do not have marking of ESD control–related items agreed with cus-
tomers should also consider whether marking of items for ESD control purposes would be
of benefit or is in practice happening in their ESD control program. If there is a need or
if markings are used in practice, then this should be documented. Common examples are
marking of packaging and ESD control equipment. Some organizations also decide to mark
documents related to ESDS devices (e.g. data sheets or component lists) or even the ESDS
devices themselves (e.g. PCBs) to assist identification.

10.5.16 References
It is normally useful to list the references used in preparing the document, including the
standards used and their user guides, and any in-house procedures, reports, test procedures,
or other documents specified.

10.6 Evaluating ESD Protection Needs


Standard ESD control precautions applying ESD control equipment by rote address most
common and well-known ESD risks, but they do not necessarily address all ESD risks. Every
step of processes where ESDS devices are handled should be scrutinized for possible ESD
risks. This is discussed in Chapter 9.

10.7 Optimizing the ESD Control Program


10.7.1 Costs and Benefits of ESD Control
Modern ESD standards allow the flexibility to optimize an ESD program for maximum
effectiveness while providing only the equipment and control practices necessary for ESD
10.7 Optimizing the ESD Control Program 339

protection. It is possible to use nonstandard or tailored ESD control methods in EPAs where
ESDS devices are handled by less common processes. The choice of ESD control mea-
sures used governs the cost of ESD control measures used in the EPAs and ESD protective
packaging used outside those areas. Documentation, training, and compliance verification
programs are also essential parts of ESD control. Trade-offs can be made between these
costs in order to optimize the ESD control program, although a high level of expertise may
be needed to do this successfully (Smallwood et al. 2014).
The first and most obvious benefit of an ESD control program is to protect ESDS devices
during processes and maintain the level of ESD damage at an acceptably low level. There
are also other benefits that could be part of the rationale for ESD control. Maintenance of an
ESD control program to a recognized international standard could be an important part of
the organization’s quality program. It may even be essential to prequalification of the orga-
nization as a provider of product to some customers. If a customer audits the organization,
a favorable impression of ESD control (whether the customer has a good understanding of
the topic or not) can be a vital step in the path to obtaining an order. An adverse impres-
sion and disagreements over the adequacy of ESD control can provide a significant block
to customer acceptance.
The main cost of inadequate ESD control is often thought to be damage to ESDS devices
handled in the organization’s processes. This cost, for many organizations, is far from clear
or quantifiable as failure analysis often is not done to the level that ESD failures are detected.
Other costs can also be significant but are often not considered. These can include prod-
uct failures, unreliability or drift in characteristics, additional test and rework expenditure,
delays to shipments, or need to overstock frequently failing items. Some of the most expen-
sive costs can be those related to customer service – cost of reaction to customer complaints
and failures at the customer’s site, impaired product or company reputation, and a possible
result in reduction of sales.
The real cost/benefit balance of an ESD control program depends on the time and
resources spent on all the aspects of ESD control. Optimizing the ESD control program
should consider all the desired benefits and areas of cost and determine an appropriate
balance between them. This means that trade-offs between investment in the different
aspects are possible and probably necessary.

10.7.2 Strategies for Optimization


10.7.2.1 Minimization of the EPA
A major contribution to ESD protection may be made by minimizing handling of ESDS
devices in an unprotected state. As well as minimizing the opportunity for ESD to occur,
it has other advantages. For example, a smaller EPA usually requires less equipment. This
also leads to reduced overhead of compliance verification testing overhead.

10.7.2.2 Choice of EPA Boundary


Including too many parts of a process in an EPA can result in processes that are incompati-
ble with ESD control being included in the EPA. This can then make it difficult or impossible
to comply with the ESD control program requirements.
Nevertheless, EPAs must be defined for all activities in which unprotected ESDS devices
are handled.
340 10 How to Develop an ESD Control Program

So, it is worth carefully considering where the EPA boundary should be placed in order to
enclose all necessary activities while minimizing the EPA size and excluding unnecessary
and incompatible activities.

10.7.2.3 Minimizing Variation of the ESD Control Program


Variations are often specified in ESD control measures to minimize ESD control equipment
required in some areas. It is perfectly possible to specify considerable variation even in an
ESD control program compliant with a standard.
It can, however, be advantageous to minimize the variation for various reasons. The
impact on documentation, training, and compliance verification should be considered.
Moreover, customers viewing an ESD program are inclined to evaluate them against their
own idea of what a good ESD control program should look like. If they see that elements
are missing or different, they may jump to a conclusion that ESD control is insufficient.
Disputes or unwillingness to place work with the organization can result.
Having variations on ESD control measures used in different EPAs can result in an
increased need for ESD training, as personnel working in each area must be trained in the
specific requirements for each area. Personnel who work in several different areas may get
confused between differing requirements used in each area, and this can lead to unintended
noncompliances.
Variation of required ESD control equipment can reduce compliance verification over-
heads if the amount of equipment needing to be verified can be reduced in some areas.
If, however, the same equipment is specified in different areas but has different pass cri-
teria, confusion can arise with equipment meeting the wrong criteria being introduced into
an area. It can be difficult to notice this type of problem until picked up through subsequent
compliance verification tests.
It is inadvisable to expect workers to make decisions or change ESD control behavior for
different ESDS devices, e.g. of different sensitivity, or for non-ESD-sensitive items. Confu-
sion, wrong decisions and poor compliance habits may result. It is probably better in most
cases to design and implement an ESD program according to the requirements for handling
the most sensitive ESDS devices and apply this generally to simplify training and compli-
ance verification and reduce likely noncompliances.

10.7.2.4 Standard or Tailored ESD Control


Standard ESD control measures provide a generally recognized set of controls that are
needed and control the most common ESD risks in most cases. It is relatively easy to
implement an ESD control program with standard ESD controls with relatively little
expertise.
With additional effort and evaluation, it may be possible to devise more cost-efficient,
tailored, nonstandard controls or to leave out some controls found unnecessary. This
requires greater expertise for successful evaluation and specification of effective tailored
controls. With inadequate expertise, there is a risk that the specified ESD controls may be
inadequate.
This approach is also likely to increase the overheads of documentation and training and
the variation of ESD controls specified in different areas. Visiting customers may be more
inclined to jump to conclusions that ESD controls are inadequate.
10.8 Considerations for Specific Areas of the Facility 341

10.7.2.5 Design for Convenience


If ESD control measures are designed to be convenient, it is likely they will be followed.
If inconvenient, it is more likely they will not, and noncompliances will occur.
Design for convenience may not give the lowest equipment, documentation, training, or
compliance verification costs, but it may give a benefit in lower noncompliance levels and
associated overheads.

10.7.2.6 Who Audits the ESD Control Program?


It is possible to delegate the task of auditing the ESD program to any personnel with accept-
able expertise. Some organizations use internal personnel, whereas others delegate the task
to second- or third-party auditors or even to suppliers of ESD control equipment.
Sometimes delegation of audit can result in a conflict of interest. For example, using a
supplier to audit may result in a conflict of interest with their desire to sell their ESD control
products.
Against this, using internal auditors may result in training and compliance verification
requirements that are difficult for a small organization to resource. Under this pressure,
compliance verification may become under resourced and neglected or ineffective.

10.8 Considerations for Specific Areas of the Facility

10.8.1 The Varying ESD Control Requirements of Different Areas


Many facilities include typical stages such as Goods In, Storage, Kitting, Manufacturing,
Test, and Dispatch. Some organizations also have areas such as Research & Development.
Some of these areas (e.g. Manufacturing) are easily controlled with formal EPAs. Other
activities (e.g. Kitting) may not obviously need an EPA but nevertheless have a large effect
on compliance within the EPAs they work with. Still others (e.g. Goods In, Stores, Dispatch)
may have aspects requiring both EPA and unprotected areas working together. Research &
Development areas also often have mixed requirements for ESD control and unprotected
areas and may be difficult to organize as formal EPAs for various reasons, including accep-
tance of ESD control measures by the personnel working in the EPAs.

10.8.2 Goods In and Stores


An essential part of the Goods In function is to remove any packaging that is provided
for protection during transport rather than ESD control (“secondary packaging”). Removal
of secondary packaging should stop as soon as ESD protective packaging is encountered
and before any ESDS devices are exposed, allowing ESD protective packaging to remain
unopened. In ESD protective packaging (e.g. PCBs in bags), a symbol should be evident
on the packaging to identify its ESD protective function (see Section 8.10). This packaging
should not be opened outside an EPA. On the other hand, secondary packaging should not
be taken into an EPA.
ESDS device items that are protected against ESD damage (e.g. enclosed within equip-
ment or ESD protective packaging) can be stored under ordinary unprotected conditions.
Items that have exposed ESDS devices should be stored under EPA conditions.
342 10 How to Develop an ESD Control Program

If inspection or test of ESDS devices is required, an EPA should be provided for the
purpose. It is preferable not to use this for any non-EPA activity, as this can be a means of
introducing noncompliant materials and lax EPA procedures. An EPA bench used for other
purposes should be checked for compliance and recommissioned before use as an EPA.
Goods In and Stores areas do not necessarily need an EPA. They will need one only if
ESD protective packaging must be opened and the ESDS device handled (e.g. for counting
or inspection) in an unprotected state. If possible, it can be preferable to avoid having this
sort of activity and therefore avoid the necessity of having an EPA. If necessary, the EPA can
be as simple as a single carefully specified workstation.
There is a considerable risk that EPA workstations in Goods In and Stores areas become
contaminated with secondary packaging and other insulating materials. To minimize this
risk, EPA workstations should not be used for non-EPA activities. Personnel using these
facilities may need to be carefully trained in avoiding non-compliances and ESD risks.
Siting of EPA workstations should be carefully chosen to be well away from areas where
high-level static generators such as secondary packaging are present. As an example, an
EPA workstation that backs onto a storage rack in an uncontrolled area could be subject to
considerable ESD risk from the static generators stored on the rack.

10.8.3 Kitting
Kitting activities may or may not require use of an EPA for handling unprotected ESDS
devices. Nevertheless, any kit destined for use in an EPA will be a source of noncompli-
ant packaging or other materials and items unless carefully specified to avoid this. Tote
boxes holding kits for transport into an EPA must be specified as compliant with EPA entry.
Items included within the kit, including mechanical, consumable, or non-ESDS device
components, must be supplied in suitable ESD protective packaging materials. Insulating
materials and items must be minimized or eliminated. If not eliminated, how they are used
and where they occur in the EPA must be carefully controlled to avoid ESD risks to ESDS
devices. If noncompliant items are included in a kit, they will cause noncompliance in the
EPA to which they are delivered.

10.8.4 Dispatch
Dispatch is another area that may or may not need EPAs as well as unprotected areas.
Dispatch areas commonly have large amounts of secondary packaging materials including
insulating packing tapes and other high-level static generators. Many of the considerations
for these areas are like Stores and Goods In areas.
Another consideration is that high levels of paper or cardboard dust, fibers, and larger
particles can be generated during handling of packaging materials. These can cause con-
tamination problems for some types of products handled in the vicinity.

10.8.5 Test
Test is probably the area most likely to require some tailoring of ESD control measures
due to safety considerations if high voltages and manual handling are present. This
often means that personal grounding requirements must be carefully thought through to
maximize safety and minimize ESD risk.
References 343

10.8.6 Research & Development


Research & Development areas are perhaps notorious for contention between the desire
for ESD control and the easy freedom of engineers to move between prototyping and office
desk or computer simulation areas. Some organizations take the position that prototypes
that are not shipped to customers are not required to be handled under ESD control con-
ditions. This, however, neglects the fact that damage to ESDS devices by ESD can lead to
considerable wasted time in debugging and reworking circuitry, and ESDS devices that may
have their properties changed by ESD could result in a prototype working differently from
a final produced design. Moreover, for some organizations, each product is effectively a
one-off product produced by a prototyping process.
It is often possible to successfully specify ESD control for Research & Development areas
that will address these problems. The ESD control design of these areas will often be very
different from a production facility. The fact that highly technically competent engineers are
using the facility can be an advantage providing they “buy in” to the need for ESD control
and it is designed to be convenient for them to use. If not, it will probably be difficult to get
them to comply with the ESD control requirements.

10.9 Update and Improvement


ESD control programs are often mistakenly treated as a process that, once specified, can
be left without change for an extended period. In practice, here are several reasons why an
ESD program should be regularly reviewed.
First, auditing reveals deviation patterns that can often be addressed by updating the ESD
control program to be easier to use. Snow and Dangelmeyer (1994) gave an example where
foot straps were often incorrectly worn. Replacing them with ESD control shoes eliminated
an important source of noncompliances. Sometimes, ESDS device failures reveal weak ESD
control that must be addressed by changes in the ESD control program.
Second, processes and facilities may change over time, and new or modified ESD control
measures may need to be implemented for them.
Third, the types of ESDS devices often may change and may require modification to the
ESD control measures used. Introduction of new, more sensitive ESDS devices may intro-
duce special challenges.
Finally, update of ESD control standards may result in necessary changes to the ESD
control program to maintain compliance with the standard.
For all these reasons, it is advisable to build in a program of regular update and continual
improvement.

References

Dalziel (1972). Electrical shock hazard. IEEE Spectrum: 41–50.


Dangelmeyer, G.T. (1990). ESD Program Management. Van Nostrand Reinhold.
EOS/ESD Association Inc. (2014). ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
344 10 How to Develop an ESD Control Program

EOS/ESD Association Inc. (2015). ESD TR53-01-15. Technical Report for the Protection of
Electrostatic Discharge Susceptible Items – Compliance Verification of ESD Protective
Equipment and Materials. Rome, NY, EOS/ESD Association Inc.
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016.
Electrostatics – Part 5-1: Protection of electronic devices from electrostatic phenomena - General
requirements. Geneva, IEC.
International Electrotechnical Commission. (2019) IEC TR 61340-5-4:2019. Electrostatics -
Part 5-4: Protection of electronic devices from electrostatic phenomena – Compliance
Verification. Geneva, IEC.
Nave, C.R. and Nave, B.C. (1985). Physics for the Health Sciences, 3e. W B Saunders.
ISBN: 0 7216 1309 8.
Smallwood, J., Taminnen, P., and Viheriaekoski, T. (2014). Paper 1B.1. Optimizing investment
in ESD control. In: Proc. EOS/ESD Symp EOS-36. Rome, NY: EOS/ESD Association Inc.
Snow, L. and Dangelmeyer, G.T. (1994). A successful ESD training program. In: Proc. EOS/ESD
Symp. EOS-16. Rome, NY: EOS/ESD Association Inc. pp. 94-–94-12.
345

11

ESD Measurements

11.1 Introduction

All materials and equipment used for electrostatic discharge (ESD) control are designed
and made to have certain properties that enable them to function. These commonly are
based on the principles of replacing exposed insulating materials with noninsulating mate-
rials and establishing a ground path for any charge generated on the equipment or material
to dissipate to electrostatic discharge protected area (EPA) ground. In addition, charge on
essential insulators may need to be neutralized using an ionizer or ESD risks controlled by
other means.
Many of the facilities equipment and materials needed for use in ESD prevention are
specified in various standards worldwide. These standards establish performance criteria
for compliance, and test methods for measuring this performance. The two main standards
discussed in this book are IEC 61340-5-1 and ESD Association ANSI/ESD S20.20 (EOS/ESD
Association Inc. (2014a)). IEC 61340-5-1 has also been adopted as a national standard in
many countries, in Europe becoming the European Norm EN61340-5-1. Individual coun-
tries may have their own versions, and, in the United Kingdom, this is BS EN 61340-5-1
(British Standards Institute. (2016)).
This chapter explains how to make basic measurements for use in the assessment of
equipment for compliance with these standards. It is a guide to the main measurements
that ESD coordinators will wish to make in their facilities, giving some practical guidance
and “work-arounds” where necessary, and giving some information on the more unusual
tests and nonstandard tests that can be used where standard tests are not specified or are
unsuitable.
For clarity, the 61340-5 series terminology is used, although the Electrostatic Discharge
Association (ESDA) series test methods are in many cases nearly identical and are cross-.
In some cases, differences with the ESDA standards are noted.

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
346 11 ESD Measurements

11.2 Standard Measurements

61340-5-1 and 20.20 use a variety of basic test methods to assess the performance of ESD
prevention equipment and materials. These include
● Point-to-point resistance
● Resistance to ground or groundable point
● Surface and volume resistance of ESD protective packaging
● End-to-end resistance of a ground cord
● Personal grounding resistance tests
● Electrostatic fields and voltages
● Walk test of footwear and flooring
● ESD shielding test of shielding bags
A “groundable point” can be, for example, a stud on a bench or floor mat that is intended
to be connected to ESD ground in the final installation.

Table 11.1 Summary of test methods and their application in 61340-5-1, and corresponding ESD
Association standards.

Test method standard


IEC ESDA
Item tested

Work surfaces, racks 61340-2-3 STM4.1


Floor 61340-4-1 STM7.1
Seating 61340-2-3 STM12.1
Ionizers 61340-4-7 STM3.1
Hand tools None None
Garments 61340-4-9 STM 2.1
Wrist straps 61340-4-6 S1.1
Footwear – shoes 61340-4-3 STM9.1
Foot grounders None SP9.2
Footwear and flooring resistance measurement 61340-4-5 STM97.1
Footwear and flooring walk test 61340-4-5 STM97.2
ESD control packaging surface resistance – concentric ring 61340-2-3 STM11.11
ESD control packaging volume resistance – concentric ring 61340-2-3 STM11.12
ESD control packaging surface resistance – two-point 61340-2-3 STM11.13
resistance measurement.
ESD shielding bags 61340-4-8 STM11.31
Resistance of gloves and finger cots None SP15.1
Charge decay of tools, gloves, and finger cots 61340-2-1 None
Electrical Soldering/Desoldering Hand Tools None STM 13.1
Compliance Verification test methods – various TR 61340-5-4 TR53-01
11.3 Product Qualification or Compliance Verification? 347

Table 11.2 Other IEC test methods that can be used in ESD protective
equipment, materials, and packaging evaluation.

Corresponding
IEC standard IEC title ESDA standard

61340-2-1 Ability of materials and products None


to dissipate static electric charge
61340-2-2 Measurement of chargeability ESD ADV 11.2

The test methods will be described and demonstrated in the following sections. Simple
and clear test procedures for each measurement are given, which could be used to form the
basis of a company test procedure.
The test method standards available from IEC and ESDA at the time of writing are listed
in Tables 11.1 and 11.2. Many of the test methods in the IEC system have corresponding test
methods in the ESDA system, and vice versa, and these are indicated. It should be noted
that “correspondence” does not mean “equivalence” – there may be differences between
these documents. In some cases, the correspondence is with only part of the corresponding
document, or there may be other differences. As always, when aiming to comply with a
standard, it is necessary to refer to that standard for full and up-to-date details.

11.3 Product Qualification or Compliance Verification?


It is recognized that there are two main reasons for making measurements on equipment
and materials used for ESD control – product qualification and compliance verification.
Product qualification means the testing of ESD control products or materials to determine
whether they are suitable for use within the ESD control program. Compliance verification
refers to the tests routinely used for checking the equipment or materials are still working
as specified, during their working life. In some cases, the same test methods may be used
for both purposes.

11.3.1 Measurement Methods for Product Qualification


During product qualification test, the equipment or material will often not be tested in its
working environment. The tests used are typically designed to establish its basic character-
istics and ensure that it will fulfill its function under the expected variation of operating
conditions over its lifetime.
For this reason, items will usually be tested in the laboratory under dry atmospheric
humidity conditions as this represents the most challenging atmospheric conditions for
operation. Many test methods from the 61340-x series specify 12 ± 3% rh 23 ± 3 ∘ C, although
some specify other test conditions or leave it to the user to select suitable conditions.
The tests used may include tests under simulated working conditions and may include
the same test as is used for compliance verification.
348 11 ESD Measurements

11.3.1.1 Product Qualification of a Bench Mat


As an example, it may be required to test a bench mat type before approving it for purchase
and fitting on workstations in a facility. The objectives may be to
● Establish the basic material performance
● Ensure it will remain within limits even at low ambient humidity
● Give an idea of variation occurring between specimens of the item
IEC 61340-5-1:2016c Table 3 specifies that the installed resistance to ground shall be
<109 Ω (1 GΩ). However, many ESD Control Program Plans will also require a minimum
point-to-point resistance of the surface, as it is known that placing a charged component
on a low-resistance surface carries a risk of charged device ESD damage. Typically, the user
may select a lower limit of point-to-point resistance of, for example, 104 Ω (10 kΩ).
So, the standard specifies the following as product qualification tests:
● Resistance to groundable point (with an upper limit of 109 Ω [1 GΩ]).
● Point-to-point resistance of the surface (with an upper limit of 109 Ω [1 GΩ]).
● The test method standard specified is IEC 61340-2-3 (International Electrotechnical Com-
mission 2016a).
So, the user specifies these two tests to be performed under controlled atmosphere labo-
ratory conditions of 12 ± 3% rh 23 ± 3 ∘ C, as this will be representative of their worst-case
expected dry air conditions. Three specimens are tested after the specified conditioning time
(as required by 61340-2-3) as this gives an idea of the variation between specimens.

11.3.1.2 Example: Product Qualification of a Footwear and Floor Combination


A more complicated example is that of specification of a combination of footwear and
flooring to give the performance required by the 61340-5-1:2016 standard. This allows a
combination to be used that gives a relatively high resistance from the footwear wearer’s
body to ground of 109 Ω (1 GΩ), providing the body voltage does not rise above 100 V. IEC
61340-5-1:2016c Table 2 specifies that for product qualification, the test method standard is
61340-4-5 (International Electrotechnical Commission 2004). This gives the following:
● A test method for determining the resistance from the wearer’s body to ground via
footwear and flooring in combination (limit of resistance to ground 109 Ω [1 GΩ]).
● A “walking test” giving the voltage generated on the body (upper limit 100 V).
Note that these tests require the footwear and the flooring to be specified and tested in
combination, as intended to be used in practice. This is because the combined performance
of footwear with flooring is often dominated by factors such as contact resistance and charge
generated between the shoe sole and the floor surface.
In addition to these tests, it is wise to test that the footwear when worn will pass the usual
compliance verification test (resistance from body via footwear to metal plate).

11.3.2 Measurement Methods for Compliance Verification


Compliance verification tests tend to be simple tests specified to check that the equipment
or materials of the EPA are still working according to their intended specification. They
are often a simple system test of the equipment under operational conditions. A major
11.5 Summary of the Standard Test Methods and Their Applications 349

consideration in developing compliance verification tests is that these will need to be done
on a regular basis and test many items during an audit. A quick and easy but effective test
is required to minimize the time and effort required. Compliance verification tests are done
under the ambient atmospheric conditions in the EPA. It is good practice to make a note of
these during the test.
The 61340-5-1 and S20.20 standards have related documents IEC 61340-5-4 (Inter-
national Electrotechnical Commission 2019) and ESD TR53-01 (EOS/ESD Association
Inc. (2018a)) that specify tests adapted from other standards specifically for compliance
verification testing.

11.3.2.1 Example: Compliance Verification Test of a Bench Mat


Compliance verification test of a bench mat in use in the EPA is specified as a simple resis-
tance to ground measurement (Section 11.8.1) from the surface of the mat to the EPA earth.
In this way, the mat and its grounding system are tested in one quick measurement.

11.3.2.2 Example: Compliance Verification of Footwear and Flooring


Although a footwear and flooring combination may be specified using tests in combination
(see Section 11.3.1.2), it is usual to use simple separate compliance verification tests of the
footwear (while worn by personnel) and flooring.
The footwear resistance (tested while worn) is specified as a measurement of the resis-
tance from the wearer’s body to a metal contact plate under the foot (see Section 11.8.3.3).
This is normally done using a proprietary tester. The floor is tested separately using a peri-
odic test of resistance from the floor surface to ground (see Section 11.8.1.1).

11.4 Environmental Conditions


Compliance verification tests are normally made under normal operational ambient
temperature and humidity. If the conditions are variable, it is useful to make occasional
measurements, when possible, under the lowest-humidity conditions. These will often be
conditions under which maximum electrostatic charging will occur.
It is good practice to measure and record these at the time of the test. Humidity is espe-
cially important when high resistance (>1 GΩ) materials are measured. A change in humid-
ity, particularly below 30% rh, can have a large effect on the result with increased decay time
and higher resistance measured at lower humidity.
For ESD control product qualification tests, the items are preconditioned and measured in
the laboratory under set low humidity and temperature conditions. Dry conditions usually
give the worst-case and highest resistance measurement results. Many standards specify
12% rh 23 ∘ C as the test conditions.

11.5 Summary of the Standard Test Methods and Their


Applications
Several standard test methods are called up by IEC 61340-5-1:2016c in Tables 1–3 within
the standard. Table 11.1 summaries these standards and their corresponding standards in
350 11 ESD Measurements

the 20.20 system. Some gaps in the standard test methods are also shown. Table 11.2 gives
some additional test method standards that can be useful in the evaluation of ESD protective
equipment, materials, and packaging.

11.6 Measurement Equipment


11.6.1 Choosing a Resistance Meter for High-Resistance Measurements
Many measurements made on ESD control equipment and materials involve high-resistance
measurements. A suitable high-resistance meter is needed to make these measurements
according to the standard requirements.
Two important points to consider when buying a high-resistance meter are the test voltage
options and the resistance range. For use with 61340-5-1, the equipment must be capable
of measurement at the standard test voltages of 10 and 100 V (Table 11.3). The resistance
meter should have the capability to measure resistance from 10 times less than the lowest
resistance limit required to be measured to at least 10 times the greatest resistance required
to be measured. Many types of ESD control equipment have an upper limit of 1 GΩ spec-
ified in the standards, requiring a resistance meter that will measure up to at least 10 GΩ.
ESD packaging and garments have a specified upper resistance limit of 100 GΩ (1011 Ω).
Measuring these require a meter that can measure up to 1 TΩ (1012 Ω).
The resistance measurement result will usually vary with test voltage, being raised for a
lower test voltage. The standard test methods used in 61340-5-1 and S20.20 typically require
use of 10 V to measure resistances up to 1 MΩ. The test voltage is increased to 100 V if a
resistance >1 MΩ is found at 10 V. Always test at 10 V first, and then move to 100 V if the
result exceeds the limit.

Table 11.3 Test voltages used in resistance measurements required in 61340-5-1 and S20.20
standards.

Standards
Usage Test voltage Resistance range referenced

General resistance measurements 10 V Up to 10 MΩ IEC 61340-2-3


100 V 100 kΩ to 10 GΩ ESD STM11.11
ESD STM11.12
Packaging and garments 10 V 10 kΩ to 10 MΩ
ESD STM11.13
100 V 100 kΩ to 1 TΩ
Footwear and flooring system test 10 V a) 10 kΩ to 10 MΩ IEC 61340-4-5
The superscript 1 indicates the note applies. 100 V a) 100 kΩ to 1 TΩ ESD STM97.1

Wrist straps as worn 7 V to 30 V a) 50 kΩ to 100 MΩ IEC 61340-4-6


ESD S1.1
Footwear as worn, to metal plate 9 V to 100 V a) 50 kΩ to 1 GΩ IEC 61340-5-1
Annex A

a) Where measurements involve the human body, the resistance meter should be current limited to a
maximum of 0.5 mA (See Section 10.1.3).
11.6 Measurement Equipment 351

Where high voltages are used for measurement, always assess the safety issues and take
appropriate safety precautions. Most handheld instruments are unlikely to source danger-
ous electrical currents, but check your instrument to evaluate any risks.

11.6.2 Low-Resistance Meter for Soldering Iron Grounding Test


For measurement of grounding of a soldering iron bit, a low-resistance meter is needed,
capable of measurement of resistance in the range 0–100 Ω. Many multimeters have a con-
tinuity range that can make this measurement.

11.6.3 Resistance Measurement Electrodes


IEC 61340-5-1 uses as the basis of many of its resistance measurements electrodes speci-
fied in standards IEC 61340-2-3 and 61340-4-1 (International Electrotechnical Commission
2015b).
Electrodes may be found on the market that conform to a variety of current and historical
standards, as well as others that do not conform to standards. Different electrode systems
in general may give different results, although the differences in some cases may not be
very great or significant. The 2.5 kg resistance measurement electrodes shown in Table 11.4
and Figure 11.1 compliant with IEC 61340-2-3, IEC 61340-4-1, ESD STM2.1, ESD STM4.1
(EOS/ESD Association Inc. (2017)), ESD STM7.1 (EOS/ESD Association Inc. (2013c)), and
ESD STM12.1 (EOS/ESD Association Inc. (2013d)) are commonly available and suitable.
Although the electrodes of ESDA 2.1, 4.1, 7.1, and 12.1 are specified as 2.27 kg, these fall
within the tolerance of the 2.5 kg IEC specification. Electrodes to all these standards are for
convenience referred to as 2.5 kg electrodes in this chapter.

Table 11.4 Comparison of characteristics of 2.5 kg resistance measurement electrodes from


various standards.

ESDA STM2.1
ESDA STM4.1
ESDA STM7.1
Characteristic 61340-4-1 61340-2-3 EN100015-1 ESDA STM12.1

Mass (kg) 2.5 ± 0.25 or 2.5 ± 0.25 2.5 ± 0.5 2.27 ± 0.06
5.0 ± 0.25a)
Diameter (mm) 65 ± 5 63.5 ± 1 75 63.5 ± 0.25
b)
Electrode surface 60 ± 10 50–70 50–70
hardness (Shore A)
Resistance (electrode <1 kΩb) <1 kΩb) <1 kΩc)
placed on metal sheet)
Notes:
a) According to 61340-4-1:2003, the 2.5 kg electrode is to be used for measurements on hard nonconformable
surfaces. The 5 kg to be used for measurements on all other surfaces. The first is applicable to most appli-
cations under 61340-5-1.
b) Conductive rubber pad need not be used with conformable (e.g. textile) surfaces.
c) Measured at 10 V between two electrodes placed on metallic surface.
352 11 ESD Measurements

Figure 11.1 An example of 2.5 kg electrodes according to IEC 61340-2-3 and ESD STM11.11.

Organizations in Europe that have been involved in ESD control for many years often
have EN100015 (CENELEC 1992) electrodes. The EN100015 electrodes may give different
results due to their different diameter and lack of conductive rubber face material but may
give results that correlate with the other electrodes. The lack of rubber face material means
that they might not make good contact with hard surface materials. These electrodes are
not recommended for making measurements according to 61340-5-1.

11.6.4 Concentric Ring Electrodes for Packaging Surface and Volume


Resistance Measurement
One option for measurements of surface or volume resistance of packaging materials
according to IEC 61340-5-3 (International Electrotechnical Commission 2015d) and
ANSI/ESD S541 (EOS/ESD Association Inc. (2018c)) requires use of a concentric ring
electrode (Table 11.5, Figures 11.2, and 11.3). This consists of an inner circular electrode
and an outer ring electrode surrounding it. The surface resistance is measured between the
inner and outer ring electrodes when the electrode is placed on a surface. For standard
electrodes, the surface resistance is approximately a factor of 10 less than the material
surface resistivity.
For volume resistance measurement, the material under test is placed on a flat metal
electrode. The concentric ring electrode is placed on the material under test. The resis-
tance is measured through the material between the flat metal electrode and the inner ring
electrode. The outer ring electrode can be used as a guard ring to prevent surface leakage
affecting the result.
11.6 Measurement Equipment 353

Table 11.5 Comparison of concentric ring surface resistance measurement electrodes from
various standards.

Characteristic 61340-2-3:2016 ESDA STM11.11-2015b

Mass (kg) 2.5 ± 0.25 2.27 ± 0.0567


Electrode surface hardness (Shore A) 50–70 50–70
Inner electrode diameter (mm) 30.5 ± 1 30.48 ± 0.64
Outer electrode inner diameter (mm) 57 ± 1 57.15 ± 0.64
Outer electrode width (mm) 3 ± 0.5 3.18 ± 0.25
Soft conductive material volume resistivity (Ω cm) <10

d2 = 57 mm d1 = 30 mm

Figure 11.2 Concentric ring electrode contact area according to IEC 61340-2-3 and ESD STM11.11
and ESD STM11.12.

Figure 11.3 Examples of concentric ring electrode.


354 11 ESD Measurements

11.6.4.1 Conversion of Surface Resistance to Surface Resistivity for a Concentric Ring


Electrode
Some standards quote material surface resistivity rather than surface resistance. If surface
resistivity is required, the surface resistance reading must be multiplied by a factor that is
related to the electrode geometry. For the concentric ring electrodes of IEC 61340-2-3 and
ESD STM11.11, the factor is 10. The surface resistivity is 10 times surface resistance.
More precisely, the conversion for a concentric ring electrode is given by IEC 61340-2-3
R 𝜋(d1 + g)
𝜌s = s
g
where 𝜌s is the surface resistivity (Ω), Rs is the surface resistance (Ω), d1 is the diameter of the
inner electrode (m), and g is the gap between inner and outer electrodes (m) (Figure 11.2).
For the electrodes specified in 61340-5-1, we have g = 0.0135 and d1 = 0.03, which gives
𝜌s = 10.1Rs
For the miniature two-point probe or point-to-point resistance measurement with two
2.5 kg electrodes, the factor for conversion to surface resistivity has not been established.

11.6.4.2 Conversion of Volume Resistance to Volume Resistivity


For a homogenous sheet sample material of thickness h (m) and electrode diameter d1 (m),
the volume resistance Rv can be converted into the volume resistivity 𝜌v of the material
according to IEC 61340-2-3.
𝜌v = 𝜋Rv d21 ∕4h
ESD STM11.12 (EOS/ESD Association Inc. (2015c)) gives an equivalent conversion for-
mula, where the inner electrode area A is specified to be 7.1 cm2 and the specimen thickness
t cm as
𝜌v = Rv A∕t

11.6.5 Two-Point Probe for Packaging Surface Resistance Measurements


A handheld two-point probe is specified in IEC 61340-2-3 and ESD STM11.13 (EOS/ESD
Association Inc. (2015d)). It consists of two 3.2 mm diameter spring loaded probes 3.2 mm
apart that are held by spring action in contact with the surface being measured with a force
of 4.6 ± 0.5 N (Figure 11.4). The probes are faced with conductive rubber contact material
with Shore hardness 50–70. A correlation factor for the results obtained with this electrode
system compared to the concentric ring probe and surface resistivity has not been deter-
mined (see Section 11.8.4).

11.6.6 Footwear Test Electrode


In testing footwear, an electrode is needed to contact the sole of the footwear. This should
be a metal plate sufficiently large for the whole of the foot to be placed upon it (Figure 11.5).
Some standards may specify the size of the plate to be used. It should have a connector to
allow connection of test leads to other equipment.
In proprietary footwear testers, the footwear test electrode is usually incorporated into
the tester equipment.
11.6 Measurement Equipment 355

Figure 11.4 Two-point probe electrode according to IEC 61340-2-3 and ESD STM11.13.

Figure 11.5 An example of a footwear test electrode.

11.6.7 Handheld Electrode


A handheld electrode is used in several types of measurement of grounding of personnel
(Sections 11.8.3.2, 11.8.3.3 and 11.8.3.4). It can also be used for in-service system measure-
ments of gloves or tools (Section 11.9.7.1, 11.9.7.3) and walk test of footwear and flooring
(Section 11.8.9).
A hand touch electrode in the form of a metal plate is often built into equipment such as
simple pass/fail resistance testers of wrist straps or footwear. The tool test electrode (Section
11.6.8) can also be used in this way.
A handheld electrode specified in IEC 61340-4-5, IEC 61340-4-9 (International Elec-
trotechnical Commission 2016b), ESD STM97.1 (EOS/ESD Association Inc. (2015f)), ESD
STM97.2 (EOS/ESD Association Inc. (2016c)), and ESD STM2.1 is a metal rod or tube
about 25 mm diameter and at least 75 mm long that has a connector for connection of
356 11 ESD Measurements

Figure 11.6 Example of standard handheld electrodes (right) and a nonstandard electrode (left).

leads to test equipment (Figure 11.6). In most cases, the handheld electrode size and shape
will have little effect on the measurement results, providing there is sufficient contact area
with the skin of the hand.

