Sei sulla pagina 1di 511

Journal of ASTM International

Selected Technical Papers

STP 1522

Flammability and Sensitivity


of Materials in Oxygen-Enriched
Atmospheres

12th Volume

JAI Guest Editors


Hervé Barthélémy
Theodore Steinberg
Christian Binder
Sarah Smith

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Journal of ASTM International
Selected Technical Papers STP1522
Flammability and Sensitivity
of Materials in Oxygen-Enriched
Atmospheres: 12th Volume

JAI Guest Editors:


Hervé Barthélémy
Theodore A. Steinberg
Christian Binder
Sarah Smith

ASTM International
100 Barr Harbor Drive
PO Box C700
West Conshohocken, PA 19428-2959

Printed in the U.S.A.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
ASTM Stock #: STP1522
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Library of Congress Cataloging-in-Publication Data

ISBN: 978-0-8031-7508-2
ISSN: 0899-1308

Copyright © 2009 ASTM INTERNATIONAL, West Conshohocken, PA. All rights


reserved. This material may not be reproduced or copied, in whole or in part, in
any printed, mechanical, electronic, film, or other distribution and storage media,
without the written consent of the publisher.
Journal of ASTM International „JAI… Scope
The JAI is a multi-disciplinary forum to serve the international scientific and
engineering community through the timely publication of the results of original
research and critical review articles in the physical and life sciences and engineering
technologies. These peer-reviewed papers cover diverse topics relevent to the
science and research that establish the foundation for standards development within
ATSM International.
Photocopy Rights
Authorization to photocopy items for internal, personal, or educational classroom
use, or the internal, personal, or educational classroom use of specific clients,
is granted by ASTM International provided that the appropriate fee is paid to ASTM
International, 100 Barr Harbor Drive, P.O. Box C700, West Conshohocken, PA
19428-2959, Tel: 610-832-9634; online: http://www.astm.org/copyright.
The Society is not responsible, as a body, for the statements and opinions expressed
in this publication. ASTM International does not endorse any products represented
in this publication.
Peer Review Policy
Each paper published in this special issue was evaluated in accordance with the JAI
review process.
The quality of the papers in this publication reflects not only the obvious efforts of
the authors and the technical editors, but also the work of the peer reviewers. In
keeping with long-standing publication practices, ASTM International maintains the
anonymity of the peer reviewers. The ASTM International Committee on
Publications acknowledges with appreciation their dedication and contribution of
time and effort on behalf of ASTM International.
Citation of Papers
When citing papers from this publication, the appropriate citation includes the
paper authors, ‘‘paper title’’, J. ASTM Intl., volume and number, Paper doi, ASTM
International, West Conshohocken, PA, Paper, year listed in the footnote to the
paper. A citation is provided as a footnote on page one of each paper.

Printed in Bridgeport, NJ
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by January, 2010
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Foreword
THIS COMPILATION OF THE JOURNAL OF ASTM INTERNATIONAL
(JAI), STP1522, on Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres: 12th Volume, contains only the papers
published in JAI that were presented at a symposium in Berlin, Germany,
on October 7, 2009–October 9, 2009 and sponsored by ASTM Committee
G04 on Compatibility and Sensitivity of Materials in Oxygen Enriched
Atmospheres.
The JAI Guest Editors are Hervé Barthélémy, Air Liquide Corporate,
Paris, France, Theodore Steinberg, Queensland University of Technology,
Brisbane, Australia, Christian Binder, BAM Federal Institute for
Materials Research and Testing, Berlin, Germany, and Sarah R. Smith,
NASA White Sands Test Facility, Las Cruces, NM, USA.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Contents
Overview .......................................................................... vii
Keynote—Materials Selection for Oxidizer Service—Successes and Challenges
J. W. Slusser ......................................................... 1
Promoted Ignition-Combustion Behavior of Cobalt and Nickel Alloys in Oxygen-
Enriched Atmospheres
J. F. Million, A. V. Samant, and R. Zawierucha ................................ 10
Defining the Flammability of Cylindrical Metal Rods Through Characterization of the
Thermal Effects of the Ignition Promoter
D. Lynn, T. Steinberg, K. Sparks, and J. M. Stoltzfus ........................... 21
A Proposed Qualitative Framework for Heterogeneous Burning of Metallic Materials:
The “Melting Rate Triangle”
N. R. Ward and T. A. Steinberg ............................................ 38
Ignition of Contaminated Aluminum by Impact in Liquid Oxygen—Influence of Oxygen
Purity
E. Werlen, F. Crayssac, O. Longuet, and F. Willot .............................. 49
Verification of the ASTM G-124 Purge Equation
K. E. Robbins, S. E. Davis, and S. D. Herald ................................. 68
Determination of Burn Criterion for Promoted Combustion Testing
K. M. Sparks, J. M. Stoltzfus, T. A. Steinberg, and D. Lynn ...................... 80
Promoted Ignition Testing of Metallic Filters in High-Pressure Oxygen
G. A. Odom, G. J. A. Chiffoleau, B. E. Newton, and J. R. Fielding ................. 96
Oxygen Compatibility of Brass-Filled PTFE Compared to Commonly Used Fluorinated
Polymers for Oxygen Systems
S. D. Herald, P. M. Frisby, and S. E. Davis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Ignition Sensitivity of Nonmetallic Materials in Oxygen-Enriched Air „NITROX…: A Never
Ending Story in SCUBA Diving?
C. Binder, T. Brock, O. Hesse, S. Lehné, and T. Tillack . . . . . . . . . . . . . . . . . . . . . . . . . . 124
The Importance of Quality Assurance and Batch Testing on Nonmetallic Materials Used
for Oxygen Service
C. Binder, K. Arlt, T. Brock, P. Hartwig, O. Hesse, and T. Tillack . . . . . . . . . . . . . . . . . . . 133
Identification and Quantification of Combustion Products Released by Non-Metallic
Materials Used for Medical Oxygen Equipment
M. Carré, H. Barthélémy, J. Bruat, S. Lombard, O. Longuet, and J. P. Schaaff . . . . . . . . 144
Determination of Time Required for Materials Exposed to Oxygen to Return to
Reduced Flammability
S. Harper, D. Hirsch, and S. Smith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Effect of Oxygen Concentration on Autogenous Ignition Temperature and Pneumatic
Impact Ignitability of Nonmetallic Materials
S. Smith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
Liquid Oxygen Rotating Friction Ignition Testing of Aluminum and Titanium with
Monel® and Inconel® for Rocket Engine Propulsion System Contamination Investigatio
S. F. Peralta, K. R. Rosales, and J. M. Stoltzfus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
Electrical Arc Ignition Testing for Common Handheld Electrical Devices in Oxygen-
Enriched Atmosphere
K. Sparks, T. Gallus, and S. Smith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
High Pressure Quick Disconnect Particle Impact Tests
K. R. Rosales and J. M. Stoltzfus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
Advanced Crew Escape Suits „ACES… Particle Impact Test
K. R. Rosales and J. M. Stoltzfus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Good Practices for Avoiding Fires in Steel Mill Oxygen Systems
E. T. Forsyth, B. E. Newton, G. J. A. Chiffoleau, and B. Brophy . . . . . . . . . . . . . . . . . . . 266
Oxygen Fire Hazards in Valve-Integrated Pressure Regulators for Medical Oxygen
E. T. Forsyth, B. E. Newton, G. J. A. Chiffoleau, and B. Forsyth . . . . . . . . . . . . . . . . . . . 285
Sealed Aluminum Cavity Reactions when Submerged in Pure O2 Reboiler Sump
W. P. Schmidt, M. Cawthra, P. A. Houghton, R. H. McDonald,Jr., R. J. Sherwood, and
S. J. Wieder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Vacuum Superinsulated Liquid Oxygen Piping and Vessels
A. Colson, E. Werlen, and H. Barthélémy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
Adiabatic Compression Testing—Part I: Historical Development and Evaluation of Fluid
Dynamic Processes Including Shock-Wave Considerations
B. E. Newton and T. Steinberg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
Adiabatic Compression Testing—Part II: Background and Approach to Estimating
Severity of Test Methodology
B. E. Newton, G. J. A. Chiffoleau, T. Steinberg, and C. Binder . . . . . . . . . . . . . . . . . . . . 362
Tribocharging of Particle Contaminants Evaluated as an Ignition Source in Oxygen-
Enriched Environments
A. Oza, S. Ghosh, and K. Chowdhury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
An Approach to Understanding Flow Friction Ignition: A Computational Fluid Dynamics
„CFD… Study on Temperature Development of High-Pressure Oxygen Flow Inside
Micron-Scale Seal Cracks
J. D. Hooser, M. Wei, B. E. Newton, and G. J. A. Chiffoleau . . . . . . . . . . . . . . . . . . . . . . 429
The Rate-Limiting Mechanism for the Heterogeneous Burning of Cylindrical Iron
Rods
N. R. Ward and T. A. Steinberg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450
Chlorine Trifluoride Exposure Testing and Oxidizer Reactivity Results
J. VanOmmeren . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
Author Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495

Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Overview
This is the twelfth Special Technical Publication (STP) originating from
ASTM Committee G04 focusing on the Flammability and Sensitivity of Ma-
terials in Oxygen-Enriched Environments. As in the past STPs, the twelfth
volume expands upon the objectives that have been carried forward since
the first STP was published in 1983. These objectives include:
• Review the current research on polymers and metals ignition and com-
bustion;
• Overview principles of oxygen systems design and issues related to
materials compatibility with oxygen; contribute to the knowledge on
the most current risk management concepts, practices, approaches,
and procedures used by individuals and organizations involved in the
design, use, retrofitting, maintenance, and cleaning of oxygen systems;
• Review accident/incident case studies related to oxygen systems and
oxygen handling procedures;
• Provide the most current data related to the flammability and sensi-
tivity of materials in oxygen-enriched atmospheres to designers, users,
manufacturers, and maintainers of oxygen components and systems
and to support Committee G04’s Technical and Professional Training
Course on Fire Hazards in Oxygen Systems and Oxygen Systems Op-
eration and Maintenance;
• Discuss enhancement, development, and use of standards sponsored
by ASTM Committee G04 on Compatibility and Sensitivity of Materi-
als in Oxygen Enriched Atmospheres;
• Provide a readily accessible reference addressing oxygen compatibility.
The twelfth Volume consists of a group of peer-reviewed publications
from the Journal of ASTM International that were also presented at Com-
mittee G04’s Twelfth International Symposium held in Berlin, Germany in
October 2009. The volume contains 29 papers on topics related to ignition
and combustion of non-metals, ignition and combustion of metals, oxygen
compatibility of components and systems, analysis of ignition and combus-
tion, failure analysis and safety, and includes aerospace and industry oxy-
gen applications.
The keynote address at the Twelfth International Symposium was pre-
sented by Joseph Slusser, Senior Engineering Associate—Materials, Air
Products and Chemicals, Inc, Allentown, PA USA. He has been involved
with materials selection for oxygen and other strong oxidizers since joining
Air Products and Chemicals, Inc in 1980 and has been a member of Commit-
tee G04 since 1997. Mr. Slusser has been active in Committee activities, is
currently the Chairman of Committee G04, and also serves on the ASTM
Committee on Standards. In addition, he has been involved with several
CGA task groups related to oxygen compatibility issues and has published
several papers on oxygen and nitrogen trifluoride compatibility topics.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by vii
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Eight papers focus on ignition and combustion of metals. A study is pre-
sented by Praxair on the promoted Ignition of cobalt and nickel alloys in
oxygen enriched atmosphere. They found that although nickel and cobalt
alloys are generally considered as having superior combustion resistance in
oxygen—enriched atmospheres, flammability testing is necessary prior to
service as there may be considerable variation in their promoted ignition-
combustion resistance. Queensland University of Technology presents two
papers; one where they show that heat transfer is the rate-limiting mecha-
nism for the heterogeneous burning of cylindrical iron rod and, in the other,
introduce the concept of the “melting rate triangle” to aid in the discussion of
burning metallic materials. The fourth paper in this section presents results
on the ignition of contaminated aluminum by mechanical impact in liquid
oxygen. There are two papers, one submitted by NASA White Sands Test
Facility and one by Queensland University of Technology that, together, de-
fine and validate the burn criterion used in the recent rewrite of ASTM Com-
mittee G04’s G124 standard on burning metallic materials. The last two pa-
pers in this section are submitted by Wendell Hull & Associates; one a test
method developed to quantify the ignitability of metallic materials and the
other presenting the test result for metallic filters for use in high-pressure
oxygen.
Six papers are related to the use of non-metallic materials in oxygen-
enriched atmospheres. The first, submitted by NASA Marshall Space Flight
Center, looks at the compatibility of brass-filled PTFE as compared to other
fluorinated polymers. The second and third papers in this section, submitted
by BAM, first investigate the ignition of nonmetallics in NITROX mixes as
encountered in typical SCUBA diving applications followed by a discussion
on the importance of quality assurance and batch testing in the qualification
of nonmetallic materials for oxygen service. The fourth paper on nonmetal-
lics, submitted by Air Liquide, provides information on the type and quan-
tity of combustion products produced when nonmetallic materials burn in
oxygen systems. The final two papers in this section are submitted by NASA
White Sands Test Facility with the first evaluating the time it takes for non-
metallics exposed to oxygen to no longer be considered “oxygen enriched”,
and the second looking at the effect of oxygen concentration on the AIT and
ignitability by pneumatic impact of nonmetallic materials.
Eight papers focus on issues related to oxygen compatibility of compo-
nents and systems. Four papers are by NASA White Sands Test Facility; the
first is a study on a frictional heating of contaminated metals in a liquid
oxygen rocket engine propulsion system, the second evaluated the arc igni-
tion characteristics of common handheld devices in oxygen-enriched atmo-
spheres and, both the third and fourth looked at particle impact ignition in a
high pressure quick disconnect and in the Advanced Crew Escape Suits. Two
papers submitted by Oxygen Safety Consultants and Wendell Hull & Asso-
ciates evaluate the fire hazards in steel mill oxygen systems and within
valve-integrated pressure regulators. The seventh paper in this section
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by viii
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
looks at reactions within a sealed aluminum cavity within a reboiler while
the final paper discusses vacuum super insulated liquid oxygen piping and
vessels.
The final section of the STP contains six papers related to specific appli-
cations (or other areas). Two papers presented here by Wendell Hull & As-
sociates, look in detail at adiabatic compression testing with the first pre-
senting a background on this important testing and the second providing a
proposed framework within which the severity of the test can be related to
real components in service. The next two papers are an attempt to help
clarify the mechanism behind flow friction ignition. The first provides a dis-
cussion on the importance of tribiocharging of particle contaminants while
the other uses CFD analysis as an explanation of this ignition phenomena.
Next a submission from NASA Marshall Space Flight Center verifies the
purge equation used in the G124 metals standard. Finally, some work is
presented on Chlorine Trifluoride reactivity and testing.
The twelfth Volume on Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres provides a diverse source of new information
to air separation industry, oxygen manufacturers, manufacturers of compo-
nents for oxygen and other industrial gases service, manufacturers of mate-
rials intended for oxygen and other industrial gases service, and users of
oxygen and oxygen-enriched atmospheres aerospace, medical, industrial
gases, chemical processing, steel and metals refining, as well as to military-
commercial-recreational diving.
Hervé Barthélémy
Air Liquide International Fellow
Paris, France
Theodore A. Steinberg
School of Engineering Systems
Faculty of Built Environment and Engineering
Queensland University of Technology
Brisbane, Queensland, Australia
Christian Binder
BAM
Berlin, Germany
Sarah Smith
NASA, WSTF
Las Cruces, NM

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by ix
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102637
Available online at www.astm.org/JAI

Joseph W. Slusser1

Keynote—Materials Selection for Oxidizer


Service—Successes and Challenges

ABSTRACT: Improper selection of materials of construction for oxygen and


other strong oxidizer applications can result in sudden catastrophic failure.
Many available materials are flammable and will burn in oxidizer systems if
sufficient ignition energy is provided. Because these combustion events
occur suddenly, they cannot be monitored by periodic inspection. Also, a
definitive cause after such an event is often difficult to determine because
much of the evidence is destroyed. Materials selection and system design
must be correct from the beginning. The key is to ensure the ignition ener-
gies imposed on the materials during operation of the system do not chal-
lenge the materials. Or, to state it differently, a designer must ensure that the
materials selected have adequate compatibility to withstand the operating
conditions 共including transient conditions such as startups兲 without igniting.
This paper will provide an overview of the science of oxygen compatibility
and how Committee G04 efforts have helped improve safety. It will also
comment on the technical issues the Committee may address in the future
and a concern, shared with many similar organizations, of declining industry
participation.
KEYWORDS: oxygen, oxygen compatibility, oxidizer, flammability

Introduction

ASTM G04 Committee was formed almost 35 years ago to address specific
concerns of materials selection for oxygen service and is the only international
organization devoted to this important issue. The Committee has responded to
this need by developing numerous standards and publications that accomplish
the following goals:
共1兲 Describe how to test materials to determine flammability limits,

Manuscript received June 29, 2009; accepted for publication August 19, 2009; published
online September 2009.
1
Senior Engineering Associate–Materials, Air Products and Chemicals, Inc., 7201
Hamilton Blvd., Allentown, PA 18195.
Cite as: Slusser, J. W., ⬙Keynote—Materials Selection for Oxidizer Service—Successes
and Challenges,⬙ J. ASTM Intl., Vol. 6, No. 10. doi:10.1520/JAI102637.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 1
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
2 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

共2兲 Explain how to test components to determine if they can survive a


strong ignition event 共i.e., are fault tolerant兲,
共3兲 Provide guidance to properly design systems,
共4兲 Promote research to further our understanding of materials–oxidizer
compatibility issues, and
共5兲 Liaison with other organizations, such as the European Industrial
Gases Association 共EIGA兲 and Federal Institute for Materials Research
and Testing 共BAM兲2 in Europe and the Compressed Gas Association
共CGA兲 and National Fire Protection Agency 共NFPA兲 in the United States
to promote oxidizer safety.
Most of the Committee’s work has been focused on oxygen, as it is the most
popular strong oxidizer used in industry. Other oxidizers such as F2, NF3, and
ClF3 are now also receiving attention. In general, the comments and practices
for oxygen safety also apply to these other strong oxidizers.
The title of Committee G04 is Compatibility and Sensitivity of Materials in
Oxygen Enriched Atmospheres, the key terms being compatibility and sensitiv-
ity. The Committee has used slightly different definitions for compatibility in
various standards, but in general compatibility refers to the inherent resistance
to ignition and combustion of a material in oxygen service. Materials with good
compatibility are expected to be highly resistant to ignition, while those with
low or poor compatibility would be expected to be relatively easy to ignite. The
Committee has not specifically defined sensitivity, but commonly sensitivity re-
fers to how a change in a material’s physical condition 共e.g., a decrease in
thickness兲 or a change in operating condition 共e.g., increase in pressure兲 can
affect the resistance of that material to ignition. Highly sensitive materials can
experience a considerable change in ignition resistance 共compatibility兲 with
changes in physical condition or operating conditions.