11.6.8 Tool Test Electrode


A tool test electrode can be used for testing the resistance to ground of tools (Section 11.9.5)
and soldering irons (Section 11.9.6.2).
A tool test electrode can be a simple metal plate having a means of connection to the test
equipment (Figure 11.7). The plate must be electrically isolated from any parallel ground
paths. This can be done by supporting it on an insulating support plate or feet.

11.6.9 Metal Plate Electrode for Volume Resistance Measurements


A metal plate electrode is needed for volume resistance measurements on packaging mate-
rials (Section 11.8.5). The electrode must be sufficiently large to support the packaging
material under test and must be made of a material such as stainless steel that does not

Figure 11.7 A simple tool test electrode (left) and underside (right) showing insulating feet.
11.6 Measurement Equipment 357

oxidize. Aluminum is not suitable due to its formation of an alumina surface layer that can
affect measurements.

11.6.10 Insulating Supports


Insulating supports are needed in some tests to prevent parallel conduction paths through
the supporting material. The main requirement for these is that the support should be large
enough to completely support and isolate the test item and at least 10 times higher resis-
tance than expected for the item under test. So, for example, an insulating support for a
packaging material that could have resistance up to over 100 GΩ (1011 Ω) should itself have
resistance over 1 TΩ (1012 Ω). Insulating supports can often conveniently be made from
common insulating plastic or rubber sheets. The surface and volume resistance should be
verified as described in Sections 11.8.4 and 11.8.5.

11.6.11 ESD Ground Connectors


In resistance to ground tests, it is necessary to connect one terminal of the test meter with
ESD ground. A suitable connector for this may depend on what ESD ground is defined.
In cases where mains electrical safety earth is used as ESD ground, proprietary connectors
may be used to connect to the earth pin (Figure 11.8).

11.6.12 Electrostatic Field Meters and Voltmeters


11.6.12.1 Electrostatic Field Meter–Based Instruments
Many electrostatic voltmeters are electrostatic field meters, calibrated to read voltage when
positioned at a certain distance from a large flat conducting target surface. These have a
wide field of view. The voltage reading is reduced for small objects and can be influenced
by nearby charged surfaces or fields.

Figure 11.8 An example of a 0 Ω mains safety earth connector.


358 11 ESD Measurements

Figure 11.9 Electrostatic field meters. (left) Induction type and (center and right) two field mill
type instruments.

The apparent voltage reading increases if the meter is brought closer to the surface. A
correction can be made if the distance is known, but it is normally most convenient to make
readings with the field meter/voltmeter held at the calibrated distance. Some meters have
a means of gauging the correct distance, e.g. converging lights or pillars that are placed in
contact with the surface.
These instruments operate by measuring the charge induced on a capacitive sensor plate
by an electrostatic field. The instrument must be grounded correctly when measurements
are made.
Some types of simple low-cost induction type instrument sense the field by charge
induced on a simple metal plate (Figure 11.9, left). These tend to suffer from drift. It is
important to zero these instruments before each measurement, with the sensitive aperture
shielded by a grounded surface or in a zero-field region, before each measurement is made.
“Field mill” type instruments have a rotating mechanical shield interrupting the electro-
static field to the sensor (Figure 11.9, center and right). These instruments are self-zeroing
and do not suffer from drift as much as the simple instruments do.
Field meters used to measure surface voltage of insulators for evaluation of risks in
61340-5-1 and 20.20 should be calibrated to read voltage at a distance of 2.5 cm (1 in.).
Meters calibrated at other distances will give different voltage readings.

11.6.12.2 Noncontact Electrostatic Voltmeters


True noncontact electrostatic voltmeters are available, based for example on vibrating reed
sensor technology (Figure 11.10). These must usually be held within a specified range of
distance several millimeters from the surface. Within this range, the voltage reading is rea-
sonably unaffected by distance from the surface. The instrument measures a surface voltage
from a relatively small area close to the voltmeter sensor. They can be used to measure the
voltage on moderately small items or areas around a centimeter dimension.
11.6 Measurement Equipment 359

Figure 11.10 A handheld noncontact electrostatic voltmeter. The sensitive tip contains a vibrating
reed sensor.

11.6.12.3 Contact Electrostatic Voltmeters


Ultra-high input resistance and ultra-low input capacitance contact voltmeters are now
available (Figure 11.11). These can be used to measure voltage on small conductive or insu-
lative items or surfaces. The sensor tip is touched to the surface of the item being measured.

11.6.13 Charge Plate Monitors (CPM)


Ionizers are tested using a charged plate monitor (CPM). A CPM can also be used for making
measurements of charge decay (e.g. of tools or gloves), body voltage in a walk test, and as a
means of estimating charging of objects.
A CPM has a metal plate that can be charged to a high voltage. It has an electrostatic volt
meter monitoring the voltage on the plate (Figure 11.12) and a means of charging the plate
to a known voltage (usually a little over 1000 V).
For ionizer performance measurements, the CPM is placed in a position simulating the
normal working position of an ESD-sensitive (ESDS) board or component. After charging
to an initial voltage, the time taken for the voltage on the metal plate to be neutralized
by the ionizer is measured. After the voltage neutralization time, the CPM monitor will
typically still show a residual reading. This is due to ionizer offset voltage, caused by ion
output imbalance. This offset voltage is also an important characteristic of the ionizer.
There are many different CPMs on the market, ranging from clip-on attachments to elec-
trostatic field meters to sophisticated dedicated instruments with built-in decay time and
offset measurement capability.
To make measurements according to the ESD STM 3.1 or 61340-4-7 (International Elec-
trotechnical Commission 2017) standards, a CPM designed according to the requirements
of these standards is required. This must have a plate of size 15 × 15 cm that has a capaci-
tance of 20 ± 2 pF (Figure 11.13).
360 11 ESD Measurements

Figure 11.11 A contact voltmeter. Source: D E Swenson. The sensing tip has extremely high
impedance.

Ion stream Charged plate


Electrostatic voltmeter

High-voltage source

Figure 11.12 Principle of CPM operation.

Many CPMs on the market are not built according to these standards and have a smaller
plate size and capacitance. They are often smaller and lighter than a standard CPM and are
used for functional and comparative measurements with ionizers. These may often give
different results to the standard CPM and should be correlated to a standard CPM for com-
pliance verification measurement against standard ionizer requirements.
In practice, most organizations will use a nonstandard CPM for comparative and func-
tionality measurements on ionizers in their working positions in a workstation. In some
cases, a smaller sized nonstandard CPM plate can better represent the smaller area of a
typical ESDS device.
11.7 Common Problems with Measurements 361

Figure 11.13 Example of a CPM built according to the IEC 61340-4-7 and ESD STM3.1 standard.
Note the large 15 × 15 cm plate. Handheld electrode and voltmeter attachment accessories are also
shown. Source: D E Swenson.

11.7 Common Problems with Measurements

11.7.1 Humidity
It is good practice to measure the ambient relative humidity when making electrostatic
measurements. High ambient humidity can cause insulating materials to show lowered
surface resistance <1011 Ω (100 GΩ). The instrument may be measuring the resistance of
the water layer adsorbed on the material. If in doubt, repeat the measurement under dry
(<30% rh) conditions.

11.7.2 Accidental Measurement of Parallel Paths


It is easy to accidentally “short circuit” the object you are trying to measure by a lower
resistance path. One common mistake is to hold the test probes and/or object under test in
the hands – and inadvertently measure the resistance of a parallel conduction path through
the body!
Some objects such as garments must be placed on a highly insulating surface for test
to avoid the underlying surface affecting the measurement. The surface may need to be
cleaned before use – for example with a material such as isopropyl alcohol wipes, which
lift off the contamination. The alcohol evaporates quickly but allows the surface to dry for
a reasonable time before use.
Some test leads and probes may not be as good insulators as would be desired. If the insu-
lating parts of these are held in the hand or placed on a dissipative surface, leakage to the
362 11 ESD Measurements

hands may affect the measurement. This is usually only a problem with high-resistance
measurements above about 1010 Ω (10 GΩ) but can be worse under high-humidity condi-
tions. Sometimes test leads can be further isolated for use in high-resistance measurements
by threading them through a clean polyethylene tube.
Sometimes it is necessary, e.g. for resistance to ground measurement, to use very long
leads at least on one side of the meter. Connect the long lead on the earthy side of the meter
to reduce leakage problems.

11.8 Standard Measurements Specified by IEC 61340-5-1


and ANSI/ESD S20.20

11.8.1 Resistance to Ground


Resistance to ground is a basic measurement used in many situations to evaluate the like-
lihood of charge dissipating quickly to ground. A 2.5 kg electrode is placed on the surface
to be tested. A high-resistance meter is connected between the electrode and ESD ground
(Figure 11.14).

11.8.1.1 Resistance to Ground from a Work Surface or Floor


Figure 11.15 shows measurement of resistance to ground of a work surface. The same
method is used for a floor, cart (trolley), or storage rack.

Equipment required:
● One 2.5 kg resistance measurement electrode.
● Connector to ESD ground (0 Ω resistance).
● 10/100 V high-resistance meter with test leads.
● Work surface or floor to be tested.

High-resistance meter
10V / 100V

2.5 kg resistance
measurement
electrode

Surface of item
measured
ESD ground

Figure 11.14 Measurement of resistance to ground.


11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 363

Figure 11.15 Measurement of resistance to ground from a work surface.

Procedure:

● Connect the resistance meter and place electrode on surface.


● Measure result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.

Common Problems
ESD ground is often, but not always, mains electrical safety earth. Make sure you have
identified the correct ESD ground before making connection. The ESD ground should be
specified in the ESD Control Program Plan document.
If the ESD earth is mains electrical safety earth, it is usually convenient to measure the
resistance from the electrode to an ESD earth connector. Many of these have an internal
1 MΩ resistor. If present, the value of this must be subtracted from the measurement result.
If the measured item has resistance very much greater than 1 MΩ, this added resistance
may be neglected. To avoid confusion, it is often better to reserve ground cords and earth
connectors that do not have these built-in resistors for use in ESD measurements.
If you check resistance to ground back to a convenient mains electrical earth point, don’t
forget to check that the earth point is really connected back to safety earth. An extension
lead can be found to be unplugged or the electrical earth had somehow disconnected!
A higher than expected resistance result can be due to dust and dirt or other contamina-
tion accumulated on the surface. This is an indication that the cleaning regime may need
to be improved.
364 11 ESD Measurements

Resistance meter
10V / 100V

ESD ground

ESD control floor

Figure 11.16 Compliance verification resistance to ground measurement of seating.

If the surface is dusty, it can be wiped with a dry cloth or paper towel before repeating
the measurement. If the surface is sufficiently dirty to require cleaning with a liquid clean-
ing material, it must be allowed to dry before measurement; otherwise, the measurement
results may be affected. Even a small amount of moisture on the surface will lower the
measured resistance results.

11.8.1.2 Compliance Verification of Seating in the EPA


This simple resistance to ground measurement is used for compliance verification test of
seating in the EPA including the ground path through the flooring. It is done while the seat
is standing on the ESD control floor at the EPA workstation (Figure 11.16).

Equipment required:
● One 2.5 kg resistance measurement electrode.
● ESD earth connector.
● 10/100 V high-resistance meter with test leads.
● Sample chair for testing.

Procedure:
● Place electrode on chair seat.
● Connect the resistance meter between electrode and ESD earth.
● Measure result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.

11.8.1.3 Qualification of Seating


In product qualification of seating, we want to measure the characteristics of the chair alone
separate from the flooring system. Also, we want to know that all relevant parts of the chair
such as the back and the arms are also suitable for ESD control use. So, the resistance mea-
surement is made to a groundable point, e.g. an electrode placed under the wheel or foot
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 365

Resistance meter
10V / 100V

Separately test
resistance to ground
point and each
conductive wheel
Groundable point

Steel plate under wheel

Insulating floor surface resistance >1010 Ω

Figure 11.17 Product qualification measurement of seating.

(Figure 11.17). Resistance to several wheels or feet may be tested to see how many provide
groundable points.
To isolate the chair from any conduction paths through the floor, the chair must be placed
on an insulating surface that has a resistance considerably higher than the resistance
expected of the chair. This is a test that is best done in a test workshop and not in the EPA.

Equipment required:
● One 2.5 kg resistance measurement electrode.
● 10/100 V high-resistance meter with test leads.
● Insulating support >1010 Ω.
● Metal plate electrode (for under wheel).
● Sample chair for testing.

Procedure:
● Set the chair on the insulating support. Place metal plate electrode under a grounding
wheel or foot.
● Place the resistance measurement electrode on chair seat.
● Connect the high-resistance meter.
● Measure result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.

Common Problems
It is usually wished to test parts of the chair such as the seat back or arms, which may not
be horizontal or have large flat areas on which an electrode can be easily balanced. Some
ingenuity is required to get around these problems. If the chair has an easily connected
groundable point, then the matter is simplified. The resistance from seat back to groundable
point may be measured with the chair laid on its back on an insulating support plate.
366 11 ESD Measurements

If the chair has an exposed metal part of the chassis that is confirmed to be connected
reliably to the groundable point, then the resistance of arms and back can be measured to
this metal part in subsequent measurements.

11.8.2 Point-to-Point Resistance


Point-to-point resistance has in current practice replaced surface resistivity measurement
for evaluating surfaces such as bench mats, rack and cart surfaces, and garments. The
method consists of a simple resistance measurement between two electrodes placed on the
surface (Figures 11.18 and 11.19).
Measurements can be made in several locations and different orientations to evaluate
if there is variation or directionality in the material surface resistance characteristic. This
can happen with contaminated or worn surfaces and nonhomogenous materials. Measure-
ments should be made across surface features such as garments seams to detect whether
continuity is broken by these boundaries.

11.8.2.1 Point-to-Point Resistance of Work Surface


During product qualification of a work surface, a point-to-point resistance test will give an
evaluation of the resistance of the surface material (Figure 11.19). An ESD program will
often give a minimum surface resistance requirement to prevent charged device ESD risks.
Making several measurements in different areas and orientations of electrodes will indi-
cate the variability of the surface resistance and whether there is any directionality. IEC
61340-2-3 specifies that the electrodes should be placed at least 250 mm apart center to
center. This may not be practical in some cases (e.g. measurements on garments).

Resistance meter
10V / 100V

2.5 kg

Conductive
rubber

300 mm

Figure 11.18 Point-to-point resistance measurement.


11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 367

Figure 11.19 Measurement of point-to-point resistance of a work surface material.

Equipment required:
● Two 2.5 kg resistance measurement electrodes.
● 10/100 V high-resistance meter.
● Sample work surface.

Procedure:
● Place the electrodes on work surface at least 250 mm apart.
● Connect the resistance meter.
● Measure result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.
The measurement can be repeated in different locations and orientations to detect
whether the material has significant variation with position or direction.

11.8.2.2 Point-to-Point Resistance Measurements on ESD Garments


The fabric and structure of ESD control garments can be measured using simple point-
to-point measurements both for product qualification and for compliance verification. IEC
61340-4-9 and ESD STM2.1 specify tests for ESD control garments. Measurement point to
point across seams verifies that the fabric conductive fibers make contact across the seams
368 11 ESD Measurements

Figure 11.20 Measurement of resistance point to point of a garment across panels. The garment
has been placed on a slab of insulating support material.

(Figure 11.20). Cuff-to-cuff measurements verify the connection of the garment materials
via the materials and seams between the cuffs.

11.8.2.3 Simple Point-to-Point Measurement on the Garment Fabric


The garment material to be tested must be placed on an insulating support having a surface
resistance at least 10 times the expected resistance of the garment material.
Some garments have a groundable point, for connection of a ground strap. In this case,
a resistance to groundable point measurement should be measured to include the garment
seams.

Equipment required:
● 2 off 2.5 kg resistance measurement electrodes.
● 10/100 V high-resistance meter with test leads.
● Garment specimen to be tested.
● Insulating support.

Procedure:
● Place the insulating support on a working surface. Lay out the garment on the insulating
support, if possible, with a single layer of fabric.
● Place the electrodes on separate panels of the garment a convenient distance apart.
● Connect the high-resistance meter.
● Measure resistance result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.
Making several measurements in different areas of the material and orientations of elec-
trodes will indicate the variability of the surface resistance, and whether there is any direc-
tionality.
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 369

Common Problems
Many garment materials have high point-to-point resistance. Measurements above 1010 Ω
(10 GΩ) tend to be unreliable, and results are dependent strongly on humidity.
If the insulating mat is not sufficiently insulating, the measurement may be affected, and
the resistance result reduced. Check the point-to-point resistance of the insulating support
before the test to make sure it’s at least 10 times the expected resistance of the garment.

Measurement from Garment Cuff to Cuff


Cuff-to-cuff measurement tests the garment sleeve and garment body panel resistance and
the connection between these as a system (Figure 11.21). Connection can be made to the
inside of the sleeve to include the area that contacts the user’s arms. This measurement
should be done for garments that are designed to make contact with the wearer’s skin at
the cuffs. Alternatively, connection can be made to the outside of the cuffs (Figure 11.22).

Equipment required:
● 2 off 2.5 kg resistance measurement electrodes.
● 10/100 V high-resistance meter with test leads.
● Garment specimen to be tested.
● Insulating support.
● Insulating cuff inserts. (These can be made from sheet insulating plastic material.)

Procedure:
● Place the insulating support on a working surface. Lay out the garment on the insulating
support.
● Place the electrodes on separate panels of the garment.
● Connect the high-resistance meter.

Figure 11.21 Measurement of cuff-to-cuff resistance of a garment, (above) with garment hanging
or (below) with garment resting on an insulating support. Source: D E Swenson.
370 11 ESD Measurements

Figure 11.22 Contacting (left) the inside or (right) the outside of a garment cuff with a 2.5 kg
electrode. The garment is placed on an insulating support. An insulating separator is used when
contacting the outside of the garment sleeve.

● Measure resistance result at 10 V.


● If R > 1 MΩ, measure result again at 100 V.
● Note the result.

Measurement of Sleeve-to-Sleeve Resistance of a Garment Using Hanging Clamps


An alternative means of connecting to the garment cuffs is by hanging clamps (Figure 11.23)
suspended from an insulating frame.

Equipment required:
● Two clamp electrodes suspended from an insulating frame.
● 10/100 V high-resistance meter with test leads.
● Garment specimen to be tested.

Procedure:
● Connect the clamp electrodes to the garment cuffs and suspend the garment.
● Connect the high-resistance meter to the clamp electrodes.
● Measure resistance result at 10 V.
● If R > 1 MΩ, measure result again at 100.
● Note the result.

Measurement of Resistance to Groundable Point of a Garment


For garments that have a groundable point that is connected to ESD ground, the resistance
should be measured between a 2.5 kg electrode placed on the material (cuffs and panels)
and the groundable point. If the cuffs are designed to ground the garment by contact with
the wearer’s body, the cuff body contact areas should be considered groundable points,
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 371

Figure 11.23 Measurement of cuff to cuff resistance of a garment hung from clamps. Source: D E
Swenson.

and the resistance should also be measured to these. The resistance should be measured
between a 2.5 kg electrode placed on the material panels and the inner surface of the cuffs
(Figures 11.21 and 11.22).

11.8.3 Personal Grounding Equipment Tests


11.8.3.1 End-to-End Resistance of a Ground Cord
End-to-end resistance is a simple resistance measurement used for items such as wrist strap
cords. This measurement is normally easy, providing there is a suitable means of making
the connections (Figure 11.24).

Equipment required:
● 10/100 V resistance meter with appropriate leads and connectors
● Ground cord to be tested
372 11 ESD Measurements

Figure 11.24 Measurement of end-to-end resistance of a wrist band cord.

Procedure:
● The wrist strap cord to be tested is connected to the resistance meter using appropriate
connectors.
● Note the result.

Common Problems
Do not be tempted to hold the cord ends in contact with test leads with the fingers. If you
do this, you risk connecting your body in parallel with the cord. This would then give an
incorrect result. Also, make sure any connectors used do not contact a conductor such as
an ESD control bench surface.

11.8.3.2 Measurement of Wrist Straps and Cords as Worn


This test is often done using a proprietary wrist strap checker giving a simple pass/fail ver-
dict (Figure 11.25). It can also easily be done using a resistance meter (Figure 11.26).

Equipment required:
● Handheld electrode
● 10/100 V resistance meter with test leads
● Wrist strap and cord to be tested

Procedure:
● The subject to be tested should wear the wrist strap in contact with their skin.
● The wrist strap cord to be tested is connected to one terminal of the resistance meter.
● The second terminal of the resistance meter is connected to the handheld electrode.
● The subject holds the electrode. Note the result.
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 373

Figure 11.25 Simple proprietary wrist strap checkers make verification easy. The pass/fail limits
of these checkers must correspond with the requirements of the ESD control program.

Figure 11.26 A handheld electrode used with suitable resistance meter allows checking of wrist
strap as worn.
374 11 ESD Measurements

Common Problems
If a wrist strap checker is used, the pass/fail limit of the checker must be selected to corre-
spond with the limits specified in the ESD control program.
Many wrist strap checkers also give a “low fail” if the resistance is less than a minimum
value. This “low fail” limit must correspond to the lower limit specified in the ESD control
program.
Note that current versions of 61340-5-1 and 20.20 do not specify a low limit for wrist strap
resistance, (see Section 7.5.3). Nevertheless, many organizations will often wish to specify
minimum resistance from body to ground for safety in the presence of high voltages or other
reasons.

11.8.3.3 Measurement of Personnel Grounding Through Footwear to Foot Plate


Electrode
This test measures the resistance from the body to a ground electrode formed by a metal
plate. It is this measurement that is commonly performed by footwear test boxes found at
the entrance of an EPA but can also easily be done with a resistance meter.

Equipment required:
● Handheld electrode
● 10/100 V resistance meter with test leads
● ESD footwear
● Foot plate electrode, big enough to accommodate an entire foot
● Insulating support mat

Procedure:
● The foot plate is placed on the insulating support mat and connected to the resistance
meter. The second terminal of the resistance meter is connected to the hand touch elec-
trode.
● The subject should wear the footwear to be tested.
● The subject stands with one foot on the foot plate and one foot on the insulating mat.
● The subject holds the handheld electrode.
● Note the result.
● repeat test for the other foot.

Common Problems
If a proprietary footwear checker is used, make sure that the upper and lower pass/fail limits
correspond with those specified in the ESD control program.

11.8.3.4 Measurement of Resistance from Person to Ground


This measurement is useful in checking the complete installed ground path from person
to ground, whether it be via wrist strap (Figure 11.27) or footwear and flooring or via a
groundable garment (Figure 11.28). It is based on IEC 61340-4-5 and ANSI/ESD STM97.1. It
can be done with an operator at a workstation with their personal grounding in its operating
state, with little interruption to their work (Figure 11.27). It shows whether the personnel
resistance to ground is within the required limits, as a system including the workstation
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 375

Figure 11.27 Measurement of resistance from the person’s body to ground.

Figure 11.28 Measurement of resistance to ground when grounded via a groundable garment.
Source: D E Swenson.
376 11 ESD Measurements

Figure 11.29 Checking the resistance from person to ground using a portable wrist strap checker.

earth bonding point or EPA floor. The test can also be done with a portable wrist strap
checker (Figure 11.29) or footwear checker, providing the pass/fail thresholds correspond
to the ESD control program requirements.

Equipment required:
● Handheld electrode
● 10/100 V resistance meter with test leads
● ESD ground connector (0 Ω resistance)

Procedure:
● The person to be tested should wear their personal grounding equipment (wrist strap
or ESD control footwear). The wrist strap should be connected to ground in its normal
working position. For footwear-flooring grounding, the person should stand on the ESD
control floor to be tested.
● Connect one side of the resistance meter to ESD ground via the ESD ground connector.
● Connect the other side of the resistance meter to the handheld electrode.
● The subject holds the electrode. Note the result.

Common Problems
The test gives an easy way of checking personal grounding as a system in operation. Don’t
forget to check that the ESD earth point connected to the resistance meter is also connected
back to ESD earth!
The test does not distinguish between individual grounding systems if more than one is
in operation (e.g. simultaneous grounding through wrist strap and footwear-flooring).
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 377

Sometimes, personnel who have dry skin do not easily make good connection with wrist
straps. This effect may be most evident just after the wrist strap is first put on and before a
sweat moisture layer has formed between the wrist and the wrist strap band. Sometimes it
can be necessary to augment the moisture layer, for example by using a skin lotion.
Contamination of the footwear sole or floor surface can lead to high-resistance results
for footwear-flooring grounding. If necessary, clean these and retest. If liquid cleaners are
used, make sure they are fully dried before retesting.

11.8.3.5 Resistance to Ground of an Earth Bonding Point


It is also necessary to verify the resistance to ground of the earth bonding points, e.g. on
workstations.

Equipment required:

● 10/100 V resistance meter


● 0 Ω ESD ground plug
● Earth bonding point to be tested

Procedure:

● Connect one side of the resistance meter to ground via a 0 Ω ground plug.
● Connect the other side of the meter to the earth bonding point under test.
● Record the result.

Common Problems
This measurement is usually straightforward.

11.8.4 Surface Resistance of Packaging Materials


The surface resistance of ESD protective packaging material is commonly measured in three
ways. Note that it is surface resistance, not resistivity, as it is not corrected for the electrode
form.

● Using a concentric ring electrode


● Using a miniature two-pin point-to-point electrode
● Using two 2.5 kg resistance measurement electrodes in a point-to-point measurement

These three electrode systems often do not give the same results. ESD packaging materials
can have highly variable surface resistance at different positions, and the electrodes respond
differently to these variations. The concentric ring electrodes tend to give the lower range
of resistance of the material around the circumference of the electrode (Smallwood 2017).
It is not directional in response.
The two-pin point-to-point electrode tends to give a result about a factor of four greater
than the concentric ring electrode, for a homogenous material of the same surface resis-
tivity. It also is highly influenced by any small areas of high-resistance material under the
point electrodes and is directional in response. So, this electrode gives greater variation in
results if the material is variable in resistance. It is more likely to indicate the upper range
of resistance of the material under the electrode points.
378 11 ESD Measurements

Figure 11.30 Measuring the surface resistance of an ESD protective packaging material. The
sample is placed on an insulating base material.

A point-to-point measurement made with closely spaced 2.5 kg electrodes can give similar
results to the concentric ring electrodes but has a directional response.

11.8.4.1 Surface Resistance of Packaging Measured Using a Concentric Ring Electrode


Surface resistance of large flat packaging surfaces can be measured using a concentric ring
electrode (Figure 11.30).

Equipment required:
● 10/100 V resistance meter with test leads
● Concentric ring electrode
● ESD protective packaging material to be tested
● Insulating support

Procedure:
● Place the sample packaging upon the insulating support.
● Place the concentric ring electrode on the packaging surface.
● Connect the resistance meter to the electrode.
● Measure the resistance at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.

Common Problems
In practice, with high resistance, materials readings do not fully stabilize in a short time.
In this circumstance, the reading should be taken after an appropriate electrification time,
e.g. 15 seconds after application of the test voltage.
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 379

Figure 11.31 Measurement of surface resistance of small or curved ESD packaging using a
two-pin probe.

The electrode needs a flat surface larger than the electrode. For profiled or curved surfaces
and small items, the surface resistance is not correctly measured as the electrode contact
area is reduced.

11.8.4.2 Point-to-Point Resistance of Small Packaging Items


Small or curved packaging surfaces can be measured using a two-pin surface resistance
probe according to IEC 61340-2-3 or ESD STM11.13 (Figure 11.31).

Equipment required:
● Two-pin probe electrode with test leads
● 10/100 V resistance meter
● Insulating support
● ESD protective packaging to be tested

Procedure:
● Connect resistance meter to the electrode.
● Place the sample packaging upon the insulating support.
● Place the electrode in contact with the packaging surface.
● Measure result at 10 V.
● If R > 1 MΩ, measure the result again at 100 V.
380 11 ESD Measurements

Common Problems
ESD protective packaging materials can have considerable variation in their surface resis-
tance at different positions. This measurement is not directly equivalent to the concentric
ring electrode or point-to-point 2.5 kg resistance measurement electrode methods and will
give different results. This two-point electrode tends to give results showing the resistance
of highest resistance of the material under the pins (Smallwood 2017, 2018). As such, the
results may be highly variable compared to results found with the concentric ring electrode.
This may be an advantage or a disadvantage depending on the purpose of the test.
For a homogenous material, the two-point probe may give results about 2.5–5 times
higher than concentric ring electrodes.

11.8.4.3 Point-to-Point Resistance of Packaging Using 2.5 kg Resistance Measurement


Electrodes
The surface resistance of large packaging items can be measured using 2.5 kg resistance
measurement electrodes (Figure 11.32). This method is useful only on large flat packaging
surfaces.
The sample under test should be placed on an insulating support. The point-to-point resis-
tance of the support material should be at least 10 times the maximum acceptable resistance
of the sample under test.

Figure 11.32 Measurement of surface resistance ESD packaging using two 2.5 kg solid resistance
measurement electrodes.
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 381

The electrodes should be placed close together without risking touching each other to
give results similar to the concentric ring electrode.

Equipment required:

● Two 2.5 kg resistance measurement electrodes


● 10/100 V resistance meter with test leads
● Insulating support
● ESD protective packaging to be tested

Procedure:

● Connect resistance meter to the electrodes.


● Place the sample packaging upon the insulating support.
● Place the electrodes close together on packaging surface.
● Measure result at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.

Common Problems
This measurement is not directly equivalent to the concentric ring electrode or two-point
probe methods and will give different results. The results vary with the distance between
the electrodes. This arrangement is most likely to give results similar to the concentric ring
electrodes when the two electrodes are separated by only a few millimeters.

11.8.5 Volume Resistance of Packaging Materials


The volume resistance of a packaging material gives the resistance from the outer sur-
face to the inner surface. This can be measured using a concentric ring electrode. In this
case, the test voltage is applied using a metal plate electrode under the packaging material
(Figure 11.33). Resistance is measured to the inner electrode of a concentric ring electrode
placed on top of the material under test. The outer electrode ring can be used as a grounded
guard ring to eliminate surface conducted currents or can be left disconnected.

Equipment required:

● 10/100 V resistance meter with test leads


● Concentric ring electrode
● Metal plate electrode
● ESD protective packaging material to be tested
● Insulating support

Procedure:

● Place the metal plate electrode upon the insulating support.


● Place the sample packaging upon the metal plate electrode.
● Place the concentric ring electrode on the material surface to be measured.
382 11 ESD Measurements

Figure 11.33 Volume resistance measurement using a concentric ring electrode (left). The
electrodes are shown on the right. This concentric ring electrode requires a weight to give the
correct total mass.

● Connect the resistance meter to the electrodes. The applied voltage is normally connected
to the metal plate electrode, and the current sense terminal is connected to the inner ring
electrode. The outer ring electrode is optionally connected to the resistance meter guard
terminal, if available.
● Measure the resistance at 10 V.
● If R > 1 MΩ, measure result again at 100 V.
● Note the result.

Common Problems
In practice, with high resistance materials, the readings do not fully stabilize in a short time.
In this circumstance, the reading should be taken after an appropriate electrification time,
e.g. 15 seconds, after application of the test voltage.

11.8.6 ESD Shielding of Bags


Shielding packaging is defined as an enclosure that limits the passage of ESD current and
energy to the device within the bag, such that the maximum energy from a 1000 V human
body model (HBM) discharge is reduced to less than a limit specified in IEC 61340-5-3 or
ESD S541. It is tested by application of a standard 1 kV HBM ESD to the outside of the
enclosure, and measurement of the transient that appears at the position normally occupied
by a sensitive component (Figure 11.34).
The shielding test is rather specialist and is normally beyond the means of ESD packaging
users. It is normally performed as a type test on bags by manufacturers.
A capacitive sensor probe is placed in a central position in the bag under test. The sensor
is connected via a 500 Ω wide bandwidth resistor to a current sensor probe and fast digital
storage oscilloscope. The bag is placed on an earthed electrode, and an upper electrode is
placed on top of the capacitive sensor outside the bag. The upper electrode is connected to a
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 383

ESD generator
1kV HBM
Electrode
500 Ω

ESD shielding bag Capacitive Current probe


under test probe

Electrode
Oscilloscope

Figure 11.34 ESD bag shielding measurement.

standard HBM ESD generator. The applied waveform is similar in specification to that used
to test ESD withstand voltage of components.
When an HBM ESD is applied to the upper electrode, a small impulse is detected by the
capacitive electrode. A transient current flow through the resistor and current sensor and is
recorded by the digitizing oscilloscope. The energy, W, in the detected transient is calculated
from the digitized current samples I over n samples covering the duration the waveform.
n

W = 500 I 2 dt
0

11.8.7 Evaluation of ESD Shielding of Packaging Systems


There is no measurement test method for ESD shielding of packaging systems other than
bags. This is in part because such systems are highly variable in format. It would be difficult
to design a test method that would be generally usable with all packaging types.
ESD shielding packaging systems can be evaluated by the procedure given in the flow
chart of Figure 11.35. The surface resistance of the inner and outer packaging surfaces,
and the volume resistance, can be measured using the methods of Section 11.8.4 and
Section 11.8.5.

11.8.8 Measurement of Ionizer Decay Time and Offset Voltage


For ionizer qualification, ionizers can be placed in a measurement arrangement with a CPM
in which the position of the CPM and ionizer are standardized. For compliance verification
and measurement of ionizer performance in the workplace, the CPM should be placed in
positions on the workstation representative of likely positions of the items requiring charge
neutralization (Figure 11.36). Standard practices ANSI/ESD SP3.3 (EOS/ESD Association
Inc. (2016a)) and ANSI/ESD SP3.4 (EOS/ESD Association Inc. (2016b)) specify that the
CPM plate should face the direction of the air flow from the ionizer. In some cases, it may
be more realistic to have the CPM plate in an orientation that imitates the target to be neu-
tralized in practice.
384 11 ESD Measurements

Start

Yes
Is inner surface
insulative?

No

Is outer surface Yes


insulative?

Not acceptable as
No
ESD protective packaging
Yes
Is there a high resistance
barrier material or air gap
between outer and inner
surfaces?
Acceptable as ESD
shielding packaging No

Not acceptable as
ESD shielding packaging

Figure 11.35 Flowchart procedure for evaluation of ESD shielding properties of packaging.

Ionizer
CPM measurement unit

CPM plate

Figure 11.36 Measuring the decay time of an ionizer. The CPM plate is positioned simulating the
normal position of items requiring neutralization.
11.8 Standard Measurements Specified by IEC 61340-5-1 and ANSI/ESD S20.20 385

100 1200
0 1100
–5 0 5 10 15 20 25 1000
–100
900
CPM voltage (V)

CPM voltage (V)


–200
Offset 800
–300 Decay time from voltage
negative voltage 700
–400 Offset
600 voltage
–500
500
–600 400 Decay time from
–700 positive voltage
300
–800 200
–900 100
–1000 0
–5 0 5 10 15 20 25
–1100 –100
Time (seconds) Time (seconds)

Figure 11.37 Decay time and offset voltage of an ionizer.

Standard tests require an initial CPM plate voltage of over 1000 V and measures the decay
time from 1000 V to a final voltage of 100 V (Figure 11.37). The offset voltage is measured
after the plate has reached a reasonably constant voltage sometime after the decay curve.
Some fluctuation of the offset voltage is normal.

Equipment required:
● Charge plate monitor
● Ionizer under test

Procedure:
● Set up the ionizer in its intended working position.
● Set up the charge plate monitor in a position typical of items requiring neutralization.
● Charge the plate and observe the decay of voltage.
● Measure the decay time between the initial voltage and the final voltage required by the
test.
● Note the residual “offset” voltage after it has stabilized.

11.8.8.1 Common Problems


The effectiveness of ionizers is dependent on position, drafts, and orientation of the ionizer
and CPM. It may be advisable to try different CPM positions and orientations and evaluate
these effects.
Standard practices SP3.3 and SP3.4 specify that the CPM plate should face the direction
of the air flow from the ionizer. In many cases, it may be more realistic to have the CPM
plate in a position simulating the position of a typical target to be neutralized. For example,
neutralization of a printed circuit board (PCB) resting on a surface by an ionizer airstream
from the side may be better simulated by a CPM plate lying parallel to the surface rather
than one at right angles to the surface.
386 11 ESD Measurements

11.8.9 Walk Test of Footwear and Flooring


Body voltage measurement while walking is a test used mainly for qualification of the per-
formance of ESD control footwear and flooring system.
Body voltage measurements can be made using a CPM or a purpose-built electro-
static body voltage measurement instrument (Figure 11.38). An ordinary multimeter
or voltmeter cannot be used as the input resistance would be too low (typically 10 MΩ)
effectively grounding the body through the meter. It is useful to have the CPM or body
voltage measurement instrument connected to a recording system such as a computer for
convenient display and logging of the results. If this is available, recordings of body voltage
can be made for future reference and analysis. The subject is given a handheld electrode
that connects them via a wire to the CPM plate or instrument.

Equipment required:
● CPM or body voltage measurement instrument
● Handheld electrode and long wire
● Computer or recording equipment (optional)

Procedure:
● Set up the CPM in a convenient position.
● Connect the long wire to the hand-held electrode and CPM plate.
● Ask the subject to hold the electrode and walk around. (Note: ESD STM97.1 requires a
specific step pattern to be used.)
● Monitor the voltage readings and record the “peaks” of positive or negative polarity.
● Calculate the average of the five highest peaks (if required).

11.8.9.1 Common Problems


It can be difficult to record peaks and valleys without a chart recorder, computer, or some
other graphical means of recording and display. Some equipment is available that has
built-in decay curve timing functions.
Digital instruments often update their display every half or one second. Peaks are easily
missed when monitoring these displays.

Figure 11.38 Measurement of body voltage of personnel (left) using electrostatic voltmeter
instrument and (right) using CPM.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 387

11.9 Useful Measurements Not Specified by IEC 61340-5-1


and ESD S20.20
Although they may be covered by standards, the following test methods are not specified
by the current versions of the 61340-5-1 and ESD S20.20 ESD standards for measurements
of equipment and materials used in ESD control. Nevertheless, they can be useful in ESD
control work, especially in situations where standard measurements are not appropriate.
They may also be used to give additional information about the performance of equipment
or materials.

11.9.1 Electrostatic Fields and Voltages


Measurements of electrostatic fields and voltages are an important way of checking that
all static measures are operating correctly and of detecting any unforeseen static sources.
The most common type of instrument used is an electrostatic field meter calibrated to read
voltage on large flat surfaces (Section 11.6.12.1). Other types of instrument for measurement
of voltages include noncontact and contact electrostatic voltmeters (Sections 11.6.12.2 and
11.6.12.3).

11.9.2 Measurement of Electric Fields at the Position of the ESDS


An electrostatic field meter can be used to measure the electrostatic field at the likely posi-
tions of the ESDS (Figure 11.39).

Equipment:
● Electrostatic field meter
● Grounding wire (if required)
● Workstation position to be evaluated

Figure 11.39 Measurement of electrostatic fields in the region where an ESDS device may be
present. The field meter is held at different positions and orientations to look for electrostatic field
sources and measure the strength of fields.
388 11 ESD Measurements

Procedure:
● Ground the electrostatic field meter. (Many field meters can be grounded simply by being
held in the hand of a grounded person.)
● Move the field meter around the region in which the ESDS may be situated and monitor
the readings.
● Note any high field readings and the positions of their sources.
When high field readings are noted, the field meter can be used to home in on the source
for further evaluation.

11.9.2.1 Common Problems


Many meters do not give a direct reading of electrostatic field but are calibrated to read
surface voltage of a large flat surface held at a set distance from the field meter. A meter
calibrated in this way can still be used to detect electrostatic fields and determine whether
they are greater or less than a required level, e.g. 5 kV m−1 . To do this, the voltage reading
that is given by a field of 5 kV m−1 must be known.
To a first approximation, this reading can be easily obtained by simple calculation using
the calibration distance d of the field meter. The voltage reading V E equivalent to a field E
is given by
VE = Ed
So, for example, the JCI 140 field meter shown in Figure 11.41 is calibrated to read surface
voltage with the meter 10 cm from the field source. If the required field limit is 5 kV m−1 , the
corresponding meter reading is 0.1 × 5000 = 500 V. When the field meter is used to evaluate
electrostatic fields, if the meter reading remains less than this value, the electrostatic field
is less than 5 kV m−1 .
The voltage reading equivalent to the field limit will be different for other meters of dif-
ferent calibration distance. For example, for a meter having a calibration distance of 2.5 cm
(1 in.), the voltage reading for a 5 kV m−1 field would be 0.025 × 5000 = 125 V.

11.9.3 Measurement of Surface Voltages on Large Objects Using an


Electrostatic Field Meter Calibrated as a Surface Voltmeter
Many electrostatic voltmeters or “static detectors” are electrostatic field meters that have
been calibrated to read voltage measured on a large flat surface when held at a specified
distance from the surface (Figures 11.40 and 11.41).
These meters give a response reflecting the effect of the charge over a wide area of the
surface. While the meters will give readings from small or curved surfaces, the reading will
in this case be less than the true surface voltage.
It is essential to know the calibration distance at which the voltmeter must be held from
the surface being measured. In Figure 11.40, the meter has two pins that are held against
the target to keep the meter at the correct distance.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 389

Calibration
distance Voltage (V) = electrostatic field (E) x distance (d)

Electrostatic
field meter
Charged
surface
voltage V

Many instruments are designed


to be earthed via the hand of an
earthed user

Figure 11.40 Measuring electrostatic voltages on a planar target surface using an electrostatic
field meter calibrated as a voltmeter.

Figure 11.41 Using a field meter calibrated as a voltmeter to measure electrostatic surface
voltages on a charged object.
390 11 ESD Measurements

Equipment:
● Electrostatic voltmeter
● Grounding wire (if required)
● Objects or material to be tested

Procedure:
● Ground the electrostatic voltmeter. (Many meters can be grounded simply by being held
in the hand of a grounded person.)
● Hold the voltmeter the correct calibration distance from the surface to be measured.
● Take a reading.

Common Problems
The electrostatic voltmeter must be grounded, or it will not read the target voltage correctly.
Any voltage on the meter is added to the voltage read on the target surface. The reading will
also change if not held at the correct distance from the surface being measured. If too close,
the reading will be too high.
Do not hold a sample in your hand for measurements – if it conducts electricity, any
charge would be conducted to your body. If your body is not grounded, any voltage on your
body could appear also on the sample.
The voltage on small objects, curved objects, insulators, or small isolated conductors is
not correctly measured. The voltage on insulators and small isolated conductors changes
with increased capacitance due to the presence of the meter.

11.9.4 Measurement of Voltage on Devices or Small Conductors


The voltage on a small item can be measured using a noncontact voltmeter (Figure 11.42)
or ultra-high input resistance contact voltmeter (Figure 11.43).

11.9.4.1 Measurement Using a Non-contact Voltmeter


A noncontact electrostatic voltmeter can be used to measure the voltage on a small item
under test.

Equipment:
● Noncontact electrostatic voltmeter
● Grounding wire (if required)
● Objects or material to be tested

Procedure:
● Ground the electrostatic voltmeter. (Many meters can be grounded simply by being held
in the hand of a grounded person.)
● Hold the voltmeter within the correct calibration distance range from the surface to be
measured.
● Note the reading.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 391

Figure 11.42 Measurement of voltage of a small item using a noncontact electrostatic voltmeter.

Figure 11.43 Measurement of voltage of a small item using a contact electrostatic voltmeter.
Source: DE Swenson.
392 11 ESD Measurements

11.9.4.2 Measurement Using a Contact Voltmeter


A high impedance contact electrostatic voltmeter can be used to measure the voltage on an
item under test without significantly discharging it.

Equipment:
● Contact electrostatic voltmeter
● Grounding wire (if required)
● Objects or material to be tested

Procedure:
● Ground the electrostatic voltmeter. (Many meters can be grounded simply by being held
in the hand of a grounded person.)
● Touch the voltmeter tip to the surface to be measured.
● Note the reading.

11.9.5 Resistance of Tools


11.9.5.1 Resistance from Tool Tip to Handle
This test measures the resistance of a tool from the point that contacts the ESDS device to
the handle (Figure 11.44).

Figure 11.44 Measurement of resistance of hand tool.


11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 393

Equipment required:
● 10/100 V high-resistance meter
● Test leads and clips
● Tool to be tested

Procedure:
● Connect the test leads and clips to the resistance meter.
● Connect one test lead to the tool bit.
● Connect the second test lead to the tool handle.
● Measure result at 10 V.
● If the resistance is >1 MΩ, measure result again at 100 V.

Common Problems
One problem with this test is that it can be difficult to make good electrical contact with
a handle made from hard high-resistance material. Contact can be improved by wrapping
a conductive self-adhesive metal tape around the handle to provide a larger area contact.
Measuring the tool as held in the hand gets around this problem.

11.9.5.2 Resistance to Ground of Handheld Tool


A more convenient tool resistance to ground system test can be done by a grounded user
holding the tool in their hand. A resistance meter is connected to ground, with its second
terminal connected to a touch plate (Figure 11.45). To test the tool system resistance to
ground, the user holds the tool handle and touches the bit to the touch plate and reads the
resistance to ground from the meter. Holding the tool in the hand makes a good connec-
tion to the handle and a realistic test. A test electrode can be made from any convenient
metal item isolated from ground by an insulating support. Avoid touching the tool tip, or
the resistance of the handle is short circuited.
This test is a convenient system test that can easily be done by a grounded user at a
workstation. It directly confirms the tool is grounded within required resistance limits. It is
important to remember that the resistance result includes the resistance of the operator’s
ground path as well as the tool resistance.

Equipment required:
● 10/100 V resistance meter
● Wrist strap or footwear-flooring grounding for the test person
● Tool test electrode

Procedure:
● Connect one terminal of the resistance meter to ground.
● Connect the second terminal of the resistance meter to the tool test electrode.
● Wear the grounded wrist strap and hold the tool.
394 11 ESD Measurements

Figure 11.45 Test of resistance to ground of hand-held tool. The person holding the tool can be
grounded in their usual manner.

● Touch the tool to the test electrode.


● Measure the result at 10 V.
● If the resistance is >1 MΩ, measure result again at 100 V.

Common Problems
If the user’s hand touches the metal tool bit, the resistance of the handle is bypassed by the
resistance of the user’s hand, and a false low resistance reading is obtained.

11.9.6 Resistance of Soldering Irons


11.9.6.1 Resistance to Groundable Point of Soldering Iron Tip
The resistance from a soldering iron tip to the earthing point (e.g. earth pin on the mains
plug) can be measured while the iron is not in use (Figure 11.46). The resistance meter used
must be capable of low-resistance measurements.

Equipment required:
● Low-resistance meter
● Test leads and clips

Procedure:
● Connect one terminal of the resistance meter to the soldering iron groundable point.
● Connect the second terminal of the resistance meter to the soldering iron tip.
● Measure the resistance result.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 395

Figure 11.46 Measurement of the resistance to groundable point of a soldering iron bit.

Common Problems
Corrosion of the bit surface contact with the iron can give variable results in this test. But,
detection of this is one of the reasons for doing the test!

11.9.6.2 Resistance to Ground of a Soldering Iron Tip


The resistance from soldering iron tip to ground can be easily measured while in use,
using a tool test electrode (Figure 11.47). The resistance meter used must be capable of
low-resistance measurements. A 0 Ω earthing connector must be used to connect to the
EPA ground.

Equipment required:

● Low-resistance meter
● Test leads and 0 Ω earthing connector
● Tool test electrode

Procedure:

● Connect one terminal of the low-resistance meter to EPA ground.


● Connect the second terminal of the resistance meter to the tool test electrode.
● Touch the soldering iron bit to the tool test electrode.
● Measure the result.

Common Problems
Corrosion of the bit contact with the iron can give variable results in this test – but then
detecting this part of the reason for doing the test!
396 11 ESD Measurements

Figure 11.47 Measurement of resistance to ground of a soldering iron tip while in use.

11.9.7 Resistance of Gloves or Finger Cots


It is often useful to measure the resistance through a glove while it is worn to determine
whether handheld objects will be isolated from the grounded hand. There are several con-
venient ways of doing this. A standardized way of doing so is specified in ANSI/ESD SP15.1
using a constant area and force electrode (CAFE) and measuring the resistance between
the CAFE and a handheld electrode (Figures 11.48 and 11.49). Gloves can also be evaluated
using a charge plate monitor (see Section 11.9.8.2).

11.9.7.1 Measurement of Resistance Through a Glove Using a Handheld Electrode


A person wearing a wrist strap and glove can measure the system resistance through
the glove and wrist strap using a resistance meter, as shown in Figure 11.50. This is the
same measurement method as given for measurement of wrist straps as worn (see Section
11.8.3.2), except that in this case the test subject wears the glove under test.

11.9.7.2 Testing Resistance Through a Glove Using a Wrist Strap Tester


If the glove has low enough resistance such that it falls within the pass range of a wrist strap
tester, this can be used to check the entire wrist strap – hand – glove system (Figure 11.51).
This will give a simple pass-fail result rather than a system resistance result. This is the
same measurement method as given for measurement of wrist straps as worn (see Section
11.8.3.2), except that in this case the test subject wears the glove under test.

Common Problems
Typical wrist strap testers have a fixed upper “fail” limit, often 35 MΩ. This upper limit
must correspond to the upper limit specified in the ESD control program for the resistance
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 397

Figure 11.48 Handheld (left) and CAFE (right) electrodes. Source: D E Swenson.

Figure 11.49 Measurement of resistance through glove to ground using CAFE electrode. Source: D
E Swenson.
398 11 ESD Measurements

Figure 11.50 Measurement of resistance through a glove using a handheld electrode.

Figure 11.51 Testing of system resistance through a glove and wrist strap using a wrist strap
tester. Source: D E Swenson.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 399

Figure 11.52 Measurement of resistance to ground through a glove using a handheld electrode.

of glove when worn. If the upper limit accepted by the ESD program for the glove is higher
than the “fail” threshold of the tester, gloves having resistance above the fail threshold are
“failed” incorrectly. If the upper limit accepted by the ESD program for the gloves is lower
than the “fail” threshold of the tester, gloves having resistance above the acceptable limit of
the ESD program, but below the “fail” threshold of the tester, are “passed” incorrectly.

11.9.7.3 Measurement of Resistance to Ground Through a Glove Using a Handheld


Electrode
A system resistance through the glove to ground can be measured by an operator who
is already grounded by either a wrist strap or footwear and flooring (Figure 11.52). The
resistance measurement is made between a handheld electrode and ESD ground. This is
the same measurement method as given for measurement of wrist straps as worn (see
Section 11.7.8.4).

11.9.8 Charge Decay Measurements


We are often interested in knowing whether charge can dissipate quickly enough to prevent
static build-up. There are many types of charge decay measurements that have been devised
for different purposes. They can use widely different equipment and principles of operation.
Charge decay test methods are not, at the time of writing, specified as standard test meth-
ods used by the ESD control standards 61340-5-1 and ESD S20.20. The IEC 61340-2-1 (Inter-
national Electrotechnical Commission 2015a) Section 4.4 gives a contact test method for
charge decay measurement using a CPM. This technique gives useful tests for tools, gloves,
or finger cots, as shown in Sections 11.9.8.1–11.9.8.3.
In these tests, the charged pate monitor plate is charged to a defined starting level V i ,
often 1000 V. The voltage on the plate is observed to drop (decay) during the test toward
400 11 ESD Measurements

Initial voltage Vi

Voltage
curve

Decay time

Decayed
voltage Vf

0 Time t

Figure 11.53 Charge decay time.

zero. This typically gives a quasi-exponential voltage curve (Figure 11.53). The time taken
for the voltage to reach a defined level V f , often chosen to be 100 V, is measured.
Slow decay curves can sometimes be monitored manually, but faster decay curves may
need a digitizing oscilloscope or chart recorder to aid measurement. The time taken for
the voltage to decay from the chosen initial voltage (V i in Figure 11.53) to the chosen final
voltage V f , is taken as the decay time.
The user should select the final voltage as appropriate for their purpose. Often a fraction
of the initial voltage (e.g. 0.1 Vi ) or a “hazard threshold voltage” (e.g. 100 V) is specified.
Typically, the charge decay time result will depend on the initial test voltage, perhaps
varying in some cases by orders of magnitude. Longer decay times are usually found for
lower test voltages.
A commercial CPM can be conveniently used for charge decay tests. This is often limited
in test voltage to the 1000 V and final voltage of 100 V designed for use with ionizer testing.
Many charge decay test methods are in effect indirect ways of comparing the resistance
to ground of the item measured. The resistance to ground of the item bleeds the charge
from the plate capacitance Cp . When touched by a handheld tool, the resistance to ground
of the tool Rt is connected in parallel with the plate capacitance Cp . The voltage decays with
a time constant Rt Cp . The decay time measured from 1000 to 100 V is approximately 2 Rt
Cp . Where the resistance is very high, this technique can be more reproducible than using
a direct resistance measurement method.

11.9.8.1 CPM Charge Decay for Tools


A handheld tool held by a grounded operator can be tested using the charge decay method
(Figure 11.54) using a CPM. This method works well with high-resistance tool handles.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 401

Figure 11.54 Measurement of charge decay time of tools. The tool is touched to the CPM plate
and the voltage decay time is measured. Both the operator and CPM are grounded.

Equipment required:
● Charge plate monitor
● Wrist strap or footwear – flooring grounding of operator
● Sample tools for test

Procedure:
● The operator must be grounded.
● Charge the CPM to 1000 V.
● The operator holds the tool in their hand but does not touch the tool blade. The tool blade
is then touched to the CPM.
● Observe and time the decay to 100 V.
It is helpful if the decay curve is recorded, as this can reveal the true charge decay char-
acteristics of the tool and be used to document the test for product qualification records.
Some examples of typical waveforms are given in Figures 11.55–11.57. A tool that has a low
resistance will give a fast drop in voltage to zero (Figure 11.55). A tool with high-resistance
handle will show a longer, slower decay time (Figure 11.56). An initial fast drop may be
seen in this case due to increasing capacitance and charge sharing as the tool bit contacts
the CPM plate.
402 11 ESD Measurements

1200
Tool touches CPM plate
1000
CPM voltage (V)

800

600

400

200

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (seconds)

Figure 11.55 Decay curve from tool that has low-resistance handle.

1200
Tool touches CPM plate
1000
CPM voltage (V)

800

600

400

200

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (seconds)

Figure 11.56 Decay curve from tool that has intermediate-resistance handle.

1200
Tool touches CPM plate
1000
CPM voltage (V)

800

600 Drop due to charge sharing between CPM plate


and tool
400

200

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (seconds)

Figure 11.57 Decay curve from tool that has insulative handle and shows capacitive voltage
reduction.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 403

If the tool handle is insulative, the plate voltage does not reduce to zero in an acceptable
time. Again, there may be a step in voltage as the tool bit contacts the plate as the charge
stored on the plate is shared with the tool capacitance (Figure 11.57).

Common Problems
The metal tool blade should not be touched with the fingers, as the resistance of the handle
will be bypassed.
Typical digital displays on CPM instruments may update only every half-second or so.
Short decay times of a second or less are difficult to measure without monitoring the plate
voltage on an oscilloscope type display.

11.9.8.2 Charge Decay of Gloves and Finger Cots


A charge decay test like that for tools using a CPM can also be used for gloves and finger
cots (Figure 11.58).

Equipment required:

● Charge plate monitor


● Grounded person for test
● Sample glove or finger cot for test

Procedure:

● Charge the CPM to 1000 V.


● The grounded person wears the glove/finger cot under evaluation and touches the CPM
through the glove/finger cot.
● Observe the decay time to 100 V.
● Remove the finger from the plate. Observe whether the plate charges due to contact with
the glove material.

Figure 11.58 Measuring the charge decay of a glove when worn. The gloved finger is brought into
contact with the charged plate and the decay observed.
404 11 ESD Measurements

Figure 11.59 Charge decay system test of tool held in gloved hand.

Common Problems
Typical digital displays on CPM instruments may update only every half-second or so. Short
decay times of a second or less are difficult to measure without monitoring the plate voltage
on an oscilloscope type display.
When the gloved finger is removed from the CPM plate, the plate may become charged.
This may indicate unacceptable charging of the plate by the glove material.

11.9.8.3 System Test of Glove and Handheld Tool


A system test can be done of grounding of a handheld tool via a gloved hand (Figure 11.59),
using the tool charge decay test method of Section 11.9.8.1. A decay curve like Figures 11.55
or 11.56, with decay time within accepted values, indicates acceptable performance.

Common Problems
Make sure the fingers do not accidentally contact the tool bit, short-circuiting the handle.

11.9.9 Faraday Pail Measurement of Charge on an Object


11.9.9.1 The Faraday Pail
In its simplest form, the Faraday pail is a metal pail isolated from ground that can be used
with a coulombmeter or electrostatic voltmeter to measure charge on an item placed within
the pail. The pail must be sufficiently large to fully contain the item. A charge is induced
on the pail that is equal to the total net charge on the item placed within the pail. Sim-
ple unshielded Faraday pails can be improvised from common metal storage containers
(Figure 11.60).
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 405

Figure 11.60 Unshielded Faraday pail on CPM plate. This pail was purchased in the kitchen
department of a local store!

Simple unshielded Faraday pails are prone to errors from induced charge due to fields
from nearby charged insulators, personnel, or voltage sources. This can be reduced by con-
taining the measurement pail isolated within an earthed metal screen (Figure 11.61).
Where a Faraday pail is used with a coulombmeter, the charge is measured directly by
the coulombmeter.
Where a Faraday pail is used with a CPM or electrostatic voltmeter, the charge gives a
voltage rise on the pail that is measured using the voltmeter or CPM. For comparative mea-
surements (e.g. comparing the charging of an item when handled by two different glove
types), it may be sufficient to compare the voltages produced on the pail. If the actual charge
values Q is required, it can be calculated, knowing the capacitance Cp of the pail in its mea-
surement arrangement and the voltage result V, from the relation.
Q = Cp V
When used with a voltmeter or CPM, a Faraday pail must usually be zeroed by momentary
connection to ground to bring the pail to zero volts. When used with a coulombmeter, the
coulombmeter must be zeroed before each measurement.

11.9.9.2 Measurement of Electrostatic Charging of Items Handled with Gloves or


Finger Cots Using a Faraday Pail
It can be useful to measure the charging of ESDS items handled while wearing gloves or fin-
ger cots. IEC/TR 61340-2-2 (International Electrotechnical Commission 2000) gives general
guidance on chargeability testing.
406 11 ESD Measurements

Figure 11.61 Shielded Faraday pail used with a coulombmeter.

The principle of these tests is that a grounded person, wearing the glove or finger cots
under test, handles the item and then places them into a Faraday pail (see Section 11.9.9.1)
or on to a CPM (see Section 11.9.9.3).
If an unscreened Faraday pail is used, electrostatic fields from charged clothing, gloves,
or other insulating items in the vicinity can strongly affect the measurement result. In this
case, make sure that all potentially charged items are moved well away from the measure-
ment area and that the operator moves away before the measurement result is noted. If a
screened Faraday pail is used, this is much less affected by nearby electrostatic fields.

Equipment required:
● Faraday pail connected to a voltmeter or charge measurement instrument
● Wrist strap grounder
● Sample glove or finger cot for test

Procedure:
● Wear the wrist strap, and ground it.
● The subject holds the product in their hand and handles it in a representative way.
● If necessary, zero the Faraday pail and measurement instrument.
● Place the product in the Faraday pail.
● Move the hand well away from the Faraday pail.
● Note the charge reading.
11.9 Useful Measurements Not Specified by IEC 61340-5-1 and ESD S20.20 407

Common Problems
Touching the Faraday pail can cause errors due to charging or discharging of the pail by
contact.
An unshielded Faraday pail is prone to errors due to induced charge from nearby electro-
static field sources such as charged clothing, the user’s body, or insulators. Proximity to the
user’s hand, body, or other conductors can increase capacitance and reduce the readings
obtained.
Placing a charged ESDS in a metallic Faraday pail or on a metal CPM plate can cause
charged device ESD that could damage the ESDS. This can be prevented by lining the Fara-
day pail or CPM plate with static dissipative material having surface resistance >10 kΩ.

11.9.9.3 Evaluation of Charging of an Item Using a CPM Plate


Indicative and comparative evaluation of charging of an item can be made using a CPM.
In this case, the charged item is placed on the CPM plate instead of into the Faraday pail
(Figure 11.62). The CPM plate is first discharged by grounding it before placing the charged

Figure 11.62 Measurement of product charging using a CPM.


408 11 ESD Measurements

product on it. The voltage on the CPM after placing the product on it is indicative of the
charge on the product. Different gloves can be compared using repeated tests handling the
same item.
The charge induced on the CPM plate is an unknown fraction K of the charge on the item
measured. If the CPM plate capacitance Cp is known, the charge Q on the item measured
can be estimated from the CPM plate voltage V from the relation.
Q = KCp V
The unknown factor K must usually be assumed to be K = 1 unless it can be somehow
evaluated.

Common Problems
Touching the CPM plate can cause errors due to charging or discharging of the pail by
contact.
The CPM plate is prone to errors due to induced charge from nearby electrostatic field
sources such as charged insulators. Proximity to the user’s hand, body, or other conductors
can increase capacitance and reduce the readings obtained.
Placing a charged ESDS device on a metal CPM plate can cause charged device ESD that
could damage the ESDS. This can be prevented by lining the CPM plate with static dissipa-
tive material.

11.9.10 ESD Event Detection


ESD event detectors react to radio frequency electromagnetic radiation emitted by ESD
in the vicinity. They can be used to detect ESD occurring nearby, including in working
automated equipment. Some detectors have switched settings to distinguish between
charged device and human body ESD. Examples of handheld ESD detectors are shown in
Figure 11.63.

Equipment required:
● ESD event detector

Procedure:
● Switch on the ESD event detector and place it near the operation being investigated.
● Watch the response of the event detector during the process. Look for coincidence of an
ESD event with a contact made between a conductor and the ESDS.

Common Problems
ESD event detectors typically detect ESD from almost any source in the vicinity. These could
be from contactors or switches operating in equipment, room lighting being switched, or
other sources. Most of these sources are irrelevant to potential damage to the ESDS. It can
be difficult to distinguish between potentially damaging ESD events and irrelevant ESD
events. Where the ESDS can be observed through the process, look for coincidence between
an ESD event and a contact between the ESDS device and a conductor. If the ESDS cannot
References 409

Figure 11.63 Examples of handheld ESD event detectors.

be seen during the part of the process where ESD was noted, it may be necessary to review
the process steps while not in operation.

References

British Standards Institute. (2016) BS EN 61340-5-1. Electrostatics - Part 5-1: Protection of


electronic devices from electrostatic phenomena - General requirements.
EOS/ESD Association Inc. (2013a) ANSI/ESD S1.1-2013. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Wrist Straps. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2013b) ANSI/ESD STM2.1-2013. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Garments. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2013c) ANSI/ESD STM7.1-2013. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Floor Materials – Resistive
Characterization of Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2013d) ANSI/ESD STM12.1-2013. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Seating – Resistance
Measurement. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2014a) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
410 11 ESD Measurements

EOS/ESD Association Inc. (2014b) ANSI/ESD STM9.1-2014. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Footwear – Resistive Characterization.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015a) ANSI/ESD STM3.1-2015. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Ionization. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2015b) ANSI/ESD STM11.11-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items – Surface Resistance Measurement of
Static Dissipative Planar Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015c) ANSI/ESD STM11.12-2015. ESD Association Standard for
Protection of Electrostatic Discharge Susceptible Items. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015d) ANSI/ESD STM11.13-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Two-Point Resistance
Measurement. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015e) ANSI/ESD S13.1-2015. Provides electrical soldering/
desoldering hand tool test methods for measuring current leakage, tip to ground reference point
resistance, and tip voltage. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2015f). ANSI/ESD STM97.1-2015. ESD Association Standard Test
Method for the Protection of Electrostatic Discharge Susceptible Items – Floor Materials and
Footwear – Resistance Measurement in Combination with a Person. Rome, NY, EOS/ESD
Association Inc.
EOS/ESD Association Inc. (2016a) ANSI/ESD SP3.3-2016. Standard Practice for the Protection
of Electrostatic Discharge Susceptible Items – Periodic Verification of Air Ionizers. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016b) ANSI/ESD SP3.4-2016. Standard Practice for the Protection
of Electrostatic Discharge Susceptible Items – Periodic Verification of Air Ionizer Performance
Using a Small Test Fixture. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016c) ANSI/ESD STM97.2-2016. Standard Test Method for the
Protection of Electrostatic Discharge Susceptible Items – Floor Materials and Footwear – Voltage
Measurement in Combination with a Person. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2017) ANSI/ESD STM4.1-2017. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Worksurfaces – Resistance
Measurements. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2018a) ESD TR53-01-18. Technical Report for the Protection of
Electrostatic Discharge Susceptible Items – Compliance Verification of ESD Protective
Equipment and Materials. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2018b) ANSI/ESD STM11.31-2018. ESD Association Standard Test
Method for Evaluating the Performance of Electrostatic Discharge Shielding Materials – Bags.
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2018c) ANSI/ESD S541-2018. Packaging Materials for ESD Sensitive
Items. Rome, NY, EOS/ESD Association Inc.
European Committee for Electrotechnical Standardisation (CENELEC). (1992) EN 100015-1.
Basic specification. Protection of electrostatic sensitive devices. Harmonized system of quality
assessment for electronic components. Basic specification: protection of electrostatic sensitive
devices. General requirements. Brussels, CENELEC.
References 411

International Electrotechnical Commission. (2001) IEC 61340-4-3:2001. Electrostatics –


Part 4-3: Standard test methods for specific applications – Footwear. Geneva, IEC.
International Electrotechnical Commission. (2000) IEC 61340-2-2:2000. Electrostatics –
Part 2-2: Measurement methods - Measurement of chargeability. Geneva, IEC.
International Electrotechnical Commission. (2004) IEC 61340-4-5:2004. Electrostatics –
Part 4-5: Standard test methods for specific applications – Methods for characterizing the
electrostatic protection of footwear and flooring in combination with a person. Geneva, IEC.
International Electrotechnical Commission. (2014) IEC 61340-4-8:2014. Electrostatics –
Part 4-8: Standard test methods for specific applications – Electrostatic discharge
shielding – Bags. Geneva, IEC.
International Electrotechnical Commission. (2015a) IEC 61340-2-1:2015. Electrostatics -
Part 2-1: Measurement methods - Ability of materials and products to dissipate static electric
charge. Geneva, IEC.
International Electrotechnical Commission. (2003 and 2015b) IEC
61340-4-1:2003+AMD1:2015 CSV. Electrostatics -
Part 4-1: Standard test methods for specific applications - Electrical resistance of floor coverings
and installed floors. Geneva, IEC.
International Electrotechnical Commission. (2015c) IEC 61340-4-6:2015. Electrostatics -
Part 4-6: Standard test methods for specific applications - Wrist straps. Geneva, IEC.
International Electrotechnical Commission. (2015d) IEC 61340-5-3:2015. Electrostatics -
Part 5-3: Protection of electronic devices from electrostatic phenomena - Properties and
requirements classification for packaging intended for electrostatic discharge sensitive devices.
Geneva, IEC.
International Electrotechnical Commission. (2016a) IEC 61340-2-3:2016. Electrostatics.
Part 2-3: Methods of test for determining the resistance and resistivity of solid materials used to
avoid electrostatic charging. Geneva, IEC.
International Electrotechnical Commission. (2016b) IEC 61340-4-9:2016. Electrostatics -
Part 4-9: Standard test methods for specific applications – Garments. Geneva, IEC.
International Electrotechnical Commission. (2016c) IEC 61340-5-1:2016. Electrostatics -
Part 5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Electrotechnical Commission. (2017) IEC 61340-4-7:2017. Electrostatics -
Part 4-7: Standard test methods for specific applications – Ionization. Geneva, IEC.
International Electrotechnical Commission. (2019) IEC TR 61340-5-4:2019. Electrostatics -
Part 5-4: Protection of electronic devices from electrostatic phenomena – Compliance
Verification. Geneva, IEC.
Smallwood, J. (2017). A practical comparison of surface resistance test electrodes. J. Electrostat.
88: 127–133.
Smallwood J. (2018) Paper 4B3. Comparison of surface and volume resistance measurements
made with standard and nonstandard electrodes. In: Proc. EOS/ESD Symp. EOS-40 Rome,
NY, EOS/ESD Association Inc.
412 11 ESD Measurements

Further Reading

EOS/ESD Association Inc. (2012) ANSI/ESD STM4.2-2012. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – ESD Protective Worksurfaces – Charge
Dissipation Characteristics. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016) ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (1999) ESD TR15.0-01-99. Standard Technical Report for the
Protection of Electrostatic Discharge Susceptible Items-ESD Glove and Finger Cots. Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2019) ANSI/ESD SP15.1-2019. ESD Association Standard Practice
for the Protection of Electrostatic Discharge Susceptible Items – In-Use Resistance Measurement
of Gloves and Finger Cots. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2019) ANSI/ESD STM9.2-2019. ESD Association Standard for the
Protection of Electrostatic Discharge Susceptible Items – Footwear – Foot Grounders Resistive
Characterization. Rome, NY, EOS/ESD Association Inc.
International Electrotechnical Commission. (2018) IEC TR 61340-5-2:2018. Electrostatics – Part
5-2: Protection of electronic devices from electrostatic phenomena - User guide. Geneva, IEC.
Smallwood, J. (2005). Standardisation of electrostatic test methods and electrostatic discharge
prevention measures for the world market. J. Electrostat. 63 (6–10): 501–508.
Vermillion R. (2016) Testing methods for ESD control packaging products. Controlled
Environments. Available from: https://www.cemag.us/article/2016/02/testing-methods-esd-
control-packaging-products [Accessed 6th June 2019].
413

12

ESD Training

12.1 Why Do We Need ESD Training?


Training is needed so that all personnel who have some role in electrostatic discharge (ESD)
control know enough to fulfill their role successfully and reliably. For effective ESD control,
unaccompanied entry into active electrostatic discharge protected areas (EPAs) should be
restricted to personnel who have received sufficient ESD training to make sure that they
will maintain ESD control procedures according to their role.
According to Snow and Dangelmeyer (1994), untrained personnel account for most
deviations from compliance, while trained personnel account for few noncompliances.
When employees received training, the number of deviations in those work areas reduced
dramatically.
Personnel may need some form of ESD training for many reasons such as
● To recognize ESD-sensitive (ESDS) devices and EPAs, ESD packaging, and ESD control
equipment
● To know what ESD control equipment, materials, and procedures to use
● To understand why they need to use them
● To understand where and when to use them
● To know how to test personal wrist straps and footwear and other ESD control equipment
according to their role
● To recognize and, if possible, prevent common noncompliances
● To know how to avoid creating noncompliances
● Any safety issues (e.g. high voltage precautions) or use of correct personal protective
equipment
● To know what to do when the equipment fails
● To know who to go to for advice
● New techniques, processes, and equipment when they arise
● Specific ESD control requirements or unusual practices used in exceptional EPAs or sit-
uations
● Basic electrostatics and ESD knowledge appropriate to their needs
Few organizations have proven documented ESD failure data, and even fewer can trace
failure to specific ESD control issues. Where loss of ESD control happens, ESD damage is a
risk rather than a certainty. Damage occurs only when ESD occurs to an ESDS device and
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
414 12 ESD Training

has sufficient strength and is from a potentially damaging source. Only a small proportion
of the ESDS device handled will fail, and the failures may be identified during test at a later
stage.
Personnel rarely then get the firsthand feedback that loss of ESD control causes ESD fail-
ures. Static electricity and ESD are rarely noticeable by sight or sound, and neither can they
be felt unless at very high levels. Furthermore, static electricity is not always present in sig-
nificant levels – it appears and disappears as materials are handled and moved, and even
in response to the weather! There are many ways in which static charge is harmlessly dis-
sipated before it can cause damage – ESD may occur only when the many factors involved
conspire to prevent static charge dissipation.
If ESD damage occurs, it is usually detected at a test stage long after the damage was
caused. Even if it is recognized as an ESD failure, the action or control failure that caused
it is usually not obvious.
These factors tend to promote skepticism that ESD damage is a real issue and that ESD
control measures are necessary and can make a real difference.
ESD training often seems to be trying to convince trainees of a scarcely believable sce-
nario, and overselling this can be self-defeating (McAteer 1980). Nevertheless, the challenge
is to give personnel the understanding they need to know when, where, and how to use ESD
control equipment and procedures effectively. They understand why they are doing this and
believe that it is important. An ungrounded person is probably the greatest ESD source in
manual handling processes. A trained person correctly using ESD control equipment and
procedures and preventing noncompliances is arguably a most effective first line of defense
against ESD risks.