Committee Successes
In its 35 years of operation, the Committee has enjoyed many successes, includ-
ing some of the following examples:
• International collaboration—sponsored 12 symposiums, five of which
have been held outside the United States;
• Standardization—developed 21 standards covering metals and nonmet-
als testing, component testing, design guides, cleaning procedures, and
fire incident studies 共see Appendix for a listing of G04 standards兲;
• Other oxidizers—addressed issues with fluorine and other nonoxygen
oxidizers in recent years;
• Research—helped fund research efforts in metals and nonmetals flam-
mability in oxygen and NF3; and
• Training—developed very successful training programs on oxygen com-
patibility, which has been completed by more than 2500 industry pro-
fessionals since it was initiated in 1990.
The Committee has been a forum where those interested in the important

2
Bundesanstalt für Materialforschung und -prüfung.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SLUSSER, doi:10.1520/JAI102637 3

topic of oxygen safety, especially the safe use of materials in oxygen and other
strong oxidizers, can meet to discuss issues of common interest. It has included
industrial gas companies, valve and equipment suppliers, aerospace industry
groups and NASA, testing laboratories, consultants, and researchers. Commit-
tee G04 members have worked with many other organizations, most notably
CGA, NFPA, EIGA, and BAM, to distill the oxygen safety information developed
in the Committee to practical industrial guides. One of the most important
documents resulting from these endeavors is the oxygen piping standard jointly
published by CGA and EIGA 关1,2兴.

Materials Selection for Oxygen Service

Materials selection for any service is always a mix of economic consideration


共cost and availability兲, strength 共adequate for the design conditions兲, resistance
to the service environment 共corrosion and erosion兲, and loss of properties with
time 共creep in metals and aging of polymers兲. However, materials selection for
oxygen also has a unique safety feature compared to other services in that
many of the materials of construction are flammable and, if ignited, will likely
result in sudden and catastrophic failure. Because of the sudden nature of
oxygen system failures, periodic inspections are not effective. The system must
be designed and operated properly.

The Fire Hazard

Several materials flammability generalizations have been known since the early
days of the oxygen industry, and these generalizations continue to influence the
design of oxygen systems today. Polymer use is limited because they are rela-
tively easy to ignite. Essentially all plastics, elastomers, oils, and greases are
flammable in oxygen systems 共many are also flammable in air兲, especially the
hydrocarbon-based polymers like natural rubber and polyethylene. Polymers
are easier to ignite 共that is, require less ignition energy and have lower ignition
temperatures兲 in pure oxygen than in air and easier still as the oxygen pressure
and temperature increases. Broadly speaking, the least flammable 共most com-
patible兲 materials are those with the lowest heat of combustion such as fluori-
nated materials like PTFE.
Although metals as a class are more difficult to ignite than polymers, many
metals, including carbon steel, alloy steel, stainless steel, and aluminum, are
also flammable in oxygen. Some metals that are essentially nonflammable in-
clude copper alloys 共e.g., brass and bronze兲, nickel, and nickel-copper alloys
共e.g., Alloy 400 or Monel兲. In general, a metal’s flammability is related to its heat
of oxide formation 共analogous to the heat of combustion of polymers兲; those
that evolve a large amount of heat such as iron, chromium 共a major constituent
of stainless steel兲, and aluminum are flammable. Those with low heats of oxide
formation, primarily copper and nickel, have low flammability.
With these generalizations, the oxygen industry quickly established several
“good practices” as follows:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
4 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

• Clean oxygen systems to remove hydrocarbon fluids 共such as machining


oil兲 and particulate 共such as metal fines兲.
• Limit the use of plastics, elastomers, and lubricants and, when needed,
use the most compatible ones consistent with the other design needs
such as strength.
• Limit oxygen velocities to limit particle velocity 共and ignition energy兲 of
any remaining particles and those that may be generated during service.
• Recognize that components that are exposed to direct impingement
共such as tees兲 or high pressure drop 共such as control valves兲 are in more
severe service and many require the use of nonflammable metallurgy.
Carbon steel and other less compatible alloys could be used for rela-
tively benign components 共i.e., straight runs of pipe兲.
The Committee’s work, to a large extent, has been to better define the ma-
terial limits. We have learned that aluminum flammability is strongly reduced
by low levels of specific impurities in the oxygen 关3兴, stainless steel flammability
is influenced by thickness 关4兴, and even Monel, commonly felt to be the most
compatible metal, is flammable in thin sections such as filters 关5兴. We also know
much more about the fundamental aspects of metals combustion 关6–8兴. Much
of this work was first reported in G04 symposia.
We are also beginning to explore how to safely use plastic structural mate-
rials such as tubing and pipe in low pressure oxygen and enriched air systems
for such applications as breathing gas and aqua agriculture. Other oxidizers are
beginning to receive serious attention. Papers in earlier symposia have been
presented on NF3 and F2 关9,10兴, and this symposium will have a paper on ClF3
关11兴.

Designing with Flammable Materials

A key issue is how to safely use materials that are flammable in oxygen service.
Test methods such as ASTM Standard G124-95 关12兴 have enabled us to deter-
mine “threshold pressures” 共the pressure below which a metal alloy in a given
configuration will not support combustion even in the presence of a strong
igniter兲. Threshold pressures are not intrinsic material properties but functions
of thickness and test conditions such as temperature. However, economics and
practicality argue that many components must use alloys at pressures well
above their threshold pressure, that is, at conditions where they are flammable.
In addition, some components, such as O-rings and gaskets, must be made
from nonmetals that are typically less compatible than metals. How do we use
these materials safely?
We need to understand the ignition mechanisms that can operate in the
system and how design feature can influence these mechanisms. Simply expos-
ing common materials, including plastics, to ambient temperature oxygen,
even high pressure oxygen, will not cause ignition. Some energy must be im-
parted to the material to initiate the combustion reaction, that is, ignite the
material. ASTM Standard G88-05 关13兴 design guide has a comprehensive dis-
cussion of ignition mechanisms. Some common ignition mechanisms are par-
ticle impact in flowing systems, heating due to adiabatic compression of oxygen
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SLUSSER, doi:10.1520/JAI102637 5

gas 共applies mostly to polymers兲, friction rubbing in rotating equipment, and


impact loading as in the closing of a check valve.
We also must consider the kindling chain. This chain seeks out the most
flammable material 共such as a polymer or contamination兲, and if that ignites, it
promotes the burning of other more critical components. For example, ignition
of a grease contaminant could ignite a polymer seat that, in turn, ignites the
valve stem and, finally, the valve body, leading to a breach of the system. This is
why cleaning is so important; it removes one of the most vulnerable links in the
chain—the most flammable materials 共oils and fine metal particles兲—and helps
prevent the kindling chain from initiating. Committee G04 has several cleaning
related standards, and both EIGA and CGA have published guides on cleaning.
Cleaning is the first line of defense in safe operation of oxygen and oxidizer
systems.
The industry also practices what is viewed as “defensive design” of compo-
nents and systems. An excellent document to review in this regard is ASTM
Standard G88-05 design guide. This effort requires knowledge of oxygen mate-
rial compatibility properties and the effect of design features on the ignition
process and the kindling chain. For instance, if a valve with tight shut off is
needed, a nonmetallic seat, possibly PCTFE3 or nylon, would be required.
PCTFE has better compatibility properties than nylon but may be a toxicity
concern in breathing gas systems; under the latter circumstances, nylon would
be used. In either case, the mass of the seat would be minimized 共to reduce heat
release if it ignites兲 and heat sunk4 into the body or stem of a highly compatible
material such as bronze or Monel. If a lower compatibility material is used,
such as nylon, greater scrutiny must be given to the valve design and system
operating conditions to minimize the potential for seat ignition and to ensure
the valve can accommodate combustion of the seat without propagating to a
large fire and breach of the system. The latter issue, the robustness of the
component is the emphasis of ASTM Standard G175-03 关14兴 test method.
In systems such as an oxygen pipelines or piping systems, the approach is
much the same—identify the components that could be exposed to severe ser-
vice, high turbulence, or direct impingement and take added precautions with
these. This could be upgrading metallurgy to more compatible alloys 共Monel
instead of carbon steel兲 or, for impingement, use of carbon steel following in-
dustry impingement velocity guidelines. The intent is to use highly compatible
alloys where ignition energies could be high or to limit the ignition energy of
particles by limiting velocity.
For rotating equipment such as compressors, there is always the potential
for high frictional ignition energy from a rub. The common approach is to
simply barricade this equipment so that if a rub should occur and ignite the
equipment, personnel are protected.
In short, most oxygen and oxidizer systems are designed with “flammable”
materials because it is usually not practical to use all nonflammable materials.

3
Polychlorotrifluoroethylene.
4
Heat sunk refers to embedding a polymer piece in metal
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
to help dissipate heat.
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
6 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

This is done safely by understanding the interactions of materials compatibility,


system and component design, and ignition mechanisms.

Some Thoughts on the Future


As this symposium shows, there continues to be much work done on oxygen
compatibility issues of metals and nonmetals. However, the “nuts and bolts” of
the Committee’s operation—revising existing standards, writing new standards,
updating other publications, etc.—has tended to fall to a shrinking group of
individuals. The current global economic downturn has exacerbated this trend.
Newer communication technologies, such as web meetings, will help increase
participation, and ASTM is aggressively pursuing these technologies. However,
we also need to communicate the importance of oxygen safety and G04’s efforts
to the management of our individual organizations. This is especially impor-
tant as industries reduce staff and re-prioritize the efforts of their engineering
and research groups.
The Committee’s efforts will likely expand beyond its primary focus, which
has been materials compatibility in gaseous oxygen 共GOX兲 systems. GOX sys-
tems and materials combustion in GOX will continue to be a chief focus. How-
ever, several aligned areas may grow in importance in the future.
• Other oxidizers such as NF3, F2, O3, and ClF3 will continue to receive
more attention.
• The area of minimum ignition energy will be studied. The only ignition
energy related standards we have are the liquid oxygen and GOX impact
test standards. More attention will be given to developing standards to
measure ignition energy.
• There is a need for a guide on the safe use of plastic tubing and piping in
low pressure oxygen and enriched air, which the Committee has started
to address with a draft document. This would apply to medical and
waste water treatment among other uses.
• The Committee will likely give more attention to cryogenic oxidant sys-
tems.
No doubt, there are many other topics and issues that will surface in the
coming years and will need to be addressed by the oxygen/oxidizer compatibil-
ity community. Oxygen use is safe and has become safer because of the efforts
of ASTM G04 Committee. This is an endeavor that will and must continue in
the future.

Acknowledgments
The writer is greatly appreciative of Air Products for allowing him to partici-
pate in Committee G04 and serve as chairman of the committee for the past six
years. Many individuals have broadened the writer’s understanding of oxygen
compatibility—for this is a team sport—but he must mention Barry Werley,
specifically, for always being willing to share his insight and knowledge of com-
patibility issues. Barry Werley is a former Chair of the G04 Committee and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SLUSSER, doi:10.1520/JAI102637 7

Lead Oxygen Compatibility Specialist for Air Products; although he is now


retired, he is still very active within the group.

APPENDIX–COMMITTEE G4 STANDARDS

G04.01 共Test Methods Subcommittee兲 Standards 共All Are Standard Test Methods兲
D2512-95共2002兲e1 Compatibility of Materials with Liquid Oxygen
共Impact Sensitivity Threshold and Pass-Fail
Techniques兲
G72-01 Autogenous Ignition Temperature of Liquids
and Solids in a High-Pressure
Oxygen-Enriched Environment
G74-01 Ignition Sensitivity of Materials to Gaseous
Fluid Impact
G86-98a共2005兲 Determining Ignition Sensitivity of Materials
to Mechanical Impact in Ambient Liquid
Oxygen and Pressurized Liquid and Gaseous
Oxygen Environments
G122-96共2002兲 Evaluating the Effectiveness of Cleaning
Agents
G124-95共2003兲 Determining the Combustion Behavior of
Metallic Materials in Oxygen-Enriched
Atmospheres
G125-00 Measuring Liquid and Solid Material Fire
Limits in Gaseous Oxidants
G144-01共2006兲 Determination of Residual Contamination of
Materials and Components by Total Carbon
Analysis Using a High Temperature
Combustion Analyzer
G175-03 Evaluating the Ignition Sensitivity and Fault
Tolerance of Oxygen Regulators Used for
Medical and Emergency Applications
G04.02 共Recommended Practices Subcommittee兲 Standards
G63-99 Standard Guide for Evaluating Nonmetallic
Materials for Oxygen Service
G88-05 Standard Guide for Designing Systems for
Oxygen Service
G93-03e1 Standard Practice for Cleaning Methods and
Cleanliness Levels for Material and
Equipment Used in Oxygen-Enriched
Environments
G94-05 Standard Guide for Evaluating Metals for
Oxygen Service
G114-06 Standard Practices for Evaluating the Age
Resistance of Polymeric Materials Used in
Oxygen Service

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
8 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

G120-01 Standard Practice for Determination of Soluble


Residual Contamination by Soxhlet Extraction
G121-98共2004兲 Standard Practice for Preparation of Contaminated
Test Coupons for the Evaluation of Cleaning Agents
G127-95共2000兲 Standard Guide for the Selection of Cleaning Agents
for Oxygen Systems
G128-02e1 Standard Guide for Control of Hazards and Risks in
Oxygen Enriched Systems
G131-96共2002兲 Standard Practice for Cleaning of Materials and
Components by Ultrasonic Techniques
G136-03 Standard Practice for Determination of Soluble
Residual Contaminants in Materials by Ultrasonic
Extraction
G145-96共2001兲 Standard Guide for Studying Fire Incidents in Oxygen
System
G04.03 共Terminology兲 Standard
G126-00 Standard Terminology Relating to the Compatibility
and Sensitivity of Materials in Oxygen Enriched
Atmospheres

References

关1兴 “Industrial Practices for Gaseous Oxygen Transmission and Distribution Piping
Systems,” Pamphlet G-4.4, Compressed Gas Association, Arlington, VA, 2003.
关2兴 “Oxygen Pipeline Systems,” Report No. IGC Doc13/02/E, European Industrial
Gases Association, Avenue Des Arts 3-5, B-1210 Brussels, Belgium, 2002.
关3兴 Werley, B. L., Barthélémy, H., Gates, R., Slusser, J. W., Wilson, K. B., and Zaw-
ierucha, R., “A Critical Review of Flammability Data for Aluminum,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Sixth Volume, ASTM
STP 1197, D. D. Janoff and J. M. Stoltzfus, Eds., ASTM International, West Con-
shohocken, PA, 1993, pp. 300–348.
关4兴 Zabrenski, J. S., Werley, B. L., and Slusser, J. W., “Pressurized Flammability Limits
of Metals,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Fourth Volume, ASTM STP 1040, J. M. Stoltzfus, F. J. Benz, and J. S.
Stradling, Eds., ASTM International, West Conshohocken, PA, 1989, pp. 178–194.
关5兴 Schadler, J. L. and Stoltzfus, J. M., “Pressurized Flammability Limits of Selected
Sintered Filter Materials in High-Pressure Gaseous Oxygen,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Sixth Volume, ASTM STP
1197, D. D. Janoff and J. M. Stoltzfus, Eds., ASTM International, West Consho-
hocken, PA, 1993, pp. 119–132.
关6兴 Glassman, I., “Combustion Fundamentals of Low-Volatility Materials in Oxygen-
Enriched Atmospheres,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Fifth Volume, ASTM STP 1111, J. M. Stoltzfus and K. McIl-
roy, Eds., ASTM International, West Conshohocken, PA, 1991, pp. 7–25.
关7兴 Wilson, D. B., Steinberg, T. A., and DeWit, J. R., “The Presence of Excess Oxygen in
Burning Metallic Materials,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. New-
ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SLUSSER, doi:10.1520/JAI102637 9

关8兴 Lanyi, M. D., “Discussion on Steel Burning in Oxygen 共From a Steelmaking Met-
allurgist’s Perspective兲,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. New-
ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关9兴 Gugliemini, C. J., Kadri, S. H., Martrich, R. L., Slusser, J. W., Vora, J., Werley, B.
L., and Woytek, A. J., “Flammability of Metals in Fluorine and Nitrogen Trifluo-
ride,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Seventh Volume, ASTM STP 1267, D. D. Janoff, W. T. Royals, and M. V. Gunaji,
Eds., ASTM International, West Conshohocken, PA, 1995, pp. 107–127.
关10兴 Newton, B. E. and Chiffoleau, G. J. A., “Materials Compatibility with Nitrogen
Trifluoride NF3 ,” J. ASTM Int., Vol. 3, No. 5, 2006, paper ID JAI 13552.
关11兴 VanOmmeren, J., “Chlorine Trifluoride Exposure Testing and Oxidizer Reactivity
Results,” J. ASTM Int., Vol. 6, No. 9, 2009, paper ID JAI 102215.
关12兴 ASTM Standard G124-95, “Standard Test Method for Determining the Combustion
Behavior of Metallic Materials in Oxygen-Enriched Atmospheres,” Annual Book of
ASTM Standards, ASTM International, West Conshohocken, PA, 2009, Vol. 14.04.
关13兴 ASTM Standard G88-05, “Standard Guide for Designing Systems for Oxygen Ser-
vices, Annual Book of ASTM Standards, ASTM International, West Conshohocken,
PA, 2009, Vol. 14.04.
关14兴 ASTM Standard G175-03, “Standard Test Method for Evaluating the Ignition Sen-
sitivity and Fault Tolerance of Oxygen Regulators Used for Medical and Emer-
gency Applications,” Annual Book of ASTM Standards, ASTM International, West
Conshohocken, PA, 2009, Vol. 14.04.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102230
Available online at www.astm.org/JAI

Joseph F. Million,1 Anand V. Samant,2 and Robert Zawierucha3

Promoted Ignition-Combustion Behavior


of Cobalt and Nickel Alloys in
Oxygen-Enriched Atmospheres

ABSTRACT: Promoted ignition-combustion behavior has been used to de-


scribe a situation in which a substance with low oxygen compatibility ignites
and supports the combustion of a more combustion-resistant material. Pre-
vious work has been reported on the investigation of this phenomenon as it
relates to carbon steel, stainless steels, and a number of significant engi-
neering alloys in the nickel, cobalt, and copper families. Reported in this
paper are the results of promoted ignition-combustion tests of three cobalt
based alloys in oxygen-enriched atmospheres at pressures ranging from
2.86 to 34.5 MPa. Included in the test program were Stellite 6, a commonly
used hard facing alloy and two high strength cobalt alloys—MP35N and
Elgiloy. Also reported in this paper are the results of promoted ignition-
combustion tests of four nickel based alloys in oxygen-enriched atmo-
spheres at pressures up to 34.5 MPa. Nickel alloys evaluated in the program
included Hastelloys B-3 and W, RA 330, and a cast variant of the Hastelloy C
type, CW6M. Hastelloy B-3 and CW6M are used in chemical process indus-
try applications where corrosion is of concern. RA 330 is a significant el-
evated temperature alloy. Hastelloy W is a filler metal used for dissimilar
metal welds.
KEYWORDS: metals, flammability, promoted ignition-combustion,
PICT behavior, Stellite 6, MP35N, Elgiloy, Hastelloy B-3, Hastelloy W,
RA 330, CW6M

Manuscript received November 13, 2008; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
Development Associate, Materials Engineering R&D, Praxair, Inc., P. O. Box 44,
Tonawanda, NY 14151-0044.
2
Manager, Materials Engineering R&D, Praxair, Inc., P. O. Box 44, Tonawanda, NY
14151-0044.
3
Independent Consultant, 25 Shearer Ave., East Aurora, NY 14052.
Cite as: Million, J. F., Samant, A. V. and Zawierucha, R., ⬙Promoted Ignition-Combustion
Behavior of Cobalt and Nickel Alloys in Oxygen-Enriched Atmospheres,⬙ J. ASTM Intl.,
Vol. 6, No. 10. doi:10.1520/JAI102230.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 10
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
MILLION ET AL., doi:10.1520/JAI102230 11

Introduction
Within recent years, the technical literature pertinent to the combustion behav-
ior of metals in gaseous oxygen has been significantly enriched, primarily due
to the activities of ASTM Committee G-4. A sampling of the literature shows
that both static and flowing test systems have been used 关1–18兴. This paper is a
follow-up to our earlier work, which dealt with the promoted ignition-
combustion behavior of carbon steel, stainless steels, and selected engineering
alloys from various alloy families.
The metals evaluated in this program vary considerably in alloy composi-
tion, capability, and cost. They were evaluated due to their potential use in
varied applications where oxygen-enriched atmospheres might be encountered.
Some of these are as follows:
• Electronics;
• Hazardous waste;
• Elevated temperatures;
• Dissimilar metal transition joints;
• Glass melting;
• Marine environments;
• Heavy industrial environments; and
• Elevated temperature air enrichment.
The cobalt and nickel alloys are frequently very expensive and selected for
specific applications. Hard facing, high strength, corrosion resistance, weld
filler for dissimilar metals, and high temperature service are the specific niches
handled by various alloys treated in this technical paper.

Alloy Description
The following cobalt alloys were evaluated:
• MP35N 共UNS No. R30035兲—MP35N has the nominal composition 35Co
35Ni 20Cr 10 Mo. It is a precipitation hardenable alloy, which has ex-
ceptional corrosion resistance in harsh environments. This is a relatively
expensive alloy, which is used in premium niche applications.
• Elgiloy 共Phynox兲 共UNS No. R30003兲—Elgiloy has the nominal composi-
tion 40Co 25Cr 17.5Fe 15.5Ni 7Mo. It is a precipitation hardenable alloy,
which is highly corrosion resistant and it is commonly used in body
implants. Outside of medical applications, it sees service as a spring
material in select temperature applications. This is a relatively expensive
alloy, which is used in premium niche applications.
• Stellite 6 共UNS No. W73006兲—Stellite 6 has the nominal composition of
55.5Co 29Cr 4.5W 3Fe 3Ni 1.5Si 1.0Mo 1Mn 1C. This is a very common
hard facing alloy, which may be applied to wear surfaces on valves or
other components in abrasive services. It also sees service in the form of
valve stems.
The following nickel alloys were evaluated:
• CW6M 共UNS No. N30107兲—the nominal composition of CW6M is 60Ni
l8.5Cr l8.5Mo and it is a cast version of the original Hastelloy C compo-
sition, which has been specially modified for casting. Typically, the alloy
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
12 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

is procured per ASTM/ASME Specifications A/SA 494. Alloys of the Has-


telloy C type and their cast analogs are widely used in the chemical
industry because they provide corrosion resistance over a wide range of
reducing and oxidizing conditions. Seawater corrosion resistance and
resistance to chlorides are excellent. CW6M may be more commonly
encountered outside of the United States.
• Hastelloy B-3 共UNS No. N10675兲—the nominal composition of Hastelloy
B-3 is 65Ni 30Mo 5共Co, Cr, Fe, W兲. This is a complex third generation
derivative of the original Hastelloy B composition. The major applica-
tions of Hastelloy B alloys involve services in hydrochloric acid, sulfuric
acid, and pure phosphoric acid. Hastelloy B-3 was specifically formu-
lated to provide improved thermal stability. Due to the low chromium
levels, its resistance to oxidizing conditions is inferior to austenitic
stainless steel and Hastelloys of the C and G series. However, in certain
moderately elevated temperature applications, the thermal expansion
characteristics of Hastelloy B-3 may result in its consideration for cer-
tain applications.
• RA 330 共UNS No. N08330兲—this nickel alloy is in some quarters thought
of as a stainless steel and has a nominal composition of 42Fe 35Ni 19Cr
1Cu 1Si. Its predominant usage is in elevated temperature applications.

Experimental

Promoted Ignition-Combustion, Significance of Test Approach


The “extended fire triangle” was modified by Slusser and Miller 关1兴 to take into
account the specific concerns or aspects of the industrial gas business as they
pertain to oxygen-enriched atmospheres. It shows that system contamination
has a direct impact on two major ignition mechanisms, i.e., “energy release
from 2nd material” also known as “promoted ignition” and also “particle im-
pact.” Promoted ignition-combustion test techniques have been utilized to de-
fine flammable and nonflammable zones for engineering alloys of significance
to the industrial gas business.
For a fire to occur, fuel, oxygen, and an ignition mechanism are required.
Successful oxygen systems minimize fuel and eliminate ignition mechanisms.

Test Vessels
Two test systems were predominantly used to conduct the test programs. For
pressures up to 10.4 MPa 共1,500 psig兲, a flow tester was used. At pressures from
10.4 MPa 共1,500 psig兲 to 34.5 MPa 共5,000 psig兲 a 100-mm diameter static tester
was used; although, some tests using the static tester were conducted at lower
pressures. It should be noted, however, that the flow test approach is required
for testing at relatively low pressures to prevent oxygen starvation that result
from combustion events. The oxygen flow rate is nominally 20 slpm. Schemat-
ics of the static tester and flow tester are discussed in previous publications
关3,4,7,16兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
MILLION ET AL., doi:10.1520/JAI102230 13

Oxygen-Nitrogen Gas Mixtures


The cylinder gas mixtures used in this study were synthetic blends of oxygen
and nitrogen with nominal oxygen concentrations of 99.7 %. Actual oxygen
concentrations as measured via gas chromatography ranged between 99.55 and
99.83 %.

Promoters
The promoters utilized are identical to those used in previous investigations
关3,4,7,16–18兴. The function of these promoters is to provide an overwhelming
ignition source. Basically, they are as follows:
• Static tester: 0.5 g of Mobil 600 oil and 0.15 g of iron wire wound at the
bottom of the rod; and
• Flow tester: 0.15 g of iron wire wound at the bottom of the rod.

Sample Dimensions
The sample rods measured 0.3175 cm in diameter by 10 cm long. Welding filler
metals, wrought bar stock, and cast rods were the sources of the sample rods.

Alloy Chemical Compositions


Table 1 shows the actual chemical compositions of the alloys tested. Heats of
combustion 共⌬H兲 have been calculated based on the summed products of the
weight percent of each element in the alloy 共from chemical analysis兲 and the
element’s theoretical heat of combustion 共⌬H兲 in cal/g.

Burn Criteria and Promoted Ignition-Combustion Transition Curve


Data generated in this study exhibited behavior elements described within the
promoted ignition-combustion transition curve 共PICT兲. In previous work 关7兴,
investigators reported that flammability experiments have shown the existence
of the PICT phenomenon and a schematic of the phenomenon is presented in
that work.
With the specimen diameters and promoters chosen, plotted test results
exhibit an upper shelf where combustion propagation is arrested, a lower shelf
where complete combustion always occurs, and a transition zone. All test
points reported as burns in this paper lost over 3 cm in sample length as a
consequence of sample ignition and burn propagation. Some of the alloys
which were tested, however, exhibited transitional behavior even at the highest
pressure tested.

Results/Observations

Cobalt Alloys
The test results for the cobalt alloys are tabulated in Table 2. This table includes
Elgiloy, MP35N, and Stellite 6.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
14 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1—Actual chemical composition of test alloys.

Element
共wt/%兲 MP35N Elgiloy Stellite 6 CW6M Hastelloy B-3 Hastelloy W RA 330
Al 0.08 0.07 ... ... 0.43 ... ...
C 0.006 ... 1.26 0.02 0.004 0.07 0.020
Cb 0.07 ... ... ... ⬍0.05 ... ...
Co 31.73 39.33 57.57 ... ⬍0.10 ⬍0.05
Cr 20.1 20.08 28.80 18.30 1.53 5.68 16.90
Cu 0.02 ... ... ... 0.02 0.01 0.04
Fe 0.12 15.75 2.76 2.80 1.50 5.18 45.01
Mn 0.01 1.84 0.10 0.46 0.64 0.64 2.09
Mo 10.23 7.12 0.17 18.00 28.24 24.30 0.03
Ni 36.84 15.31 2.81 59.44 66.94 64.37 35.38
P ⬍0.003 0.002 ... 0.001 0.003 0.004 0.015
S 0.009 ⬍0.001 ... 0.001 0.002 0.002 0.015
Si 0.06 0.5 1.21 0.97 ⬍0.02 0.27 0.50
Ta 0.01 ... ... ... ... ... ...
Ti 0.76 ... ... ... ... ... ...
V 0.01 ... ... ... ⬍0.01 ... ...
W 0.01 ... 4.82 ... 0.18 0.01 ...
Zr 0.01 ... ... ... ... ...
Others ⬍0.50
⌬H cal/g 1,432 1,535 1,611 1,522 1,197 1,358 1,652

Elgiloy

Table 2 shows test results for Elgiloy up to 10.4 MPa 共1,500 psig兲 in the flow
tester. No burns occurred with this alloy up to a test pressure of 3.20 MPa 共450
psig兲. At pressures of 3.55 MPa 共500 psig兲 and 5.27 MPa 共750 psig兲 the alloy
exhibited a mix of burns and no burns indicative of transitional behavior. The
exact location of the lower shelf where burns always occur is not certain and is
test method dependent but it appears to be above 6.9 MPa 共1,000 psig兲 in the
static tester. A limited number of test specimens in a suitable diameter pre-
cluded further investigation.

MP35N

Data in Table 2 indicates that no burns, in ten tests, occurred irrespective of test
chamber with MP35N at a pressure of 10.4 MPa 共1,500 psig兲. Above that pres-
sure a mix of burns and no burns occurred up to the highest test pressure of
34.5 MPa 共5,000 psig兲 in the static tester.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
MILLION ET AL., doi:10.1520/JAI102230 15

TABLE 2—Flammability data for cobalt alloys.

Pressure Elgiloy MP35N Stellite 6

psig MPa Flow Static Flow Static Flow Static


400 2.86 0B/5NB
450 3.20 0B/5NB

500 3.55 1B/4NB 0B/5NB


650 4.58
750 5.27 3B/2NB 0B/5NB

850 5.96
900 6.31
1,000 6.9 5B/0NB 0B/4NB 2B/3NB

1,200 8.38 1B/4NB


1,500 10.4 5B/0NB 1B/0NB 0B/5NB 0B/5NB 2B/1NB

2,000 13.8 3B/2NB


3,000 20.7 5B/0NB 2B/3NB
5,000 34.5 3B/2NB
Note: B—burn; NB—no burn; double lines denote threshold pressure above which burns
occurred or no tests were conducted.

Stellite 6
Stellite 6 data in Table 2 indicates that no burns occurred in tests conducted up
to 5.27 MPa 共750 psig兲 in the flow tester. Above that, at pressures up to 10.4
MPa 共1,500 psig兲, transitional behavior with a mix of burns and no burns oc-
curred in the flow tester.

Nickel Alloys
Test results for the nickel alloys are found in Table 3, which contains the data
for CW6M, Hastelloy B3, Hastelloy W, and RA 330.

CW6M
The CW6M alloy, a cast version of the original Hastelloy C, was the most
combustion-resistant alloy as reported on in this document. At pressures up to
13.8 MPa 共2,000 psig兲 the CW6M alloy exhibited no burns in the static tester. At
a pressure of 20.7 MPa 共3,000 psig兲 the alloy burned in all tests in the static
tester.

Hastelloy B-3
At pressures up to 5.27 MPa 共750 psig兲 Hastelloy B-3 experienced no burns in
the flow tester. Above this, tests conducted at pressures of 6.9 MPa 共1,000 psig兲
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
16 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 3—Flammability data summary for nickel alloys.

Pressure CW6M Hastelloy B-3 Hastelloy W RA 330

psig MPa Flow Static Flow Static Flow Static Flow


250 1.83 0B/3NB 0B/5NB
500 3.55 0B/5NB 0B/5NB 0B/3NB 0B/3NB 0B/5NB
625 4.41 0B/5NB

750 5.27 0B/5NB 0B/3NB 0B/3NB 3B/2NB

1,000 6.9 0B/5NB 1B/4NB 0B/5NB 0B/3NB 3B/2NB

1,250 8.72
1,500 10.4 0B/5NB 1B/4NB 0B/3NB 3B/1NB

2,000 13.8 0B/5NB 1B/4NB 1B/2NB

3,000 20.7 5B/0NB 3B/0NB 3B/0NB


5,000 34.5 3B/0NB 3B/0NB
Note: B—burn; NB—no burn; double lines denote threshold pressure above which burns
occurred or no tests were conducted.

up to 13.8 MPa 共2,000 psig兲 resulted in a mix of burns and no burns 共flow and
static testers兲. Test pressures at 20.7 MPa 共3,000 psig兲 and above resulted in
burns in all tests in the static tester.

Hastelloy W
Hastelloy W did not experience any burns per the test criteria at pressures up to
10.4 MPa 共1,500 psig兲 in the flow tester. At 13.8 MPa 共2,000 psig兲 a mix of burns
and no burns resulted in the static tester. Pressures at 20.7 MPa 共3,000 psig兲 and
above resulted in specimen burns in all tests in the static tester.

RA 330
RA 330 flammability tests were conducted using only the flow tester and re-
sulted in no burns per established criteria at pressures up to 4.41 MPa 共625
psig兲. A mix of burns and no burns occurred in tests conducted within the 5.27
MPa 共750 psig兲 to 10.4 MPa 共1,500 psig兲 range.

Alloy Ranking and Heat of Combustion


Engineering alloys with low heats of combustion due to high levels of copper,
nickel, and cobalt usually tend to have high resistance to combustion in
oxygen-enriched environments. However, as requirements for properties such
as oxidation resistance, corrosion resistance, and high strength are introduced,
increasing levels of elements such as chromium, molybdenum, aluminum, ti-
tanium, etc., may be required. This change in alloy chemistry will raise the fuel
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
MILLION ET AL., doi:10.1520/JAI102230 17

TABLE 4—Threshold pressure and heats of combustion.

Threshold Pressurea Heat of Combustion

Alloy MPa psig cal/g


CW6M 13.8 2,000 1,522
MP35N 10.4 1,500 1,432
Hastelloy W 10.4 1,500 1,358
Hastelloy B-3 6.9 750 1,197
Stellite 6 5.27 750 1,611
RA 330 4.41 625 1,652
Elgiloy 3.20 450 1,535
a
The threshold pressure is the highest pressure at which no burns occurred based on
both flow tester and static tester results. The test pressure limit for the flow tester is 10.4
MPa 共1,500 psig兲.

value that may affect combustion resistance. What follows in Table 4 ranks
alloys from least flammable to most flammable based on threshold pressure in
a promoted ignition-combustion test in 99.7 % oxygen. The corresponding
heats of combustion for individual alloys are shown.
In this alloy set, the correlation between heat of combustion and threshold
pressure is not very strong. However, it should also be stated that with the
exception of Hastelloys B-3 and W, which are Ni-Mo alloys, the metallurgy of
the seven alloys do not really have common characteristics besides belonging
either to cobalt or nickel alloy families.

Relationship of Data to Other Work


The CW6M data is of interest because of past work done on variants of the
Hastelloy C family such as Hastelloy C-4, C-22, and C-276 as well as casting
analogs such as CW2M and CX2MW. Essentially, the CW6M results fit well with
the past data on the previously cited alloys. See Refs 7 and 18.

Other Comments
It has generally been the conventional wisdom that nickel or cobalt alloys have
superior combustion resistance in oxygen-enriched atmospheres. The wide
range of performance of the cobalt and nickel alloys in the investigation dem-
onstrates the need to characterize even the cobalt and nickel alloys via pro-
moted ignition-combustion testing prior to service in oxygen-enriched atmo-
spheres. The relatively poor performance of Elgiloy needs further study.
Further work is being considered.

Summary
In this investigation the following major observations were made:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
18 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

共1兲 The correlation between heat of combustion and threshold pressure


was not very strong. From this, it was concluded that heats of combus-
tion are not suitable as predictors of combustion resistance.
共2兲 Cobalt and nickel alloys, even though they are often thought of as es-
pecially combustion resistant in oxygen-enriched atmosphere, need to
be characterized via tests such as the promoted ignition-combustion
test prior to service in oxygen-enriched atmospheres. Considerable
variation in promoted ignition-combustion resistance may occur be-
tween different cobalt and nickel alloys.
共3兲 The results of this study are in agreement with the binary alloy work
described in Ref 14.