12.2 Training Planning

The International Electrotechnical Commission (IEC) 61340-5-1 and American National


Standards (ANSI)/ESD S20.20 standards require that an ESD Training Plan is written to
document the ESD program’s training needs. Regardless of whether compliance with a stan-
dard is required, it is advisable to have a documented ESD Training Plan covering all aspects
of ESD training.
Snow and Dangelmeyer (1994) stated that the goal of training is to reduce deviations
to a minimal or zero level. They found that most employees are willing to comply with
procedures when offered the right training at the right time. They applied three principles
of the psychology of training and learning.

● Train only to affect a measurable change in work behavior


● Motivate students to improve learning
● Consider that students tend to forget information and skills that are not used regularly

Personnel who have different roles often need different training levels or content to fulfill
those roles. The Training Plan should document the following:

● The personnel who need ESD training


● The types and content of training required by different personnel
12.3 Who Needs Training? 415

● The need for training before handling ESDS


● The frequency of refresher training
● Policy for maintaining training records and where the records are kept
● Tests and methods used to ensure trainees have understood and can apply their training

12.3 Who Needs Training?


Any person who has a role in ESD control or needs to enter an EPA probably needs some
form of ESD training. This may be simple (e.g. do’s and don’ts list for visitors) or may need
considerable depth (e.g. principles and practice of ESD control, and compliance with stan-
dards, for the ESD coordinator).
As a minimum, all personnel who handle ESDS must have sufficient training to ensure
they understand and use the equipment and procedures required to prevent ESD damage
from occurring. They should receive this training before they start work that involves enter-
ing the EPA and handling ESDS devices.
Personnel also usually need periodic refresher training for various reasons. This is in part
to remind them of the ESD control practices they must use. Refresher training can also
help increase understanding and improve reliability of applying ESD control in practice.
Refresher training courses can also be an opportunity to update, explain, and communi-
cate any changes to the ESD control practice that may have been made since the previous
training was given.
Personnel who have particular job roles may also need specific training according to their
role. The specific training may be simpler than the basic ESD training course (as in the case
of a visitor, who may need simple instructions such as what to do, what not to do, and how
to wear and test personal grounding equipment), it may be more complex (e.g. for an ESD
coordinator or auditor), or it may be just different (e.g. for a cleaner). There may be many
reasons why personnel other than those who enter an EPA or handle ESDS devices may
also need some form of ESD training specific to their role. Some examples of role specific
ESD training are given next.
The ESD coordinator and other personnel involved in developing the ESD control pro-
gram may benefit from in-depth training in the principle and practice of ESD control and
compliance with ESD control standards. Update training may include review of standards
updates, trends in ESDS threats and ESD control techniques and practices, or attending
conferences and seminars.
Personnel who check, test, and audit an ESD control program may need specific training
in ESD measurements as well as auditing practices.
Cleaners who go into an EPA need specific instruction regarding things they must not
do, as well as on use of any specific cleaning techniques and materials they must use with
floors, bench surfaces, or other ESD control equipment.
Managers and supervisors who have budget responsibility for ESD control or who may
need to go into an EPA may need training specific to their role. For successful ESD control
implementation, management must be convinced that ESD is a real issue and the control
measures are necessary (McAteer 1980). Sadly, nearly 40 years after McAteer’s paper, this
can still be a challenge.
416 12 ESD Training

Managers who need to enter EPAs may also need brief instructions on the use of foot
straps, garments, or other ESD control equipment they will use in the EPA, as well as on
any activities or actions they must avoid (e.g. touching ESDS). If they accompany visitors
into the EPA, they may need instruction on supervision of these visitors.
Trainers who need to develop and present effective ESD training may themselves need
training on the specific ESD control practices current in their organization. They may also
need sufficient understanding of ESD control principles and practice to answer trainee
questions during training sessions. They may need training on how to present effective
demonstrations relevant to electrostatics and the ESD control program.
Personnel who purchase ESD-related items may need an overview of ESD control
practices and their impact on product specification and procurement. If the ESD control
program complies with a standard, they may need a working knowledge of the standard
requirements for sourcing ESD control equipment.
Subcontractors who go into an EPA may need either specific training according to their
activities or instructions on any areas and activities they must avoid.
Visitors who go into an EPA should normally be accompanied by trained personnel to
ensure they do not do anything that might compromise ESD control. Nevertheless, they may
need brief instructions on the use of foot straps, garments, or other ESD control equipment
they will use in the EPA, as well as on any activities or actions they must avoid (e.g. touching
ESDS).
Skeptical engineers who do not believe in ESD control can be among the most resistant
and unreliable in using ESD control measures. They can also have a negative effect on
others around them, reducing their confidence in the need for ESD control. Conversely,
engineers who have good understanding of ESD issues and control measures can be a
great asset in implementing an ESD control program, helping to develop an effective
ESD control program and helping others understand the importance and how to use ESD
control equipment.
One way of summarizing the roles of personnel that need training, and the training they
will need, is to present them as a matrix (Table 12.1). This can help plan training courses
and their content, as well as who will receive them.

Table 12.1 An example of a matrix of ESD training and personnel roles.

Training type
ESD Use of EPA Testing EPA Principles and practice ESD control
Personnel role awareness equipment equipment of ESD control for managers
√ √
Operators
√ √ √
Supervisors

Managers
√ √ √
Audit and test
personnel
√ √ √ √ √
ESD coordinators
12.4 Training Form and Content 417

12.4 Training Form and Content


12.4.1 Training Goals
Training should address an observable outcome that will help the attendee do their job
better. An example is to wear a wrist strap or ESD control foot straps correctly and test them
(Snow and Dangelmeyer 1994). Other training goals may be to explain how ESD damages
devices or recognize what materials might carry charge. The success of training may be
measured as how well the attendee does a task or answers questions.
Training course attendees are better motivated if they understand the purpose of the train-
ing and see the relevance to their work. Motivations can include
● Building a better or more reliable product
● Getting credit for doing a better job
● Avoiding frustration and possible blame when devices fail without known cause
● Reducing rework
● Avoiding the cost of failures
Skepticism and disbelief in ESD can be a real demotivating factor. Conversely, demonstra-
tion of real evidence of ESD issues in various forms can be highly motivating for attendees,
stimulating interest and a desire to find solutions to the issues. The demonstrations should,
where possible, be appropriate to the attendee’s job role. Various demonstrations are further
discussed in Sections 12.6 and 12.7, although it is best if the trainer devises demonstrations
tailored to the course and attendee’s roles and workplaces.
Snow and Dangelmeyer (1994) found that telling attendees how the learning goals will
help each person do a better job and build a better product gave positive motivation. Telling
when goals are met also helped to motivate.
Training content should as far as possible be designed to address the needs of the person-
nel being trained and the practices and processes of the facility. Some training topics may
be generally of interest to most personnel. Other topics may be of interest only to personnel
from specific job roles.
Topics that may be of more general interest might include
● Recognition of ESDS
● Recognition of ESD protected areas
● Recognition of ESD packaging that might contain ESDS devices
● Awareness of the need to control ESD
● Knowledge of who the ESD coordinator is
● ESD control procedures used in the facility
● The need for and use of ESD control equipment
● How to test personal wrist straps and footwear
● Common noncompliances and how to avoid creating them
● What to do (or who to notify) if failures or noncompliances are found
● New techniques, processes, and equipment when they arise
● Basic electrostatics and ESD knowledge (appropriate to their job role)
418 12 ESD Training

Role or area-specific ESD training might include topics such as

● ESD awareness for manager, including the possible financial cost/benefit and other
impact of ESD damage and ESD control
● ESD measurement test procedures, for personnel who check ESD control equipment and
EPAs
● ESD control, safe working, and special procedures for personnel who work in
high-voltage or other areas with special safety issues
● Audit techniques and audit of ESD control procedures
● ESD coordinator training and knowledge development and standards update
● Cleaning regime cleaning materials and practice for personnel who clean within
the EPA
● Do’s and don’ts for visitors, and guidance for personnel who accompany visitors
● Instructions for contractors working in the EPA
● Instructions for facility maintenance activities in the EPA

12.4.2 Initial Training


Initial ESD training should typically cover topics appropriate to the ESD control program
requirements, EPA discipline, and person’s job role. For a trainee handling ESDS and work-
ing in the EPA, this might include topics such as the following:

● What is static electricity?


● How and when does ESD occur?
● What problems come from static electricity and ESD, and why do we need to avoid them?
● Recognition and correct use of ESD control packaging
● What is secondary packaging, and how do we recognize it?
● What is an EPA, and how do we recognize one?
● Don’t bring secondary packaging into an EPA.
● Don’t open ESD control packaging outside an EPA.
● Charged people are a major source of ESD damage in manual handling, and personal
grounding is an extremely important ESD control measure.
● Use and test of wrist straps and ESD footwear.
● Understanding what unnecessary insulators are, their recognition and why we don’t want
them in an EPA.
● Don’t take unnecessary insulators into an EPA, and remove them if you find them in the
EPA.
● ESD control equipment, its recognition, and how to use it.
● For portable equipment such as tools and gloves, recognition of ESD control versions and
the importance of not bringing ordinary versions into the EPA.
● The importance of not bringing personal items, which might bring undefined risks, into
an EPA.
● Don’t remove or store clothing near unprotected ESDS devices.
● Any safety precautions necessary in the EPA.
12.4 Training Form and Content 419

12.4.3 Refresher Training


Refresher training gives an opportunity to reinforce key aspects of ESD control, especially
anything that has been noted to regularly go wrong. It also provides an opportunity to give
more depth or detail to topics and updates or changes that may have occurred in the ESD
control program. Trainees can be encouraged to give feedback or discuss issues and make
suggestions for improvements.
Refresher training can be a good way to review things that have arisen as regular noncom-
pliances. However, persistent noncompliances may not always be best addressed by more
or improved training. It may be an indication that an existing ESD control is inconvenient
or difficult to use, or at least may not be the best solution. As an example, Snow and Dan-
gelmeyer (1994) reported that a persistent noncompliance related to use of foot straps was
eliminated by moving to the use of ESD control shoes.

12.4.4 Training Methods


Any form of ESD training can be used, varying from group watching of videos to one-to-one
hands-on practical training. The main consideration is that it should be effective in com-
municating the topics to be covered.

12.4.4.1 Video, Computer or Internet-Based Training


A video or film or computer or internet-based courses can be effective and cost-effective for
training large or small numbers of personnel or a single person, for example on induction
into their place of work.
Online resources such as YouTube videos can provide some valuable material for train-
ing purposes, but they can be very variable in their quality, veracity, and relevance to the
organization’s facility and processes.
There is a risk that commercially available training may be too general and not appear
to the trainee to relate to the activities and processes they work with in practice. A generic
training course may be ineffective if it appears to be irrelevant and difficult to relate to their
work. The course should appear authentic and realistic with no exaggeration of the issues.
As the detailed ESD control measures often vary with the organization’s ESD control
program, it is often better that discussion of detailed control measures is avoided in generic
training. Universal and common measures such as wrist strap and ESD control footwear
use can be usefully covered.
Short videos can be a good way of teaching basic skills such as putting on and testing a
wrist strap or ESD control footwear, with the trainee then practicing the skill (Snow and
Dangelmeyer 1994).

12.4.4.2 Instructor-Led Training


Perhaps the most effective means of training is instructor led, either in a group or on a
one-to-one basis. Classroom or training facility–based practical training can be very effec-
tive for a group, and on-the-job training may be productive for individuals or small numbers
of trainees.
The instructor should have a good knowledge ESD control theory and practice and the
organization’s ESD control procedures and practices, to a level above the level they are to
420 12 ESD Training

teach. If the necessary expertise does not exist in-house, it should be sourced outside the
organization. This might include using an external trainer with appropriate expertise to run
a suitable course.
A commercial course prepared and presented by a third-party instructor can be a good
way of providing broad-based awareness training or general expertise such as ESD measure-
ments or ESD coordinator level training. A disadvantage of generic commercial training is
that it may not align well to the organization’s specific ESD control program, processes, and
facilities.
An in-house course prepared and presented by an expert instructor (in-house or external)
can be an excellent way of providing training that is closely aligned to the organization’s
ESD program and facilities and include details specific to them.
High-level training, e.g. for the ESD coordinator, may be available only as external
courses. Attendance of seminars, conferences, or symposia should be considered as a
means of updating knowledge of current standards, trends and ESD control techniques,
available equipment, know-how, and expertise.

12.4.5 Supporting Information


It is likely that a substantial amount of information on ESD, standards, and training mate-
rials will be collected, both for ESD program development and for training purposes. This
should be stored and kept available to personnel as appropriate. These could include

● ESD course handout booklets, CDs, videos, and computer-based presentations


● Written procedures and instructions
● Copies of the organization’s ESD control procedures
● Copies of standards related to ESD control
● Books, conference proceedings or papers, and other white papers
● Magazine or other articles
● Reports of studies made in-house or by third parties to investigate ESD issues and ques-
tions or in support of ESD program development
● ESD control product data sheets
● ESDS susceptibility data of components

The location and availability of this information and resources should be publicized so
that personnel can refer to it.

12.4.6 Training Considerations


According to Snow and Dangelmeyer (1994), five stages should be considered in the training
process.

● Preparation
● Delivery
● Instructor-led demonstrations
● Hands-on learning
● Follow-up assessment and training
12.4 Training Form and Content 421

Many books have been written on the subject of creating and delivering successful presen-
tations. The best presentations use a mixture of verbal and visual elements. Points should
be presented in a logical sequence. Each point should lead to following points or build on
previous points. I have found that short videos, demonstrations, or audience participation
(discussions or activities) are all useful ways of keeping interest in a presentation high.

12.4.6.1 Preparing a Presentation


In preparing a presentation, two fundamental questions must be asked. Who is the audi-
ence, and what type and level of background knowledge can they be assumed to have? What
are the learning points that they must take from the presentation? From this, the presen-
tation material, including demonstrations, can be developed. The presentation should be
made appropriate to the audience in terms such as topics of interest, technical content and
level, and depth.
Many presenters find that about one slide per minute of presentation time gives a rea-
sonable estimate of a presentation duration. For some, more time per slide is needed. Don’t
forget, however, to make time allocation for demonstrations, videos, discussion, or another
attendee participation.
If the presentation is to attendees who are unfamiliar with the venue and instructor, it
may be necessary to start with some introductions and introductory information. This could
include introduction of the presenter, location of the facilities, emergency fire exits, and
whether a test alarm is expected. It is a worthwhile precaution to ask attendees to switch
off their mobile phones (When presenting, this is a useful reminder to switch mine off also).
For a course with a small number of attendees, it can be useful to ask the attendees to give
a short sentence of introduction to themselves also.
A long course can be split into specific topics or themes. It is good to vary the presen-
tation and engage the attendees by using demonstration and participation where possible.
Questions and discussion can be encouraged. Discussing issues and situations arising in
the attendee’s workplace can be valuable in demonstrating relevance, providing interest-
ing practical learning points and helping their understanding. It can, however, be a mixed
blessing, leading to over-run if not carefully controlled!
If a training session is long, decide if breaks and refreshments will be necessary. Breaks
can also be useful for giving the presenter the opportunity to talk to attendees on a
one-to-one basis. This can be useful for evaluating whether the attendees have adequately
received and understood the material up to that point. Attendees often take to opportunity
to ask questions at breaks on topics that they have not had courage to ask in the more
public setting of the course.
While it is good to be well prepared and familiar with the presentation material, it can be
counterproductive to be over rehearsed or appear to be presenting a memorized speech.

12.4.6.2 Presentation and Attendee Participation


The delivery skills of the presenter make a large contribution to the engagement of the
audience. These days, there are many good books on presentation skills that can give guid-
ance and ideas for this. Dressing appropriately so that both the presenter and the audience
feel comfortable is important. (An audience of managers might expect the presenter to be
in a suit, but “smart casual” may be better for an audience of EPA workers). Try to avoid
422 12 ESD Training

Figure 12.1 The author presenting a course.

anything that might be a distraction or a barrier to communication. Try to make the presen-
tation as interesting and enjoyable as possible for the audience and yourself (Figure 12.1).
Eye contact, or an illusion of eye contact, is important, but maintaining unnatural eye
contact can make the recipient uncomfortable. In a small group, it may be possible and
important to make occasional eye contact with each attendee in a relaxed and random order.
In a larger group, there are usually a few attendees around the room who are obviously more
receptive and engaged than others. Keeping eye contact with some of these can give the
illusion of keeping eye contact with the whole audience. Keep eye contact moving around
the room to avoid making anyone feel uncomfortable! If the presenter feels self-conscious
about eye contact, looking at different points behind the audience can help give the illusion
of it.
The presenter should face the audience whenever possible and try to avoid turning their
back. This can be difficult when the display screen is behind the presenter, and it may be
necessary to point to parts of it. If possible, the presenter should avoid blocking the view of
the screen or demonstrations. Achieving this may require some thought into the arrange-
ment of the room and presenting area when setting up for the presentation. Before starting,
walk around the room and look at the presenter’s area from the point of view of the entire
audience, sitting at selected points, to look from attendee eye height.
Classes where learners can participate, discuss, and engage are more likely to be found
interesting and remembered by the audience. Demonstrations can be very valuable, espe-
cially if the audience participates in them. As learners best remember what they practice
frequently, it is valuable to have attendees practice the skills they will need to use. It is
best to immediately follow up this learning by practice in the workplace. Surprisingly, even
demonstrations that go wrong can entertain and make the course more memorable. As elec-
trostatic demonstrations can be fickle, I explain early on that at least one will probably go
12.4 Training Form and Content 423

wrong, but I don’t know which one. This creates a sense of interested anticipation! When
one does go wrong, its failure can often be used to reinforce some other learning point!
Learning is often facilitated by working in a small group and encouraging questions and
discussion. I find that attendee questions often bring up very interesting points specifically
relevant to the questioner. Questions indicate that the attendee is thinking about the course
material presented and trying, in their mind, to apply it in their work.
Unless a course is very short or the audience is large, I prefer to start by briefly introducing
myself and then asking each attendee to introduce themselves and their job role in a couple
of sentences. This helps “break the ice” and starts the attendees speaking and contribut-
ing. It also helps me understand my audience for later formulating answers to questions in
appropriate terms and level.
Early on in ESD awareness courses I ask the question, “Who here has experienced elec-
trostatic shocks in everyday life?” I then go on to ask typical circumstances they have felt
shocks and to explain that the shock is a form of electrostatic discharge. I encountered very
few people who claim not to have experienced an electrostatic shock. Getting the audience
to respond to this serves several useful purposes.
● Normally, everyone has experienced a static shock at some time, and everyone can
respond – this again helps establish the habit of audience participation.
● It generates interest.
● It shows they have themselves experienced ESD, and it is a matter of everyday experience.
● It gives me the opportunity to explain that to feel a shock the body voltage must be over
2000 V for most people, but many components can be damaged by ESD at lower body
voltages we do not feel shocks (this is later substantiated by demonstrating body voltage).
Specific points can often be illustrated by relevant anecdotes from real life. If an attendee
has an illustrative anecdote from their experience, this can be even better, bringing the
point into relief and immediate relevance to other attendees. Occasionally, attendees have
brought to the discussion some evidence of ESD damage in their experience – this can be
particularly valuable.
When preparing electronic slide presentations, the visual style, type face, color selection,
etc. should be selected for clarity. Transition effects should be used sparingly as they can
be distracting. Establishing a consistent slide style helps give a professional presentation
image. Be careful to make sure that text and information low down on a slide can be seen
be all the audience.

12.4.6.3 Hands-On Learning


For some types of learning, there is no substitute for hands-on exercises for the attendees.
Examples are wearing and testing of wrist straps or making ESD measurements and tests
for compliance verification. The exercise normally consists of demonstration of the actions
to be learned, followed by the learner trying the actions for themselves, and practicing until
sufficiently familiar and proficient. The demonstrator can also evaluate the proficiency of
the learner during the exercise.
In some cases (e.g. testing of wrist straps and footwear), the learning exercises may be
best done in the workplace. In other cases, the exercise can be done in a training facility as
part of a course.
424 12 ESD Training

12.4.6.4 Follow-Up Assessment and Training


It is necessary to evaluate the attendee’s understanding or ability to perform tasks as taught
on the course, during or after completion of the course. This evaluation is a requirement of
modern ESD control standards. This can be any of the following, for example:
● A practical evaluation (e.g. observation of wearing and testing a wrist strap correctly, or
testing a workstation and recording results)
● Question and answer (this can be an interactive discussion or written questions)
● Multiple-choice questionnaire
● Some other form of evaluation

12.4.7 Public Tutorials and Courses


Public tutorials and courses are available from a wide variety of sources such as the ESD
Association, IPC, industry groups and organizations, and specialist ESD control consul-
tants, trainers, and equipment suppliers. Some of these, such as the ESD Association tuto-
rials provided in association with its symposia, can lead to certification qualifications.
The level of these courses ranges from basic ESD awareness and control practice to
in-depth ESD practitioner level courses. Many of the industry provider courses are at the
operator basic ESD awareness and control level. Some providers offer courses at the user’s
site as well as open public events. Some providers offer courses up to ESD coordinator
(program manager) level, including measurements and auditing.

12.4.8 Qualifications and Certification


There are currently a few ESD-related qualifications and certification providers. At the time
of writing, the ESD Association is probably the foremost provider of ESD control personnel
certification at various levels.
Many independent training companies give ESD training that includes a certificate of
attendance of the course. This should not be confused with certification – a certificate
of attendance merely shows that a person has attended a course. The certificate of
attendance is usually issued without an examination of the attendee’s knowledge and
understanding.
Certification provides confirmation that a person meets a level of knowledge and
problem-solving ability (Newberg 2017). Achieving certification will usually require some
extensive training and testing. For the certified professional, it provides credibility in the
industry, demonstrating knowledge, experience, and competency. It is a form of profes-
sional development and can improve performance and confidence. It can differentiate
the individual from those who are not certified in terms of technical skill. It can create
opportunities for career advancement and increased earnings. In some organizations,
it may give an advantage over noncertified personnel in competition in recruitment
applications.
At the time of writing (2017), the ESD Association offers the following:
● TR53 Technician Certification
● ESD Certified Professional Program Manager Certification
12.4 Training Form and Content 425

● Device Stress Testing Certification


● ESD Certified Professional – Device Design
Each certification course requires attendance to a number of day or half-day courses,
which are normally given in association with ESD Association Symposia around the world.
As an example, the ESD Certified Professional Program Manager course covers the follow-
ing topics:
● ESD basics
● How-tos of in-plant ESD auditing and evaluation measurements
● Ionization issues and answers
● Packaging principles
● ESD standards overview
● Device technology and failure analysis overview
● Electrostatic calculations
● Cleanroom considerations
● System-level ESD/electromagnetic interference (EMI), including principles, design, trou-
bleshooting, and demonstrations
● ESD program development and assessment (ANSI/ESD S20.20 Seminar)
After attending the courses, the candidates can take an examination that is also provided
at the symposiums and, if successful, leads to award of the certificate.
Alternatively, certification can be gained through National Association for Radio and
Telecommunications Engineers (iNARTE) examination. iNARTE provides certifications for
qualified engineers and technicians in the fields of telecommunications, electromagnetic
compatibility/interference (EMC/EMI), product safety (PS), ESD control, and wireless sys-
tems installation.
The iNARTE ESD certification program was implemented in conjunction with the ESD
Association in the 1990s. Certification can be obtained at the engineer or technician level
(iNARTE 2017). Assessment includes providing a record of the candidate’s relevant experi-
ence and passing of an iNARTE examination.
IPC is a trade association connecting electronics industries. Among other activities, it
provides standards and training for the electronics industries and is widely used across the
world.
IPC provides ESD control certification courses via the online IPC learning management
system, IPC Edge (IPC 2017). IPC developed these courses with the EOS/ESD Association
to train operators and trainers on ESD controls and best practices. This online ESD Certifi-
cation Program allows users to validate their knowledge and skills by passing an exam on
ESD principles.

12.4.9 National and International ESD Groups and Electrostatics Interest


Organizations
Many countries have national ESD or electrostatics interest organizations. There are also
ESD Association chapters around the world e.g. in Texas, Philippines, India, and Korea
(https://www.esda.org/membership/local-chapters/Accessed:Dec.2017). Some national
specialist ESD interest organizations are given in Table 12.2.
426 12 ESD Training

Table 12.2 Some regional or national ESD interest organizations.

Organization Region Activities (languages) Web site

ESD Association North Corporate or individual www.esda.org


America membership, Standards, https://www.facebook.com/
International symposia, conferences, EOSESDAssociationInc
tutorials, white papers
(English)
ESD Association Korea Membership, symposia, http://www.esd.or.kr
Korea Chapter conferences, tutorials https://www.facebook.com/
(Korean) KOREA-Chapter-of-ESD-
Association-186248624800065
STAHA Scandinavia Membership, symposia, http://www.staha.fi
conferences, tutorials
(Finnish and English)
ESD Forum Germany Membership, symposia, http://www.esdforum.de
conferences, tutorials
(German and English)
Italian ESD Italy Membership, symposia, http://www.esditaly.com/
Assoc. conferences, tutorials esditaly.html
(Italian and English)
Dutch ESD-EMC Netherlands Membership, symposia, http://www.emc-esd.nl
Assoc. conferences, tutorials
(Dutch and English)
ESD Association China Membership, training. http://chinaesd.org.cn
China (Chinese)
Japan ESD Japan Seminars, exhibitions, http://www.jesda.org
Association publications (Japanese)
Industry Council International Invited industry http://www.esdindustrycouncil
on ESD Target membership, white .org
Levels papers(English)

Some organizations are interested more generally in electrostatics in industrial processes,


and others in academic research (Table 12.3). These organizations often also have an inter-
est in ESD control in electronics manufacture as well as measurements and many other
areas of electrostatics. They often include ESD control in their conferences or publications.

12.4.10 Conferences
Many of the organizations listed in Tables 12.2 and 12.3 organize conferences worldwide
that may include some ESD-related papers. Their current activities may be found via their
web sites.
The organizations listed in Table 12.2 are most likely to arrange specialist ESD
control–related conferences and publish papers in their proceedings.

12.4.11 Books, Articles, and Online Resources


There are a range of books available on ESD-related topics. Many of these are concerned
mainly with design of on-chip ESD protection, e.g. Amerasekera and Duvvury (2002),
12.4 Training Form and Content 427

Table 12.3 Some general electrostatics interest organizations.

Organization Region Activities Contact

Electrostatics Society N. America Membership, conferences, http://www.electrostatics


of America International newsletter .org
Working Party Static Europe Industrial and academic
Electricity in Industry committee.
International Conference every
four years.
Institute of Physics UK Membership, Dielectrics, http://www.iop.org/
and Electrostatics Group activity/groups/subject/
International Conference every de/page_65558.html
four years

Table 12.4 Some magazines, journals, and online resources that publish ESD-related articles.

Publication Activities Contact

InCompliance Online magazine https://incompliancemag.com


Interference Technology Online magazine https://interferencetechnology.com
Controlled Environments Online magazine www.cemag.us
Evaluation Engineering Online magazine https://www.evaluationengineering.com
Microelectronics Peer reviewed https://www.journals.elsevier.com/
Reliability Journal microelectronics-reliability
IEEE Transactions on Peer reviewed http://ieeexplore.ieee.org/xpl/RecentIssue
Device and Materials research Journal .jsp?punumber=7298
Reliability
IEEE Transactions on Peer reviewed http://ieeexplore.ieee.org/xpl/RecentIssue
Electron Devices research Journal .jsp?punumber=16
IEEE Transactions on Peer reviewed http://ieeexplore.ieee.org/xpl/RecentIssue
device and materials research Journal .jsp?punumber=7298
reliability
Journal of Electrostatics Peer reviewed https://www.sciencedirect.com/science/
research Journal journal/03043886

Wang (2002), and Voldman (2004). There are several books on ESD control in the man-
ufacturing environment, e.g. Dangelmeyer (1999), McAteer (1990), Welker et al. (2006).
Unfortunately, few of these are up to date although their content may still be highly
relevant. Because of their age, they are usually not well aligned with current ESD control
standards. Welker et al. (2006) is notable in that it focuses on ESD control in the clean room
environment.
There are several online magazines that occasionally publish good-quality articles on
the subject (Table 12.4). Some academic journals that from time to time publish papers
on ESD-related topics are also listed here.
428 12 ESD Training

12.5 Electrostatic and ESD Theory


12.5.1 The Pros and Cons of Theory
For those who understand it, inclusion of some level of electrostatics and ESD theory and
simple circuit diagrams can be helpful and informative. Conversely, for attendees who do
not understand it, theory can be daunting and off-putting and can reduce their confidence
that they can understand the course. So, theory and circuit diagrams must be used sparingly
and with caution.
In many courses, the attendees are from many backgrounds and job roles. While many in
my courses do not understand even simple electronic circuits, there are normally a few who
do and appreciate this type of explanation. Normally I include some level of theory in simple
nonelectronic explanation terms. Except in basic awareness courses, I often also include a
simple circuit explaining electrostatic charging and charge dissipation, appropriate to the
level and audience of a course. I may spend more, or less, time on theory depending on the
level of interest expressed by attendees and my understanding of their background gleaned
from introductions given at the beginning of the course. If necessary, use of the whiteboard
and discussion can supplement the core presentation material.

12.5.2 A Technical and Nontechnical Explanation of Electrostatic Charging


In electronic terms, the simple circuit of Figure 2.1 and Figure 12.2 is easily understood by
attendees with some electrical circuit knowledge including Ohm’s law. Charge generation is
represented by a current generator I and flows through a resistance to ground R. Ohm’s law
gives the voltage V produced as V = IR. It can easily be seen that the higher the resistance,
the higher the voltage produced. In a higher-level course, this is used to explain why we
determine an upper limit of resistance to ground in order to determine a maximum voltage
produced by electrostatic charging. The capacitance C represents charge storage. With no
current generation, the charge stored in the capacitance drains through the resistance, and
the voltage drops away with a characteristic charge decay time given by the product RC.
If the resistance R is very high (as with insulating materials), the voltage generated by
even a small current I can be very high. This explains why we specify an upper limit of
resistance to limit voltage build-up. For very high resistance, the decay time constant RC
also becomes long, showing that charge and voltages can remain for long periods.
A water analogy can be used to explain electrostatic charging to nontechnical people
(Figure 12.2). In this explanation, charge is analogous to water, and a basin represents
charge storage. The capacity of the wash basin is analogous to the capacitance of the circuit.
The level of the water in the basin is like the voltage due to the charge. The water flow into,
and out of, the basin is analogous to electrostatic charging and charge dissipation current
flow. The basin has a drain that allows water to escape and a tap that provides water input.
The size of the drain is analogous to the electronic resistance to ground. A small drain allows
only slow water escape, like a high resistance in a circuit.
Clearly with no water input and the drain open, there is no water level in the basin. Most
people will accept that for a given water flow into the basin, the water level will depend on
the size of the drain hole and rate at which water can escape. If the plug is in the drain hole
12.6 Demonstrations of ESD Control–Related Issues 429

Voltage V = IR

CV = Q Water flow in
dQ/dt = I
C
Charge
Capacitance
generation R Stored water level
(current source) Resistance
Plug

Basin
Drain

Water flow out

Figure 12.2 Water analogy of a simple model of electrostatic charging that is easily understood
by many nontechnical people.

(analogous to an insulator blocking charge dissipation), even a small water flow in (such as
a dripping tap) produces a significant water level that can remain for a long time.
Finally, the height of water in a basin for a given charge depends on the shape of the basin.
A basin that has a smaller footprint will give a higher water level, and a basin with a larger
footprint will give a lower voltage level for a given amount of water. This is analogous to
the same charge giving higher voltage for smaller storage capacitance.

12.6 Demonstrations of ESD Control–Related Issues

12.6.1 The Role of Demonstrations


Demonstrations can be very effective for showing the reality of ESD-related issues and
bringing into focus the need for ESD control. To be most effective, demonstrations should
emphasize the facts that the audience are most interested in – for example, for managers,
the cost of ESD damage and rework, possible return on investment, and impact on product
quality, yield, reliability, or reputation may be of most interest. For production personnel,
the impact on need for rework may be more relevant.
I have found that most people respond well to demonstrations of real static electricity and
how ESD can arise. These usually work best if they use materials that are commonly found
in the uncontrolled workplace, such as packaging materials, engineering plastics and tapes,
and isolated conductors. This shows directly why these materials should be excluded from
EPAs where possible.

12.6.2 Demonstrating Real ESD Damage


McAteer (1980) found that to help people believe in ESD damage, they needed to see a
device being damaged. A convincing authenticity is provided by showing real ESD damage
examples relevant to the organization.
430 12 ESD Training

McAteer used samples of different types of ESDS to demonstrate damage, including a


metal packaged operational amplifier, metal-oxide-semiconductor (MOS) device, and a
small assembly containing a MOS device. He used a curve tracer to show changes to the
device parameters resulting from ESD produced during the demonstration. He commented
that this type of demonstration persuaded management that the ESD problem was real.
They were then anxious to discuss seriously the costs of static failures versus the costs of
prevention.
An indirect way of showing the effect of ESD control on failures is if changes in prod-
uct yield can be correlated with ESD control measure compliance. Snow and Dangelmeyer
(1994) found that yield charts showing a decline in failures correlated with increased ESD
control are convincing evidence. Sometimes, increases in failure rates can be correlated
with a drop in atmospheric humidity, especially below 30% rh. Where this occurs, it is a
strong indicator that ESD damage is occurring.
Unfortunately, many organizations are not able to do failure analysis to a level that could
reliably identify ESD damaged devices. In this situation, it is difficult if not impossible to
provide convincing firsthand evidence of ESD damage to the organization’s components,
assemblies, or product.
There are, however, many documented accounts of ESD damage given in literature such
as the EOS/ESD Symposium Proceedings, journals such as Microelectronics Reliability or
IEEE Transactions (Table 12.4), and books such as Dangelmeyer (1999) and McAteer (1990).
Selected examples of these can be used to support ESD training and help present a case for
the importance of ESD control.
Unfortunately, it is likely that some people will remain unconvinced of the risk and reality
of ESD damage, and importance of ESD control, without direct evidence of ESD damage
occurring in their processes.
Sometimes in an ESD control course, some attendees have been able to relate their own
company experience of ESD damage. When this happens, their contribution can be very
valuable, as a “real-life” anecdote from a fellow attendee’s experience can be much more
convincing than second-hand and possibly old accounts from an instructor, other organi-
zations, or research publications.

12.6.3 The Cost of ESD Damage


Unfortunately, few seem to have a real understanding of the failure levels and cost of ESD
damage in their organization. This is partly because the cost of failure analysis to a level
that could confirm ESD damage can be high. In my experience, most organizations do not
do failure analysis to this level and so do not have the basic information required to estimate
the cost of ESD damage. There are a few published studies of this topic, and these can be
quite instructive. These can give useful information for inclusion in a course.
Helling (1996) gave an interesting and useful account of estimating cost of failures and
the return on investment in ESD control. His data and examples can be useful to quote to
help make points on this topic. His data confirms the following useful general points:

● The cost of failures tends to increase as the product goes through the production process
stages.
12.7 Electrostatic Demonstrations 431

● The most expensive failures are usually those that occur at the customer’s site.
● Cost/benefit indications of over 1:10 can be found for an effective ESD control program.
In a course including relevant attendees such as managers or QA personnel, I often ask
the hypothetical question – what is the likely approximate cost of a failure or unreliability at
the customer’s site for your product? This usually provokes some interesting discussion. In
many cases, the answer to this question alone makes it worth taking ESD control seriously.
Many people will not be convinced by accounts of the cost of ESD damage found in other
factories, especially if the sources are old. There is no doubt that if it is available or can be
determined, real current data on the cost of ESD damage at the attendee’s own facility will
be far more convincing. In the absence of this data, it can be as valuable to demonstrate cases
where component failure levels, although not analyzed and proven due to ESD, nevertheless
fell when ESD controls were improved.