Acknowledgments
The writers acknowledge the invaluable contributions of Joseph Wegrzyn,
Daniel Hamberger, and Ronald D’Agostino for the experimental work; also
Tonie Litmer in the preparation of this manuscript.

References

关1兴 Slusser, J. W. and Miller, K. A., “Selection of Metals for Gaseous Oxygen Service,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 812, B. L. Werley, Ed., ASTM International, West Conshohocken, PA, 1988,
pp. 167–191.
关2兴 Benz, F. J., Shaw, R. C., and Homa, J. M., “Burn Propagation Rates of Metals and
Alloys in Gaseous Oxygen,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Second Volume, ASTM STP 910, ASTM International, West
Conshohocken, PA, 1986, pp. 135–152.
关3兴 McIlroy, K., Zawierucha, R., and Drnevich, R. F., “Promoted Ignition Behavior of
Engineering Alloys in High-Pressure Oxygen,” Flammability and Sensitivity of Ma-
terials in Oxygen-Enriched Atmospheres: Third Volume, ASTM STP 986, D. W.
Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp. 85–104.
关4兴 McIlroy, K., and Zawierucha, R., “The Effects of Testing Methodology on the Pro-
moted Ignition-Combustion Behavior of Carbon Steel and 316L Stainless Steel in
Oxygen Gas Mixtures,” Flammability and Sensitivity of Materials in oxygen-
Enriched Atmospheres: Fourth Volume, ASTM STP 1040, J. M. Stoltzfus, F. J. Benz,
and J. S. Stradling, Eds., ASTM International, West Conshohocken, PA, 1989, pp.
38–53.
关5兴 Steinberg, T. A., Rucker, M. A., and Beeson, H. D., “Promoted Combustion of Nine
Structural Metals in High Pressure Gaseous Oxygen: A Comparison of Ranking
Methods,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
sphere: Fourth Volume, ASTM STP 1040, J. M. Stoltzfus, F. J. Benz, and S. S. Stra-
dling, Eds., ASTM International, West Conshohocken, PA, 1989, pp. 54–75.
关6兴 Zabrenski, J. S., Werley, B. L., and Slusser, J. W., “Pressurized Flammability Limits
of Metals,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Fourth Volume, ASTM STP 1040, J. M. Stoltzfus, F. J. Benz, and S. S.
Stradling, Eds., ASTM International, West Conshohocken, PA, 1989, pp. 178–194.
关7兴 Zawierucha, R., McIlroy, K., and Mazarella, R. B., “Promoted Ignition-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
MILLION ET AL., doi:10.1520/JAI102230 19

Combustion Behavior of Selected Hastelloys® in Oxygen Gas Mixtures,” Flamma-


bility and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Fifth Volume,
ASTM STP 1111, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West
Conshohocken, PA, 1991, pp. 270–287.
关8兴 Stoltzfus, J. M., Lowrie, R., and Gunaji, M. V. “Burn Propagation Behavior of Wire
Mesh Made from Several Alloys,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres: Fifth Volume, ASTM STP 1111, J. M. Stoltzfus and
K. McIlroy, Eds., ASTM International, West Conshohocken, PA, 1991, pp. 326–337.
关9兴 Bryan, C. J., Stoltzfus, J. M., and Gunaji, M. V., “An Assessment of the Flammabil-
ity Hazard of Several Corrosion Resistant Metal Alloys,” Flammability and Sensi-
tivity of Materials in Oxygen-Enriched Atmospheres: Sixth Volume, ASTM STP 1197,
D. D. Janoff and J. M. Stoltzfus, Eds., ASTM International, West Conshohocken,
PA, 1993, pp. 112–118.
关10兴 Steinberg, T. A., Wilson, D. B., and Benz, F. J., “Microgravity and Normal Gravity
Combustion of Metals and Alloys in High Pressure Oxygen,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Sixth Volume, ASTM STP
1197, D. D. Janoff and J. M. Stoltzfus, Eds., ASTM International, West Consho-
hocken, PA, 1993, pp. 133–145.
关11兴 Key, C. F., Lowery, F. S., Darby, S. P., and Libb, R. S., “Factors Affecting the
Reproducibility of Upward Propagation Pressure Thresholds of Metals in Gaseous
Oxygen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997, pp. 71–92.
关12兴 Janoff, D. and Pedley, M. D., “Configurational Effects on the Combustion of Sev-
eral Alloy Systems in Oxygen-Enriched Atmospheres,” Flammability and Sensitivity
of Materials in Oxygen-Enriched Atmospheres: Eighth Volume, ASTM STP 1319, W.
T. Royals, T. C. Chou, and T. A. Steinberg, Eds., ASTM International, West Con-
shohocken, PA, 1997, pp. 147–156.
关13兴 Steinberg, T. A. and Stoltzfus, J. M., “Combustion Testing of Metallic Materials
Aboard the NASA Johnson Space Center’s KC-135,” Flammability and Sensitivity of
Materials in Oxygen-Enriched Atmospheres: Eighth Volume, ASTM STP 1319, W. T.
Royals, T. C. Chou, and T. A. Steinberg, Eds., ASTM International, West Consho-
hocken, PA, 1997, pp. 170–188.
关14兴 Tayal, M., Wilson, D. B., and Stoltzfus, J. M., “Influence of Alloying Additions on
the Flammability of Nickel-Based Alloys in an Oxygen Environment,” Flammabil-
ity and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Eighth Volume,
ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Steinberg, Eds., ASTM Inter-
national, West Conshohocken, PA, 1997, pp. 189–202.
关15兴 Werley, B. L. and Hansel, J. G., “Flammability Limits of Stainless Steel Alloys 304,
308, and 316,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997, pp. 203–224.
关16兴 Zawierucha, R. and Million, J. F., “Promoted Ignition-Combustion Behavior of
Engineering Alloys at Elevated Temperatures and Pressures in Oxygen Gas Mix-
tures,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. Newton, and H. D. Beeson,
Eds., ASTM International, West Conshohocken, PA, 2000, pp. 119–133.
关17兴 Samant, A. V., Zawierucha, R., and Million, J. F., “Thickness Effects on the Pro-
moted Ignition-Combustion Behavior of Engineering Alloys,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Tenth Volume, ASTM STP
1454, T. A. Steinberg, H. D. Beeson, and B. E. Neton, Eds., ASTM International,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
20 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

West Conshohocken, PA, 2003.


关18兴 Zawierucha, R., Samant, A. V., and Million, J. F., “Promoted Ignition-combustion
Behavior of Cast and Wrought Engineering Alloys in Oxygen-Enriched Atmo-
spheres,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Tenth Volume, ASTM STP 1454, T. A. Steinberg, H. D. Beeson, and B. E.
Neton, Eds., ASTM International, West Conshohocken, PA, 2003.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 7
doi:10.1520/JAI102253
Available online at www.astm.org/JAI

David Lynn,1 Ted Steinberg,2 Kyle Sparks,3 and


Joel M. Stoltzfus4

Defining the Flammability of Cylindrical Metal


Rods Through Characterization of the
Thermal Effects of the Ignition Promoter

ABSTRACT: All relevant international standards for determining if a metallic


rod is flammable in oxygen utilize some form of “promoted ignition” test. In
this test, for a given pressure, an overwhelming ignition source is coupled to
the end of the test sample and the designation flammable or nonflammable is
based upon the amount burned, that is, a burn criteria. It is documented that
共1兲 the initial temperature of the test sample affects the burning of the test
sample both 共a兲 in regards to the pressure at which the sample will support
burning 共threshold pressure兲 and 共b兲 the rate at which the sample is melted
共regression rate of the melting interface兲; and, 共2兲 the igniter used affects the
test sample by heating it adjacent to the igniter as ignition occurs. Together,
these facts make it necessary to ensure, if a metallic material is to be con-
sidered flammable at the conditions tested, that the burn criteria will exclude
any region of the test sample that may have undergone preheating during
the ignition process. A two-dimensional theoretical model was developed to
describe the transient heat transfer occurring and resultant temperatures
produced within this system. Several metals 共copper, aluminum, iron, and
stainless steel兲 and ignition promoters 共magnesium, aluminum, and
Pyrofuze®兲 were evaluated for a range of oxygen pressures between 0.69
MPa 共100 psia兲 and 34.5 MPa 共5,000 psia兲. A MATLAB® program was uti-

Manuscript received November 28, 2008; accepted for publication May 19, 2009; pub-
lished online July 2009.
1
Ph.D. Candidate, School of Engineering Systems, Queensland Univ. of Technology,
GPO Box 2434, Brisbane, Queensland 4001, Australia.
2
Professor, School of Engineering Systems, Queensland Univ. of Technology, GPO Box
2434, Brisbane, Queensland 4001, Australia.
3
NASA Test and Evaluation Contract, NASA White Sands Test Facility, Las Cruces, NM
88004.
4
Special Program Manager, NASA White Sands Test Facility, Las Cruces, NM 88004.
Cite as: Lynn, D., Steinberg, T., Sparks, K. and Stoltzfus, J. M., ⬙Defining the
Flammability of Cylindrical Metal Rods Through Characterization of the Thermal
Effects of the Ignition Promoter,⬙ J. ASTM Intl., Vol. 6, No. 7. doi:10.1520/JAI102253.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 21
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
22 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

lized to solve the developed model that was validated against 共1兲 a published
solution for a similar system and 共2兲 against experimental data obtained
during actual tests at the National Aeronautics and Space Administration
White Sands Test Facility. The validated model successfully predicts tem-
peratures within the test samples with agreement between model and ex-
periment increasing as test pressure increases and/or distance from the pro-
moter increases. Oxygen pressure and test sample thermal diffusivity were
shown to have the largest effect on the results. In all cases evaluated, there
is no significant preheating 共above about 38°C/100°F兲 occurring at distances
greater than 30 mm 共1.18 in.兲 during the time the ignition source is attached
to the test sample. This validates a distance of 30 mm 共1.18 in.兲 above the
ignition promoter as a burn length upon which a definition of flammable can
be based for inclusion in relevant international standards 共that is, burning
past this length will always be independent of the ignition event for the igni-
tion promoters considered here兲.
KEYWORDS: promoted ignition, metal combustion, heat conduction,
thin fin, promoted combustion, burn length, burn criteria, flammability,
igniter effects, heat affected zone

Nomenclature

Bia ⫽ lateral Biot number Bia = hRⴱ / k


BiT ⫽ tip Biot number BiT = hTL / k
G ⫽ geometry factor 共=L / Rⴱ兲
H ⫽ ratio of convective heat transfer coefficients 共=hT / h兲
h ⫽ convective heat transfer coefficient-lateral surface
hT ⫽ convective heat transfer coefficient tip
J0 ⫽ zero-order Bessel function
J1 ⫽ first-order Bessel function
k ⫽ thermal conductivity of test sample
L ⫽ cylindrical rod length
R ⫽ dimensionless rod radius 共=Rⴱ / L兲
Rⴱ ⫽ radius of cylindrical rod
Tⴱ共xⴱ , rⴱ , tⴱ兲 ⫽ transient temperature
T共x , r , t兲 ⫽ dimensionless transient temperature 关=共Tⴱ共xⴱ , rⴱ , tⴱ兲 − Tⴱ⬁兲 / 共Tⴱm
− Tⴱ⬁兲兴
Tⴱm ⫽ material melting temperature
Tⴱ⬁ ⫽ ambient chamber temperature
t ⫽ dimensionless time 共=␣tⴱ / L2兲
tⴱ ⫽ time
x,r ⫽ dimensionless coordinates 共x = xⴱ / L, r = rⴱ / L兲
x , rⴱ

⫽ coordinate system 共x is along rod and r is radial兲
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 23

␣ ⫽ thermal diffusivity of sample


␩n , ␤m ⫽ nth and mth positive roots of transcendental equations,
respectively

Introduction

A number of test methods are currently used to evaluate the flammability of


metallic materials, including those developed by the National Aeronautics and
Space Administration 共NASA兲 关1兴, ASTM 关2兴, and International Organization for
Standardization 共ISO兲 关3兴. In each of these tests, a cylindrical rod of length, L,
and radius, Rⴱ, is vertically mounted in a test chamber and ignited at the bot-
tom end, typically through the use of an ignition promoter that is resistively
heated to initiate burning. The promoter heats the end of the test sample to its
melting point then detaches 共melts off兲 from the sample. The material is con-
sidered flammable, under the specific environmental conditions being evalu-
ated, if the burning continues after the effects of the igniter have dissipated.
This burning independent of the igniter has historically been indicated by the
sample burning more than some predefined length 共that is, burning continues
along the test sample for more than some arbitrary length兲.
The particular method of ignition used, however, means that during the
ignition process energy is transferred to the test material, resulting in some
preheating of the sample. Past work 关4,5兴 has demonstrated that an increase in
test sample temperature can potentially affect the flammability of a test sample
and the rate at which the sample is consumed, that is, raising the initial tem-
perature of the sample can lead to the sample burning at a lower test pressure
and being consumed 共burning兲 faster. Additionally, the material properties of
the test sample being evaluated, as well as the different characteristics of the
various promoter types typically used, will influence the results obtained 共that
is, the extent of the test sample that is preheated by the igniter兲. Since these
standard tests are commonly used to relatively rank different metallic materi-
als, dissimilar thermal diffusivities of the test material 共for a similar igniter兲
will lead to differences in the extent of the preheated zone and could, therefore,
potentially change the ranking of one metallic material relative to another.
These reasons make it necessary to ensure that any definition of flammability
that is used based upon a test sample’s burn length preclude any region that is
preheated by the promoter during the ignition process.
The work presented here outlines the development of a theoretical model
and numerical simulation using MATLAB® to predict the temperature distribu-
tion in a test sample as a result of preheating by the igniter. Based on this
model, a prediction of the transient heat transfer that occurs during the igni-
tion process is made. The objective of the current work is to validate the devel-
oped model and then characterize the resulting temperature distribution pro-
duced in the test sample from the ignition promoter. Characterization of this
resultant temperature distribution is then used to provide a definition of when
a sample can be considered flammable 共i.e., it is burning independent of the
igniter used兲 for incorporation into relevant international standards.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
24 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Background

According to all the relevant international standards in this area, for a test
sample to be considered flammable, under the specific environmental condi-
tions being evaluated, the burning must be “self-sustaining.” Currently, the
NASA test specifies this is true if one 共of ten兲 specimen burns more than 30 mm
共1.18 in.兲 关for a sample that is at least 102 mm 共4 in.兲 pretest length兴, the ISO
test specifies this is true if one 共of ten兲 specimen burns more than halfway 共150
mm for a sample that has 305 mm pretest length兲 and the ASTM test specifies
this is true if complete consumption of the sample 共up to the sample holder兲
occurs in at least one 共of five兲 test 共for samples that are ⬃150 mm pretest
length兲. These burn lengths were arbitrarily chosen and a goal of the current
work is to better characterize the actual length a test sample needs to be con-
sumed for the burning to be considered self-sustaining 共and, therefore, the
sample to be considered flammable by a standard test兲. The ASTM test stan-
dard is, as a matter of course, currently being reviewed and updated, with some
focus on clarifying the true length a test sample must burn for it to be consid-
ered flammable; that is, burning independently of the ignition event.
The general burning characteristics of metals subjected to promoted igni-
tion testing have been the focus of numerous studies. Past work has typically
used two parameters, the threshold pressure 共TP兲 and regression rate of the
melting interface 共RRMI兲 to characterize a metallic material’s propensity to
support burning under a given set of experimental conditions. The TP is the
lowest pressure 共of those tested兲 that will support burning of a sample and the
RRMI is the rate at which the melting interface 共surface between the liquid
molten ball formed on the end of the rod during burning and the solid metal
rod兲 moves along the rod while burning. These parameters provide an indica-
tion of a metallic material’s absolute and relative flammability making compari-
sons between different metals possible.
In particular, the heat affected zone 共HAZ兲 and other thermal effects pro-
duced by the igniter have received some, albeit limited, attention. Figure 1
shows a typical test sample, posttest, and how conductive heating affects the
sample internally. Previous work has shown that metallic material flammability
is directly affected by the temperature of the sample being tested 关6–13兴. One
limited study has suggested that no significant preheating effects are present
further than ⬃20– 30 mm from the base of the sample 关see Fig. 2共b兲兴 关14兴.
There are, however, limited experimental or theoretical results available to in-
dicate the exact effect of temperature on metallic material flammability 关8–13兴.
The limited results available use TP and RRMI as indicators of the effects of
temperature on burning. This past work shows that, in general, an increase in
the temperature of the test sample results in an increase in the RRMI 关8,10–13兴
and a decrease in the TP 关8,9,13兴. Though these are the general trends reported
on in these publications, it is very important to note that these changes in TP
and RRMI were not 共given the limited data presented兲 observed for relatively
minor increases in sample temperature 关e.g., 100– 250° C 共210– 480° F兲兴, rather
they were observed for major increases in the test samples ambient tempera-
ture 关often temperature values much greater than 250– 300° C 共480– 570° F兲兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 25

Unaffected Region -
Long, striated grains consistent
with wire-drawing process during
manufacture.
(horizontal lines are scratches)

Heat Affected Zone (HAZ) -


Grains become progressively
larger towards the hotter
region of the sample.

Melting Region -
Grains become stretched-out as the
sample liquefies or re-solidifies.

Oxide Inclusions

Melted and Re-Solidified Region

Attached Oxide

FIG. 1—Cross section of a typical 3.2-mm 共0.125 in.兲 diameter sample, post test 关23兴.