12.7 Electrostatic Demonstrations

12.7.1 The Value of Electrostatic Demonstrations


Static electricity cannot be sensed unless it is at very high levels, so it is difficult for people
to really understand that it is at work around us much of the time. Using an instrument to
reveal the static electricity makes it real to the audience in a way that talking about it and
explaining it cannot.
Electrostatic voltages and fields can be revealed with an electrostatic voltmeter or field
meter. These can be used as the basis of many simple demonstrations that reveal how static
electricity works and can bring real understanding to attendees.

12.7.2 The Pros and Cons of Demonstrations


There are several ways in which demonstrations can be effective as a training method. This
is perhaps particularly true of teaching how static electricity works and the effectiveness
of some ESD control products. I have found static electricity demonstrations to be most
effective if they use materials and objects representative of items that are normally found
in a poorly controlled working area. A successful demonstration then shows not only how
static electricity works but helps recognition of what sort of materials cause problems. Even
experienced people can find that a good demonstration helps them really understand elec-
trostatic action and ESD issues. Developing reliable electrostatic demonstrations can also
be a great way of educating the trainer! Some of my most reliable demonstrations have
arisen out of adversity and failure of a less reliable demonstration. The demonstration of
Figure 12.5 was invented out of need when other demonstrations failed under high humid-
ity conditions in Shanghai!
One problem with live static electricity demonstrations is that they are rather fickle and
can unexpectedly fail or give surprising results. This is especially true under high humidity
conditions, and demonstrations are best done (where possible) under moderate or dry air
conditions. Nevertheless, with practice and selection of materials, demonstrations can be
found that are reasonably reliable even at 70% rh. As a precaution, I normally tell my course
432 12 ESD Training

attendees that it’s likely that at least one demonstration will not work, and I don’t know up
front which it will be. This helps to create a sense of interested attention and anticipation!
The most convincing and relevant demonstrations are often those made using materials
that are commonly found in a poorly controlled workplace. For this reason, I generally
avoid unusual props like balloons (I once did see balloons in an EPA workstation; I was
told it was the operator’s birthday!) or Van der Graaf generators. The following examples
give some demonstrations that I have used in my electrostatics training.

12.7.3 Useful Equipment for Demonstrations


The heart of good electrostatic demonstrations is a good electrostatic field meter or elec-
trostatic voltmeter to show the presence and magnitude of electrostatic voltages and fields.
Field mill instruments are more useful than induction field meters as they avoid problems
due drift and do not need zeroing. Beware that nearby charged insulating materials can
cause unexpected effects.
While a stand-alone voltmeter or field meter works well for one-to-one interactions or
small groups, for a larger class it is better to use an instrument that has an output that can
be fed into a suitable display (Figure 12.3). PC-based oscilloscopes such as Picoscope can
be used to display the changes in field or voltage over time and project the display onto a
large screen display or via a data projector.
A metal plate mounted on a good-quality insulating handle or stand can be used for
demonstration of triboelectrification and induced voltages on conductors, equipotential
bonding, and many other effects.
A selection of insulating materials and known reliable static generators should be col-
lected, especially those that tend to be found as noncompliant items in the workplace.

Figure 12.3 Demonstrating a charged insulator, in this case a slab of expanded polystyrene foam.
12.7 Electrostatic Demonstrations 433

Materials may need to be replaced from time to time as their charging and charge-holding
properties tend to diminish with handling. This is due to contamination with oils, salts, and
perspiration from the hands.
A selection of ESD control packaging can be useful. Low-charging materials such as pink
polythene should be kept in a package separately from the other demonstration materials.
This is because they tend to contaminate other materials with their antistat and can make
demonstrations of electrostatic charging rather unreliable.
It can be useful to have an electric hairdryer or hot air gun in the demonstration kit to dry
out materials under high-humidity conditions. This can then be used as a teaching point,
demonstrating how humidity affects electrostatic charging and activity.
An ESD detector can be useful for demonstrating that ESD occurs during experiments.
This should be selected to make a loud enough sound when ESD occurs, to be easily heard
by all trainees.

12.7.4 Showing How Easy It is to Generate Electrostatic Charge


The main learning point in this demonstration is that many insulating materials charge
easily and give strong electrostatic fields. Highly insulating materials tend to charge during
handling and hold their charge for long periods. This can be shown using an electrostatic
field meter held in a stand (Figure 12.3).
I have often used a plastic document file divider for this demonstration. This just rests
in, or on, a pile of demonstration materials until the moment of demonstration. Lifting the
material and holding it in front of the electrostatic field meter shows that it is already highly
charged. Many plastic packaging materials can also be used for this demonstration. With
suitably chosen materials, this demonstration is reliable under most except the most humid
conditions.
Another reliable demonstration is the high voltage produced by ordinary tapes such as
packing tape on stripping from the reel. This is revealed by stripping a length of tape from
the reel and holding it in front of the field meter.

12.7.5 Understanding Electrostatic Fields


The experiment of Figure 12.3 can also be used to show the variation of electrostatic field
with distance from the field source. The field meter reading increases as the charged item
is move toward the field meter and then drops dramatically as it is moved away.
Depending on the level of course, the instructor can then explain some important learn-
ing points, such as the following:

● The field meter is calibrated to read surface voltage but correctly does so only when taking
readings from a large flat surface a defined distance away.
● Electrostatic field drops rapidly with increasing distance from the source.
● The closer a field source is to the ESDS, the more concerned we will be about it.
● If an insulator is found to have greater voltage level than a defined risk level, the risk can
be reduced by keeping it a sufficient distance away from any ESDS. This is the basis of
the field and voltage limits given in the standards.
434 12 ESD Training

12.7.6 Understanding Charge and Voltage


People tend to think that voltages are at the root of electrostatics and stay constant unless the
item becomes more charged or charge leaks away. This demonstration can be used to show
that it is charge that is at the root of electrostatics and that voltages change, appear, and
disappear as materials are moved around – even though the charge present is not changed.
I have often used a training booklet with an insulating plastic cover in this demonstration
(Figure 12.4). When the booklet is held in front of the field meter with the cover closed,
only low or zero electrostatic voltage is shown. The instructor can explain that the cover is
already charged by contact with the cover, but no voltage is visible as the charges on the
cover and paper are in balance and are close together and so cancel in their effect.
When the cover is opened, a high voltage appears on the cover material. (A high voltage
does not appear on the paper as it is a sufficiently conductive material for the charge to
escape somewhere.) After closing the cover again, the voltage has normally largely disap-
peared again as the charge has been rebalanced.
The fact that the cover material is highly charged can also be used to demonstrate electro-
static attraction. Usually the first page of the booklet adheres to the cover due to attractive
charge forces.
Another rather interesting demonstration can be made with a document in an insulating
plastic document holder (Figure 12.5). With the document in the holder, little voltage or
field can be seen by the field meter. If the document is then partially withdrawn from the
holder, a voltage appears on the part from which the paper is withdrawn. If the document
is reinserted, the voltage disappears.
These demonstrations support the following teaching points:
● Charge is already separated between the plastic and the paper, but no voltages show until
it is separated by moving the paper and plastic and charges apart.
● The voltages disappear when the paper and plastic are moved back together.

Figure 12.4 Using a clear plastic covered booklet to show how voltages appear and disappear
when the cover is opened and closed.
12.7 Electrostatic Demonstrations 435

Figure 12.5 Using a document holder to show how voltage appears when the document is pulled
out and disappears when it is returned.

I usually neglect two other factors that are present in these experiments, for the sake of
reducing possible confusion. First, presence of a conducting material nearby can reduce
the apparent voltage produced by a charge. This is a second reason the voltage disappears
when the paper is reinserted into the holder. Second, the paper is normally sufficiently
conducting that any charge on it can move to and from the body of the demonstrator while
they are performing the demonstration.
I have occasionally done these experiments while wearing highly insulating rubber
gloves. If this is done, it can often be shown that the paper is positively charged on
separation while the plastic is negatively charged.

12.7.7 Tribocharging
Tribocharging can be demonstrated with conductors or insulating materials. As many peo-
ple, even those more familiar with electrostatics, believe that conductors cannot generate
charge, it can be useful to demonstrate tribocharging of a conductor. One way of doing this
is to use a metal plate mounted on a highly insulating handle (Figure 12.6). The field meter
is positioned to show the voltage on the plate and display it.
A suitable charged plate monitor can also be used to do these experiments. The metal
plate, however, can be easier for trainees to relate to metal items and printed circuit board
(PCB) used in the workplace. The instructor can make the point that what happens to the
plate can be expected to happen to the conductors on a PCB isolated from ground.
In this arrangement, the plate can be tribocharged by rubbing with various materials or
other actions. I have typically used the following:

● Negative charging of the plate by flicking it with a wool cloth


● Positive charging of the plate by flicking it with a rubber glove
● Charging of the plate by stripping tape previously applied to it

I have also used this arrangement to demonstrate the charging effect of processes, for
example use of a cooling spray can.
Under humid air conditions, the moisture layer condensing on the insulating handle can
be sufficient to allow slow discharge of the plate. In these conditions, a hair dryer or hot air
gun can be used to temporarily dry the insulator and prevent discharge. This can itself be a
436 12 ESD Training

Figure 12.6 Tribocharging a metal plate.

useful teaching point, showing that the discharge of the plate across the insulator is due to
moisture on the surface that is removed by drying.

12.7.8 Production of ESD


Once the metal plate is charged, it is easy to demonstrate ESD using an EMI-based ESD
detector (Figure 12.7). These instruments detect ESD from the EMI radiated from the ESD.
Some emit an audible signal as well as updating a count on each ESD detected.
With the plate charged, a discharge can be initiated by touching it with a ground wire. At
the same time as ESD occurs, the voltage on the metal plate can be seen to drop instanta-
neously to zero.

Figure 12.7 Demonstration of ESD generation.


12.7 Electrostatic Demonstrations 437

Possible teaching points in this experiment include the following:

● ESD occurs when a charged conductor is touched by another conductor.


● This is true even if the conductor that touches it is not grounded.
● The metal plate acts similarly to the conductors of a PCB or other ESDS device. So, ESD
can occur if a charged ESDS device contacts another conductor or is grounded.

The conductor used to touch the plate and generate ESD does not have to be a ground wire
and does not have to be grounded. ESD can also be demonstrated by touching with a tool,
or even a grounded ESD tool held in the hand of a grounded instructor. Even experienced
trainees can sometimes be surprised that ESD can occur when a correctly grounded person
handling or working on ESDS using ESD tools. The explanation is of course that the ESD
can occur if the ESDS is itself charged. ESD can occur when a person touches an ESDS if
either the person is charged or the ESDS is charged, or both.

12.7.9 Equipotential Bonding and Grounding


Following the demonstration of production of ESD on touching in the experiment of
Section 12.7.8, it can be demonstrated that ESD is prevented by equipotential bonding
(Figure 12.8). First it is shown that if a person touches the metal plate with an ESD tool,
ESD occurs because the person and the plate are at different voltage. ESD is registered on
the ESD detector.
The demonstrator can then connect themselves to the metal plate using a wrist strap.
No connection is made to ground. Usually, some voltage is indicated on the metal plate
as the demonstrator moves around and generates body voltage. Nevertheless, no ESD is

Figure 12.8 Demonstration of equipotential bonding.


438 12 ESD Training

produced when the demonstrator touches the plate with the tool. The explanation is that
the demonstrator and the plate are at the same voltage and so ESD does not occur.
This experiment supports the following teaching points:
● ESD occurs when two conductors touch and there is a voltage difference between them.
● If there is no voltage difference between the conductors, no ESD occurs when they touch.
● Once an equipotential bonding connection is made, no ESD can occur between the con-
nected conductors. However, at the point that the connection is made, ESD is likely to
occur as the conductors will probably be at different voltages.
This experiment can lead into a discussion explaining grounding as a form of equipo-
tential bonding where all conductors are bonded to earth. It can also be explained that
in situations where grounding is not possible equipotential bonding can be used to con-
trol ESD sources. Current ESD control standards often consider equipotential bonding and
grounding to earth both to be acceptable forms of “grounding.”

12.7.10 Induction Charging


The metal plate mounted on an insulating handle can be used to show induced voltages on
conductors due to electrostatic fields. The metal plate is initially discharged to zero volts in
the absence of any electrostatic fields. When a charged insulator is moved closer, the voltage
on the conductor changes and can reach surprisingly high levels (Figure 12.9).
ESD can then be made by touching the metal plate with a ground wire. An audible ESD
detector can be used to demonstrate the occurrence of ESD. With the ground wire removed,
the insulator is moved away, and the plate voltage and the plate can be seen to go to an
opposite polarity voltage.
This experiment supports the following teaching points:
● The voltage on an electrically isolated conductor changes in response to nearby electro-
static fields. The induced voltage changes as the insulator is moved closer or farther away.
The plate is not necessarily charged although it can be at high voltage.

Figure 12.9 Induction charging demonstrated using a charged plate monitor (left) voltage
induced on plate by nearby charged insulator, ready for ESD (middle). ESD occurs when charge
moves from plate on grounding; (right) plate is left charged to opposite polarity when insulator is
removed, ready for another discharge. Source: D. E. Swenson.
12.7 Electrostatic Demonstrations 439

● The charged insulator does not need to touch the conductor for the voltage to be induced.
● The induced voltage is negligible if the insulator is kept sufficiently far away.
● ESD occurs if the conductor is touched by another conductor or ground wire when at a
different voltage. After touching with the ground wire, the plate is now charged, although
at this point it has no voltage, until the field source is moved away.
● After the ground wire is removed and the insulator moved away, the plate can be seen to
be charged as it achieves a high opposite polarity voltage.

12.7.11 ESD on Demand – The “Perpetual ESD Generator”


Once it has been shown that an electrostatic field induces a voltage on a conductor, it can
be demonstrated that repeated ESD can be made just by moving the conductor in the field
and touching it with another conductor.
A charged insulator is used to induce a voltage on the isolated metal plate (Figure 12.10).
An ESD detector is used to detect the ESD generated by touching the metal plate with a
metal item such as a metal tool. The field meter shows the voltage changes occurring on
the metal plate.
After each ESD is made, the insulator is moved in position, and the voltage on the metal
plate changes. It is now ready to provide another ESD on touching with the tool. ESD occurs
even if the demonstrator is grounded and using an ESD tool to touch the plate.

Figure 12.10 The “perpetual ESD generator.” The metal plate rises in voltage due to electrostatic
field changes. A conductor contacting the plate initiates ESD.
440 12 ESD Training

This experiment supports the following teaching points:

● The voltage on an isolated conductor changes in response to nearby electrostatic fields.


The voltage changes as the insulator is moved around.
● ESD can be generated on touching the conductor with another conductor at different
voltage.
● If the field strength is changed, e.g. by moving the relative positions of the plate and the
insulator, the voltage again changes, and ESD can again occur.
● This situation could occur in the workplace if a conductor on an ESDS device or PCB is
touched by a metal tool or other item in the presence of an electrostatic field.

The effects shown in Sections 12.7.10 and 12.7.11 are often the main reason for control of
electrostatic fields and insulators in an EPA. Electrostatic fields set up the conditions under
which ESD is more likely to occur when the ESDS touches another conductor. The stronger
the field, the greater the concern and risk of ESD damage.

12.7.12 Body Voltage and Personal Grounding


The voltage developed on the body while walking can be demonstrated using the same
arrangement of metal plate with the plate voltage measured by an electrostatic voltmeter
as above. The demonstrator connects themselves to the metal plate via a long wire and
handheld electrode. Alternatively, a charged plate monitor (CPM) (see Section 11.6.13) can
be used for the demonstration (Figure 12.11). The CPM has the same physical arrangement
as the metal plate and electrostatic voltmeter but has calibrated performance.
With the demonstrator’s body connected to the plate, any voltage developed on the body
is now shown on the field meter. For most ordinary footwear and flooring combinations, it
can immediately be seen that significant voltages are generated on the body although the
person has no idea these voltages are present.
Occasionally the floor material in the training facility is a low-charging material that does
not give much voltage on the people walking on it. It is therefore wise to also have available
a known high-charging material (e.g. sample of nylon carpet) that can be compared with
the local floor material in the training facility. This also has the useful function of showing
that the body voltages generated vary with the floor material.
In a small enough class, each trainee can then be asked to try the experiment, holding the
handheld electrode and walking around. I try to do this experiment about halfway through
the presentation, as relief from sitting and listening and to get all attendees moving about.
Most of them will show some level of voltage, even if they are wearing ESD control footwear
(providing the flooring is not an ESD control floor or noninsulative material and the air
humidity is not too high!).
I usually ask the trainee who produces the highest voltage to help demonstrate grounding
using a wrist strap. (Grounding via footwear and flooring could also be demonstrated using
an ESD floor mat and ESD foot straps.) The trainee confirms that they are generating a high
voltage in the walking experiment. They are then asked to wear a grounded wrist strap and
move around in the same way. The body voltage is shown to be zero while wearing the wrist
strap. The effect of wrist strap failure can be demonstrated by disconnecting the wrist strap
12.7 Electrostatic Demonstrations 441

Figure 12.11 Demonstrating body voltage while walking, using a charged plate monitor. The
computer displays the body voltage waveform.

while the trainee is moving around. The body voltage is seen to reappear when the wrist
strap is disconnected.
This experiment supports the following teaching points:

● Body voltage of hundreds of volts are usually generated by ordinary actions such as walk-
ing in daily life in unprotected areas. The person has no idea these voltages are present
unless they are high enough to cause electrostatic shocks. Different footwear and flooring
combinations give different body voltage results.
● The body voltage is reduced to less than 100 V by wearing a wrist strap.
● If the wrist strap connection fails, the body voltage reappears. A continuous connection
is required to control body voltage.
442 12 ESD Training

If a similar experiment is done using ESD footwear and flooring, the body voltage is often
greater than when grounded by wrist strap. Using different combinations of footwear and
flooring, it can be demonstrated that body voltage typically is different for each combina-
tion. In a higher-level course, this can lead to a discussion of qualification of footwear and
flooring combinations with regard to body voltage generation, and the necessity of a walk
test. It can be demonstrated that body voltage is not usually well controlled if ESD control
footwear is used without ESD control flooring or the flooring used without ESD control
footwear.

12.7.13 Charge Generation and Electrostatic Field Shielding of Bags


ESD packaging is a poorly understood area of ESD control. Many people do not realize that
different types of packaging materials can have very different combinations of properties.
They often incorrectly believe all ESD packaging gives adequate ESD protection for ESDS
devices.
A simple demonstration can show differences between electrostatic charging and elec-
trostatic field shielding properties of ESD bags. Unfortunately, it is not so simple to demon-
strate differences in ESD shielding properties in a course. I normally compare samples of
ordinary polythene, pink polythene, black polythene, and metallized ESD shielding bags. I
normally introduce the test of ESD bags by doing the test with an ordinary polythene bag
(previously selected to give a good strong electrostatic charging response!).
The differences can be demonstrated by inserting an electrostatic field meter into an ESD
bag as if it were an ESDS device. The response of the field meter shows electrostatic charge
generation and fields experienced within the bag (Figure 12.12). This experiment works
best with a “field mill” type of instrument.
The field meter is placed within the bag. With a polythene bag, the field meter is normally
already showing high electrostatic fields at this point.
With the field meter within the bag, the bag can be handled, and the bag material moved
about. Any tendency to generate electrostatic fields and charge is indicated by the field
meter.

Figure 12.12 Placing a field meter inside an ESD bag gives a means of showing differences in
electrostatic charge generation and electrostatic field shielding.
12.7 Electrostatic Demonstrations 443

A highly charged insulator can then be brought nearby. If the electrostatic field shielding
provided by the bag type is poor, then strong indication of field is shown by the field meter.
If the electrostatic field shielding is good, then little indication of field is shown.
It is more convincing if the field meter is then taken out of the bag, and the charged
insulator brought nearby to show that it is indeed charged and gives a strong electrostatic
field.
When the demonstration is done with pink polythene bags under dry (<30% rh) air con-
ditions, the bag often shows strong electrostatic charging and little electrostatic field shield-
ing. It can be explained to the trainees that this material relies heavily on atmospheric
moisture for its low-charging properties. Under dry conditions, its performance can be little
different from an ordinary polythene bag. Aging and loss of antistat can also give this loss
of performance.
Under humid air conditions (>50% rh) pink polythene gives little static charging and can
appear to give good electrostatic field shielding. After demonstrating this, the bag can be
dried using a hair drier or hot air gun. After drying, the experiment is repeated, and the bag
will often show strong electrostatic charging and little electrostatic field shielding.
Black polythene bags give little electrostatic charging and show good electrostatic field
shielding in this test and can be shown to be independent of humidity.
Metallized ESD shielding bags usually show little electrostatic charging in this demon-
stration. They often show a small electrostatic field appearing within the bag in the elec-
trostatic field shielding test. Heavily crumpled bags may show reduced electrostatic field
shielding.
This experiment supports the following teaching points:
● Different types of ESD packaging have different electrostatic charging and electrostatic
field shielding properties.
● Ordinary polythene bags generate strong charges and are transparent to electrostatic
fields.
● Pink polythene bags have variable properties. With sufficient humidity, they give low
electrostatic charging and may also give good field shielding. Under low humidity, or
when old, they may appear little different from ordinary polythene.
● Black polythene bags show low electrostatic charging and good electrostatic field shield-
ing.
● Metallized ESD shielding bags show low electrostatic charging and good electrostatic
field shielding, although a small residual electrostatic field often appears within the bag.

12.7.14 Insulators Cannot Be Grounded


People often think that insulators can be grounded. It is easy to demonstrate that grounding
does not work for an insulator (Figure 12.13).
A rigid charged insulator such as a plastic tray is placed in front of the electrostatic field
meter so that the field due to its charge is shown. A ground wire is then clipped to the tray.
The electrostatic field does not disappear. The demonstrator explains to the trainees that
the charge cannot move on the insulator; therefore, it cannot move to the ground wire to
go to earth.
444 12 ESD Training

Figure 12.13 Demonstrating that an insulator cannot be successfully grounded.

This experiment supports the following teaching points:

● Charge does not move around quickly on insulating materials and cannot move to the
ground wire.
● Insulators cannot be successfully grounded.

12.7.15 Neutralizing Charge – Charge Decay and Voltage Offset of Ionizers


A means is often needed of reducing electrostatic fields and voltages from charged essential
insulators. Ionizers provide one means of doing this by neutralizing the charge that creates
the field.
This demonstration can follow naturally from demonstration that an insulator cannot be
grounded (see Section 12.7.14). A charged insulator is placed in front of the field meter,
showing a strong electrostatic field or high voltage (Figure 12.14). A bench fan ionizer is
then directed toward the insulator. The electrostatic field can then be seen to gradually
reduce toward zero. After some time, the surface voltage has reached a near constant level.
If a chart recorder type oscilloscope or similar display is used in this experiment, the time
taken to reduce the voltage to a low level can be measured. The final voltage remaining
on the insulator surface can also be measured. It can be explained to the trainees that this
offset voltage is due to imbalance in the quantity of negative and positive ions produced by
the ionizer.
In this experiment, it can also be shown that the time taken for charge decay can vary
greatly with factors such as distance from the ionizer and orientation of the ionizers.
The demonstration can be repeated using a metal plate on insulating handle as the target.
The metal plate can be charged by induction or from a voltage source. Using a metal plate,
two further demonstrations can then be done.
First, with the ionizer neutralizing the voltage on the charged plate, a residual offset volt-
age is shown. Temporarily grounding the plate with a wire demonstrates the offset voltage
and gives detectable ESD. On removal of the ground wire, the voltage on the metal plate
12.7 Electrostatic Demonstrations 445

Figure 12.14 Demonstrating charge decay and offset voltage of an ionizer using an insulating
material.

can be seen to rise again to the offset voltage. It can then be explained any item within the
region around the ionizer becomes charged to the offset voltage. This could, if too high, cre-
ate ESD risks. The offset voltage may vary with age and condition of the ionizer, and regular
test and maintenance is necessary to be sure this is kept within acceptable levels.
In a second experiment a charged insulator can be moved around near the metal plate
(Figure 12.15). The metal plate can be seen to vary in induced voltage in response to this

Figure 12.15 Demonstrating inability of an ionizer to keep a metal plate at low voltage in the
presence of fast-changing electrostatic fields or charging.
446 12 ESD Training

insulator. It can be shown that an ionizer does not necessarily keep the voltage on an iso-
lated conductor to an acceptable level if the electrostatic field or charging conditions create
voltage faster than the ionizer can neutralize them.
This experiment supports the following teaching points:

● Ionizers take time to neutralize charge and voltage. It may be necessary to wait for volt-
ages to be reduced to an acceptable level.
● For many types of ionizer, the neutralized voltage is near zero but does not reach zero
due to the offset voltage.
● Ionizer offset charges items to the offset voltage.
● Neutralization does not necessarily maintain the voltage on an isolated conductor at a
low level in fast-changing conditions.

12.8 Evaluation
12.8.1 The Need for Evaluation
It is important to make sure that trainees have understood and remember key training
points. It is also important to ensure they can use key equipment correctly and, in some
cases such as personal grounding, test them correctly. This means that some form of test
(paper or practical) will be needed to cover key points. The IEC 61340-5-1 and ANSI/ESD
S20.20 standards require this as part of their ESD Training Plan (International Electrotech-
nical Commission 2016; EOS/ESD Association Inc 2014).

12.8.2 Practical Test


Competence in some activities are perhaps best evaluated using a practical test or
“on-the-job” observation. An example is testing personal grounding (wrist strap and/or
footwear) and correctly recording the result. The actions to be taken if a “fail” is obtained
should also be tested.

12.8.3 Written Tests


Written tests could be, for example, in the form of a simple multiple-choice questionnaire.
Questions should address each of the most important teaching points of the training.
Evaluation of performance of a number of attendees can form useful feedback as to which
points are commonly poorly understood. The training course can then be modified if nec-
essary, to improve understanding in these areas.
Higher-level courses (e.g. certification) often include formal written examination. These
can be “open book” or “closed book.”
Computer or internet-based courses often have built-in tests that evaluate the under-
standing of the trainee.
References 447

12.8.4 Pass Criteria


Whatever the testing method, a pass level must be established. This could be achievement
of a certain mark in a paper test or successful demonstration of an action or procedure in a
practical test. Records should be kept of all tests and stored in a convenient place for future
reference and evidence.

References

Amerasekera, A. and Duvvury, C. (2002). ESD in Silicon Integrated Circuits, 2e. Wiley. ISBN: 0
470 49871 8.
Dangelmeyer, T. (1999). ESD Program Management, 2e. Clewer. ISBN: 0-412-13671-6.
EOS/ESD Association Inc. (2014) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
Helling, K. (1996). ESD protection measures – return on investment calculation and case study.
In: Proc. EOS/ESD Symp. EOS-18, 130–144. Rome, NY: EOS/ESD Association Inc.
INARTE (2017) Electrostatic Discharge Control Certification. Available from: https://inarte
.org/certifications/inarte-electrostatic-discharge-control-esd-certification [Accessed 7th
Dec. 2017]
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
IPC (2017) IPC Announces New ESD Control Certification Courses on IPC EDGE Courses
provide professional credentials for Certified ESD Trainers (CETs) and ESD Certified
Operators (ECOs). Available from: http://www.ipc.org/ContentPage.aspx?Pageid=IPC-
Announces-New-ESD-Control-Certification-Courses-on-EDGE [Accessed 7th Dec. 2017]
McAteer, O.J. (1980). An Effective ESD awareness training program. In: Proc. EOS/ESD Symp,
189–191. Rome, NY: EOS/ESD Association Inc.
McAteer, O. (1990). Electrostatic Discharge Control. San Francisco, CA: McGraw-Hill. ISBN:
0-07-044838-8.
Newberg C (2017) Certification. Available from: https://www.esda.org/certification [Accessed
5th Dec. 2017].
Snow, L. and Dangelmeyer, G.T. (1994). EOS-16 94-1 – 94-12. A successful ESD training
program. In: Proc. EOS/ESD Symp. Rome, NY: EOS/ESD Association Inc.
STAHA Association. Available from: http://www.staha.fi/index_files/STAHA_leaflet_2013.pdf
[Accessed 5th Dec. 2017].
Voldman, S.H. (2004). ESD Physics and Devices. Wiley. ISBN: 0-470-84753-0.
Wang, A.Z.H. (2002). On-Chip ESD Protection for Integrated Circuits. Klewer Academic Press.
Welker, R.W., Nagarajan, R., and Newberg, C. (2006). Contamination and ESD Control in
High-Technology Manufacturing. Wiley-Interscience, IEEE Press. ISBN-10: 0 471 41452 2,
ISBN-13: 978 0 471 41452 0.
448 12 ESD Training

Further Reading

Baumgartner, G. ESD demonstrations to increase engineering and manufacturing awareness.


In: Proc. EOS/ESD Symp. EOS-18, 156–166. Rome, NY: EOS/ESD Association Inc.
449

13

The Future

13.1 General Trends

The future, particularly in a fast-moving world like the electronics industry, has even more
capacity to surprise than the workings of electrostatics and electrostatic discharge (ESD).
Attempting to predict it might seem to be doomed to failure. To the engineers who first dis-
covered ESD damage, it must have been unexpected. Many of those who followed showed
disbelief, and skepticism remains rife in the industry to this day. Although the future can
never be accurately predicted, some general short- and medium-term trends seem likely
to continue. New technologies can of course disrupt the picture at any time. Nevertheless,
it’s likely that the future reader of this chapter may be surprised and even amused by the
difference between my predictions and the intervening reality!
Electronic components will continue to become smaller in their internal feature dimen-
sions, and this will tend to make them more susceptible to ESD damage. New device tech-
nologies will continue to be developed. Some of these will be more susceptible, and some
less susceptible to ESD.
New device packages and circuit assembly technologies will become available and more
widely used. Some may even replace current technologies, as through-hole components
have been widely replaced by surface-mounted technology today.
ESD control programs will continue to evolve, as will the standards that support them. As
low ESD withstand (very sensitive) components become more widely handled, ESD control
programs will need to develop to handle them, both in manual and automated processes.
I hope that ESD withstand data will become more widely published on component data
sheets or available to the ESD control professional on demand. Unfortunately, many com-
ponent manufacturers as yet seem unwilling to make this data easily available.
Knowledge and understanding will be more important – leading to greater need for ESD
training at high level and audit, as well as specific aspects such as ESD-related measure-
ments.
Some of these topics are discussed more fully in the following paragraphs.

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
450 13 The Future

13.2 ESD Withstand Voltage Trends


13.2.1 Integrated Circuit ESD Withstand Voltage Trends
The ESD Association publishes from time to time an “ESD Technology Roadmap” (ESD
Association 2010, 2013, 2016c). This reviews and predicts the trends in the integrated cir-
cuit (IC) industry for the forthcoming few years and their impact on ESD control practices
and trends in device testing. The roadmap focuses on the IC manufacturing industry and
does not cover devices such as magnetoresistive heads, LEDs, lasers and photodiodes, and
thin film transistor flat-panel displays. The 2016 roadmap does look at the impact of novel
technology trends such as 2.5d and 3d ICs, micro bumps, and on-package I/O. These trends
are highly dependent on technology changes as time goes on.
Typically, the roadmap gives graphs showing the human body model (HBM) and charged
device model (CDM) ESD withstand ranges from the mid-1970s and projected over a few
years into the future (Figures 13.1 and 13.2). The ranges are a projection by engineers from
leading semiconductor manufacturers. The maximum represents what is possible due to
technology scaling, and the minimum arises from meeting circuit performance demands,
which often require reduced ESD protection.
The graphs show that between the late 1970s when ESD became a widespread issue and
the mid-1990s, ESD withstand voltages increased in general. This is because with awareness
of ESD issues devices were designed to be more ESD robust, in many cases with built-in
on-chip ESD protection networks.

HBM Roadmap (Typical Min – Max)


6 kV

4 kV

De facto
Target Level
Current Target Projected Target
Level Level
2 kV

1 kV
500 V
0V
1975 1980 1985 1990 1995 2000 2005 2010 2015 2020
Copyright © 2016 ESD Association

Reproduced Figure 1, from the Electrostatic Discharge (ESD) Technology Roadmap, May 2016, by
permission of The EOS/ESD Association, Inc.

Figure 13.1 HBM ESD withstand projections. Source: ESDA (2016c).


13.2 ESD Withstand Voltage Trends 451

CDM Roadmap (Typical Min – Max)


1500 V

1250 V
De facto
Target Level

1000 V
De facto
Target Levels

750 V Current Target Projected


Level Target
Level
500 V

250 V

125 V
0V
1975 1980 1985 1990 1995 2000 2005 2010 2015 2020

Copyright © 2016 ESD Association


Reproduced Figure 1, from the Electrostatic Discharge (ESD) Technology Roadmap, May 2016, by
permission of The EOS/ESD Association, Inc.

Figure 13.2 CDM ESD withstand projections. Source: ESDA (2016c).

From the mid-1990s, withstand voltages fell again as technology pressures made achiev-
ing high ESD withstand voltage increasingly difficult.
In the 2000s a group of semiconductor manufacturers joined together to recommend new
ESD target levels (Industry Council 2010a,b, 2011). The motivation was to set ESD protec-
tion requirements for the safe handling of ICs within electrostatic discharge protected areas
(EPAs), while responding to the increasing difficulty in achieving high ESD withstand volt-
age in the face of technology scaling and other technology changes in IC design. The result
was to recommend reducing target HBM ESD withstand voltage to 1000 V and machine
model (MM) ESD withstand voltage to 30 V (Industry Council 2011). This was followed by
a recommendation to reduce CDM ESD withstand voltage target level to 250 V (Industry
Council 2010a). They concluded that “basic” ESD controls were adequate for protection
of devices with ESD withstand voltage down to 500 V HBM. “Detailed” ESD controls are
required for handling devices with ESD withstand voltages less than 500 V HBM.
The Industry Council for ESD Target Levels recommended HBM ESD withstand target
levels are reduced to 1 kV. This is probably only a first step – it seems likely that the HBM
ESD withstand target level will be further reduced to 500 V, with a corresponding reduction
in CDM ESD withstand target level. CDM target levels have already been reduced to 250 V,
and the roadmap suggests that a target of 125 V CDM will be needed in the future. Some
high-performance devices already have ESD withstand voltages as low as 100 V HBM.
The 2016 roadmap also gives an expanded view of HBM and CDM ESD withstand trends
from 2010 until beyond 2020 (Figures 13.3–13.6). These show that as time goes on, a greater
proportion of ESD-sensitive (ESDS) devices will have ESD withstand voltages between
500 and 100 V HBM, and <200 V CDM (Figures 13.5 and 13.6). In Figures 13.3 and 13.4,
452 13 The Future

HBM Forward Looking Roadmap (Typical Min – Max)

Current Target
Level
Projected Target
Level

≥2 kV

HBM Control Methods


1 kV Estimated Levels

Basic Handling
(500 V)
500 V
S20.20 (100 V)

0V
Custom Controls
2010 2015 2020 Region (<100 V)

Reproduced Figure 2, from the Electrostatic Discharge (ESD) Technology Roadmap, May 2016, by
permission of The EOS/ESD Association, Inc.

Figure 13.3 Technology Roadmap expanded 2010 and beyond HBM ESD withstand projections.
Source: ESDA (2016c).