Theoretical Description
The transient heat transfer occurring in a cylindrical-pin fin is of interest in a
number of applications. Chapman, cited in Ref. 关15兴, initially examined the
transient heat transfer in annular fins. Yang 关16兴 and Aziz 关17兴 then examined
the heat transfer occurring in straight and annular fins, respectively, with both
fins exposed to a periodic variation in base temperature. Chu et al. 关15兴 per-
formed a two-dimensional analysis of cylindrical-pin fins and these studies en-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
26 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Additional Thermocouples
TC-8: 93.98mm
TC-9: 144.78mm
2R* TC-10: 195.58mm

TC-7: 43.94mm
Test Sample

TC-6: 31.75mm

TC-5: 26.16mm
TC-4: 21.08mm Sample Rod
TC-3: 16.26mm Peened Material
L
TC-2: 11.68mm
Thermocouple
TC-1: 7.62mm
High Temperature
Cement
Dual Bore
r* Ceramic Tubing
0.00
(0,0)
Promoter

(a) (b) (c)

FIG. 2—Schematic of sample configuration used in 共a兲 model coordinate system, 共b兲
experimental configuration, and 共c兲 thermocouple attachment.

abled Su and Hwang 关18,19兴 to extend the work to a two-dimensional model to


account for convective effects from the pin sides and tip. Finally, Chang et al.
关20兴 modified the model to allow for nonconstant spatial and time boundary
conditions occurring at the base of the pin fin.
The analytical solutions produced by Su and Hwang and Chang et al. are
slightly extended and used in the current work. The current work is solved by
the combined method of separation of variables and principle of superposition.
The resulting model produced describes the temperature distribution, as well
as heat transfer along the test sample, as determined by a number of physically
relevant boundary conditions and appropriate initial condition. These condi-
tions and the final equations for this model are presented and discussed in
more detail, along with the adaptation of the model to the system being studied
in this work. Figure 2 shows a schematic of the configuration of the cylindrical
rod, as is typically used in the various standardized tests and as is used in the
theoretical model presented here.
Assuming cylindrical symmetry, the equations describing the temperature
distribution and heat transfer within the test system as well as the initial and
boundary conditions for a constant base temperature are:
Two-dimensional heat transfer equation

⳵ T共t,x,r兲 ⳵2T共t,x,r兲 ⳵2T共t,x,r兲 1 ⳵ T共t,x,r兲


= + + 共1兲
⳵t ⳵ x2 ⳵ r2 r ⳵r
Initial condition: t = 0

For all 共x,r兲 T共0,x,r兲 = 0 共2兲


Boundary conditions: t ⬎ 0

x=0
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
T共t,0,r兲 = 1 共3兲
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 27

⳵ T共t,1,r兲
x=1 + BiTT共t,1,r兲 = 0 共4兲
⳵x

⳵ T共t,x,0兲
r=0 =0 共5兲
⳵r

⳵ T共t,x,R兲
r=R + GBiaT共t,x,R兲 = 0 共6兲
⳵r
Eqs 1–6 are solved by standard techniques to produce the solution for the time
independent 共base兲 boundary condition, as shown in Eq 7. This is used to de-
termine the temperature distribution found in the cylindrical test specimen at a
specific time and location, 共t, r, and x兲 for specific material properties 共␣兲 and
convective properties 共h and hT兲.

T共t,x,r兲 = 兺

共 ␩
2Bi J 共␩ Gr兲
a 0 n

n Bia 兲J0共␩n兲
2
+ 2 再 ␩nG cosh关␩nG共1 − x兲兴 + BiT sinh关␩nG共1 − x兲兴
␩nG cosh共␩nG兲 + BiT sinh共␩nG兲


n=1
⬁ 2 2 2兲t
2␤2me−共␤m+␩nG sin共␤ x兲
− 兺
m=1 共 ␤ 2
m + ␩ 2 2
n G 兲关 ␤ m − cos 共 ␤ m 兲
m
sin共␤m兲兴
共7兲

The above solution is conditional on satisfying the following two transcenden-


tal functions:
␩nJ1共␩n兲 − BiaJ0共␩n兲 = 0 共8兲

␤m cot共␤m兲 = − BiT 共9兲


It is only the preheating of the sample rod during the ignition stages of a test
that is of relevance in this work, thus making the time an ignition promoter is
burning an important parameter in the model. It is during this critical “igniter
burn time” 共tcrit兲, that the bottom of the sample rod undergoes a rapid rise in
temperature from ambient to melting temperature 共corresponding to the spe-
cific test material being evaluated兲. Promoted ignition tests have demonstrated
that, depending upon the promoter used 共aluminum, magnesium, or Pyrofuze®
igniter wire兲 and the gas pressure of the test environment, tcrit can vary from
between approximately 0.1 s up to about 0.6 s 关21兴. During this initial tcrit, the
base temperature of the test sample is modeled to vary exponentially with time
between the ambient temperature and the material melting temperature, and is
approximated by setting the dimensionless base temperature boundary condi-
tion 共Eq 3兲 to be equal to
x=0 T共t,0,r兲 = 1 − e−bt 共3‘兲
where, b is a suitable time constant such that the base reaches roughly the
sample’s melting temperature over the time period, tcrit.
Using Duhamel’s method 共see Ref. 关22兴兲, the solution for a time dependent
共base兲 boundary condition can be expressed as in Eq 10, in agreement with the
work of Chang et al. 关20兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
28 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

T共t,x,r兲 = T共0,0,r兲␾共t,x,r兲 + 冕 t

␶=0
dT共␶,0,r兲
d␶
␾共t − ␶,x,r兲d␶ 共10兲

where ␾共t , x , r兲 represents the time independent solution, as found in Eq 7, and


the time dependent functions are evaluated at a fixed time t = ␶.
Applying the exponential boundary condition 共Eq 3⬘兲 and substituting Eq 7
into Eq 10, yields

冋 册再

2BiaJ0共␩nGr兲
T共t,x,r兲 = 兺 共␩2n + Bi2a兲J0共␩n兲
共1 − e−bt兲

再 冎
n=1

␩nG cosh关␩nG共1 − x兲兴 + BiT sinh关␩nG共1 − x兲兴



␩nG cosh共␩nG兲 + BiT sinh共␩nG兲

兺再 冎

2␤2msin共␤mx兲

共␤2m + ␩2nG2兲关␤m − cos共␤m兲sin共␤m兲兴

冎冎
m=1

⫻ 再 b− 共␤2m
b
2 2 关e
+ ␩ nG 兲
−共␤2m+␩2nG2兲t
− e−bt兴 共11兲

Eq 11 specifies the temperature distribution in the cylindrical test specimen for


a given location, at a given time for specific material properties 共␣兲 and convec-
tive properties 共h and hT兲. From this system of equations, appropriate boundary
and initial conditions, the temperature at a given time and point within or on
the cylindrical test piece can be specified 共and compared to experimental mea-
surements兲.

Model Solution
A model has been developed to describe the transient heat transfer along a
cylindrical metallic “pin fin” caused by promoted ignition at the beginning of
an experiment. This model is used to provide the temperature distribution
along the rod during the ignition process prior to the test sample burning in an
ongoing fashion.
From Eq 11, it can be seen that to solve for a given temperature, two
infinite sums need to be determined. As a result, a convergence criteria needs to
be specified. To replicate the process used by Su and Hwang 关18兴, the summing
process was terminated when the absolute value of the last term in the series
was less than 1 ⫻ 10−6.
The presence of the hyperbolic cosines in the equations produced excep-
tionally large numbers 共Ⰷ1.78⫻ 10308兲, which necessitated the use of the Sym-
bolic Math Toolbox™, since normal MATLAB® floating point precision can
handle only numbers of absolute magnitude less than approximately 1.78
⫻ 10308. This Symbolic Math Toolbox™ allows the use of variable precision
accuracy, in turn allowing the developed set of equations to be readily solved.
Using the developed model, and the thermophysical parameters of a given
test material and environmental conditions present, the system allows the pre-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 29

FIG. 3—Solution of developed model by 共a兲 this work—MATLAB®, and 共b兲 Su and
Hwang 关18兴.

diction of the temperature distribution and transient heat transfer characteris-


tics within the test sample. Figure 3 shows the MATLAB® solution of the devel-
oped equations, compared to the results of Su and Hwang 关18,19兴. As Fig. 3
shows, the surface temperature distribution produced by the MATLAB® model
in the present work is in excellent agreement with the previously published
work, confirming the validity of the program and solution approach allowing it
to be applied to the rod ignition phenomena being studied here.

Experimental System
To further validate the model, experimental work was conducted to allow com-
parison of model predictions to actual temperatures recorded along a sample
rod while ignition occurs at the bottom of the sample. The experimental work
was conducted at NASA White Sands Test Facility 共WSTF兲 关14兴. Temperature
measurements were obtained from a number of thermocouples placed at vary-
ing intervals along the rod. Figure 2共b兲 shows a typical experimental thermo-
couple configuration utilized.
Two sets of experimental data are shown in Fig. 4 describing the tempera-
ture at each of the thermocouples 关shown in Fig. 2共b兲兴, for a copper rod with a
magnesium and aluminum promoter in 34.5 MPa 共5,000 psia兲 and 3.45 MPa
共500 psia兲 oxygen. For the higher pressure case 共34.5 MPa/5,000 psia兲, there is
little heating of the test sample evidenced above about 10 mm 共0.4 in.兲 from the
top of the promoter. However, for the lower pressure case 共3.45 MPa or 500
psia兲 some preheating of the sample occurs, though significant preheating ef-
fects 关heating to ⬎150° C 共300° F兲兴 are only recorded in the 12–17 mm 共0.6–0.8
in.兲 closest to the top of the promoter. Above this position 共12–17 mm, 0.6–0.8
in.兲, there is negligible heating of the test sample from the ignition event.

Results and Discussion


The experimental work included three metal test samples 共copper, Monel 400
and 316 stainless steel兲, three igniter materials 共Pyrofuze®, aluminum and mag-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
30 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Experimental versus model temperatures for a copper sample at 共a兲 34.5 MPa
共5,000 psia兲 with magnesium promoter and 共b兲 3.45 MPa 共500 psia兲 with aluminum
promoter.

nesium though most testing was only performed with aluminum and magne-
sium promoters兲, and four test pressures 共0.69 MPa/100 psia, 3.45 MPa/500
psia, 6.9 MPa/1,000 psia, and 34.5 MPa/5,000 psia兲. These test materials provide
a good range of thermal diffusivities and the ignition promoters provide a good
range of tcrit’s, thus allowing a robust validation of the model’s predictions.
Table 1 provides the thermophysical parameters used for each material in the
model to generate the theoretical results.
The ignition promoter used and the gas pressure of the test environment
have been observed to affect the length of the ignition period, tcrit 关4,14兴. A table
of typical tcrit values as a function of these parameters, used in the solution of
the theoretical model to compare to the experimental measurements, is pro-
vided in Table 2. These values are obtained through experimental observations.
As expected, as the pressure increases, the tcrit value decreases. Also, due to its
low mass, when using just the Pyrofuze® igniter wire, a relatively faster burn
time at a given pressure than either the aluminum or magnesium igniters is
observed.
Figure 4 shows a comparison of model predictions to experimental results
for a 关3.2-mm 共0.125 in.兲 diameter兴 copper test sample in 共a兲 34.5 MPa/5,000
psia oxygen with a magnesium promoter and 共b兲 3.45 MPa/500 psia oxygen
with an aluminum promoter, respectively. There is little radial variation in tem-
perature within the rod and the figure shown is for the predicted surface tem-
perature on the rod 共at r = R兲. Also indicated on Fig. 4 is the time the promoter
burned, 共for the pressure the test was conducted at兲 tcrit, as given in Table 2. As
can be seen from these figures, except for the thermocouple immediately adja-
cent to the ignition promoter, there is good agreement between the experimen-
tal results and the model predictions. Additionally, agreement between the
model and the experimental results clearly increases with distance from the
promoter. In all cases evaluated here, excellent agreement is obtained between
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Summary of thermophysical properties useda 关24,25兴.

Material/Chamber Property Cu Al Fe 304 SS

Downloaded/printed by
Thermal conductivity 共k兲, W / m · K, 398 247 80.4 13.9
共Btu· ft/ h · ft2 · ° F兲 共228兲 共143兲 共46.2兲 共8.0兲
Melting point共Tm兲, °C, 共°F兲 1,083 657 1,538 1,398
共1,981兲 共1,215兲 共2,800兲 共2,550兲
Density 共␳兲, kg/ m3, 共lb/ in3兲 8,930 2,699 7,870 7,700
共0.323兲 共0.098兲 共0.284兲 共0.278兲
Specific heat 共Cp兲 J / kg· K, 共Btu/ lb· ° F兲 385 900 447 377
共0.092兲 共0.215兲 共0.107兲 共0.09兲
Thermal diffusivity 共␣ = k / Cp␳兲, m2 / s, 共ft2 / s兲 1.16⫻ 10−4 1.02⫻ 10−4 2.29⫻ 10−5 4.79⫻ 10−6
共1.23⫻ 10−3兲 共1.09⫻ 10−3兲 共2.44⫻ 10−4兲 共5.14⫻ 10−5兲
Convective coefficient—lateral 共h兲, W / m2 · K,
共Btu/ h · ft2 · ° F兲 3.5 共0.616兲

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Convective coefficient—tip 共hT兲, W / m2 · K,
共Btu/ h · ft2 · ° F兲 3.5 共0.616兲
Sample length 共L兲, m, 共in.兲 0.3175 共12.5⬙兲
Sample radius 共s兲, m, 共in.兲 0.0016 共0.063⬙兲
a
The values used in Table 1 are approximated averages only; though they will vary with the changing temperature and pressure

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


experienced by the sample, this variation is small and the average value shown in this table is the value used.
LYNN ET AL., doi:10.1520/JAI102253 31
32 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 2—Summary of tcrit values used in model as a function of promoter material and
oxygen pressure.

tcrit 共s兲 for Igniter for Different


Oxygen Pressures—MPa 共psia兲

0.69 3.5 6.9 35.0


Promoter Material 共100兲 共500兲 共1,000兲 共5,000兲
Magnesium 0.7 0.55 0.37 0.14
Aluminum 0.8 0.57 0.37 0.14
Pyrofuze® ¯ 0.20 ¯ ¯

model and experiment for distances greater than approximately 14 mm 共0.6 in.兲
from the top of the promoter.
The poor agreement between the model predictions and experimental mea-
surements for the thermocouple closest to the promoter, and to a lesser extent
the other thermocouples, is likely due to a number of real effects that are not
incorporated in the theoretical model. These effects include 共in perceived order
of importance兲:
共1兲 The use of constant convective heat transfer coefficients 共h and hT兲
along/on the end of the test sample. Clearly evidenced from the video
recordings of the results is the fact that during the ignition of the pro-
moter, large convective mixing is induced; it is not uniform over the
length of the test sample and this mixing is most pronounced near the
bottom of the sample 共where the promoter is located兲.
共2兲 The use of constant thermophysical parameters in the model 共that is,
the use of average material properties not changing with temperature
or pressure兲. This would include the conductivities, densities, and spe-
cific heats 共and hence, thermal diffusivities兲 for the metals.
共3兲 Nonuniform burning of the promoters. When a promoter is ignited it is
assumed it is burning uniformly. The model averages this effect out by
considering the heating of the promoter region from ambient to the
rod’s melting point to occur over the time tcrit.
共4兲 Not accounting for the physical interaction of the burning promoter
and end of test sample that is being melted.
共5兲 Noting that the thermocouples are actually just below the surface of
the test sample at different depths 共but are compared to model predic-
tions for the surface兲.
Because of these effects, agreement between the model and the experimen-
tal measurements in the region immediately adjacent 共above兲 the promoter
would be expected to be 共and is兲 limited. However, model predictions and ex-
perimental measurements are in good agreement and consistent above a small
region 共⬃14 mm兲 just adjacent to the promoter and it is here that the model
can be applied and is useful in determining what region of the test sample is
within the HAZ while the promoter is still attached. Clarification of the extent
of this HAZ will assist in defining when a sample is flammable 共that is, ongoing
burning independent of the ignition event兲.
These results validate the predictions of the thermal model developed, es-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 33

FIG. 5—Model temperatures predicted at 共a兲 15 mm 共0.59 in.兲 and 共b兲 30 mm 共1.18 in.兲
above promoter for 3.2-mm 共0.125 in.兲 diameter copper, stainless steel, iron, and alu-
minum test samples in 0.69 MPa 共100 psia兲 oxygen 共tcrit = 0.75 s兲.

pecially at distances greater than 14 mm 共0.6 in.兲 above the promoter. Further
presentation and discussion of the experimental results obtained during this
program and their meaning is provided in the publication by Sparks et al. 关14兴.
The model can be used to predict 共for many different sample materials, test
pressures, and igniter types examined here兲 the approximate extent of the HAZ
produced while the promoter is melting the end of the test sample. The extent
of this HAZ can then be used 共coupled with a factor-of-safety if required兲 to
develop a definition of flammability that is independent of the ignition event
and any associated heating of the test sample. That is, if a test sample burns a
distance greater than the extent of the HAZ produced by the promoter, it can be
considered to be burning independently of the ignition event and, conversely, a
test sample that burns a distance less than this value cannot be considered as
flammable.
As stated, the model can be used to study the effects of various sample and
environmental properties on the extent of the HAZ produced during ignition of
the test samples. Of particular interest here is how far up the rod the preheating
occurs and the extent of which the HAZ changes with the material thermal
diffusivity. Figures 5–7 show, for a test sample in oxygen at 0.69 MPa 共100 psia兲,
3.45 MPa 共500 psia兲, and 34.5 MPa 共5,000 psia兲, a comparison of the tempera-
ture within the test sample at 15 mm 共0.59 in.兲 and 30 mm 共1.18 in.兲 above the
promoter, respectively, for the metallic materials given in Table 1. These figures
provide a good range of thermal diffusivities typically encountered and confirm
that since few metallic materials have a higher thermal diffusivity than copper
共and none significantly so兲, estimates for the extent of the HAZ will be largest
when they are based on a copper test sample.
As expected, and clearly illustrated in Figs. 5–7: 共1兲 as test pressure in-
creases, the burning time of the promoter is shorter 共tcrit兲 and the preheating
effects within the test sample decrease, that is, there is a smaller HAZ created
within a test sample at higher pressures; 共2兲 as the thermal diffusivity of the
metal sample increases, the extent of the HAZ increases; and 共3兲 if general
estimates are to be made regarding the extent and magnitude of the HAZ, a low
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
34 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Model temperatures predicted at 共a兲 15 mm 共0.59 in.兲 and 共b兲 30 mm 共1.18 in.兲
above promoter for 3.2-mm 共0.125 in.兲 diameter copper, stainless steel, iron, and alu-
minum test samples in 3.45-MPa 共500 psia兲 oxygen 共tcrit = 0.56 s兲.

oxygen test pressure combined with a test material with a high thermal diffu-
sivity should be used.
Figure 7 shows that at 34.5 MPa 共5,000 psia兲, for all of the materials con-
sidered here, there are no heating effects within the test sample from the igni-
tion above about 15 mm 共⬃0.59 in.兲 from the top of the promoter. Figure 6
shows that at 3.45 MPa 共500 psia兲, for all of the materials considered here, there
are no heating effects within the test sample from the ignition above about 30
mm 共1.18 in.兲 from the top of the promoter. However, for tests at 3.45 MPa 共500
psia兲 the sample rod was found to reach between 90– 150° C 共⬃200– 300° F兲 at
a distance of 15 mm 共0.59 in.兲 from the top of the promoter 共during the ignition
period兲, depending upon the thermal diffusivity of the sample rod. Figure 5共a兲
shows that at low pressures 共0.69 MPa/100 psia兲 at 15 mm 共0.59 in.兲 above the

FIG. 7—Model temperatures predicted at 共a兲 15 mm 共⬃0.59 in.兲 and 共b兲 30 mm 共1.18
in.兲 above promoter for 3.2-mm 共0.125 in.兲 diameter copper, stainless steel, iron, and
aluminum test samples in 34.5-MPa 共5,000 psia兲 oxygen 共tcrit = 0.14 s兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 35

top of the promoter, there is significant heating of the test sample occurring for
both the copper 共up to ⬃232° C / 450° F兲 and aluminum 共up to ⬃121° C / 250° F兲
samples while the promoter is burning 共up to tcrit = 0.75 s兲. However, Fig. 5共b兲
shows that at 30 mm 共1.18 in.兲 above the top of the promoter, there is no
significant heating 共⬍38° C / 100° F兲 of the test sample 关at those same low pres-
sures 共0.69 MPa/100 psia兲兴 occurring for any of the metals considered.
Using the results for copper 共high thermal diffusivity兲 at low pressure 共0.69
MPa/100 psia兲 provides the “worst-case” estimates for the extent of the HAZ
produced by an ignition promoter while attached to the end of the test sample.
These results clearly indicate that for all metal test samples considered 共copper,
aluminum, iron, stainless steel兲, for all promoters evaluated 共Pyrofuze®, mag-
nesium, aluminum兲, and for test pressures over 0.69 MPa 共100 psia兲, the HAZ
produced is restricted to a region within about 30 mm 共1.18 in.兲 from the top of
the promoter. That is, no heating above about 38° C 共100° F兲 will occur within
a test sample for any combination of these parameters while the promoter is
attached. As stated earlier, using the limited experimental data available, no
changes in either TP or RRMI are expected for a material preheated to this level
共or even twice this temperature兲. This implies any burning of a standard test
sample for more than 30 mm 共1.18 in.兲 is the result of self-sustained burning
independent of the ignition event for the promoter types investigated.

Conclusions

A two-dimensional theoretical model was developed to describe the transient


heat transfer occurring and resultant temperatures produced within a cylindri-
cal metal rod when an ignition promoter is ignited on the bottom of the rod in
gaseous oxygen. Several metals 共copper, aluminum, iron, and stainless steel兲
and promoters 共magnesium, aluminum, and Pyrofuze®兲 were evaluated for a
range of oxygen pressures between 0.69 MPa 共100 psia兲 and 34.5 MPa 共5,000
psia兲. A MATLAB® program was utilized to solve the developed model that was
validated against 共1兲 a published solution for a similar system and 共2兲 against
experimental data obtained during actual tests at NASA WSTF. The validated
model successfully predicts temperatures within the test samples with agree-
ment between model and experiment increasing as test pressure increases
and/or distance from the promoter increases. Oxygen pressure and test sample
thermal diffusivity were shown to have the largest effect on the results with
increases in sample thermal diffusivity or decreases in oxygen pressure produc-
ing a larger HAZ within the test sample. In all cases evaluated, there is no
significant preheating 共above about 38° C / 100° F兲 occurring at distances
greater than 30 mm 共1.18 in.兲. This validates a distance of 30 mm 共1.18 in.兲
above the ignition promoter as a burn length upon which a definition of flam-
mable can be based for inclusion in relevant international standards 共that is,
burning past this length will always be independent of the ignition event for the
ignition promoters considered here兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
36 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

References

关1兴 “National Aeronautics and Space Administration 共NASA兲, Interim NASA Technical
Standard NASA-STD-共I兲-6001A, “Flammability, Offgassing, and Compatibility Re-
quirements and Test Procedures for Materials in Environments that Support Com-
bustion,” National Aeronautics and Space Administration, Washington, D.C.,
2008.
关2兴 ASTM G124, “Standard Test Method for Determining the Combustion Behaviour
of Metallic Materials in Oxygen-Enriched Atmospheres,” Annual Book of ASTM
Standards, ASTM International, West Conshohocken, PA, 1995.
关3兴 International Organization for Standardization 共ISO兲, ISO 14624-4, Space
Systems–Safety and Compatibility of Materials—Part 4: Determination of Upward
Flammability of Materials in Pressurized Gaseous Oxygen or Oxygen-Enriched Envi-
ronments, International Organization for Standardization, Geneva, Switzerland,
2003.
关4兴 Steinberg, T. A., “Metals Combustion at High-Pressure Oxygen in Normal and
Reduced Gravity: Model and Experiment,” Ph.D. thesis, New Mexico State Univ.,
Las Cruces, NM, 1990.
关5兴 De Wit, J. R., Steinberg, T. A., and Stoltzfus, J. M., “Igniter Effects on Metals
Combustion Testing,” Flammability and Sensitivity of Materials in Oxygen-Enriched
Atmospheres: Ninth Volume, ASTM STP 1395, Steinberg, T. A., Newton, B. E., and
Beeson, H. D., Eds., ASTM International, West Conshohocken, PA, 2000, pp. 190–
203.
关6兴 Zaweirucha, R., and Million, J. F., “Promoted Ignition-Combustion Behaviour of
Engineering Alloys at Elevated Temperatures an Pressures in Oxygen Gas Mix-
tures,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Ninth Volume, ASTM STP 1395, Steinberg, T. A., Newton, B. E., and Beeson, H. D.,
Eds., ASTM International, West Conshohocken, PA, 2000, pp. 119–133.
关7兴 Sato, J. and Hirano, T., “Fire Spread Limits Along Metal Pieces in Oxygen,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Vol-
ume, ASTM STP 986, Schroll, D. W., Ed., ASTM International, West Consho-
hocken, PA, 1988, pp. 158–173.
关8兴 Engel, C. D., Herald, S., and Davis, E., “Promoted Metals Combustion at Ambient
and Elevated Temperatures,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Eleventh Volume, ASTM STP 1479, Hirsch, D. B., Zaw-
ierucha, R., Steinberg, T. A., and Barthelemy, H. M., Eds., ASTM International,
West Conshohocken, PA, 2006, pp. 51–61.
关9兴 Slockers, M. J. and Robles-Culbreth, R., “Ignition of Metals at High Temperatures
in Oxygen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eleventh Volume, STP 1479, Hirsch, D. B., Zaweirucha, R., Steinberg, T. A.,
and Barthelemy, H. M., Eds., ASTM International, West Conshohocken, PA, 2006,
pp. 62–79.
关10兴 Sato, J., “Fire Spread Rates along Cylindrical Metal Rods in High Pressure Oxy-
gen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Fourth Volume, ASTM STP 1040, Stoltzfus, J. M., Benz, F., and Stradling, J. S.,
Eds.,ASTM International, West Conshohocken, PA, 1989, pp. 162–177.
关11兴 Sato, J. and Hirano, T., “Behaviour of Fires Spreading Along High-Temperature
Mild Steel and Aluminum Cylinders in Oxygen,” Flammability and Sensitivity of
Materials in Oxygen-Enriched Atmospheres: Second Volume, ASTM STP 910, Ben-
ning, M. A., Ed., ASTM International, West Conshohocken, PA, 1986, pp. 118–134.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
LYNN ET AL., doi:10.1520/JAI102253 37

关12兴 Benz, F. J., Shaw, R. C., and Homa, J. M., “Burn Propagation Rates of Metals and
Alloys in Gaseous Oxygen,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Second Volume, ASTM STP 910, Benning, M. A., Ed., ASTM
International, West Conshohocken, PA, 1986, pp. 135–152.
关13兴 ASTM Committee G4 High Temperature Metals Combustion Test Program, per-
sonal communication, 2008.
关14兴 Sparks, K., Stoltzfus, J. M., Lynn, D. B., and Steinberg, T. A., “Determination of
Pass/Fail Criteria for Promoted Combustion Testing,” J. ASTM Int., Submitted for
publication, ASTM International, West Conshohocken, PA, 2009.
关15兴 Chu, H. S., Chen, C. K., and Weng, C. I., “Transient Response of Circular Pins,”
ASME J. Heat Transfer, Vol. 105, 1983, pp. 205–208.
关16兴 Yang, J. W., “Periodic Heat Transfer in Straight Fins,” ASME J. Heat Transfer, Vol.
94, 1972, pp. 310–314.
关17兴 Aziz, A., “Periodic Heat Transfer in Annular Fins,” ASME J. Heat Transfer, Vol. 97,
1975, pp. 302–303.
关18兴 Su, R. J. and Hwang, J. J., “Transient Analysis of Two-Dimensional Cylindrical Pin
Fin with Tip Convective Effects,” Heat Transfer Eng., Vol. 20, 1999, pp. 57–63.
关19兴 Su, R. J. and Hwang, J. J., “Analysis of Transient Heat Transfer in a Cylindrical Pin
Fin,” J. Thermophys. Heat Transfer, Vol. 12, 1998, pp. 281–283.
关20兴 Chang, W., Chen, U., and Chou, H., “Transient Analysis of Two-Dimensional Pin
Fins with Non-Constant Temperature,” JSME Int. J., Ser. A, Vol. 45, 2002, pp.
331–337.
关21兴 Steinberg, T. A., Wilson, D. B., and Benz, F. J., “Modeling of Al and Mg Igniters
Used in the Promoted Combustion of Metals and Alloys in High Pressure Oxygen,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Sixth
Volume, ASTM STP 1197, Janoff, D. D. and Stoltzfus, J. M., Eds., ASTM Interna-
tional, West Conshohocken, PA, 1993, pp. 183–195.
关22兴 Carslaw, H. S. and Jaeger, J. C., Conduction of Heat in Solids, 2nd ed., Oxford
University Press, London, 1959.
关23兴 National Aeronautics and Space Administration 共NASA兲, “Evaluation of the Zone
That Is Heat-Affected by Standard Promoters in Metals Testing,” 2006 共unpub-
lished兲.
关24兴 Davis, J. R., Metals Handbook, 2nd ed., ASM International, Materials Park, Ohio,
1998.
关25兴 Peckner, D. and Bernstein, I. M., Handbook of Stainless Steels, 2nd ed., McGraw-
Hill, New York, 1977.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 6
doi:10.1520/JAI102270
Available online at www.astm.org/JAI

Nicholas R. Ward1 and Theodore A. Steinberg1

A Proposed Qualitative Framework for


Heterogeneous Burning of Metallic Materials:
The “Melting Rate Triangle”

ABSTRACT: This paper presents a proposed qualitative framework to dis-


cuss the heterogeneous burning of metallic materials, through parameters
and factors that influence the melting rate of the solid metallic fuel 共either in
a standard test or in service兲. During burning, the melting rate is related to
the burning rate and is therefore an important parameter for describing and
understanding the burning process, especially since the melting rate is com-
monly recorded during standard flammability testing for metallic materials
and is incorporated into many relative flammability ranking schemes. How-
ever, while the factors that influence melting rate 共such as oxygen pressure
or specimen diameter兲 have been well characterized, there is a need for an
improved understanding of how these parameters interact as part of the
overall melting and burning of the system. Proposed here is the “melting rate
triangle,” which aims to provide this focus through a conceptual framework
for understanding how the melting rate 共of solid fuel兲 is determined and regu-
lated during heterogeneous burning. In this paper, the proposed conceptual
model is shown to be both 共a兲 consistent with known trends and previously
observed results, and 共b兲 capable of being expanded to incorporate new
data. Also shown are examples of how the melting rate triangle can improve
the interpretation of flammability test results. Slusser and Miller previously
published an “extended fire triangle” as a useful conceptual model of ignition
and the factors affecting ignition, providing industry with a framework for
discussion. In this paper it is shown that a melting rate triangle provides a
similar qualitative framework for burning, leading to an improved understand-
ing of the factors affecting fire propagation and extinguishment.

Manuscript received December 8, 2008; accepted for publication April 17, 2009; pub-
lished online May 2009.
1
School of Engineering Systems, Queensland University of Technology 共QUT兲, Brisbane,
Queensland 4001, Australia.
Cite as: Ward, N. R. and Steinberg, T. A., ⬙A Proposed Qualitative Framework for
Heterogeneous Burning of Metallic Materials: The “Melting Rate Triangle”,⬙ J. ASTM
Intl., Vol. 6, No. 6. doi:10.1520/JAI102270.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 38
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102270 39

KEYWORDS: heterogeneous burning, melting, heat transfer, rate-


limiting mechanism, model

Introduction

This paper presents a proposed conceptual framework for analysis and discus-
sion of the many parameters that affect the melting rate of a solid metallic
material during heterogeneous burning. The paper provides a description of
the proposed model, the melting rate triangle, and shows that it is consistent
with known trends in experimental data. Potential applications are also de-
scribed, which include the analysis of results from standardized flammability
tests and also the use of these test data in industrial scenarios.

Background

The regression rate of the melting interface 共RRMI兲 is defined as the velocity at
which the external 共visible兲 boundary of the solid/liquid interface 共SLI兲 pro-
ceeds along the test specimen during a standard promoted ignition test for
metal flammability. The RRMI is closely related to the melting rate of solid fuel
and is thus a highly relevant parameter relating to the heterogeneous burning
of metallic materials. RRMI is often used as a secondary indicator of relative
flammability, especially in distinguishing between metallic materials with the
same threshold pressure 关1兴. Therefore, RRMI is an important parameter
within metal flammability testing and oxygen system fire safety, and any further
insight into the factors affecting RRMI is relevant and valuable. In a significant
contribution to the understanding of metallic material performance in oxygen
service, Slusser and Miller 关2兴 defined the “extended fire triangle,” which de-
scribed the many factors affecting the propensity for ignition. The analysis was
qualitative, but provided a highly useful conceptual framework within which
existing and future knowledge could be incorporated. In a similar way, this
paper presents the melting rate triangle, which aims to provide a conceptual
framework to improve the understanding of the factors affecting RRMI.
RRMI is known to be dependent on many factors, including specimen
shape, size, composition, configuration, and orientation, oxygen pressure and
concentration, and many other parameters 关3–6兴. In a work presented concur-
rently by Ward and Steinberg 关7兴, it is shown that heat transfer across the SLI
is the mechanism that limits melting and burning. Further, the heat transfer
process is shown specifically to be limited by the available contact surface area
for heat transfer between the burning droplet and the solid metal rod.
In this paper, the process of heat transfer across the SLI is considered in a
broader context and it is compared with the well-known trends regarding
RRMI. Arguably, the two most characterized trends observed for burning cylin-
drical metallic rods are that as 共1兲 oxygen pressure increases or 共2兲 test speci-
men diameter decreases, or both, RRMI increases 关3–6,8,9兴. Significantly, varia-
tions in SLI shape and surface area do not account for either of these trends.
For cylindrical rods burning in an upward-burning configuration in normal
gravity, the SLI shape is independent of pressure and is generally assumed to
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
40 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

remain planar and perpendicular for all pressures. This means that RRMI can
vary with test pressure but without any alteration of the SLI shape 共or surface
area兲. Also, while the total SLI surface area clearly changes with rod diameter,
this does not account for the observed dependence of RRMI on diameter be-
cause 共for cylindrical rods in upward-burning configurations in normal gravity兲
the SLI surface area and rod cross-sectional area are equal for all rod sizes. In
this way, the dependency of RRMI on both pressure and diameter indicates
that, while heat transfer across the SLI surface area is rate limiting for RRMI,
there are other parameters that strongly affect melting and burning. The con-
ceptual framework for discussion of heterogeneous burning presented in this
paper is therefore intended to provide a holistic view of the system that is
consistent with both the rate-limiting mechanism and also the important influ-
ence 共on RRMI兲 of other factors including pressure and specimen diameter.

The Melting Rate Triangle

The melting rate triangle is a conceptual representation of the many factors


that affect RRMI during the heterogeneous burning of cylindrical metallic rods.
In constructing the melting rate triangle, self-sustained heterogeneous burning
is considered as a three-part process by which
1. conditions within the burning droplet such as extent of reaction, oxy-
gen availability, and heat of combustion 共and many other factors兲 de-
termine the heat flux input into the SLI, which represents the potential
of the burning droplet to cause melting of the solid. In the context of
the melting rate triangle, this process is referred to as heat source.
2. The heat flux from the burning droplet is transferred across the surface
area of the SLI, which determines the total heat transfer rate into the
solid metal. In the melting rate triangle, this is referred to as heat trans-
fer.
3. Heat transferred to the solid metal causes melting at a rate determined
by the cross-sectional area, material properties, and initial temperature
of the solid test specimen, which results in the observed RRMI. This
process is referred to as heat sink.
In this way, the RRMI can be influenced by parameters that relate to any of the
three processes, by, for example, 共1兲 factors that alter the heat flux that is pro-
duced within the burning droplet 共changes in the potential to cause melting兲,
共2兲 the surface area of the SLI, which limits the amount of heat flux that enters
the solid, or 共3兲 altered energy requirements for melting the solid. This is not
only consistent with heat transfer across the SLI being the rate-limiting mecha-
nism but also implies that any known 共or unknown兲 factors that influence other
aspects of the burning system may also affect RRMI. The melting rate triangle
incorporating these processes is presented in Fig. 1. The melting rate triangle
was specifically developed to represent burning iron and it likely applies gener-
ally for heterogeneously burning metallic materials.
The three sides of the melting rate triangle represent the three basic pro-
cesses of heterogeneous burning 共as defined previously兲 and the three corners
represent critical parameters 共circled兲. The critical parameters are the maxi-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102270 41

FIG. 1—The proposed melting rate triangle.

mum droplet temperature, TD, the surface area of the SLI, ASLI, and the RRMI,
␷RRMI. During self-sustained burning, the heat flow and the interpretation of the
model proceed in a clockwise direction. The critical parameters have been
highlighted as they each help relate the preceding process to the next. These
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
42 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

processes and critical parameters are described in more detail in the following
sections.