CDM Forward Looking Roadmap (Typical Min – Max)

Current Target
Level Projected Target
Level
500 V

CDM Control Methods


Estimated Levels

250 V S20.20 (200 V)

Process Specific
125 V Measures * (100 V)

Charge / Discharge
Measurements at
0V each process step**
(50 V)
2010 2015 2020
* - Include process specific measures to avoid charging or discharge
** - Include process specific measures to avoid charging and discharge

Copyright © 2016 ESD Association

Figure 13.4 Technology Roadmap expanded 2010 and beyond CDM ESD withstand projections.
Source: ESDA (2016c).
13.2 ESD Withstand Voltage Trends 453

Figure 13.5 Projected change in distribution of


HBM withstand voltage among ICs. Source:
ESDA (2016c). > 1000 V

> 1000 V

500 – 1000 V

500 – 1000 V
< 500 V
< 500 V

2010–2015 By 2020
Reproduced Figure 3, from the Electrostatic Discharge (ESD) Technology Roadmap, May 2016, by
permission of The EOS/ESD Association, Inc.

Figure 13.6 Projected change in distribution of


> 500 V
CDM withstand voltage among ICs. Source: > 500 V
ESDA (2016ac).
250 – 500 V

250 – 500 V
125 – 250 V

125 – 250 V
< 125 V
< 125 V

2010–2015 By 2020
Reproduced Figure 7, from the Electrostatic Discharge (ESD) Technology Roadmap, May 2016, by
permission of The EOS/ESD Association, Inc.

the control levels achieved by an ESD control program compliant with ANSI/ESD S20.20
are also shown. Below these levels (100 V HBM and 200 V CDM), special “custom” and
process-specific ESD control measures are usually required.
Looking further than 2020, the roadmap suggests that the range of HBM ESD withstand
voltages may not change so much, but the distribution of devices within the range may
change, with a greater proportion falling in the 500–1000 V and 100–500 V withstand volt-
age ranges. Similarly, a greater proportion is likely to fall within the 250–500 V, 125–250 V,
and <125 V CDM withstand voltage ranges.
The drivers for these reductions in ESD withstand are rooted in technology changes. The
internal device dimensions are reducing into the 22 and 18 nm range, and this tends to
lead to inherently more sensitive internal circuitry. Increasing numbers of devices include
high-speed input/output (I/O) pins that will need to operate into the 10–30 Gbit range of
data rate. Radio frequency circuits are becoming ever more widespread. These device pins
often tolerate only minute capacitance added by ESD protection networks, before perfor-
mance is impaired. Increased device performance often comes at the expense of reducing
ESD withstand voltage.
The MM ESD withstand test is already being phased out, and future ESD withstand char-
acterizations will be normally in terms of HBM and CDM withstand voltages.
The CDM ESD susceptibility is also reduced with increasing package size. The roadmap
states that current packages of 3000 pins in land grid array (LGA) or ball grid array (BGA)
with high-speed I/O at 22 nm barely meet a CDM target of 125 V.
454 13 The Future

13.2.2 Other Component ESD Withstand Voltage Trends


While the ESDA roadmap deals with the ESD withstand voltage trends for integrated cir-
cuits, there are many types of transistor, diode, surface acoustic wave device (SAW), micro-
electromechanical system (MEMS), magnetoresistive (MR) sensor, and other devices that
are susceptible to ESD damage. Some of these are among the most sensitive devices handled
in the manufacture of electronic systems. These devices often have little or no ESD protec-
tion built into the device. It is likely that these will continue to be among the most ESDS
devices handled. Their ESD withstand voltages will continue to depend on the particular
device and technology.

13.2.3 Availability of ESD Withstand Voltage Data


There is currently scope for the improvement of the availability and publication of device
ESD withstand data. I hope that device manufacturers will eventually publish this data, for
example on device data sheets, as a matter of course.

13.2.4 Device ESD Withstand Test


MM ESD withstand voltage test is already being dropped from routine device characteri-
zation. The MM test is seen as giving little useful additional information over HBM ESD
characterization (Industry Council 2011).
HBM and CDM ESD withstand voltage test practices may continue to be developed to
more easily test high pin count and small pin spacing devices and counter increasing test
times and costs (ESD Assoc. 2016c). Statistical sampling methodologies are likely to be used
for high pin count devices. These changes will entail regular update of the relevant device
ESD test standards and test equipment. As devices become more sensitive, introduction of
lower voltage test levels may be needed.
The human metal model (HMM) test will continue to be developed for testing device pins
that can be exposed to system-level ESD. One area of research is improving the repeatability
of this test. Transient latch up (TLU) and other tests may also be further developed. Use of
transmission line pulsing as a technique for characterizing ESD protection capability will
also be further developed, with standards emerging for this type of testing.

13.3 ESD Control Programs and Process Control

13.3.1 ESD Control Program Development Strategies


While some organizations will be content to apply standard ESD control measures in ESD
control programs according to standards such as IEC 61340-5-1 and ANSI/ESD S20.20, oth-
ers will feel the need to tailor their ESD program to a greater extent. Reasons for this may
include, for example, the following:
● Handling of low ESD withstand voltage ESDS devices
● Use of processes having less common ESD risks
13.3 ESD Control Programs and Process Control 455

● A desire to optimize the ESD control program for best effectiveness and return on invest-
ment
● A desire to achieve very low risk of ESD damage (e.g. for high value or high reliability
product)
Even in the absence of these motivations, the continuing trends in reduction of device
ESD withstand voltages confirm a need for continuing improvement of ESD control proce-
dures and compliance in processes handling these devices.
The Industry Council (2011) commented that ESD Control Programs range across three
levels.
● Little or no ESD control
● Basic ESD control programs
● Detailed ESD control programs
It is likely that organizations will need to move toward better ESD controls as
time goes on, and more organizations will move into the “detailed ESD controls”
category.
Organizations that have little or no ESD control are likely to experience ESD losses even
with devices of 2 kV HBM or greater ESD withstand voltage. Without adequate personal
grounding, body voltage can easily exceed 2000 V. Even if simple ESD control measures may
be present, ESD training and compliance verification are often minimal or absent. It is no
coincidence that the key requirements of modern ESD control standards include documen-
tation of the ESD program, an ESD Control Product Qualification Plan, an ESD Training
Plan, and a Compliance Verification Plan. Omission of any of these means that the ESD
program is likely to be ineffective.
● Lack of an ESD control program document means that ESD control measures will be
poorly reproduced and key measures may become omitted.
● Lack of an ESD Control Product Qualification Plan can lead to purchase of equipment
that does not effectively fulfill its purpose in the long term.
● Lack of a Compliance Verification Plan means that ESD control equipment failures are
unlikely to be discovered and remedied.
● Lack of an ESD Training Plan means that personnel are likely to remain insufficiently
aware of the procedures and equipment required to be used in ESD control.
With the increasing proportion of components that have low ESD withstand voltage,
organizations that currently pay minimal attention to ESD control are likely to find that
increasing losses will be experienced. This will demonstrate the need for establishment of
an effective ESD control program having at least good basic control measures.

13.3.2 A Basic ESD Control Program


A basic ESD control program implements the standard ESD control measures with little or
no redundancy (Industry Council (2011). The principles of ESD control are that
● ESDS devices are handled only within an EPA in which ESD risks are controlled.
● Outside the EPA, components are protected within ESD protective packaging.
456 13 The Future

The techniques used to control ESD risks within the EPA are that
● All conductors especially personnel who might contact ESDS devices are equipotential
bonded, preferably to earth.
● Unnecessary insulators are removed from the vicinity of ESDS devices.
● Insulators that are necessary to the process or product are evaluated for ESD risk. Unac-
ceptable risks are reduced by some means to an acceptable level. Risks are normally
evaluated in terms of electrostatic fields and voltages.
For long-term success, it must include documentation and implementation of the ESD
program, an ESD Control Product Qualification Plan, an ESD Training Plan, and a Compli-
ance Verification Plan.
The current ESD control program standards IEC 61340-5-1:2016 and ANSI/ESD
S20.20-2014 are designed to protect devices down to 100 V HBM and 200 V CDM. ESD
from charged conductors is limited to 35 V.
A basic ESD control program well implemented can be very effective, providing any fail-
ures of ESD control equipment are quickly detected and remedied.

13.3.3 Detailed ESD Control Program


Detailed ESD control programs add redundancy to the ESD control program and may use
constant monitoring systems to make sure that failures are quickly detected (Industry
Council 2011). Redundancy helps prevent single equipment failures causing ESD failures,
as a second control measure may also help control the ESD risk.
Handling devices with ESD withstand voltages below 500 V HBM typically requires a
more careful level of ESD control, as control equipment failures will often quickly lead to
ESD failures occurring. Handling devices with ESD withstand voltages below 100 V HBM
takes an ESD control program out of the range of standards such as IEC 61340-5-1 and
ANSI/ESD S20.20 and into a region where more specific controls are needed.
The ESD roadmap predicts that tighter limits will be required for ESD control, and possi-
bly more frequent compliance verification. A reduction in CDM ESD withstand voltage is
expected due to increasing speed, package sizes, and complexity with multichip packages
such as 2.5D and 3D technology.
As device ESD withstand voltages reduce, increased care, advanced, and nonstandard
ESD controls will be increasingly necessary. Processes will need to be evaluated with care
to identify specific ESD threats. These may need to be countered using specifically designed
control measures. ESD risk will be determined less by checklist with fixed limits and more
by application an engineering approach by skilled ESD coordinators (Jacob et al. 2012).
The current standard design limits of 100 V HBM, 200 V CDM ESD withstand, and ±35 V
on isolated conductors may need to be reduced as lower ESD withstand voltage devices
become more common.

13.3.4 Human Body ESD


Control of human body ESD using a wrist strap or ESD control footwear and flooring is
a mature and well-understood topic. These solutions work well, providing users correctly
implement standard ESD control requirements. This must include
13.3 ESD Control Programs and Process Control 457

● Use of wrist straps when seated


● Qualification of footwear and flooring by resistance from person to ground, for each com-
bination of footwear and flooring used
● Qualification of footwear and flooring by a body voltage walk test, for each combination
of footwear and flooring used
● Compliance verification of footwear and flooring by resistance from person to ground,
for each combination of footwear and flooring used

At the time of writing, many users have yet to adequately adopt adequate footwear and
flooring compliance verification procedures as per the standards. Use of walk tests is partic-
ularly important when handling low HBM ESD withstand voltage devices, as the objective
is to keep body voltage below a low level related to the HBM withstand voltage of the
devices handled. Studies have shown that body voltage generated is, for some combina-
tions of footwear and flooring, badly predicted by the resistance of the footwear and flooring
individually and even in combination.
Working with low ESD withstand devices may require some users adopt lower body volt-
age limits to reflect the ESD withstand voltage of the devices they handle. For example, an
upper limit for body voltage of 50 V might be selected when working with 50 V HBM ESD
withstand devices.

13.3.5 ESD Between ESDS and Conductive Items


ESD between ESDS and conductive items include “two-pin” ESD from charged ungrounded
conductive items and “one-pin” ESD between the ESDS and another conductive part at a
different voltage. A “two-pin” ESD event is one in which the ESD current enters the device
through one pin and passes out through another pin. Susceptibility to “two-pin” ESD from
charged ungrounded conductive items was formally represented by the MM ESD withstand
test. This test is now considered redundant and is falling out of use, in part because it typi-
cally gives failures similar to the HBM test but with about 1/30th the withstand voltage or
more. A part that shows 1 kV HBM ESD withstand would typically be expected to give >30 V
MM ESD withstand (Industry Council 2011). This can be justified by theoretical arguments
as well as practical experience.
A “one-pin” ESD event is one in which the ESD is directly between one pin of the ESDS
device and another conductive object, with no other ESDS pin being involved in the dis-
charge. Susceptibility to a “one-pin” ESD event is represented by the CDM ESD suscepti-
bility test.
As automated assembly increases, the threat from human body ESD is reduced except in
manual parts of processes. The threat from ESD between the ESDS devices and conductive
parts correspondingly represents a greater proportion of ESD risks.
At present in the ESD control standards, the threat of ESD from charged conductors is
controlled by two requirements.

● Conductors that make contact with ESDS devices must be grounded if possible.
● Conductors that make contact with ESDS devices but cannot be grounded must have the
voltage difference between the conductor and ESDS limited to ±35 V.
458 13 The Future

These requirements address both one-pin and two-pin ESD threats. The charged device
ESD threat is also addressed by limitation of electrostatic field in the vicinity of the ESDS
device. Both types of ESD damage can also be prevented by preventing contact between the
ESDS devices and low-resistivity conductive items.
In practice, it is difficult to measure and verify the voltage difference between the ESDS
device and an ungrounded conductor. To do so normally requires measurement of the volt-
age on the ESDS, measurement of the voltage on the conductor, and then calculating the dif-
ference. These can be difficult measurements to make, especially in an operating machine
process, as the ESDS device and conductor may be small low-capacitance items. Special
electrostatic voltmeters may be required to make these measurements. So, it remains to be
seen how practical these requirements are in practice. The standards also do not yet define
in a measurable way what is meant by an ungrounded conductor.

13.3.6 “Two-Pin” ESD From Charged Ungrounded Conductive Items


The requirement that the voltage difference between the conductor and ESDS device lim-
ited to ±35 V can be viewed as addressing two-pin ESD risk from conductors up to 200 pF
capacitance (the MM source capacitance). This protects devices with MM ESD withstand
as low as to 35 V, corresponding to HBM ESD withstand around 1000 V. It could be argued
that as lower voltage HBM withstand parts become more commonly used, this voltage value
might have to reduced.
In practice, the real threat from ESD from conductors is dependent on the capacitance
of the conductor. Many conductors in real situations may have lower capacitance than
the 200 pF MM value. Some may have higher capacitance. For a lower-capacitance item,
a higher voltage could be tolerated, and for higher capacitance a lower voltage may be
required, based on the lowest ESD withstand device handled. The threat may also depend
on whether the device is susceptible to ESD current, charge, or energy transferred in the
discharge (Smallwood and Paasi 2003; Paasi et al. 2003).

13.3.7 “One-Pin” ESD Between the ESDS and Another Conductive Part
Over the last decade or so, there has been considerable interest in developing a methodol-
ogy for relating process capability to device CDM ESD withstand data. This is particularly of
interest in automated processes but also in understanding ESD risks from contact between
devices and metal objects (whether grounded or not). Some researchers have used existing
CDM withstand voltage as the parameter (e.g. Steinman 2010, 2012), but other researchers
have proposed that it is necessary to record peak current or other CDM waveform data,
rather than the source voltage, for comparison with real-world conditions (e.g. Tamminen
and Viheriäkoski 2007; Tamminen et al. 2017a,b). At the same time, it has proven difficult
to make in-process measurements for comparison with CDM waveform parameters. It’s
likely that this area will continue to be researched over the forthcoming decade especially
as device CDM withstand voltages are reduced. New measurement techniques or methodol-
ogy may be needed before success is achieved in this area (Tamminen et al. 2017a,b). Device
ESD withstand measurement practices may need to be modified to record the required data.
13.4 Standards 459

13.3.8 Charged Board, Module, and Cable Discharge Events


Charged board and cable discharge events happen when a PCB contacts a conductive item
or a cable at a different voltage is connected. The PCB, or the cable, or other conductive
object may be the charged item. These events can be highly energetic compared to charged
device ESD, because the PCB or cable have high capacitance compared to devices. The
resulting damage can look more like electrical overstress than ESD (Olney et al. 2003).
Modules and subassemblies are often potted or enclosed by plastic housings that can
easily become charged by triboelectrification by rubbing during handling or contact with
packaging. These assemblies are often assumed to be immune to ESD due to the protection
afforded by the enclosure as a barrier against direct ESD. Polymer potting compounds and
enclosures are, however, “transparent” to electrostatic fields from nearby charged materials.
In an uncontrolled area, charged personnel can be an external source of electrostatic fields
when holding handheld items. Internal PCBs and other conductors can rise to very high
voltages by induction and then discharge by contact of a terminal with another conductor.
Cables can become charged by triboelectrification through contact and rubbing while on
a reel, within packaging or being moved over surfaces. They can also become inductively
charged by the electrostatic field from nearby charged insulators such as packaging materi-
als. Charged cables can have high and variable capacitance of hundreds of pF. So, a charged
cable or wiring loom is a very energetic ESD source. This is often discharged into a PCB con-
nector during assembly of a product or installation of a system. The circuitry connected to
the PCB connector may not be designed to survive high levels of ESD, particularly if they are
internal system connections rather than those exposed to the outside world (Stadler et al.
2017; Marathe et al. 2017).
These ESD risks are not specifically covered by current ESD standards, but it’s possible
that they may become so in the future.

13.3.9 Optimization
Optimization of an ESD program in terms of return on investment has long been an aim for
some ESD control practitioners. Nevertheless, compliance with standards has often been
an overriding goal, particularly in organizations where understanding of ESD controls has
been at a lower level. As greater expertise is applied in successful ESD control, that expertise
will be also applicable to optimization. More organizations will be able to achieve improved
optimization of ESD control alongside compliance with the standards.

13.4 Standards

The core ESD control standards IEC 61340-5-1 and ANSI/ESD S20.20 will continue to be
refined and developed for the foreseeable future. At their current stage of maturity, the
changes are expected to be incremental rather than major. But, experience with using the
standards, the increasing use of low ESD withstand voltage devices, and developing ESD
control approaches will be reflected in rewording and changes to the detailed requirements
of the standards. As an example, the 61340-5-1:2007 standard set a limit of electrostatic
460 13 The Future

fields at 10000 V m−1 , and this was reduced to 5000 V m−1 in the 61340-5-1:2016 update. The
latter also added some requirements for dealing with ungrounded conductors that make
contact with ESDS devices. The ANSI/ESD S20.20 often leads the way in this development,
with a new version often appearing a year or so before the 61340-5-1. New versions of the
User Guides IEC 61340-5-2 and ESD TR20.20 usually accompany the new standard versions
(International Electrotechnical Commission 2018a; EOS/ESD Association Inc 2016b).
Alongside this, some of the related test methods and other standards undergo develop-
ment and change with experience of their usage. New test methods may emerge for specific
applications. Methodologies for process evaluation are likely to be improved and developed
and may emerge as standard practice or full standards.

13.4.1 Impact on Future Standards


It is likely that changes in approach to ESD program development and process control (see
Section 13.3) may have to be addressed in future versions of the standards. What constitutes
an ungrounded conductor will need better definition in terms of measurable parameters,
e.g. resistance to ground.
It may be necessary to move from simplified “coverall” requirements to measures more
carefully tailored to the real component susceptibility and failure mode. For example, volt-
age limits on conductors could be linked to the capacitance of the conductor and ESD
HBM and CDM withstand voltage of devices handled. It may be possible to establish that
small ungrounded conductors below a capacitance limit do not pose a threat to devices
having withstand voltage above a given level. Conversely, very sensitive devices, especially
voltage-sensitive devices (e.g. low gate capacitance low voltage MOSFETs) or those that
might be damaged by high transient peak current levels may require very careful control or
elimination of contact with ungrounded conductors, especially those of low resistivity.

13.4.2 ESD Control in Automated Handling


So far, ESD control in automated handling equipment has escaped standardization,
although the ESD Association has produced the Standard Practice document ANSI/ESD
SP10.1. Part of the problem has been difficulty in providing appropriate standardized
ESD controls for automated equipment and processes. Standards apply requirements that
then are necessary for compliance, whether they are needed for successful ESD control
in a particular situation. An inappropriately specified requirement can represent an
unwelcome cost increase or may be difficult to achieve in a machine for technical reasons.
In principle, the controls given in 61340-5-1 and ESD S20.20 (grounding of conductors,
treatment of ungrounded conductors, and control of insulators) are also applicable within
the critical path of ESDS within a machine. So, it is possible that standardization could
progress in one of three ways.

● The 61340-5-1 and S20.20 documents could be developed to directly address automated
systems. Guidance for application to automated systems could be incorporated in the
61340-5-2 and ESD TR20.20 User Guide documents.
13.6 ESD-Related Measurements 461

● Separate standards could be developed covering requirements applicable to automated


equipment.
● Separate documents (Technical Report (TR), Technical Specification (TS) in the IEC sys-
tem, or Standard Practice (SP) in the ESDA system) could be further developed that give
guidance rather than requirements for compliance.

13.5 ESD Control Equipment and Materials


13.5.1 ESD Control Materials
New materials will continue to be developed for ESD control as process control require-
ments become more demanding. Techniques for evaluating material characteristics will
also evolve as understanding of ESD risks improves (Viheriäkoski et al. 2017).

13.5.2 ESD Protective Packaging


The development of new component packages and automated system handling techniques
will require new ESD protective packaging types and materials. The IEC has produced
the IEC TR 61340-5-5 technical report to review current and future requirements for ESD
packaging and related test methods (International Electrotechnical Commission 2018b).
The diversity of general ESD protective packaging types, and the materials they are made
from, will increase with time. This will generate a need for new and better ways of eval-
uating packaging materials and systems, and a general improvement in understanding of
ESD packaging effectiveness. This may lead to new requirements for ESD packaging. For
example, the performance of an ESD barrier layer could be defined in terms of volume
resistance and breakdown voltage of the package material. For some types of components,
transferred charge and induced current on the ESDS within the package, as well as energy
attenuation, could be found to be important.
Electrostatic attraction has already become an issue where small components are pack-
aged for automated handling (D E Swenson private communication 2017). This trend is
likely to continue as the number of very small components proliferates.

13.6 ESD-Related Measurements


13.6.1 ESD Protective Packaging Measurements
Development of packaging forms and materials will require development of improved test
methods for ESD protective packaging materials and systems. At the time of writing, there
is already a need to develop resistance test electrodes for use with small feature and non-
planar ESD packaging materials. Improved understanding of the necessary requirements
for ESD protective packaging for different ESDS devices and applications may also lead to
development of new test methods.
462 13 The Future

13.6.2 Voltage Measurement on ESDS Devices and Ungrounded Conductors


Voltage measurements on small ESDS devices and ungrounded conductors are challeng-
ing, especially in operating automated equipment. The different types of voltage measure-
ment equipment, their capabilities, and their limitations are currently poorly understood
within industry except among the more expert practitioners. There will be a need for more
widespread understanding and use of these techniques.
It is likely that these measurements will remain a central part of ESD risk evaluation in the
medium term, despite the realization that voltage is often not the parameter best correlated
to ESD risk. The lack of viable alternatives and established industry mind-set continues to
frustrate adoption of other parameters as measures of ESD risk.

13.6.3 Measurements Related to ESD Risk in Automated Handling Equipment


Measurements related to ESD risk in automated handling equipment under working condi-
tions is currently one of the most pressing problems for ESD control. Consequently, we can
expect to see continuing developments in this area over the next few years. It is difficult to
predict whether solutions will be found in direct or indirect measurement of ESD current,
electromagnetic radiations, or some other parameter related to the ESD.

13.7 System ESD Immunity


While it has been demonstrated that in general system ESD immunity is not related to
component ESD susceptibility (Industry Council 2010a,b, 2012), a system-efficient electro-
static discharge design (SEED) approach has been suggested. This addresses failure related
to high-level ESD applied to device pins that are connected to external interfaces such as
universal serial bus (USB) connectors, as well as “soft” failures due to conducted or radi-
ated electromagnetic interference from ESD (Stadler et al. 2017; Marathe et al. 2017). This
approach may become more widely used and further developed in the future.

13.8 Education and Training


The industry is moving out of an era in which ESD control can be specified effectively
by standard measures and recipes and toward increased specification through engineer-
ing principles, knowledge, and understanding. This will require a greater level of specialist
education and training to be available to industry practitioners. This will require a greater
availability of high-level industry courses including continual professional development.
Ideally, I hope that more universities and colleges will see value in teaching ESD control at
different levels as part of their electronics-related courses.

References

EOS/ESD Association Inc (2010). ESD Association Electrostatic Discharge (ESD) Technology
roadmap. Revised March 2010 Available from: https://www.esda.org/assets/Uploads/docs/
2013ElectrostaticDischargeRoadmap.pdf (Accessed: 15th May 2017).
References 463

EOS/ESD Association Inc. (2013). ESD Association Electrostatic Discharge (ESD) Technology
roadmap. Revised March 2013 Available from: https://www.esda.org/assets/Uploads/docs/
2013ElectrostaticDischargeRoadmap.pdf (Accessed: 10th May 2017).
EOS/ESD Association Inc. (2014). ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016a) ANSI/ESD SP10.1-2016. Standard practice for protection of
Electrostatic Discharge Susceptible Items – Automated handling Equipment (AHE). Rome, NY,
EOS/ESD Association Inc.
EOS/ESD Association Inc. (2016b). ESD TR20.20-2016. ESD Association Technical Report -
Handbook for the Development of an Electrostatic Discharge Control Program for the Protection
of Electronic Parts, Assemblies and Equipment. Rome, NY, EOS/ESD Association Inc.
EOS/ESD Association Inc (2016c) ESD Association Electrostatic Discharge (ESD) Technology
roadmap. revised 2016. Available from: https://www.esda.org/assets/Uploads/docs/
2016ESDATechnologyRoadmap.pdf (Accessed: 10th May 2017).
Industry Council on ESD Target Levels (2010a) White paper 2: A case for lowering component
level CDM ESD specifications and requirements. Rev. 2.0. http://www.esdindustrycouncil
.org/ic/en/documents/6-white-paper-2-a-case-for-lowering-component-level-cdm-esd-
specifications-and-requirements (Accessed: 10th May 2017)
Industry Council on ESD Target Levels (2010b) White paper 3: System Level ESD Part I:
Common Misconceptions and Recommended Basic Approaches. Rev. 1.0 http://www
.esdindustrycouncil.org/ic/en/documents/7-white-paper-3-system-level-esd-part-i-
common-misconceptions-and-recommended-basic-approaches (Accessed: 10th May 2017).
Industry Council on ESD Target Levels (2011) White paper 1: A case for lowering component
level HBM/MM ESD specifications and requirements. Rev. 3.0. Available from: http://www
.esdindustrycouncil.org/ic/en/documents/37-white-paper-1-a-case-for-lowering-
component-level-hbm-mm-esd-specifications-and-requirements-pdf (Accessed: 10th May
2017).
Industry Council on ESD Target Levels (2012) White paper 3: System Level ESD Part II:
Implementation of Effective ESD Robust Designs. Rev. 1.0 http://www.esdindustrycouncil
.org/ic/en/documents/36-white-paper-3-system-level-esd-part-ii-effective-esd-robust-
designs (Accessed: 10th May 2017).
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016. Electrostatics – Part
5–1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Electrotechnical Commission. (2018a) IEC TR 61340-5-2:2018.
Electrostatics – Part 5-2: Protection of electronic devices from electrostatic phenomena - User
guide. ISBN 978-2-8322-5445-5 Geneva, IEC.
International Electrotechnical Commission. (2018b) IEC TR 61340-5-5:2018. Electrostatics -
Part 5-5: Protection of electronic devices from electrostatic phenomena – Packaging systems
used in electronic manufacturing. Geneva, IEC.
Jacob, P., Gärtner, R., Gieser, H. et al. (2012). Paper 3B.8. ESD risk evaluation of automated
semiconductor process equipment – A new guideline of the German ESD Forum e.V. In:
Proc. EOS/ESD Symp. EOS-34. Rome, NY: EOS/ESD Association Inc.
Marathe, S., Wei, P., Ze, S. et al. (2017). Paper 3A.4. Scenarios of ESD Discharges to USB
Connectors. In: Proc. EOS/ESD Symp. EOS-39. Rome, NY: EOS/ESD Association Inc.
464 13 The Future

Olney, A., Gifford, B., Guravage, J., and Righter, A. (2003). Real-world charged board model
(CBM) failures. In: Proc. EOS/ESD Symp. EOS-25, 34–43. Rome, NY: EOS/ESD Association
Inc.
Paasi, J., Salmela, H., and Smallwood, J.M. (2003). New methods of assessment of ESD threats
to electronic components. In: Proc EOS/ESD Symp. EOS-25, 151–160. Rome, NY: EOS/ESD
Association Inc.
Smallwood, J. and Paasi, J. (2003). Assessment of ESD threats to electronic components and
ESD control requirements. In: Proc. Electrostatics 2003. Inst. Phys. Conf. Ser. No. 178 Section 6,
247–252.
Stadler, W., Niemesheim, J., and Stadler, A. (2017). Paper 3A.1. Risk assessment of cable
discharge events. In: Proc. EOS/ESD Symp. EOS-39. Rome, NY: EOS/ESD Association Inc.
Steinman, A. (2010). Paper 3B3. Measurements to Establish Process ESD Compatibility. In:
Proc. EOS/ESD Symp. EOS-32. Rome, NY: EOS/ESD Association Inc.
Steinman, A. (2012). Paper 2B.4. Process ESD Capability Measurements. In: Proc. EOS/ESD
Symp. EOS-34. Rome, NY: EOS/ESD Association Inc.
Tamminen, P. and Viheriäkoski, T. (2007). Paper 3B.3. Characterization of ESD risks in an
assembly process by using component-level CDM withstand voltage. In: Proc. EOS/ESD
Symp. EOS-29, 202–211. Rome, NY: EOS/ESD Association Inc.
Tamminen, P., Smallwood, J., and Stadler, W. (2017a). Paper 1B.4. Charged device discharge
measurement methods in electronics manufacturing. In: Proc. EOS/ESD Symp. EOS-39.
Rome, NY: EOS/ESD Association Inc.
Tamminen, P., Smallwood, J., and Stadler, W. (2017b). Paper 4B.2. The main parameters
affecting charged device discharge waveforms in a CDM qualification and manufacturing.
In: Proc. EOS/ESD Symp. EOS-39. Rome, NY: EOS/ESD Association Inc.
Viheriäkoski, T., Kärjä, E., Gärtner, R., and Tamminen, T. (2017). Paper 4B.3. Electrostatic
discharge characteristics of conductive polymers. In: Proc. EOS/ESD Symp. EOS-39. Rome,
NY: EOS/ESD Association Inc.

Further Reading

Fung, R., Wong, R., Tsan, J., and Batra, J. (2017). Paper 1B.3. An ESD case study with high
speed interface in electronics manufacturing and its future challenge. In: Proc. EOS/ESD
Symp. EOS-39. Rome, NY: EOS/ESD Association Inc.
Koh, L.H., Goh, Y.H., and Wong, W.F. (2017). Paper 1B.2. ESD Risk Assessment Considerations
for Automated Handling Equipment. In: Proc. EOS/ESD Symp. EOS-39. Rome, NY:
EOS/ESD Association Inc.
465

Appendix A: An Example Draft ESD Control Program

A.1 About This Plan


This appendix shows an example electrostatic discharge (ESD) program derived using the
procedure of Chapter 10.
It should not be assumed that this example is a “model” ESD program or that the sug-
gested procedures in writing an ESD control program must be followed. It is merely an
example of how things might be approached. Each organization should develop their own
ESD Control Program Plan according to their process and facility needs.
This ESD program has been derived as suggested in Chapter 10. Headings from the stan-
dard have been used to create a structure for the plans and make sure the required items are
covered. The needs of the facility and anticipated operations are also considered to make
sure that necessary equipment and convenient control measures are included. The order
of the heading may have been changed, and in some cases, headings were deleted because
they were unnecessary in an ESD control program plan.

A.2 Description of the Example Facility


The example considers a simple facility in which ESD-sensitive (ESDS) printed circuit
boards (PCBs) and assemblies are received at Goods In, removed from their packaging,
and individually repacked in bags for dispatch. In some workstations, limited test by
connection to electrical mains voltages may be done.
In this example, most ESDS devices are contained within ESD protective packaging that
is specified in customer contracts and marked as specified in the contracts. Nevertheless,
some ESD protective packaging used within electrostatic discharge protected areas (EPAs)
that is not specified in this way.

The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.


© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
466 Appendix A: An Example Draft ESD Control Program

In some cases, operators will enter the EPA and work for a short time handling unpro-
tected ESDS devices while standing. In this case, personal grounding will be done most
conveniently by wearing ESD control footwear and standing on an ESD control floor. In
other cases, operators will sit at the workstation for some time handling unprotected ESDS
devices, and a wrist strap will be required. ESD control seats will also be grounded using an
ESD control floor.
A typical EPA in the imagined facility includes

● An ESD workstation with a surface on which unprotected ESDS devices may be placed
● An ESD control floor mat for grounding standing personnel (through footwear) and ESD
control chair
● An ESD control chair
● Wrist straps for grounding seated personnel, and earth bonding points of connection of
straps
● Containers such as tote boxes
● ESD control garments (coats)

A.3 Test and Qualification Procedures

The ESD control Program Plan contains an ESD Control Product Qualification Plan and a
Compliance Verification Plan. These require several qualification procedures (QPs) and test
procedures (TPs). These are not given here but would be defined and documented including
the test equipment to be used, details of the test method, and a simple test procedure.

A.4 ESD Control Program Plan at XXX Ltd

A.4.1 Introduction
This document provides an ESD control program for XXX Ltd. in compliance with IEC
61340-5-1:2016 and ANSI/ESD S20.20-2014.

A.4.2 Scope
This ESD program covers all activities that manufacture, process, assemble, install, pack-
age, label, service, test, inspect, transport, or otherwise handle ESDS devices at XXX Ltd.
Specifically, these activities include receipt of ESD-susceptible (ESDS) PCBs and assemblies
at Goods In, removal from their ESD protective packaging for inspection, and individual
repacking in bags for dispatch.
The ESD withstand voltages of the individual types of ESDS devices are not known. The
ESD program is designed to be compliant with the current standards and capable of han-
dling electrical or electronic parts, assemblies, and equipment with ESD withstand voltages
greater than or equal to 100 V human body model (HBM), 200 V charged device model
(CDM) and ±35 V for isolated conductors.
A.5 Personal Safety 467

A.4.3 Terms and Definitions


The following definitions and terms are used in this ESD control program:

CDM Charged device model. ESD stress model that approximates the
discharge event that occurs when a charged component is quickly
discharged to another object at a different electrostatic potential.
EPA ESD protected area. Area in which an ESDS can be handled with
accepted risk of damage as a result of electrostatic discharge or
fields.
ESDS ESD-susceptible device. A sensitive device, integrated circuit, or
assembly that may be damaged by electrostatic fields or
electrostatic discharge.
HBM Human body model. ESD stress model that approximates the
discharge from the fingertip of a typical human being onto a pin of
a device with another pin grounded.
Shall “Shall” indicates a requirement of the ESD control program. If it is
stated that some aspect shall be done, and if it is not, this
constitutes a noncompliance.
Tailoring Modification of the requirements of the standard after evaluation
of the applicability of each requirement. Requirements may be
added, modified, or deleted. Tailoring decisions, including
rationale, and technical justification, shall be documented.
Static dissipative A packaging material or item that has surface resistance <100 GΩ
and >10 kΩ as per IEC 61340-5-3.
Conductive A packaging material or item that has surface resistance <10 kΩ as
per IEC 61340-5-3.
Insulator A materials or item that has surface resistance >100 GΩ as per IEC
61340-5-3.
Conductor Item or material that has surface resistance <100 GΩ.
ESD shielding packaging ESD protective packaging compliant with IEC 61340-5-3 that has
static dissipative inner and outer surfaces and a barrier layer or air
gap to prevent ESD attenuate ESD current passing from the outside
to the inside of the package.

A.5 Personal Safety

Compliance with local safety laws, regulations, and requirements shall take precedence
over ESD control requirements.
Where use of personal protective equipment (PPE) is required, this shall be specified
in line with safety requirements and, where possible, also in compliance with ESD con-
trol requirements. PPE specified for use within this ESD control program shall be specified
within this ESD Control Program Plan.
To avoid electric shock risk, personnel who might come into accidental contact with high
voltages shall be protected by control measures that limit the current flow under fault con-
ditions to less than 0.5 mA. To limit current and provide protection against electric shock
468 Appendix A: An Example Draft ESD Control Program

in case of accidental contact of the wearer with 250 VAC electric power systems, the min-
imum resistance from the wearer’s body tested to groundable point (wrist strap) or metal
plate (footwear as worn) shall be 750 kΩ (see Section A.7.2 of this plan).