Heat Source
The heat source side of the melting rate triangle is concerned with all aspects
that relate to the physical and chemical processes of burning within the at-
tached liquid droplet and is therefore the most complex and poorly understood
of the three sides. This side represents the net rate of heat generation, which is
the difference between heat generation and heat loss terms. Heat generation is
estimated here as the product of fuel mass flow rate, extent of reaction, and
heat of combustion. While fuel mass flow rate is easily characterized 共it is
simply the product of RRMI, density, and rod cross-sectional area兲, the other
two terms are more difficult to estimate. The heat of combustion is obtained
experimentally or analytically 共once an assumption is made regarding reaction
chemistry兲 关10兴, and the extent of reaction is dependent on oxygen pressure,
concentration, adsorption/absorption, solubility, and transport 共and, likely,
other factors兲. Recent work by Suvorovs 关11兴 indicated that the extent of reac-
tion is also a function of test specimen diameter. The presence of alloying ele-
ments, contaminants, or catalysts is also known to affect the rate of heat gen-
eration. Heat loss occurs through bulk mass transport, especially dripping,
which is gravity dependent. Heat is also lost to the surroundings, especially to
the solid rod by conduction and convection, and this term is therefore depen-
dent on the RRMI. All of these competing effects determine the effective droplet
maximum temperature, TD, which is identified as a critical parameter.

Heat Transfer
The heat transfer side of the melting rate triangle is concerned with all aspects
that relate to transferring thermal energy from within the burning droplet
共where it is generated兲 across the SLI and into the solid rod 共where it is dissi-
pated兲. The key parameter TD is critical in this process because it is the princi-
pal factor that determines 共or drives兲 the heat flux that enters the SLI surface.
Heat flux input into the SLI is represented by the one-dimensional conduction
equation, incorporating an effective heat transfer coefficient, ␬eff. This is as-
sumed to capture the combined effects of both conduction and convection, as
described by Steinberg and Wilson 关12兴 and Wilson and Stoltzfus 关13兴. Heat flux
is then integrated over the entire SLI surface area to determine the total heat
transfer rate to the solid metal 共although the simplifying assumption that heat
flux is constant over the SLI is used兲. This highlights the importance of SLI
surface area because it limits the heat flux that can enter the solid rod. This side
of the melting rate triangle clearly shows that any factor that changes the three-
dimensional shape 共and hence, area兲 of the SLI can alter the RRMI, such as
sample orientation 共as shown by Sato et al. 关14兴兲, test specimen cross-sectional
shape 共circular, rectangular, triangular, etc., as shown by Suvorovs et al. 关15兴兲,
configuration 共rod, sheet, mesh, etc.兲, surface finish 共e.g., threaded兲, or gravity
level 关16–18兴. Transient phenomena such as precession in reduced gravity, char-
acterized by random helical motion of the spherical molten droplet, also alter
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102270 43

SLI surface area and RRMI in a similar way. The surface tension of the molten
material at the melting temperature, TM 共the temperature at the external
boundary of the SLI兲, is noted because any change in SLI shape is dependent
on surface tension forces 共since it is the surface tension forces that act at the
solid/liquid boundary to define or alter the SLI shape兲. In this way, heat transfer
is modeled by taking the droplet temperature, TD, obtained from the heat
source side of the melting rate triangle, using this to determine heat flux, and
then integrating over the surface area of the SLI, ASLI, which is the next key
parameter.

Heat Sink
The heat sink side of the melting rate triangle is concerned with the extent to
which melting occurs as a result of heat transfer to the solid rod. Of the three
sides, this is the most easily characterized because the material properties of
the solid are well known and heat transfer in a slender rod is almost entirely
one dimensional 关19兴. The total heat transfer rate is obtained from the product
of heat flux and SLI surface area. As shown in Fig. 1, this can be related to the
RRMI, ␷RRMI, by the rod cross-sectional area 共or diameter squared for a circular
cross section兲, material properties, and initial temperature. In some normal-
gravity cases, when the SLI is planar and perpendicular to the rod centerline,
the equation shown would simplify because ASLI = ␲␾2 / 4, which is why this
equation alone does not explicitly capture the dependency of RRMI on rod
diameter. The dependence of RRMI on diameter is discussed in further detail in
the following section. Importantly, the result is that the melting rate 共repre-
sented by RRMI兲 is obtained, which, as the interpretation of the model contin-
ues, is linked back to the heat source step in the next iteration. This highlights
the closed-loop nature of the model, whereby a change in one of the sides
eventually feeds back, which is consistent with the inherent stability or self-
control that many burning systems exhibit, for example, in that the RRMI often
remains almost constant throughout a test.

Assessing the Melting Rate Triangle


An important requirement of a conceptual model, as noted by Wilson and
Stoltzfus 关13兴, is the need for it be both consistent with existing experimental
data and capable of being expanded to incorporate new information. In this
section, the melting rate triangle is compared to well-known trends reported for
the heterogeneous burning of metallic rods. These include the dependence of
RRMI on the following: test material; oxygen pressure and concentration; grav-
ity level, precession 共in reduced gravity兲, and sample orientation 共except in mi-
crogravity兲; and sample cross-sectional shape, configuration, and initial speci-
men temperature.
• Test material—Clearly, changing material properties such as density and
heat of combustion will significantly affect all sides of the melting rate
triangle, especially heat source and heat sink, which would clearly result
in a change in RRMI. A change in density, for example, will alter the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
44 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

energy requirement for melting 共heat sink side兲 and the mass-flow rate
of fuel into the droplet 共for a given RRMI; heat source side兲. Altered heat
of combustion will affect the rate of energy generation 共heat source
side兲. In this way, the melting rate triangle demonstrates how altering
the material properties will affect the rates of energy generation and
dissipation, and how this is likely to influence RRMI.
• Oxygen pressure and concentration—The melting rate triangle captures
the effects of any change in oxygen availability in the heat source side,
as this affects the extent of reaction and the overall energy release rate.
For example, the melting rate triangle predicts that a reduction in oxy-
gen availability at the reaction zone will reduce the net rate of energy
generation, causing a decrease in droplet temperature, TD 共heat source
side兲. The ensuing reduction in heat flux from the droplet to the solid,
without any change in the heat sink energy requirements, causes the
RRMI to decrease, which is consistent with the observed trend. RRMI,
which sets the mass flow rate of fuel into the droplet, will tend to de-
crease until equilibrium is reached with a lower RRMI at which the
melting rate can be sustained. In the limiting case, if the oxygen supply
is completely removed, this rapidly decreases the energy generation rate
in the heat source side of the melting rate triangle and the RRMI will
quickly tend toward zero; that is, fire propagation will cease.
• Gravity level, precession 共in reduced gravity兲, and sample orientation 共ex-
cept in microgravity兲—These are all examples of changes in SLI surface
area altering the heat transfer rate into the metal rod, causing a change
in RRMI. In these cases, the melting rate triangle captures the effect on
RRMI in the heat transfer side. Each of these factors alters the interfa-
cial geometry between solid and liquid phases 共with no change in the
energy requirement for melting and minimal changes in the net rate of
energy generation in the droplet兲. This increases the available contact
surface area for heat transfer, which increases the total heat transfer
rate into the solid resulting in faster melting.
• Sample cross-sectional shape, configuration—Changes in sample cross-
sectional shape were shown by Suvorovs et al. 关15兴 to alter RRMI in a
statistically significant way. RRMI values were compared for test speci-
mens with circular, rectangular, and triangular cross-sectional shapes
but the same cross-sectional area. They reported differences in RRMI
values that correlated 共qualitatively兲 with the extent to which the SLI
shape was altered due to the molten material climbing the sharp corners
of the rectangular and triangular rods. The correlation between the
change in SLI shape 共and surface area兲 and RRMI was confirmed by
Ward and Steinberg 关7兴. In the context of the melting rate triangle, there
is no difference in the heat source side since the test conditions were
held constant and there was no effect on droplet size or surface area,
which means that conditions within the burning droplet were likely un-
affected by the change in rod cross-sectional shape. Further, there are no
differences in the energy requirements for melting in the heat sink side
since the cross-sectional area was kept constant despite the change in
shape. Therefore, by elimination, the melting rate triangle indicates that
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102270 45

the cause of the variation in RRMI is related to the heat transfer side.
Close inspection of the test imagery reveals small differences in SLI
shape that correlated with the observed change in RRMI. This clearly
illustrates an analysis process in which the melting rate triangle can be
used to identify and clarify observed results, and also shows how the
affects of geometric changes are modeled.
• Initial specimen temperature—The influence of specimen initial tempera-
ture on melting rate was demonstrated by Sato and Hirano 关20兴 and
Engel et al. 关21兴. Rods with higher initial temperatures were shown to
exhibit higher melting rates. This effect is captured in the heat sink side,
whereby elevated initial specimen temperature reduces the amount of
energy that is required to melt the solid material. So, for no changes in
heat source or heat transfer, the altered energy requirement in the heat
sink produces a change in RRMI.
These examples demonstrate that the model is consistent with existing data and
provides a rigorous method for interpreting experimental results. It also pro-
vides a conceptual framework for the incorporation and analysis of new infor-
mation as it becomes available, for example, in further understanding the effect
of test specimen diameter on RRMI. Recent work by Suvorovs 关11兴 showed that
as diameter increases, the molten droplet becomes flooded with unburnt metal,
which reduces the extent of reaction and has the double effect of 共a兲 reducing
the amount of energy released by burning and 共b兲 sinking energy into heating
liquid metal that never burns. A lower extent of reaction at higher diameter was
confirmed through the analysis of quenched specimens. Captured in the heat
source side of the melting rate triangle, the influx of liquid metal absorbs en-
ergy within the molten droplet and reduces the net amount of energy that is
available to cause continued melting of the solid. This reduces the maximum
droplet temperature, TD, heat flux to the solid and the RRMI. Also, although the
heat transfer surface area at the SLI increases 共for larger rod diameters兲, this is
offset by the increased requirement for melting. This means that there is no net
change in the heat transfer or heat sink sides of the melting rate triangle and
the dependency of RRMI on rod diameter is due to changes in the heat source
parameters alone.
The melting rate triangle can be used to investigate either quasisteady self-
sustained burning 共such as global pressure or temperature dependencies兲 or
transient phenomena. For example, as was discussed earlier, the melting rate
triangle may provide insight into the effects of large disturbances such as ex-
tinguishment by inert gas or immersion in water. These clearly affect the heat
source side by reducing the availability of oxygen or increasing heat loss to the
surroundings, or both, breaking the feedback cycle, and causing a critical in-
terruption of the melting process. In this way, the melting rate triangle provides
insight into the processes occurring during burning and, by extension, may also
contribute to further understanding the limits of burning—ignition and extin-
guishment. The melting rate triangle therefore provides a powerful tool for
organizing, coupling, presenting, and understanding the many competing fac-
tors that influence instantaneous RRMI during heterogeneous burning, thus
creating a framework for analysis and discussion.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
46 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Applications

The melting rate triangle concept may be extended to applications beyond stan-
dardized flammability testing of upward-burning cylindrical rods. This section
discusses how the work is relevant to the application of standard flammability
test results in an industrial context. Metal flammability is not an inherent ma-
terial property but is instead dependent on factors including specimen shape,
size, composition, configuration and orientation, oxygen pressure and concen-
tration, and many other parameters 关3–6兴. Metal flammability is therefore as-
sessed experimentally in standardized tests, such as ASTM G124, “Standard
Test Method for Determining the Combustion Behavior of Metallic Materials in
Oxygen-Enriched Atmospheres.” However, given the dependence of flammabil-
ity, of which RRMI is one indicator, on so many parameters, it is difficult to
confidently apply the results of standard tests in an industrial environment. For
example, in service applications under similar conditions to a standardized test
共oxygen concentration, pressure, etc.兲, while this may imply similar burning
characteristics on the heat source side of the melting rate triangle, differences
in geometry could result in significant changes in the heat transfer and heat
sink processes and, hence, melting rate.
For example, burning liquid may interact differently with threaded sur-
faces compared to the smooth surface of a standard test specimen, altering the
interfacial geometry and SLI for similar sized droplets and possibly affecting
melting rate through a change in the heat transfer side of the melting rate
triangle. Any surface feature that facilitates increased contact between liquid
and solid phases may increase the melting rate. The SLI geometry is also
clearly different depending on whether the liquid is hanging freely from an
exposed component or pooling in a cavity, which will also affect fire propaga-
tion. Local cross-sectional area, dependent on component size and global ge-
ometry, affects the energy requirements for melting, as shown in the heat sink
side of the melting rate triangle. For example, a burning droplet may extinguish
if it lands on the flat surface of a large body 共such as a valve body兲, but the same
droplet may lead to significant fire propagation if it lands on a sharp edged
feature 共such as some internal valve components兲, where the energy require-
ments for subsequent melting and initial burning are smaller. In the context of
industrial oxygen systems, this implies, for example, that components should
be oriented with thin-walled features at the top, so that any molten material
that lands on these surfaces will, due to gravity, form a freely hanging droplet
共instead of pooling兲, which would tend to minimize the contact area between
liquid and solid phases, limiting the total heat transfer rate and making subse-
quent melting less favorable. In this example, the strategy for limiting fire
propagation is to, despite the presence of a significant a heat source in the
burning droplet and the lack of a large heat sink 共due to the local thin cross-
section兲, attempt to interrupt the melting cycle by using controllable geometric
design parameters to restrict the heat transfer side of the melting rate triangle.
In this way, the melting rate triangle provides insight into strategies to limit fire
spread. Just as the Slusser and Miller extended fire triangle 关2兴 implies that
ignition can be avoided by removing one of the three legs, the melting rate
triangle implies that RRMI can be reduced or, perhaps, even stopped 共reduced
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102270 47

to zero兲 by limiting or interrupting one of the three processes. The melting rate
triangle shows that, even in the presence of a heat source, such as the case
when a fire occurs inside an oxygen system, the melting rate can be reduced by
limiting the extent to which the melting potential is realized, by restricting the
heat transfer process and increasing the heat sink. This relates especially to the
design of thin-walled or sharp-edged features, threaded surfaces, meshes, link-
age rods, and other small valve components.

Summary
The melting rate triangle provides a framework for discussion and enables fur-
ther understanding of the net effect of the competing influences of multiple
known 共and unknown兲 parameters that are relevant to the heterogeneous burn-
ing of metallic materials. It enables a holistic view of the burning system and
contributes to an improved understanding of how different parameters interre-
late and affect the burning process under different conditions. The melting rate
triangle was shown to be consistent with existing experimental data and known
trends, and capable of incorporating new information. Application of the melt-
ing rate triangle provides insight into the results of standardized flammability
tests and may improve the process of using these results in the design and
analysis of practical oxygen components and systems in industrial scenarios.

References

关1兴 Stoltzfus, J. M., Homa, J. M., Williams, R. E., and Benz, F. J., “ASTM Committee
G-4 Metals Flammability Test Program: Data and Discussion,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 986, Vol. 3, ASTM
International, West Conshohocken, PA, 1988, pp. 28–53.
关2兴 Slusser, W. M. and Miller, K. A., “Selection of Metals for Gaseous Oxygen Service,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP
812, Vol. 1, ASTM International, West Conshohocken, PA, 1983, pp. 167–191.
关3兴 Kirschfeld, L., “Rate of Combustion of Iron Wire in Oxygen Under High Pressure,”
Arch. Eisenhuettenwes., Vol. 32, 1961, pp. 57–62.
关4兴 Samant, A. V., Zawierucha, R., and Million, J. F., “Thickness Effects on the Pro-
moted Ignition-Combustion Behviors of Engineering Alloys,” Flammabilty and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 1454, Vol. 10, ASTM
International, West Conshohocken, PA, 2003, pp. 171–191.
关5兴 Sato, J., “Fire Spread Rates along Cylindrical Metal Rods in High Pressure Oxy-
gen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres,
STP 1040, Vol. 4, ASTM International, West Conshohocken, PA, 1989, pp. 162–177.
关6兴 Sato, K., Sato, Y., Tsuno, T., Nakamura, Y., Hirano, T., and Sato, J., “Metal Com-
bustion in High Pressure Oxygen Atmosphere: Detailed Observation of Burning
Region Behavior by using High Speed Photography,” Proc. SPIE, Vol. 348, 1982,
pp. 828–832.
关7兴 Ward, N. R. and Steinberg, T. A., “The Rate-Limiting Mechanism for the Hetero-
geneous Burning of Cylindrical Iron Rods,” J. ASTM Int., paper ID JAI102269.
关8兴 Steinberg, T. A., and Stoltzfus, J. M. “Combustion Testing of Metallic Materials
aboard NASA Johnson Space Center’s KC-135,” Flammability and Sensitivity of
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
48 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Materials in Oxygen-Enriched Atmospheres, STP 1319,Vol. 8, ASTM International,