A.6 ESD Control Program


A.6.1 ESD Control Program Requirements
This ESD Control Program Plan shall cover the following:
● The requirements of the ESD program
● ESD training
● ESD control product qualification
● ESD control product compliance verification
● EPA ground and grounding and bonding systems
● Personal grounding systems
● EPA requirements
● ESD protective packaging
● Marking for ESD control and protection purposes

A.6.2 ESD Coordinator


An ESD coordinator shall be appointed. The ESD coordinator shall have responsibility for
implementing the ESD control program including establishing, documenting, maintaining,
and verifying the compliance of the program as well as updating the program in response
to changes in practices, operations, or facilities. Other personnel may be appointed as nec-
essary to assist the ESD coordinator in these tasks.

A.6.3 Tailoring ESD Control Requirements


Where necessary the requirements of the standard may be omitted or modified providing
the rationale for these is documented and referenced by this ESD control program. The
rationale for requirements added to those of the standard shall also be documented and
referenced.
Aspects of this ESD control program that are subject to tailoring include use and handling
of essential papers (see Section A.7.3.2).

A.7 ESD Control Program Technical Requirements


A.7.1 ESD Ground
Mains electrical safety protective earth is used as ESD control earth (ground).
A.7 ESD Control Program Technical Requirements 469

Table A.1 Personal grounding equipment requirements and test criteria.

ESD control item Test procedure Pass criteria Test frequency

Wrist strap and cord TP1 >750 kΩ <35 MΩ Each day before first use
when worn, wearer’s
body to cord end
groundable point
Footwear when worn, TP2 >750 kΩ <100 MΩ Each day before first use
tested between wearer’s
body and one foot placed
on metal plate
Temporary wrist strap Visual inspection Correctly connected Each day before first use
bonding point
Wrist strap bonding point TP3 <1 MΩ Monthly

A.7.2 Personal Grounding


All personnel shall be grounded when handling ESDS, either by wearing

● A wrist strap connected to an wrist strap bonding point


● ESD control footwear on both feet and standing on an ESD control floor

Personnel handling ESDS when seated shall be grounded using a wrist strap. Wrist straps
shall make good direct contact with the wearer’s skin.
The personal grounding equipment requirements and test criteria are listed in Table A.1.

A.7.3 ESD Protected Areas (EPA)


A.7.3.1 General EPA Requirements
ESDS devices shall be handled out of ESD protective packaging only when within an EPA.
ESDS devices shall be within designated packaging at all times when taken outside an EPA.
EPA boundaries shall be clearly defined and identified with suitable boundary markings
and signage at the entrances (For example Figure A.1).
EPA workstations shall not be used for purposes other than operations and processes in
which ESDS are handled.
Personnel who have completed and passed ESD awareness training may enter the EPA
unaccompanied. Personnel needing to enter the EPA who have not completed this training
shall be accompanied by a trained person at all times while within the EPA.

Figure A.1 Example of signage showing the entrance


to an EPA.
470 Appendix A: An Example Draft ESD Control Program

A.7.3.2 Insulators
The following items are considered likely to be made from insulators unless tested and
approved for use within an EPA:

● Secondary (non-ESD protective) packaging materials


● Items made of plastics
● Personal items
● Clothing
● Papers
● Furniture

All items made of insulators shall be excluded from EPAs unless they have been desig-
nated essential to the process or operations.

Tailored Requirement for Essential Papers


It is deemed essential to the inspection process that records are kept and signed off on in
paper form as inspection of ESDS devices is completed. These papers and associated pens
shall be kept and used in a designated area at least 30 cm from the position where unpro-
tected ESDS devices are handled. When not in use, papers shall be kept in static dissipative
document holders.
Essential papers shall never be brought into the EPA directly from photocopying as they
may be highly charged. Photocopied papers should be left for at least an hour under mod-
erate humidity atmosphere (>30% rh) after copying for charge to dissipate before bringing
them into the EPA.

Electrostatic Fields and Voltages


Electrostatic fields at the workstation where ESDS devices are handled shall be maintained
<5000 V m−1 measured using TP6. Any insulators or electrostatic field sources that can have
surface voltage >2 kV measured using TP7 shall be kept >30 cm from positions where ESDS
devices are handled. Any insulators or electrostatic field sources that can have surface volt-
age >125 V measured using TP7 shall be kept >2.5 cm from positions where ESDS devices
are handled.

A.7.3.3 Isolated Conductors


Any conductor that has resistance to ground >100 GΩ shall be considered isolated. Any
conductor that has resistance to ground <1 GΩ can be considered adequately grounded.
All conductors that make contact with ESDS devices shall be grounded unless covered by
ESD risk evaluation and verified voltage requirements below. Small isolated conductors that
do not make contact with ESDS devices (e.g. hand tool tips) are permitted. Where isolated
conductors make contact with ESDS devices, the ESD risk should be evaluated and suitable
ESD controls devised and documented in the ESD Control Program Plan.
Isolated conductors shall not make contact with ESDS devices unless the voltage can be
demonstrated to be controlled to < ±35 V.
A.7 ESD Control Program Technical Requirements 471

A.7.3.4 ESD Control Equipment


All equipment used in EPAs shall be qualified and approved for EPA use.

A.7.4 ESD Protective Packaging


Where ESD protective packaging is specified in customer contracts or other related docu-
mentation, the specified packaging shall be used to contain ESDS devices whenever they
are to be taken out of the EPA. In all other cases, ESD shielding packaging shall be used to
contain ESDS devices whenever they are to be taken out of the EPA.
ESD protective packaging used within the EPA may be static dissipative, conductive,
or ESD shielding. Packaging materials that have surfaces that are insulators shall not be
brought into the EPA.
Pink polythene shall not be brought into EPAs. The properties of this material are highly
dependent on age and atmospheric humidity.
Marking of ESD protective packaging shall be as specified in Section A.7.5.

A.7.5 Marking of ESD-Related Items


A.7.5.1 General
Where marking of items for ESD control purposes is specified in customer contracts or other
related documentation, the specified marking shall be used.
In other cases, where marking ESD control equipment or items for ESD control purposes
is deemed necessary, it shall be defined in this plan.

A.7.5.2 Marking of ESD Protective Packaging


Where marking of ESD protective packaging is specified in customer contracts or other
documentation, the specified marking shall be used.
In all other cases, ESD shielding packaging shall be used to contain ESDS whenever they
are to be taken out of the EPA.
The outer surface of ESD protective packaging of ESDS
devices taken out of the EPA shall be marked with a symbol
identifying it as ESD protective packaging, for example the sym-
bol given in Figure A.2. The main packaging function should be
indicated by a letter code as follows:
● S to mean electrostatic discharge shielding
● F to mean electrostatic field shielding Figure A.2 ESD
● C to mean electrostatic conductive shielding packaging
● D to mean electrostatic dissipative symbol. (Reproduced
by permission of the
EOS/ESD Association
A.7.5.3 Marking of ESD Control Equipment Used in EPAs Inc.)
ESD control equipment other than packaging that can be used inside or outside the EPA
shall be identified by markings showing their ESD control properties. The preferred mark-
ing is the ESD protective symbol shown in Figure A.2 without the letter code.
472 Appendix A: An Example Draft ESD Control Program

Table A.2 EPA equipment requirements and test criteria.

ESD control item Test method Pass criteria Test frequency

Work surface or TP4. Measurement of resistance < 1 GΩ 6 monthly


storage rack to ground from surface
Floor TP4. Measurement of resistance < 1 GΩ 6 monthly
to ground from surface
Seat TP4. Measurement of resistance < 1 GΩ 6 monthly
to ground from. surface
ESD control TP5. Point-to-point resistance of < 1 GΩ 6 monthly
garments ESD garments
ESD Protective TP8 Surface resistance < 100 GΩ 6 monthly
packaging

A.8 Compliance Verification Plan


All items required for ESD control must be tested on a regular basis. Suitable test meth-
ods and equipment used shall be defined in company test procedures based on IEC TR
61340-5-4:2019.
The test procedures used, pass criteria, and frequency of testing are given in Tables A.1
and A.2. The ambient temperature and humidity at the time of test shall be recorded. Test
records shall be kept by the ESD coordinator for a minimum of two years.
When possible, occasional checks should be made under low ambient humidity condi-
tions (< 30% rh).

A.9 ESD Training Plan


A.9.1 General Requirements of the ESD Training Plan
The personnel who need ESD training and the type of training given are defined in
Table A.3. Suitable tests shall be defined for evaluation of trainee comprehension and
ability to execute required ESD control procedures.

Table A.3 ESD training matrix.

Principles and
ESD ESD audit and practice of
Personnel awareness measurements EPA cleaning ESD control

Managers and supervisors yes


responsible for EPAs
All personnel who handle ESDS yes
or enter the EPA
ESD coordinator yes
Cleaners yes yes
ESD auditing personnel yes yes
A.9 ESD Training Plan 473

A.9.2 Training Records


Current ESD training records shall be kept by the human resources department.

A.9.3 Training Content and Frequency


A.9.3.1 ESD Awareness Training
ESD awareness training shall be given to any person who is to enter and work in the EPA,
before they first enter the EPA. Exceptions to this requirement are the ESD coordinator and
visitors accompanied by trained personnel.
Refresher ESD awareness training shall be given on an annual basis.
The ESD awareness course shall cover the following:
● Awareness of the need for ESD control
● Demonstration of basic static electricity
● Recognition of EPAs and EPA boundaries
● Recognition of EPA compliant and noncompliant materials and equipment
● Identification of ESDS
● Recognition and correct handling of essential insulators such as papers
● What to do, and not to do, in an EPA
● Use and testing of wrist straps and ESD control footwear
● Introduction to the ESD coordinator and their role
Trainees shall be tested, for example, by:
● Multiple choice questionnaire covering key ESD awareness points
● Practical use and testing of wrist straps and ESD control footwear
A pass mark of 80% shall be achieved and correct use and testing of wrist straps and ESD
control footwear demonstrated by the trainee before they are allowed to work in the EPA.

A.9.3.2 ESD Audit and Measurements Training


ESD audit and measurements training shall be given to personnel who are required to check
and test ESD control equipment, before they start these duties. Refresher training may be
required at the discretion of the ESD coordinator.
The ESD audit and measurement training shall cover the following:
● Hands-on experience of compliance verification measurement of EPA equipment
● Basic EPA auditing techniques
Trainees shall be tested by an observed practical measurement and audit practice. Correct
use of the organization’s equipment and procedures by the trainee shall be observed before
they commence audit and measurement work.

A.9.3.3 EPA Cleaning Training


EPA cleaning training shall be given to personnel who are required to clean within an EPA
facility, before they start these duties. Refresher training may be required at the discretion
of the ESD coordinator.
The EPA cleaning training shall cover the following:
● What to do and what not to do as a cleaner working in an EPA
474 Appendix A: An Example Draft ESD Control Program

● Cleaning materials and techniques for the EPA floor and bench mats
Trainees shall be tested by questionnaire and observation of practice.

A.9.3.4 Principles and Practice of ESD Control Training


Training in the principles and practice of ESD control shall be given to the ESD coordinator,
before they commence responsibility for the ESD control program. Continuous professional
development in the subject shall be undertaken.
Training in the principle and practice of ESD control will be conducted by techniques
such as the following:
● External courses according to availability
● Reading of books, articles, and other resources
● Reading of standards and associated user guides and other standardization documents
● Familiarisation with existing ESD Control Program Plans and facilities
Where available, study for relevant ESD control qualifications should be encouraged.

A.10 ESD Control Product Qualification

Unless otherwise specified, all ESD control materials, equipment used in EPAs, and ESD
protective packaging selected shall as a minimum have a data sheet or other document
showing compliance with IEC 61340-5-1 or ANSI/ESD S20.20 and requirements of this ESD
Program Plan.

Table A.4 ESD control product qualification tests and pass criteria.

ESD control item Qualification procedure Pass criterion

ESD control footwear, QP1. Resistance from person to Resistance from the wearer’s body to
while worn ground while standing on ESD ground >100 kΩ and <1 GΩ
floor
QP2. Body voltage walk test of No peaks in body voltage greater
footwear and flooring system than 100 V
Wrist strap TP1 >750 kΩ
<35 MΩ
Floor mat QP3. Resistance to groundable <1 GΩ
point of work surfaces or mats
Chair or seat TP4. Measurement of resistance <1 GΩ
to ground from surface
Work surface on which QP3. Resistance to groundable <1 GΩ
ESDS may be placed point of work surfaces or mats
QP4. Point-to-point resistance of <1 GΩ
work surface
ESD control garment TP5. Point-to-point resistance of < 1 GΩ
ESD garments
ESD Protective TP8 Surface resistance < 100 GΩ
packaging
References 475

The ESD coordinator shall maintain a list of items and equipment approved for use in
EPAs. Listed items shall be identified by manufacturer, type, or other specific information.
Further product qualification requirements and procedures for specific ESD control items
are given in Table A.4.
Proprietary wrist strap and footwear checkers used for daily checking of wrist straps and
footwear shall be verified to have pass criteria as specified in Table A.1.

References

EOS/ESD Association Inc. (2014) ANSI/ESD S20.20-2014. ESD Association Standard for the
Development of an Electrostatic Discharge Control Program for – Protection of Electrical and
Electronic Parts, Assemblies and Equipment (excluding Electrically Initiated Explosive Devices).
Rome, NY, EOS/ESD Association Inc.
International Electrotechnical Commission. (2015) IEC 61340-5-3:2015. Electrostatics.
Protection of electronic devices from electrostatic phenomena. Properties and requirements
classifications for packaging intended for electrostatic discharge sensitive devices. Geneva, IEC.
International Electrotechnical Commission (2016) IEC 61340-5-1: 2016. Electrostatics – Part
5-1: Protection of electronic devices from electrostatic phenomena - General requirements.
Geneva, IEC.
International Electrotechnical Commission. (2019) IEC TR 61340-5-4:2019. Electrostatics - Part
5-4: Protection of electronic devices from electrostatic phenomena – Compliance Verification.
Geneva, IEC.
477

Index

a ESD event detection 408


anodization 133, 140–141, 144–146, ESD protective packaging 134, 137,
148, 253 139, 143, 147–150, 242, 253,
antistatic 1, 11–13, 173, 200, 202, 260–261, 263–265
243–244, 256, 258, 335 ESD risks 133–142, 147–150
additive, see antistat handling very sensitive components
definition 11–13, 173, 243 152
footwear 200, 202 isolated conductors 141–142, 146,
packaging 243–244, 247, 250–251, 149, 134–135, 333–334
256, 258, 266–268, 272 measurements 133–135, 137–138,
antistat (antistatic additive) 105, 143, 140–142, 145, 147–152, 408
184–185, 232–233, 247, 250–251,
modification 134, 136, 138
256, 258, 266–268, 272, 433
process capability 283–286
atoms 2–3, 17–18
qualification 137–138
automated systems 45, 54, 60, 71,
strategy for ESD control 135–137
109, 128, 133–152, 218–219, 242,
trends 449, 457–458, 460–462
253, 260–261, 263–265, 283, 286,
289–290, 297–301, 333–334, 408, use of ionizers 133, 143, 146–147,
449, 457–458, 460–462 151–152, 289, 301
bearings 139–141, 145, 149 vacuum cups, chucks or pickers
compliance verification 138, 140, 143–144, 146–147
140–141, 147
conveyor belts 137, 141, 145, 147 b
critical path of ESDS 136–140, body voltage 21–22, 33, 93, 111–114,
142–143, 145–146, 149, 151–152, 130, 163, 168–169, 171, 180,
283 186–188, 190–195, 197–198,
determination and implementation 200–201, 214, 216, 285–287, 312,
of control measures 138–141, 319, 321, 328–330, 336, 348, 359,
283, 297–299 386, 390, 423, 437, 440–442,
documentation of ESD control 456–457, 474
measures 140 demonstration 440–442
ESD control measures 133, limits 168–169, 171, 190, 197, 200,
135–136, 138–147 285–287, 321, 328–329, 348, 474
The ESD Control Program Handbook, First Edition. Jeremy M Smallwood.
© 2020 John Wiley & Sons Ltd. Published 2020 by John Wiley & Sons Ltd.
478 Index

body voltage (contd.) grounding 182–184, 186–187, 362


measurement 63, 168–169, 180, purpose 210
186, 216, 386 qualification 212–213, 366
walk test 186–188, 200, 328, 330, selection 212–213
346, 348, 359, 386, 440–442, 474 types 210
breakdown 7, 253 use 210–212
of air 15, 17–18, 30–31, 284 charge 2–8, 11–15, 17–49, 97–100,
of anodised layer 253 104–113, 119–123, 141–142,
dielectric 15, 284, 292 146–147, 150–151, 161, 183, 187,
field strength 17–18, 30–31 198, 201, 203, 217, 219, 220–221,
surface charge density for 18 230–231, 233, 345–348, 357–360,
362, 383, 388–390, 399–408, 414,
c 417, 428–429, 433–439, 442–446
cable discharge events, see ESD, from a build-up, see charge, build-up
charged cable decay 11, 22–24, 43–44, 100,
capacitance 1, 14, 19–22, 24, 27, 29, 107–109, 119, 220–221, 230–231,
34–38, 48, 100, 108–109, 119, 130, 233, 245, 255, 265, 268, 300,
135, 150, 286–292, 298, 300, 336–337, 346, 359, 383, 385,
428–429, 458–460 399–404, 428, 444–445
combination of 14 dissipation 3, 13, 18–19, 122, 161,
of CPM plate 400–401, 408 187, 198, 201, 203, 244, 247, 256,
device 60–61, 150, 286, 290–292, 273, 345, 347, 362, 389, 414,
298 428–429
effect on ESD 33–37, 64, 79, 130, generation, see charge generation
135, 150, 286–291 imbalance 2–3, 18–19, 45, 107, 222,
effect on ESD susceptibility 70–71, 359, 359, 383, 385, 444
74 measurement, see measurement,
ESD model 53–54, 56, 58, 61, 64 charge
of everyday objects 19–22 neutralization 3, 18, 27, 32, 42–44,
of Faraday Pail 405, 407 105–107, 122, 143, 146, 217, 219,
to ground 19–21, 29 289, 333, 345, 359, 383, 385,
human body 20–22, 33, 54–55 444–446
isolated conductors 34, 287, 298, charge (voltage) build-up 11, 17–20,
333, 390, 458–460 37, 91–92, 99–100, 104, 107, 109,
of on-chip protection networks 69 112, 141, 147, 213, 241, 244, 299,
of printed circuit board (PCB) 21, 333, 428, 458, 459
36, 78–79, 130, 459 on cables 80–81, 130, 250, 282–283,
of tool 108, 119, 230, 300, 401, 403 296–297, 301, 459
of voltmeter 359 electrical model 19–20, 24, 99–100,
carts, trolleys and mobile equipment 112, 299, 428–429
98, 111, 113, 116–118, 182–184, on ESDS 64, 74–76, 143–144, 282,
186–187, 210–213, 274, 362, 366 286–288, 290–295, 298–299, 458
care and maintenance 211–212 on furniture 80
common problems 213 on garments 216
compliance verification 213, 362 on gloves and finger cots 120
Index 479

on module or assembly 301, 459 surface resistance of packaging


on packaging 241, 266, 273 378, 380–381
on PCB 301, 459 tools 231, 393–394, 403–404
on personnel 80, 440–442 training 125
polarity 3–4, 18–19, 435 volume resistance of packaging
on tools 119 382
charge generation 17–20, 24, 109, walk test of footwear and flooring
112–113, 141–142, 144, 183, 386
187–188, 197, 216, 218, 243–244, work surface 206, 363
247, 254, 256, 273, 287, 289, 333, wrist strap 197, 372, 374, 376
428–429, 442 compliance verification 95, 100,
current 4, 19–20, 24, 100, 113, 429 125–130, 138, 140–141, 147, 163,
leading to body voltage 112–113, 167–168, 179–181, 268, 305,
197, 287, 437, 440- 442 307–314, 318, 327–329, 334–337,
packaging 243–244, 254, 256, 273, 339–341, 346, 348–349, 364, 367,
442–444 383, 455–457, 466, 468, 472–473
triboelectric series 18–20 automated systems 138, 140–141,
charged device model (CDM) 7, 54, 147
60–68, 71–72, 78–79, 105, 109,
carts and mobile equipment 213,
119, 129–130, 133, 135, 151–152,
334–336
165, 167, 226, 282–283, 285–286,
earth bonding point 190, 328, 466,
290–293, 295–296, 298, 321, 325,
468
450–454, 456–458, 460, 466–467
environmental conditions 148,
charged plate monitor (CPM) 43,
149, 349
151–152, 220–221, 230–231, 233,
ESD control earth (ground) 183
336–337, 359–361, 383–386,
399–408, 435, 438, 440–441 floor 188, 328, 335–336, 472
common problems 125 footwear 201–202, 328–329, 469
carts and mobile equipment 213 footwear and flooring 328–329,
electrostatic field and voltage 349, 469
measurements 388, 390 garment 203, 228–229, 328,
ESD earth (ground) 183 334–335, 337, 367–368, 472
floors 188–190 gloves and finger cots 234, 337
footwear 202, 374, 386 how often should items be tested
garments 361, 369 128, 328–329, 335
gloves and finger cots 234, 396, 404 ionizers 221–222, 328, 334, 337,
ionizers 222, 385 383
measurements 361–362, 363, 365, measurements, see measurements
369, 372, 374, 376, 378, 380–382, packaging 472
385–386, 388, 390, 393–396, pass criteria, see pass criteria
403–404, 407–408 plan, see ESD control program,
packaging 252, 378, 380–381 compliance verification plan
seats 215, 365 seats 215–216, 328–329, 336, 364,
soldering irons 395 472
storage racks and shelves 209–210 soldering irons 172, 231
480 Index

compliance verification (contd.) burnout 72, 77


storage racks and shelves 207, 209, in electrostatic field 25–30, 74–75,
328, 334–335, 472 101–102, 104–105, 122, 130,
test methods 127–128, 140–141, 142–143, 291–298, 301, 321,
168–172, 327–329, 334–335, 337, 331–332, 438–440
469, 472 isolated, see isolated conductor
tools 231, 328, 335–337 ungrounded, see isolated conductor
what should be tested 127, 327 conveyor belts 137, 141, 145, 147,
why we need 126 298
wrist strap, see wrist strap, coulomb 2–6
compliance verification Coulomb’s Law 6
work surface, see work surface, current 4, 6–9, 15, 17, 19–27
compliance verification ESD, see ESD, current
conductive 1, 11–14, 24, 93, 97, 107, ion 43
114, 117, 123, 129–130, 134,
139–141, 144, 146, 150, 162, 173, d
242, 244, 247–254, 256, 258–259, decay time 11, 22–24, 44, 100,
261–263, 266–267, 270, 272–276, 106–109, 119, 142, 145, 147, 151,
321–322, 331, 335, 467, 471 168, 220–221, 230–231, 233, 245,
definition 11–13, 107–108, 134, 300, 328, 335–337, 346, 349, 359,
139, 162, 173, 242, 335, 467 383–386, 399–404, 428, 444–446
fibres or threads 203, 224–225, charge 11, 22–24, 44, 100, 106–109,
227–229, 232 119, 142, 145, 147, 151, 233, 245,
footwear 200, 202 300, 336–337, 346, 359, 399–404,
material or object 27–31, 34–35, 428, 444–446
40, 45, 53, 58, 60, 80, 93, 97, ionizer, see ionizer, decay time
107–108, 114, 117 129–130, 136, waveform 40, 43, 56, 63, 106
139–141, 144, 146, 148, 150, 162, dew point 15–16
173, 233, 282–284, 290, 296, 312, dielectric constant, see permittivity
321–322, 331 dissipative 1, 11–13, 24, 107–108,
packaging, see ESD protective 113–114, 117, 120, 130, 134, 162,
packaging, conductive 180, 225, 242, 244, 249–251,
wheels or feet 214 253–254, 256–259, 261–263,
conductivity 8–11, 24, 245, 251–254 265–268, 270, 272–274, 276,
conductor 3–5, 7–8, 11–14, 19, 21, 298–299, 312, 331, 333, 335, 467,
23–30, 32–34, 36–39, 43–47, 74, 470–471
93, 98, 101–102, 104–105, definition 11–13, 107, 134, 139,
107–115, 122, 129–130, 134–137, 162, 242, 335, 467
139–144, 146–151, 172, 181, 217, material or object 137, 141,
219, 230, 283, 285, 287–294, 144–146, 149, 150, 180, 203, 225,
296–301, 313, 321–322, 324–325, 298–299, 331, 333
329–334, 337, 429, 432, 435, packaging, see ESD protective
437–440, 446, 456–460, 462, 467 packaging, static dissipative
Index 481

e shielding, see ESD protective


earth, see also ground packaging, electrostatic field
ESD control, see ESD control earth shielding
protective 168, 171, 181–183, 189, electrostatic field meter 22, 26–29,
310, 312, 328–329, 468 221, 357–359, 387–390, 431–435,
resistance 168, 183, 190 439–440, 442–444
earth bonding point (ground point) energy 3–4, 8, 14–15, 31–32, 38,
112, 115, 139, 188, 189–193, 195, 40–41, 45–47, 56–57, 60, 63–65,
197, 203, 206, 211, 215, 229, 310, 69, 71–73, 75, 77–79, 93, 97–99,
312, 322, 327–328, 330, 334, 108–110, 130, 243, 257, 263, 270,
376–377, 466, 468 274, 285–288, 290, 298
compliance verification 190, 328, equipotential bonding 7–8, 104,
376–377 110–111, 168, 171, 181, 210, 312,
purpose 189 328, 333, 432, 437–438, 456
qualification 190, 328 ESD 1, 7–15, 17, 29–41, 281–288,
selection 190, 334 290–293, 295–299, 301–302
test 190, 377, 468 control measures, see ESD control
use 190 measures
electromagnetic interference (EMI) current 24, 30, 30–41, 46–48, 53,
7, 45, 147, 151, 214, 220–221, 255 55–65, 68–72, 75–79, 81, 93,
electrostatic attraction (ESA) 6, 97–100, 109–110, 115, 119,
41–42, 94, 217, 226–227, 248 286–287, 290–291, 293, 295,
electrostatic discharge 1, 7, 15, 17, 298–301
27, 30–42, see also ESD damage to components, see ESD
brush 7, 31 damage
common electrostatic discharge electronic models of ESD 7, 37–41,
sources 7, 32–37, 320–321 53–58, 60–64, 66, 105
control, see ESD control from charged cables 37, 81, 130,
corona 7, 30, 32, 42–43 250, 282–283, 296–297, 301, 459
propagating brush 32 from charged conductive objects
spark 7, 30–31, 39 58, 60, 101, 104–105, 108, 110,
electrostatic field 2–7, 15, 18, 22, 114, 133, 135, 282–284, see also
25–30, 51, 60–62, 67, 73–76, 79, machine model
94, 97, 101–106, 113, 118–123, from charged devices 35, 135, 139,
129–130, 135–139, 142–143, 142–144, 282–284, 286–288,
146–147, 149–150, 172, 203, 208, 290–301, see also charged device
213–214, 217–218, 220–223, model
225–227, 230, 239, 241–243, 281, from a charged module 36–37, 135,
284, 291–298, 301–302, 313, 319, 249–250, 270–271, 282, 297,
321–322, 331–333, 346, 357–358, 301–302, 459
387–388, 406, 431–435, 438–440, from a charged PCB 36, 78–79,
442–444, 446, 467, 470–471 134–135, 139, 282, 292–293, 301,
limits 172, 294, 313 459
482 Index

ESD (contd.) handling very sensitive components


from charged tools 136, 143, 129–130, 152, 286, 303
300–301 non-standard ESD sources 130
from personnel 33, 40, 45, 80, 97, principles 17, 46–47, 321–322
122, 135–137, 282–283, 285, 287, product qualification, see ESD
see also human body model control product qualification
models, see ESD models program plan, see ESD control
packaging, see ESD protective program, ESD Control Program
packaging Plan
peak current 30–40, 45, 53, 55–65, seating (chair), see seat
79, 81, 97–98, 109–110, 115, standards, see standards
285–286, 290, 293, 300–301 strategy 46–47, 123, 317, 332–334
risk of damage, see ESD damage, risk storage rack, see storage racks and
of damage shelves
sensitivity, see ESD susceptibility tools, see tools
sources 7, 17, 30–41, 53–66, 72, 74, trends 450–459
78, 80, 97, 108, 134–136, 142, 150, workstation 94–95, 104–105, 117,
281–283, 285–288, 290, 296, 305, 122, 128, 172, 182, 184–185, 189,
320–321 203–205, 209, 214, 218–219, 293,
system level, see system level ESD 322–323, 327–328, 331, 336, 342
waveform 7, 30–41, 53, 55–66, 81, work surface, see work surface
286, 295 wrist strap, see wrist strap
withstand voltage 54–55, 58, 60, ESD control earth 181–183, 310, 312,
62, 64–66, 68–71, 74, 77–78, 328–329, 468
80–82, 98, 105, 112–113, 119, bonding point, see earth bonding
129–130, 133, 135, 140, 151, 167, point
282–283, 285–288, 290, 292–293, common problems 183
295–296, 298, 303–304, 308, 321, compliance verification 183
325, 449–460 qualification 183
ESD control 1, 8, 10–13, 15, 17, role and function 181
23–24, 46, 91–130, 179–234, selection 181–182
281–314, 317–343, 449–462 ESD control measures 95, 98–130,
earth (ground), see ESD control earth 164–174, 179–234, 283–302,
(ground) 317–318, 320–343, 461, 468–472
earth bonding point, see earth in automated systems, see automated
bonding point systems, ESD control measures
equipment and materials 179–234, cart (trolley), see carts, trolleys and
297, 303, 307, 318, 320–324, mobile equipment
326–338, 340–342 conductors 172, 181, 321–322,
floor, see floor 324–325, 329, 331–334, 337
footwear, see footwear documentation 165–167, 173–174,
garment, see garment 202, 207, 216, 323–338
gloves and finger cots, see gloves and floor, see ESD control floor
finger cots garments, see ESD control garments
Index 483

gloves and finger cots, see gloves and garment 170, 203, 227–229, 329,
finger cots 366–367, 474
how to determine 321–323 gloves and finger cots 233–234
how to specify 334–338 ionizer 171, 220–221, 383
insulators, see insulators measurements, see measurements
ionizer, see ionizer packaging 262
personal grounding, see personal pass criteria, see pass criteria
grounding seat 215–216, 364–365, 474
rack, see ESD control, rack soldering irons 172, 231
return on investment 299, 305–306 storage racks and shelves 170, 209,
seating, see ESD control seating 366
tools, see ESD control tools test methods 127–128, 167–172
trends 461 tools 230, 401
ways of working 322–323, 341–343 wrist strap 169, 195, 327, 474–475
work surface, see work surface work surface 170, 194, 205–206,
ESD control product care and 327, 348, 474
maintenance 179–234 ESD control product role and purpose
carts and mobile equipment carts and mobile equipment 210
211–212 earth bonding point 189
floors 187–188 floor 183–184, 187
gloves and finger cots 234 gloves and finger cots 232
ionizer 221 ionizer 217
storage racks and shelves 208 packaging 239–241
work surface see work surface personal grounding 190
wrist strap 195 storage racks and shelves 207–208
ESD control product qualification tools 229
127, 130, 137–138, 165–172, work surface 203–204
179–180, 185–187, 194–195, wrist strap 192–193
200–201, 203–206, 209, 211–212, ESD control product selection
215–216, 220–221, 227–231, 233, 179–180, 184–185, 190, 194–195,
307, 309–311, 232–324, 326–330, 200, 204–205, 208, 214–215,
347–349, 364–367, 383, 386, 401, 219–220, 226–227, 232–233, 307
466, 468, 474–475 carts and mobile equipment
automated equipment 137–138 212–213
carts and mobile equipment 170, earth (ground) 181–182, 328–329
186–187, 211–213, 366 earth bonding points 190, 334
earth (ground) 183, 328–329 floors 184–185
earth bonding points 183, 190, 328 footwear 200
environmental conditions 180, garment 226–227
185, 349 gloves and finger cots 232–233
floors 169–170, 185–187, 474 ionizer 219–220
footwear 169, 171, 200–201, 466, packaging 269–275
468, 474–475 seat 214–215
footwear and flooring 169, 171, storage racks and shelves 208
201, 330, 348, 386, 474 tool 230
484 Index

ESD control product selection (contd.) 307–312, 314, 317–318, 323–341,


work surface 204–205 455, 466–475
wrist strap 194–195 equipment used in the EPA 172,
ESD control product use 179–180, 179–234, 297, 303, 305–308, 312,
184–187, 189–190, 193–195, 198, 320–337
200–203, 204–206, 207–208, ESD Control Product Qualification
210–212, 214, 217, 220–221, Plan 165, 167, 309–311,
223–225, 228–233, 317–320, 323–324, 327, 455–456
321–323, 333–334 ESD Control Program Plan
carts, trolleys and mobile equipment 165–168, 294, 307–312, 314,
210–212 323–324, 326, 330–334, 336–337,
earth bonding points 189–190 455–456
floor 184–187, 336 ESD Coordinator, see ESD
footwear 198, 200–202, 329 Coordinator
garment 203, 223–225, 228, 322, ESD Protected Areas (EPA), see ESD
337 Protected Area (EPA)
gloves and finger cots 232–233, ESD protective packaging, see ESD
319, 337 protective packaging
ionizer 146, 217, 220–221, 228, ESD Training Plan 165, 167,
333–334, 337 309–310, 323–324, 326, 455–456
packaging 247–250, 322, 332, 335, evaluation of compliance with
337–338, 342 standard 308–314
seat 214 checklists 308–314
storage racks and shelves 207–208
common problems 314
tools 229–231
example 465–475
work surface 204–206, 335
grounding and bonding systems
wrist strap 193–195, 330
167–171, 181–183, 309–310, 312,
ESD control program 1, 91–130, 140,
319, 321–324, 326, 328–330, 336,
157, 163, 165–174, 179, 281–314,
342, 468
317–343, 449, 453–457, 459–460,
465–475 how to develop 307–308, 317–343
basis of ESD protection 92, 129, insulators 134, 138–139, 142–143,
138–139, 317 146–147, 172–173, 284, 292–295,
benefit of 304–305, 338–339 297–299, 313, 319, 321–324,
Compliance Verification Plan 331–333, 337, 342
140–141, 147, 165, 167–168, isolated conductors 172, 287–291,
309–312, 314, 323–324, 327–328, 297–298, 300, 321–322, 324–325,
455–456, 466, 468, 472–473 329, 333–334, 337
cost effectiveness 281, 302–308, marking 167, 173- 174, 215, 231,
318, 338–340 234, 309, 312, 314, 324, 326,
cost evaluation 305–306, 340 337–338
cost of inadequate program optimizing 307–308, 338–341, 455,
302–304 459
documentation and planning 140, personnel grounding, see grounding
165–167, 173–174, 216, 294, personnel
Index 485

plans 165–168, 174, 294–295, 248, 282–284, 286, 290–296,


307–314, 317–318, 323–338 300–301, 321, 331, 348, 407–408
scope 324–325 conditions leading to 136
terms and definitions 325 by electrostatic fields, see ESD
process capability, see process susceptibility, electrostatic fields
capability failure criteria 56, 60, 62–64
return on investment 299, 302, failure mechanisms 72–73
305–307 field effect structures 67, 74, 76
safety, see safety light emitting (LED) and laser diodes
soldering and desoldering tools 67–68, 70, 76–77
172, 231, 335 magnetoresistive heads 75, 77
tailoring 166, 309–310, 314, 324, microelectromechanical devices
326, 329–330, 332, 334–335, 337, (MEMS) 70, 73, 75, 77
342 modules and system components
technical basis 165–166, 216, 317 52, 58, 78–80
technical requirements 161–162, MOSFETs 66–67, 70–74, 284,
164–166, 168–172, 282, 288, 294, 292–293, 295, 297
299, 311–313, 326–330, 468–472 passive components 71, 78
piezoelectric crystals 76
test methods 158–159, 164,
prevention 109, 116, 318
167–173, 311, 327–329, 334–337,
printed circuit boards (PCBs) and
345–409
assemblies 52, 78–79, 92–93,
trends 449, 453–459
130, 282, 286, 292–293, 301–302
update 326, 343
risk of damage 52, 60, 65, 67–68,
who is responsible 165–166, 309,
78–79, 81, 92–94, 97, 101,
318, 324, 326
104–105, 109, 115, 124–125,
ESD control standards, see standards 134–136, 139, 152, 281–286,
ESD Coordinator 99, 125, 165–166, 290–298, 300, 302, 304, 336,
308–309, 318, 324, 326, 337 413–414, 440
ESD damage 7, 33, 41, 45–46, 51–53, semiconductor junctions 68, 72–77
55, 60, 64–68, 72–79, 81, 91–94, ESD Protected Area (EPA) 91–122,
97, 101, 104–105, 109, 115–116, 124–126, 128–129, 168–174, 239,
124–125, 130, 134–136, 143–146, 241–242, 247–249, 256, 262–263,
152, 239, 241, 248, 250, 270–271, 270–276, 281, 297, 305, 307–309,
281–286, 290, 292–298, 300–304, 312–313, 317–318, 320–324,
306, 318, 331, 336–337, 339, 341, 328–332, 334, 336–342, 348–349,
343, 414, 417–418, 423, 429–430, 364–365, 451, 455–456, 467–472
440, 454–455, 458–460 boundaries 94–96, 171, 320–321,
breakdown of thin dielectric 331, 339–340, 469
72–73, 75, 77 requirements 168–174, 312–313,
burnout of device conductors or 318, 322–324, 326, 328–334, 336,
resistors 73–74, 77 340–341, 469–472
charged device 64–65, 67, 72, 79, signage 171, 320, 331, 469
97, 105, 109, 115–116, 130, what is an EPA? 93–95 171–172,
134–135, 143–145, 152, 205, 241, 317–318, 320
486 Index