West Conshohocken, PA, 1997, pp. 170–188.
关9兴 Steinberg, T. A., Wilson, D. B., and Benz, F. J., “The Burning of Metals and Alloys
in Microgravity,” Combust. Flame, Vol. 88, 1992, pp. 309–320.
关10兴 Steinberg, T. A., “The Combustion of Metals in Gaseous Oxygen,” Ph.D. thesis,
New Mexico State University, 1990.
关11兴 Suvorovs, T., “Promoted Ignition Testing: An Investigation of Sample Geometry
and Data Analysis Techniques,” Ph.D. thesis, Queensland University of Technology,
2007.
关12兴 Steinberg, T. A. and Wilson, D. B., ““Modeling the NASA/ASTM Flammability Test
for Metallic Materials Burning in Reduced Gravity,” Flammability and Sensitivity
of Materials in Oxygen-Enriched Atmospheres, STP 1395, Vol. 9, ASTM Interna-
tional, West Conshohocken, PA, 2000, pp. 266–291.
关13兴 Wilson, D. B. and Stoltzfus, J. M., “Metals Flammability: Review and Model Analy-
sis,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres,
STP 1395, ASTM International, West Conshohocken, PA, Vol. 9, 2000, pp. 469–496.
关14兴 Sato, K., Hirano, T., and Sato, J., “Behavior of Fires Spreading over Structural
Metal Pieces in High Pressure Oxygen,” American Society of Mechanical Engineers
and Japanese Society of Mechanical Engineers Thermal Engineering Joint Conference
Proceedings, 1983 共unpublished兲, Vol. 4, pp. 311–316.
关15兴 Suvorovs, T., Ward, N. R., Steinberg, T. A., and Wilson, R., “Effect of Geometry on
the Melting Rates of Iron Rods Burning in High Pressure Oxygen,” J. ASTM Int.,
Vol. 4, 2007, Paper ID JAI101004.
关16兴 Ward, N. R. and Steinberg, T. A., “Iron Burning in Pressurized Oxygen Under
Microgravity Conditions,” Microgravity Sci. Technol., Vol. 21共1–2兲, 2008, pp. 41–46.
关17兴 Ward, N. R., “The Rate-Limiting Mechanism for Heterogeneous Burning in Nor-
mal Gravity and Reduced Gravity,” Ph.D. thesis, Queensland University of Tech-
nology, 2008.
关18兴 Ward, N. R. and Steinberg, T. A., “Geometry of the Melting Interface in Cylindrical
Metal Rods Under Microgravity Conditions,” C. R. Mec., El Ganaoui, M., and,
Prud’homme, R., Eds., Vol. 5–6, 2007, pp. 342–350.
关19兴 Edwards, A. P. R., “Modelling of the Burning of Iron Rods in Normal Gravity and
Reduced Gravity,” Ph.D. thesis, The University of Queensland, 2004.
关20兴 Sato, J. and Hirano, T., “Behavior of Fire Spreading Along High-Temperature Mild
Steel and Aluminum Cylinders in Oxygen,” Flammabilty and Sensitivity of Materi-
als in Oxygen-Enriched Atmospheres, STP 910, ASTM International, West Consho-
hocken, PA, Vol. 2, 1986, pp. 118–134.
关21兴 Engel, C. D., Herald, S. D., and Davis, S. E., “Promoted Metals Combustion at
Ambient and Elevated Temperatures,” J. ASTM Int., Vol. 3, 2006, Paper ID
JAI13539.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102264
Available online at www.astm.org/JAI

Etienne Werlen,1 Frédéric Crayssac,2 Olivier Longuet,3 and


Fabien Willot3

Ignition of Contaminated
Aluminum by Impact in Liquid
Oxygen—Influence of Oxygen Purity

ABSTRACT: Impact tests in liquid oxygen and oxygen enriched liquids have
been conducted on aluminum foils 共0.2 mm thick兲 used for heat exchangers
of air separation units 共ASUs兲. Some adaptations to the ASTM Standard
D2512, “Standard Test Method for Compatibility of Materials with Liquid Oxy-
gen,” have been done to ensure the good control of oxygen content in the
cryogenic liquid and to get reproducible results in aluminum ignition with
hydrocarbon coating. Modified striker pins with a cavity on the contact sur-
face have been used. The vertical guiding and the centering of the striker
pins have been ensured by using two machined guides. Test samples were
three layers of aluminum foil, two corrugated and one flat, contaminated by
various quantities of hexadecane. The criteria for the identification of alumi-
num ignition 共positive tests兲 have been validated by performing Scanning
electron microscope surface analysis on residuals of a few impact tests per-
formed in copper cups. Criteria have also been defined to check that hexa-
decane combustion took place in negative tests. A sufficient number of tests
have been performed to plot the probability of ignition of the aluminum
samples as a function of the average thickness of hexadecane coating and
as a function of the oxygen content in the cryogenic liquid varying from 50 %
to above 99.95 %. The remaining part in the cryogenic liquid was nitrogen
and 2.3 % argon except for one series of tests where the influence of the
absence of argon could be identified. The results of these tests can be used
to assess the risk of ignition of some aluminum exchangers or structures
used in ASU.

Manuscript received December 5, 2008; accepted for publication August 19, 2009; pub-
lished online September 2009.
1
Air Liquide Engineering, Champigny-sur-Marne, 94503, France.
2
CRCD, Air Liquide R&D, Les-Loges-en-Josas, 78354, France.
3
CTE, Air Liquide Testing Center, Le Blanc-Mesnil, 93155, France.
Cite as: Werlen, E., Crayssac, F., Longuet, O. and Willot, F., ⬙Ignition of Contaminated
Aluminum by Impact in Liquid Oxygen—Influence of Oxygen Purity,⬙ J. ASTM Intl., Vol.
6, No. 10. doi:10.1520/JAI102264.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 49
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
50 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

KEYWORDS: oxygen, argon, nitrogen, cryogenic, liquid, purity,


ignition, flammability, hydrocarbon, hexadecane, thickness, aluminum,
copper, foil, layer, impact, air separation unit, exchanger, reboiler

Introduction

Thin aluminum structures are used in air separation units 共ASU兲 at cryogenic
temperatures in contact with liquid oxygen 共LOX兲 and oxygen enriched liquids.
Typical examples are structured packing for distillation sections 共refer to Ref 1兲,
brazed aluminum heat exchangers 共BAHX兲 performing dry vaporization at el-
evated pressure 共refer to Ref 2兲, and brazed aluminum reboiler/condensers op-
erating at low pressure 共refer to Ref 3兲.
Several incidents involving hydrocarbons combustion sometimes associ-
ated to aluminum combustion have been reported in air separation plants re-
boilers. A few of these incidents have involved major explosions 共one of these
accidents is described in Ref 4兲. Safety practices 共design and operation兲 have
been defined to prevent recurrences of such incidents. These are described in
the industry documents referenced above. The long experience accumulated
indicates that provided that appropriate safety practices are in place, the use of
such thin aluminum structures in ASUs is safe.
However, most of the accumulated experience corresponds to ASUs pro-
ducing standard purity oxygen 共⬃99.5 %兲. Also of interest is the safe operation
of nitrogen generators, which are a specific type of ASUs where the reboiler
operates with a liquid containing generally less than 70 % oxygen and up to 80
% for some specific process. It is described in the literature 共refer in particular
to Ref 5兲 that aluminum combustion rate in gaseous oxygen is significantly
decreased by the presence of inert impurities 共i.e., argon兲 even down to traces
concentration.
The objectives of this article are to review existing test results on aluminum
combustion propagation with regard to the oxygen content and to describe new
test results investigating the probability of ignition of thin aluminum in pres-
ence of hydrocarbons as a function of the oxygen content in the cryogenic
liquid varying from 50 % to above 99.95 %. The results of these tests can be
used to assess the risk of ignition of the abovementioned aluminum structures
used in ASUs.

Literature Review on Aluminum Combustion Propagation

Several aluminum flammability tests performed in the past and published can
provide information on aluminum combustion with regard to oxygen purity.
• Promoted combustion tests of aluminum rods and thin sheets in gas-
eous oxygen
• Promoted combustion tests of structured aluminum packing in gaseous
oxygen and partially immersed in, or irrigated, by LOX
• LOX shock tests with low purity oxygen
• Promoted combustion tests of BAHX samples in gaseous oxygen and
partially immersed in LOX
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 51

The tests referred in this literature review as “promoted combustion” tests


are propagation of combustion experiments, meaning that the promoter is se-
lected to provide enough energy to ignite and that the key observation is
whether or not combustion continues or accelerates once an initial ignition
occurs.
The LOX shock tests described in this paper and in Ref 6 are fundamentally
different because the key observation is whether or not ignition of aluminum
occurs for a given ignition energy delivered by a combination of mechanical
impact and hydrocarbon contamination combustion.
In promoted combustion tests with LOX, the violent reaction of aluminum
burning has often been referred to as a violent energy release, or VER. Al-
though the definition of a VER differs amongst the authors, it is always asso-
ciated with a shock wave 共usually characterized by physical destruction or de-
formation of the specimen and/or the test vessel兲 and with rapid aluminum
combustion. VER is used in this paper to describe test results having the above
characteristics.
The purpose of this literature review section is to show that the results
described in this paper are not contradictory with previous studies and why it is
believed that once ignited by hydrocarbons in LOX, aluminum combustion
continues and leads to VERs in both high purity LOX 共⬎99.95 %兲 and stan-
dard purity LOX 共99.5 %兲.

Promoted Combustion Tests of Aluminum Rods and Thin Sheets in Gaseous


Oxygen
The data available in the literature 共Refs 5 and 7兲 show that the threshold
pressure for propagation of combustion and the speed of combustion are sig-
nificantly affected by high oxygen purity 共even small traces of argon being a
mitigation factor兲.
In Ref 5, a threshold pressure below which aluminum combustion will not
propagate is determined as a function of oxygen purity and of rod diameter. As
an example with argon diluent, the threshold pressure for a 3.2 mm diameter
rod in 99.77 % O2 is 8 bar abs. It decreases to 3 bar abs in 99.93 % O2 and to 1
bar abs in 99.99+ % O2.
The effect of inert gases is reported to be caused by the formation of an
inert layer, which reduces the oxygen feed to the combustion area. As all the
species produced by aluminum combustion are liquid or solid, the accumula-
tion in the mass-transfer zone of gaseous inert impurities will impede the flux
of oxygen to the reaction site. It has been shown that nitrogen is less efficient to
reduce the flammability of aluminum; this would be due to the formation of
aluminum oxynitrides.
As described in Refs 8 and 9, flame speed experiments where aluminum
tubes are ignited in pressurized gaseous oxygen 共GOX兲 have shown that the
presence of LOX is making a very big difference to the flame speed: Flame
spreading rate in GOX is between 0.05 and 0.5 m/s even at elevated pressure
共up to 700 bar兲 when it can reach 73 m/s in LOX/GOX even at moderate pres-
sure 共⬍20 bar兲. However these tests do not investigate the purity effect.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
52 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Promoted Combustion Tests of Structured Aluminum Packing


• Reference 10 describes promoted combustion tests, which have been
performed on structured packing 共aluminum and other metals兲 partially
immersed in LOX and in GOX at atmospheric pressure. The ignition
was performed using thermite and reactive metal promoters. The purity
was variable. For aluminum packing in GOX, no propagation occurred,
except for the highest purity tested 共99.9+ %兲. For aluminum packing
partially immersed in LOX 共promoter in the gaseous phase兲, VERs oc-
curred for LOX purity down to 98.5 % but with a lower probability than
in ultra-pure oxygen. It is mentioned that at high oxygen concentrations
共99.9+ %兲, the specimen appeared to be near-totally consumed,
whereas at lower concentrations, the combustion appeared to destroy
only between 40 % and 60 % of the specimen.
• Tests described in Ref 11 have been performed with partially immersed
packing samples at atmospheric pressure 共commercial LOX 99.88 % O2
purity—ignition in gaseous oxygen immediately above LOX surface兲;
with irrigated packing at atmospheric pressure 共same LOX purity兲; with
partially immersed packing samples under slight pressurization 共1.36–
1.57 bar—commercial LOX 98.7 % O2 purity—ignition in gaseous oxy-
gen immediately above LOX surface兲; and with partially immersed and
irrigated packing samples under slight pressurization 共1.36–1.7 bar—
commercial LOX 99.9 % O2 purity—ignition in gaseous oxygen imme-
diately above LOX surface兲.
Among all these tests for which various promoters have been used, only
one explosion 共VER兲 with almost complete combustion is reported, with
partially immersed not irrigated packing at 1.57 bar with 98.7 % LOX
purity.
These tests illustrate the following conclusions: It is experimentally dif-
ficult to ignite irrigated packing 共no combustion is reported with irri-
gated packing兲, and if ignition is achieved, the presence of trace impu-
rities is not a mitigating factor 共one explosion with 98.7 % oxygen purity
completely destroyed the test apparatus兲.
• Reference 12 describes two series of tests using thermite and reactive
metal promoters. First series of tests with partially immersed packing
samples 共ignition in gaseous oxygen兲 have been performed with three
different grades of oxygen: Standard cryogenic oxygen 共99.8+ %兲, high
purity cryogenic oxygen 共99.99+ %兲, and ultra-pure oxygen from elec-
trolytic source 共99.999+ %兲. The publication reports that all things
being equal, only 1/3 of the tests performed with standard oxygen purity
resulted in a VER 共but combustion always occurred兲 when all the tests
with high purity and ultra-pure oxygen resulted in VERs.
Second series of tests have been performed with the same three grades
with irrigated packing and with no LOX pool underneath. No VER was
observed, only aluminum melting 共residuals slightly oxidized兲.
The publication concludes that VERs are produced by burning molten
aluminum falling at very high temperature in LOX of high or ultra high
purity.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 53

• Tests described in Ref 13 have been performed in large scale column


models equipped with aluminum structured packing and aluminum
trays. Studying the effect of purity was not the purpose of these series;
that is the reason why all the packing tests and most of the tray tests
have been performed with 99.99 % LOX.
These tests have demonstrated that it is possible to have a violent com-
bustion of irrigated packing with no LOX pool 共but that no VER occurs
in this configuration兲. Downward propagation was more likely when a
protective screen was installed above the pack equipped with the igniter
allowing ignition in vapor phase 共propagation occurred only in one case
out of five without a protective screen and in two cases out of two with
the protective screen兲.
Tests with irrigated aluminum trays 共1 mm thick兲 have demonstrated
that the ignition can result in VERs, with 99.99 % O2 and with 99.80 %
O2. With LOX hold-up on the trays, these tests do not show an impact of
purity.
• Reference 14 describes promoted combustion tests using thermite and
reactive metal promoters, which have been performed on structured
packing in O2 / N2 and O2 / Ar gaseous mixtures 共0 – 15° C兲 at various
pressures. The results are globally consistent with other promoted com-
bustion tests in GOX: The higher the pressure, the lower the purity
threshold 共O2-index兲, above which combustion propagation occurs. The
O2-index at atmospheric pressure is reported to be about 99.6 % in
O2 / N2 gaseous mixtures.
It is also reported that unpublished tests have been performed do deter-
mine the O2-index for combustion of irrigated packing in O2 / N2 mix-
tures at 1, 3, and 6 bar abs. These tests are not described in detail, but a
plot shows that the O2-index at 1 bar abs is close to 99.6 %; it decreases
to ⬃98 % at 3 bar abs and to 95–97 % at 6 bar abs. It is also mentioned
that combustion propagation with normal purity LOX 共99.9 %兲 proceeds
much smoother than with ultra high purity 共99.9999 %兲 and that the
ignition energy must be higher with normal purity LOX to obtain a
reliable combustion.

Liquid Oxygen Shock Tests with Low Purity Oxygen


LOX shock tests performed with a procedure similar to the one described in
this article but not incorporating the latest improvements are reported in Ref 6.
They show that aluminum combustion occurred with oxygen purity down to 95
%, with argon being the remaining part and with oxygen purity down to 65 %
with nitrogen being the remaining part.
These tests and the ones described further in this paper support the idea
that in LOX, aluminum flammability dependence on purity is very much lower
than in GOX.

Promoted Combustion Tests of Brazed Aluminum Heat Exchanger Samples


Tests reported in Ref 15 have been performed with BAHX samples irrigated by
LOX at atmospheric pressure, in pressurized GOX, and for the major part with
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
54 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—BAHX sample flammability 共tests reported in Ref 15—partially immersed in


LOX兲.

samples partially immersed in LOX, ignited in the vapor phase. Aluminum


combustion and VERs have only been observed with the last configuration.
This last configuration is similar to the one that resulted in VERs with
aluminum packing. The BAHX sample was partially immersed in LOX of vari-
ous purities, pressurized up to 90 bar, and ignited in vapor phase above the
LOX pool. One difference with regard to the majority of the abovementioned
packing samples ignition tests is that the promoter used for most of the BAHX
samples ignition tests was nichrome igniter with 2 g of hydrocarbons. Thermite
and reactive metal promoters typically used for packing tests do not generate
hydrocarbons reaction products, which in theory would reduce local oxygen
content and limit the combustion reaction.
Figure 1 above gives an overview of the results. It was not possible to have
aluminum combustion at atmospheric pressure; this is reported to be in con-
tradiction with air separation plants incidents where VER occurred close to the
atmospheric pressure. Above the atmospheric pressure, VER have been ob-
served with some pressure-purity relationship 共see Fig. 1兲. Between 99.4 % and
99.9 % LOX purity, no pressure-purity relationship can be identified. It is
worthwhile to mention that hydrocarbon promoted combustion tests in flowing
high purity GOX 共99.96 % purity兲 did not lead to aluminum combustion up to
14.8 bar abs.

Conclusion on Combustion Propagation


Aluminum combustion propagation and VER are two different phenomena. As
demonstrated by packing combustion tests 共see Ref 13兲, aluminum combustion
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 55

propagation does not systematically corresponds to a VER. However, it can be


assumed that when a VER is reported, aluminum combustion propagation in
LOX did take place.
From the above synthesis of the aluminum flammability tests on packing
and BAHX samples, several observations can be made.
• Promoted combustion tests show that the probability of a VER is higher
in high oxygen purity when partially immersed samples are ignited in
gaseous phase using thermite and reactive metals promoters. In gaseous
phase, it is well known that the rate of propagation is highly dependent
on the oxygen purity. Therefore, the energy released by the initial step of
combustion propagation in gaseous phase must have been highly depen-
dent on the purity. The tests have demonstrated that VER can occur
with standard purity oxygen 共down to 98.5 %兲, with a lower probability,
probably as a consequence of the less extensive initial step of combus-
tion in gaseous phase. Therefore, it is not believed that these tests are
giving a reliable indication of the probability of aluminum ignition with
hydrocarbon coating, especially at 99.5 % LOX purity and above.
• Promoted combustion tests with partially immersed samples ignited
using hydrocarbon promoter do not show pressure-purity relationship
between 99.4 % and 99.9 % LOX purity with regard to VERs. It can be
assumed that like oxygen impurities do, the hydrocarbon reaction prod-
ucts generated by the promoter are reducing the oxygen feed to the
ignition area. This phenomenon can explain the lowest sensitivity to
oxygen purity observed in promoted combustion tests, which use hydro-
carbon promoters. Based on these tests, it is believed that once ignited
by hydrocarbons in LOX, aluminum combustion continues and leads to
VERs both in high purity LOX 共⬎99.95 %兲 and in standard purity LOX
共99.5 %兲. However, the data published do not allow to rate unambigu-
ously the level of energy released by such VERs as a function of oxygen
purity.

Shock Tests in Liquid Oxygen and Oxygen Enriched Liquids

Liquid Oxygen Shock Test Bench Description


The LOX shock test bench is based on the testing method