ESD protection 317–321, 338–339, conductive 134, 139, 173, 242, 244,
341 247–254, 256, 258–259, 261–263,
basis of ESD protection, see ESD 266–267, 270, 272–276, 335
control program, basis of ESD customer requirements 269–270,
protection 272, 275, 313–314, 337–338
ESD protection measures, see ESD electrostatic field shielding 105,
control measures 123, 172, 245–248, 253, 255–257,
evaluation of needs 283–289, 262–263, 265–267, 270, 274–276,
293–294, 296–302, 319, 338 442–443
ESD protective packaging ESD shielding 105, 123, 172,
269–275, 317–318, 338–339, 341 242–243, 246, 248–249, 253–259,
limits of standard precautions 262–263, 265–268, 271, 274–276,
296–297 338, 346, 382–384
evaluation of risks 281–289, foam 242, 251, 258–259, 274–275
294–297, 319, 332–334 functions 123, 241–242, 273–275,
on chip protection 64–65, 68–71, 276
75, 79 humidity (moisture), effect of
on chip protection targets 68–70 239–240, 243, 247, 250–251, 256,
ESD protective packaging 12, 91–95, 433, 443
intimate 137, 242, 248–249, 256,
98, 100–101, 108–109, 122–124,
259, 261, 263, 273–275, 290
126, 130, 134, 137, 143, 147–149,
labels 262, 264, 276
173–174, 213, 239–276, 290,
low charging 173, 243–244,
297–298, 301–302, 306–307, 309,
249–251, 254–259, 261–262,
312- 313, 317–318, 322, 324, 326,
266–268, 272, 332, 433, 443
330, 332, 335- 339, 341–342,
marking 242, 262, 264, 272,
346–347, 350, 352–354, 356–357,
275–276, 314, 337–338
377–384, 413, 417–418, 425, 429, materials 250–254, 332, 342
433, 442–443, 455, 459, 461, 467, measurements 134, 149, 168, 244,
468–471 268–269, 346–347, 350, 352–354,
anodized aluminium 133, 140–141, 356–357, 377–383
144–146, 148, 253 moisture barrier 241, 253–255,
antistatic 173, 247, 250–251, 256, 257–258, 264–268
258, 266–268, 272, 335 opening 123–124, 256, 341–342
in automated systems 134, 137, operational environment 269–273
143, 147–149 pink polythene 245–247, 250–251,
bags 243, 245–247, 252–258, 261, 255–259, 272, 332, 433, 442–443
263–269, 271–272, 274–275, 297, properties 173, 239, 241–250, 332,
338, 341, 442–443 442–443
boxes and trays 251–252, 258–259, proximity 242, 249, 274–275
271, 342 resistance and resistivity 47, 100,
bubble wrap 254, 258 108–109, 123, 130, 134, 149, 168,
charge dissipation 122, 244, 247, 173, 239–249, 251–253, 255–256,
256, 273 258–259, 261–263, 265–271,
charging 123, 241, 243–244, 249, 273–275, 265, 268, 354–355, 338,
273–275, 433, 442 377–382
Index 487

role and purpose 239–241 ESD training 91, 95, 99–100,


selection 269–275 124–126, 138, 165–167, 171,
standards 240–244, 247, 254–259, 305–310, 318, 323–324, 326, 331,
262–269, 272, 275–276, 338 339–341, 413–447, 449, 455–456,
static dissipative 173, 213, 242, 462, 468–469, 472–474
244, 249–254, 256–259, 261–263, assessment, see ESD training
265–268, 270, 272–274, 276 evaluation
stick magazines (tubes) 252–253, automated equipment 138
261–262 books and articles 420, 426–427
systems 254, 258–259, 263, certification 424–425, 446
269–276, 326, 338 conferences 415, 420, 426–427
tape and reel 260–261 considerations 138, 414, 419–421
terminology 242–243 content 126, 414, 416–427
test 243–244, 257, 261–263, demonstrations 416–417, 420–425,
265, 267–268, 270–271, 275, 429–446
346–347, 350, 352–354, 356–357, body voltage 437, 440–442
377–383 charge and tribocharging
triboelectric charging 241, 433–436, 442–443
243–244, 247–251, 254–259, charge neutralisation 444–446
261–262, 266–268, 272–275 cost of ESD damage 430–431
types 239–241, 254–262, 271 electrostatic fields 431, 433–434,
use 239, 241–244, 247–250, 439, 442–444
318, 330–332, 335, 337–339, electrostatic field shielding
341–342 442–443
for non-ESDS items 249 equipment 432–433
outside the EPA 248–250, 330, ESD 429–430, 436–437, 439–440
339, 341–342 grounding and equipotential
within the EPA 248–249, 322, bonding 436–444
335, 338, 341–342 induction 438–439
ESD sensitive devices (ESDS) 51–82, pros and cons 431–432
92–95, 97–106, 109, 111–125, real ESD damage 429–430
128–130, 133–152, 165, 239–243, role and value of 429, 431
247–250, 257, 261–262, 264, voltage 434–437
266–276, 317–325, 329–343, 451, electrostatics and ESD theory
454–462 428–429
ESD susceptibility 7, 66–71, 135, 138, evaluation 421, 423–424, 446–447
142, 248, 269–270, 282–283, goals 417–418
285–286, 288, 294–295, 298, hands–on learning 419–420, 423
320–321, 325 initial training 167, 331, 418
discrete devices 71, 73 interest groups and organisations
effect of scaling 71 425–426
electrostatic fields 51, 74–75, methods 419–420, 431
293–296 instructor led 419–420
latent failures, see latent failures video or internet based 419
package effects 60, 67, 71, 76 online resources 419, 425–427
susceptibility test 53–66 pass criteria 447
488 Index

ESD training (contd.) compliance verification 201–202,


planning 165–167, 307–310, 318, 328–329, 349, 469
323–324, 326, 414–415, 455–456 conductive 200
presentation 420–423 qualification 200–201, 328–330,
public courses 424 348, 474
qualifications 424–425 safety 180, 186, 191–192, 198–201
refresher 126, 167, 415, 419 selection 200
supporting information 420 standard requirements 169, 171,
trends 415, 420, 449, 455–456, 462 328–330
who needs ESD training 95, 99, test 163–164, 329, 346, 348–349,
125–126, 167, 326, 331, 340, 374, 386, 418, 423, 446
415–416 types 198–199
why train? 91, 124–125, 326, 340, use 198, 200–202, 446
413–414
g
f Gauss’s Law 5
Faraday cage 27–28, 245, 275 garment 98, 121–122, 125–126, 164,
Faraday pail 30 170, 191, 203, 222–229, 322,
floor 12, 98, 111–114, 116–118, 328–329, 334–335, 337, 346, 350,
128–130, 163, 169–171, 183–189, 361, 366–371, 374–375, 416, 466,
322–323, 328–330, 334–336, 346, 472, 474
348–349, 350, 355, 362, 364–365, care and maintenance 228
374, 376–377, 386, 415, 440–442, compliance verification 203,
456–457, 466, 469, 472, 474 228–229, 328–329, 334–335, 337,
acceptance 187 367–368, 472
care and maintenance 187–188 grounding 222–229, 322, 329, 337,
common problems 188 368, 370–371, 374–375
compliance verification 188, personal grounding via, see
328–329, 334–336, 349, 386 grounding personnel, via
qualification 185, 328–330, garment
348–349 qualification 203, 227–228,
role and function 183–184, 187 328–329, 367, 474
selection 184–185, 322–323, role and purpose 222
335 selection 226–227, 337
standard requirements 169–171 standard requirements 170,
test 163, 328–329, 362–363 328–329, 335, 337
types 184–185 test 164, 328–329, 335, 337, 346,
use 184–187, 415 350, 366–371
footwear 12, 162, 180, 184, 186–187, types 223–227, 337
195, 197–202, 322–323, 328–330, use 203, 228, 329, 334, 337, 416
334, 336, 346, 348–350, 374, 386, gloves and finger cots 98, 111,
413, 417–419, 423, 440–442, 446, 120–121, 162, 192, 230, 232–234,
456–457, 466, 468–469, 473–475 297–298, 300–301, 319, 335, 337,
antistatic 200 364, 355, 396, 399, 403, 405–406,
common problems 202 408, 418
Index 489

care and maintenance 234 via floors 116–117, 181, 186, 191,
common problems 234 195, 198, 211, 214–215, 322–323,
compliance verification 234, 335, 329–330, 336, 348–349, 364,
337 376–377
grounding 231–234, 337, 319, functional earth 8, 110–111, 168,
399 171, 181, 183, 312, 328
insulating 231 gloves and finger cots, see ESD
qualification 233 control, gloves and finger cots
role and purpose 230, 232–233 hard 152
selection 232–233, 337 insulators 104, 443–444
standard requirements 335, 337 personnel, see grounding personnel
standard test methods 335, 337 qualification 181, 183, 190,
types 232, 319, 337 211–212, 215, 229, 233, 329–330
use 230–233, 297–298, 300–301, resistance to ground 13, 140–142,
319, 418 145, 147–149, 164, 166, 168, 174,
ground (grounding) 3–4, 7–8, 11–13, 181, 186–191, 193, 196–198,
98, 101, 104–105, 107–122, 200–201, 203–206, 209, 212–213,
127–130, 139–143, 145, 147, 215–216, 223, 227, 229, 231, 234,
167–171, 181–217, 223–234, 253, 287–288, 290–291, 299–300,
266, 268, 284, 288–292, 295–296, 327–330, 333, 335–337, 346,
310, 312, 322–324, 326–337, 342, 348–349, 351, 356–357, 363–364,
345–346, 348–349, 355–357, 374–377, 393–397, 399–400
362–366, 368, 370–371, 374–377, resistance to groundable point 180,
393–399, 404, 437–440, 443–444, 186–187, 194, 196, 203–204, 209,
446 212–213, 223, 227, 229, 327–328,
carts (trolleys), see ESD control, cart 335, 337, 346, 348, 365–366, 368,
(trolley) 370, 394–395
common problems 183, 363 role and function 110, 189
compliance verification 183, 190, safety, see safety, ground
195–197, 201, 203, 206, 209, 213, seats, see ESD control, seating
215–216, 228–229, 231, 310, selection 181–182, 190
328–329, 335–337, 348–349, 364, soft 152
415, 446 system 110–113, 115–116, 127, 167,
conductors 11, 139, 141–143, 147, 171, 326–327, 349, 374, 376, 393,
181, 217, 289–290, 292–293, 396, 399, 404
331–334, 436–440 tools, see ESD control, tools
elimination of ESD 110, 181, 217 types of ground 110–111, 171, 181,
equipotential bonding, see 328–329, 437–438
equipotential bonding work surfaces 113–115, 170, 180,
ESD control equipment 110–130, 182, 184, 204, 206–207, 209,
180–216, 230–234, 323–324, 327, 327–328, 335–336, 348, 363
336–337 wrist strap 139–140, 169, 191–197,
ESD control garments, see ESD 322–323, 328–330, 334, 355,
control, garment 376–377
490 Index

grounding personnel 111–113, 139, definition 11–13, 104, 107–108,


169, 189–203, 284, 287, 300, 309, 134, 162, 173
312, 319, 321–323, 326, 329–330, essential (process required)
346, 349, 371–377, 456–457 100–106, 129, 142, 146–147, 217,
electrical safety 168, 191, 193, 199, 219, 284, 293–294, 322, 331–333,
201–202, 300, 319, 322, 342 473
footwear and flooring 112–113, evaluation of 172, 332–333
164, 191, 195, 197–202, 322–323, non-essential 100–104, 129, 142,
329–330, 334, 336, 346, 348–349, 322, 293–294, 331–333
355, 374–377, 386, 456–457 reducing ESD risk from 104–107,
via garment 203, 229, 322, 329, 136, 138, 142–143, 146–147,
337, 370–371 331–333
purpose of 111–112, 189–191, 336 ionizer 27, 32, 43–45, 105–107, 122,
requirement for 111 133, 143–144, 146–147, 151–152,
seats 113, 202, 215–216, 322–323, 164, 181, 217–222, 289, 292, 298,
329, 336 301, 306, 322, 328, 333–335, 337,
wrist strap, see wrist strap 345–346, 359–360, 383–385,
444–446
h balance 42–45, 107, 217, 220–221,
Human Body Model (HBM) 7, 39, 292, 298, 328, 334–335, 337, 359,
54–58, 62–64, 112–113, 129–130, 383–385, 444–446
133, 135, 151–152, 165, 167, 190, care and maintenance 45, 221
194, 200, 226, 282–283, 285–287, common problems 222, 385
292, 295, 298, 321, 325, 329, compliance verification 221–222,
450–454, 456–460, 466 337, 383–385
corona discharge 27, 42–43,
i 217–218
induction 27–30, 93, 101, 105, 111, decay time 11, 43–44, 142, 145,
119, 122, 130, 136, 228, 247–248, 151–152, 168, 171, 220–221, 328,
270–271, 283, 291, 299, 302, 335, 337, 383–385, 444–446
438–439, 444–445 electrical 43
insulative 1, 7, 11–13, 15, 107–108, offset voltage 44, 146–147, 151,
124, 134, 139, 144, 151, 162, 206, 217–218, 220–222, 289, 328,
208–209, 213, 227, 230, 232, 239, 334–335, 337, 383–385, 444–446
242, 244, 256, 258, 261, 263, 270, passive 43
274, 284, 298, see also insulator qualification 220–221, 383
insulator 11–13, 19, 23–24, 30, 36, 47, radioactive 43
99–108, 129–130, 134, 136, rate of charge neutralization, see
138–139, 142–143, 146–147, 151, charge, neutralization
172–173, 206, 217, 219, 244–245, role and purpose 217, 322
321–322, 324, 331–333, 418, region of effective charge
428–429, 433–436, 438–440, neutralisation 44
443–446, 456, 459–460, 467, selection 219–220
470–471, 473, see also insulative standard requirements 171
air 30, 42 standards 164, 171
Index 491

test 164, 168, 171, 181, 220–222, 376–378, 380–382, 385–386, 388,
328, 335, 337, 346, 359–360, 390, 393–396, 403–404, 407–408
383–385 compliance verification 167–172,
types 218 181, 188, 190, 195–196, 201, 203,
use 146, 172, 217, 220–221, 228, 206, 209, 213, 215, 221–222,
322, 333–334 228–229, 231, 234, 268, 311, 327,
X ray 42 336, 345–349, 364, 367, 383
ions 3, 32, 42–45, 106–107, 217–219 decay time 142, 145, 151, 220–221,
ion sources 217–219 230, 233, 268, 335–336, 346, 359,
isolated conductor 27–29, 44, 47, 75, 383–385, 399–404
101, 114, 119, 122, 141–142, 146, of gloves and finger cots 337,
149, 172, 217, 219, 221–222, 231, 403–404
287–291, 297–298, 300, 321–322, of packaging 268
324–325, 329, 333–334, 337, 393, of tools 230–233, 336, 346, 359,
435, 438–440, 446, 456, 466, 470 399–403
resistance to ground 149, 288, 300, earth 168
333–334 earth bonding point 190, 327, 377
standard requirements 172, 325, electrostatic field 168, 172, 346,
333 387–388
latent failures 52, 64, 66–68, 76–77 end to end resistance of ground cord
195, 327, 346, 371–372
m EMI detection, see ESD event
Machine Model (MM) 7, 54–55, ESD current 151
58–60, 62–64, 133, 135, 283–288 ESD event 408–409
marking 95–97, 167, 173–174, 215, ESD protective packaging 163,
231, 234, 242, 245, 262, 264, 272, 167–168, 244, 268–269, 346–347,
275–276, 309, 312, 314, 320–321, 352–354, 356–357, 377–384
326, 337–338, 451, 453–454, 458, ESD shielding 268, 346, 382–384
465, 468–469, 471 Faraday pail 404–407
EPA boundary 95–97, 320, 469 floor 158, 163, 170, 181, 186, 188,
measurements 163, 180, 181, 186, 336, 346, 348–350, 362–364
188, 190, 194–196, 200–201, footwear 163, 186, 169, 201–202,
203–206, 209, 212–213, 215–216, 346, 374
220–221, 225, 227–233, 244, footwear and flooring in combination
268–269, 284–286, 289, 291–293, 164, 169, 180, 186, 201, 200, 346,
299, 311, 327, 334–336, 345–409, 348–350, 374–377, 386
458, 461–462 garment 164, 170, 203, 225,
in automated equipment 147–151 227–229, 346, 350, 366–371
body voltage 169, 180, 186, 188, gloves and finger cots 231,
190, 194–195, 201, 203, 216, 348, 233–234, 346, 355, 359, 396–399,
386 403–404
carts (trolleys) 213 humidity 347–349, 361
charge 147, 151, 404–408 ionizer 147, 151, 164, 168, 171,
common problems 188, 205–206, 221–222, 334–335, 346, 359–360,
361–363, 365, 369, 372, 374, 383–385
492 Index

measurements (contd.) surface resistance 168, 180, 188,


decay time 147, 151, 168, 171, 205–206, 209, 212–213, 225,
221, 335, 383–385 227–229, 244, 268–269, 335, 346,
offset voltage 147, 151, 168, 171, 352–354, 361, 378–381
221, 335, 383–385 surface resistivity 268–269, 352,
personal grounding tests 169, 171, 354
186, 194–196, 201–203, 216, 229, trends 461–462
346, 349, 371–377, 386 tool 230–231, 233, 346, 356, 359,
point to point resistance 145, 392–394, 400–403
148–149, 168, 181, 186, 327, 335, voltage 149–151, 289, 291–292,
346, 348, 363, 366–369, 378–380 334, 346, 357–359, 386–392
qualification 167–172, 180, 186, voltages on ESDS 136–137,
190, 194, 201, 203, 205–206, 209, 142–143, 149–151, 289, 291, 334,
212, 215–216, 220–221, 227–231, 390–393, 462
voltages on small conductors
233, 327, 347–349, 364–367, 383,
149–151, 172, 289, 291–292, 334,
386, 401
390–393, 462
resistance 144, 147–149, 152,
volume resistance 163, 168, 244,
163–164, 168, 186, 188, 194–196,
268, 335, 346, 352–353, 356–357,
201–206, 209, 212–213, 215–216,
381–382
225, 227–231, 233, 244, 268–269,
volume resistivity 168, 354
327, 335–336, 346, 348–357,
walk test for body voltage 186, 346,
361–383, 392–403
348, 386
of gloves and finger cots 396–399 work surfaces 164, 170, 174, 181,
of soldering irons 172, 394–395 204–206, 209, 335, 346, 348,
of tools 336, 392–394 362–364, 366
resistance to ground 8, 13, wrist strap 164, 169, 194–196, 462,
112–119, 140, 141–142, 147–149, 350, 355, 371–374
152, 168, 174, 186, 188, 190, 195, measurement equipment 311,
200–201, 335, 346, 348–349, 357, 350–361
362–366, 370, 375–377, 393–394, charged plate monitor (CPM)
400 151–152, 220–221, 230–231, 233,
earth bonding point 377 336–337, 359–361
resistance from person to ground concentric ring electrode 9–10,
169, 171, 348, 374–375 149, 150, 244, 346, 352–354,
seat 164, 170, 180–181, 186, 377–378, 380–382
215–216, 336, 346, 364–365 electrostatic field meters 149–150,
shelves and racks 209, 213, 346, 221, 357–358, 387–389
362, 366 electrostatic voltmeter 149–150,
soldering tools 168, 172, 231, 346, 186, 221, 334, 357–361, 386–392,
351, 356, 394–395 404–406
standard 163–164, 168–172, 186, ESD ground connectors 183,
209, 213, 215, 220–221, 228, 231, 189–190, 193, 195, 206, 357
233, 327, 345–353, 359, 361–387, footwear test electrode 354–355,
399 377
Index 493

handheld electrode 355–356, 361, equipotential bonding, see


372–374, 376, 386, 396, 398–399 equipotential bonding
high resistance meter 148, power 8
350–351, 362–365, 367–370, 374, process capability 283–297
377–378, 382, 393 charged device ESD risks 283, 287,
insulating supports 356–357, 365, 290–292
368–370, 374, 378–381, 393 critical ESDS path 283–285, 296
low resistance meter 148, 351, damage to voltage sensitive
394–395 structures 292–293
metal plate for volume resistance electrostatic fields 281, 284,
356–357, 381–382 291–298, 301–302, 313
resistance measurement electrodes essential insulators 284, 293–294,
351–352, 362–370, 377, 380 296–297
tool test electrode 356, 394–395 isolated conductors 287–291,
two point probe 354–355, 379–380 297–298, 300
microelectromechanical devices manual handling 283, 287, 290,
(MEMS) 70, 73, 75, 77 296, 300–301
structured approach 283
o
troubleshooting 296
Ohm’s law 19, 24, 40
use of ESD withstand voltage data
285–288, 290, 292–293, 296, 298,
p
303
packaging
ESD, see ESD protective packaging
r
secondary (ordinary, non-ESD
relative humidity 15–16, 18, 20,
protective) 122–123, 207–208,
24–25, 105, 113, 122–124,
234, 240, 242, 249, 251, 256, 272,
275, 341–342, 459, 470 185–189, 201–206, 209, 215–216,
papers 124, 239–240, 293, 323, 222, 224–225, 228, 287, 291–293,
331–332, 342 296, 298–299, 330, 332, 347–349,
Paschen’s Law 31 361–362, 369, 430–431, 433, 440,
pass criteria 127–128, 137, 138, 141, 443, 470–472
167, 179–180, 187, 195–197, resistance 1, 8–16, 99–100, 104,
201–203, 206, 209, 213, 215–216, 134–135, 139–142, 144–145,
220, 227- 231, 233–234, 244, 275, 147–149, 152, 180–181, 183–184,
311–312, 327–329, 334–335, 337, 186–206, 209, 212–216, 223,
340, 447 227–231, 233–236, 239–249,
permittivity 6, 20–22, 24, 100, 108 251–253, 255–256, 258–259,
of free space 6 262–263, 266–274, 284–288,
relative 6, 100 290–291, 297–301, 312, 321–322,
personal grounding, see grounding 327–330, 333, 335–338, 346,
personnel 348–357, 361–384, 392–403
potential 3–5, 7–8, 333 combination of resistances 13–14
difference 2, 245, 333 of conductor 107–108, 283, 285,
equipotential 4–5, 7–8 290, 297, 333
494 Index

resistance (contd.) use in protection against charged


contact 109, 181, 188, 198, 201, device ESD 144, 284–285,
216, 288, 290–292, 297–298 290–292, 297, 321–322, 327
earth 8, 183, 190, 192–193 volume 10, 99–100, 108, 134,
end to end 195, 327, 346, 371–372 163–164, 173, 243–244, 247, 252,
from body to ground 111–114, 163, 256, 263, 266, 268, 271, 335, 338,
171, 186–187, 191, 194, 196–198, 346, 352–354, 356–357, 381–383
200–201, 216, 287, 301, 312, resistivity 8–13, 15, 22–24, 99–100,
328–330, 374–376 108, 115, 119, 252, 255, 261,
to ground 13, 24, 108–110, 112, 265–266, 268–269, 275, 322,
114–119, 127, 140–142, 145, 352–354, 377
147–149, 164, 166, 168, 171–172, surface 9–10, 100, 108, 255, 261,
174, 181, 186–191, 193–194, 265–266, 268–269, 352–354, 377
197–198, 200–201, 204–206, volume 10–11, 13, 100, 108, 252,
212–213, 215–216, 223, 225, 231, 352–354
234, 216, 223, 299–300, 312,
327–329, 333, 335–337, 346,
s
safety 8, 110, 166, 168, 179–182,
348–349, 356–357, 362–366,
185–186, 191–193, 196, 198–202,
374–381
204–205, 208, 212, 214, 218,
of insulator 104, 107–108, 173
220–221, 227, 232–233, 297, 305,
leakage 13, 24
318–319, 322, 325, 342, 351, 357,
maximum 112, 142, 287, 299–300
374
measurement, see measurements,
in automated equipment 134–135,
resistance
142, 148
minimum material resistance 109, footwear 180, 198–202
142, 290–292, 300, 366 ground (earth) 8, 110, 111, 140,
packaging, see ESD protective 149, 168, 181–182, 357, 363
packaging, resistance grounding personnel 186, 191–192,
in parallel 14–15 202, 318–319, 374
personal grounding 186–187, scientific notation 1
191–198, 200–203, 229, 234, 287, seat 98, 113, 116, 118, 164, 170–171,
312, 328–330, 346, 371–377 180–181, 184, 186–187, 191, 202,
point to point 145, 148–149, 180, 214–216, 322–323, 328–329,
186, 203–206, 209, 212, 223, 227, 334–336, 346, 364–365, 466, 469,
229, 290, 346, 348, 366–370, 472, 474
378–380 care and maintenance 215
in series 14–15 common problems 215
surface 9, 100, 115, 130, 163–164, compliance verification 170,
173, 180, 186–187, 192, 206, 209, 215–216, 328, 365, 472
213, 228–229, 239–241, 243–245, personal grounding 113, 118, 128,
247–249, 255–256, 262–263, 202, 215–216, 322–323, 328–329,
266–269, 271, 274, 287, 290–291, 335–336
327, 335, 338, 346, 352–354, 357, qualification 170, 215–216,
361, 366, 368, 377–381 364–365, 474
Index 495

role and purpose 214 ANSI/ESD S541 241, 256, 262–263,


selection 214–215 265–266, 268–269, 276, 352, 382
test 346, 364–365 ANSI/ESD SP10.1 133–134, 136,
types 214 138–140, 149–150, 283–284
use 214, 312, 322–323, 329, 336 ANSI/ESD SP15.1 346, 396
shielding 15, 164, 168, 173, 292, 346, ANSI/ESD STM2.1 164, 170, 346,
382–384, 442–443 351, 355, 367
electrostatic field, see packaging, ANSI/ESD STM3.1 164, 171, 346,
electrostatic field shielding 361
ESD, see packaging, ESD shielding ANSI/ESD STM4.1 164, 170, 346,
SI unit 1 351
soldering and desoldering tools 168, ANSI/ESD STM4.2 170
172, 231, 335, 346, 351, 356, ANSI/ESD STM7.1 163, 170, 346,
394–396 351
compliance verification 172, 231 ANSI/ESD STM9.1 163, 169, 200,
qualification 231 346
standards 1, 8–13, 15, 93, 98–99, ANSI/ESD STM9.2 164, 169, 346
102–103, 106–108, 110–112, ANSI/ESD STM11.11 9, 163, 173,
126–129, 148, 157–174, 180, 268–269, 346, 350, 352–354
ANSI/ESD STM11.12 10, 163, 173,
182–183, 186–188, 190–196,
346, 350, 353–354
199–203, 205–206, 209, 212–216,
ANSI/ESD STM11.13 145, 149,
220, 223, 227–229, 230–231, 233,
163, 173, 346, 350, 354–355, 379
240–244, 247, 254, 259, 262–265,
ANSI/ESD STM11.31 164, 173,
268, 272, 275–276, 281–282, 284,
265–266, 268
288, 294, 299, 304–305, 307–314,
ANSI/ESD STM12.1 164, 170, 346
320, 323–343, 345–354, 359, 361,
ANSI/ESD STM13.1 172, 346, 351
367, 379, 383, 385, 387, 396, 399, ANSI/ESD STM97.1 163, 186, 346,
449, 454–461 350, 355, 374, 386
American National standards (ANSI) ANSI/ESD STM97.2 163, 186, 346,
162 355
ANSI/ESD/JEDEC STM 5.2 55 ANSI/ESDA/JEDEC JS-001 54
ANSI/ESD S1.1 164, 169, 192, 346, ANSI/ESDA/JEDEC JS-002 54
350 ASTM D257 265–266, 268
ANSI/ESD S5.6 55, 58 automated equipment 133–134,
ANSI/ESD S8.1 234 148, 449, 457–458, 460–462
ANSI/ESD S11.4 254–257, 264, CEN/TR 16832 200, 226, 233
266–267, 276 definitions 164, 324–325
ANSI/ESD S20.20 1, 8, 12, 108, development of ESD control
128–129, 134, 157–159, 162–165, standards 157–159
168–174, 180, 183, 193, 223, EN1149-3 228
230–231, 243–244, 247, 262, 275, EN16350 233
282, 288, 324, 326, 337, 345, EN100015 158
349–350, 362, 387, 399, 452–454, ESD Association 157–159, 162, 243
456, 459–460 ESD susceptibility test 55–62
496 Index

standards (contd.) MIL-STD-1686 157–158


European Norm 157, 159 MIL STD 3010 METHOD 4046
FED STD 101 Method 4046A 265–266, 276
265–266, 276 recommendations 161, 284, 290
IEC 60364-1 182 requirements 158–163, 282, 284,
IEC60749-26 54–55 287–289, 294, 299, 308, 324–339,
IEC 60749-27 54–55, 59 350, 357, 359–360
IEC 60749-28 54 requirements of IEC61340–5-1 and
IEC 61000-4-2 55, 57–58 ANSI/ESD S20.20 164–174,
IEC 61340-1 7, 15 282, 284, 287–289, 294, 309–314,
IEC 61340-2-3 10, 149, 268–269, 324–339, 350
346, 348, 350–355, 366, 379 specific meaning of words 161
IEC 61340-4-1 186, 346, 351 Standard Practice (SP) 162
IEC 61340-4-3 200, 346 Standard Test Method (STM)
IEC 61340-4-5 186, 200, 346, 348, 162–164, 346, 350–355, 359, 361,
350, 355, 374 367, 374, 379, 386
IEC 61340-4-6 192, 223, 346, 350 Technical Report 162
IEC 61340-4-8 263, 268, 346 trends 449, 454–461
IEC 61340-4-9 223, 226, 346, 355, who writes the standards
367 159–160
IEC 61340-5-1 1, 7–8, 12, 106, 108, static dissipative, see dissipative
128–129, 134, 157–159, 163–171, storage racks and shelves 94, 98, 111,
180, 183, 191, 193, 223, 230–231, 116–118, 170, 183–184, 187,
243–244, 247, 262, 275, 282, 284, 207–210, 213, 328, 334–336,
288, 290, 294, 308–309, 312–313, 341–342, 346, 364, 366
317, 324–326, 329, 331–335, acceptance 209
337–338, 345–346, 348–352, 354, care and maintenance 208
358, 362, 374, 387, 399 common problems 209–210
IEC 61340-5-2 228, 234 compliance verification 170, 207,
IEC 61340-5-3 93, 163–164, 173, 209, 328, 472
240, 262–263, 268–269, 276, 312, grounding 183–184, 336
338, 352, 382 mobile 210, 213
IEC 61340-5-4 335, 346, 349 qualification 170, 209
IEC 62631-3-1 10 role and purpose 207–208
IEC 62631-3-2 9 selection 208
International Electrotechnical test 346, 364, 366
Commission (IEC) 157 use 207–208, 342
International Standard 159, 161- system level ESD 51–53, 56–58,
162 65–66, 80–82, 462
ISO 20344 200 relationship with component ESD
ISO 20345 12, 180, 199–200, 202 withstand 80–81
JEDEC 259–260, 264 System-Efficient ESD Design (SEED)
MIL-PRF-81705 254–257, 264–266, 81–82
268 trends 462
Index 497

t high 185, 192, 202, 204–205, 208,


time constant 22–23, 25 212, 217–218, 220–221, 230, 232,
tools 93, 98, 108, 111, 114, 119, 120, 250, 271, 298, 318–319, 342, 413,
124, 191, 203, 207, 211, 229–233, 418, 438–440
283, 298, 300–301, 321–322, 328, induced 61, 67, 75, 79, 122, 130,
335–337, 346, 355–356, 359, 136, 139, 142, 195, 214, 284,
393–394, 399–404, 418, 437–440 292–293, 295, 298, 301, 321, 331
common problems 231 measurement, see measurement,
compliance verification 231, 336 voltage
role and purpose 229 offset, see ionizer, offset voltage
qualification 230 soldering iron tip 231
selection 230, 300, 301 test 350, 360, 385
test 346, 355–356, 359, 393–394, voltmeter 357–361, 386–392,
399–404 404–405, 431–432, 440
types 229–230 electrostatic, see measurement
use 229–231, 322, 335–337 equipment, electrostatic
transmission line pulse (TLP) 64–65 voltmeter
trends 415, 420, 449–462
ESD control programs 449, w
543–459 work surface 98, 113–118, 164,
ESD withstand voltage 449–454 180–181, 203–207, 209–210, 213,
standards 449, 454–462 322, 327–328, 334–336, 346, 348,
system immunity 462 362–364, 366–367, 466, 469–470,
trolleys, see carts and mobile 472
equipment acceptance 206
triboelectrification, see charge cleaning and maintenance 206
generation common problems 206–207, 363
compliance verification 170, 181,
u 206, 327–328, 349, 362–363
unprotected areas (UPAs) 91–94, qualification 170, 180, 205–206,
117, 123–124, 180–181, 208, 212, 327, 334–336, 348, 474
257, 274, 317–318, 320, 336 role and purpose 203–204
selection 204–205
v test 346, 362–364, 366–367
voltage 2–5, 7–11, 13–15, 93, 133, types 204
135–137, 139–143, 145–152, use 204–206, 327, 335
283–295, 297–302, 313, 318–319, wrist strap 158, 164, 169, 171,
321–322, 325, 328–337, 342, 346, 191–197, 201–203, 214, 216,
348, 357–360, 386–392, 401–403, 222–223, 225, 229, 287, 309, 312,
413, 418, 423, 428–429, 431–442, 318, 322–323, 328–330, 334, 346,
445–447 350, 355, 371–374, 376–377, 393,
body, see body voltage 396, 398–399, 401, 406, 413,
ESD withstand, see ESD, withstand 417–419, 423–424, 437, 440–442,
voltage 446, 466, 468–469, 473–475
498 Index

wrist strap (contd.) qualification 169, 195, 329, 474


common problems 197, 374 role and purpose 192–193
compliance verification 169, selection 194–195
195–197, 328–329, 334, 372–374, standard 164, 169
469 system 169, 171, 192–197
constant monitor 193–194 test 346, 355, 372–377, 413,
cordless 194 417–419, 423–424, 446
grounding points 139–140, types 192–194
192–193, 322, 328–330, 334, use 193–195, 312, 318, 322–323,
376–377 329–330, 334, 393–394, 396–399
maintenance 195, 330

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy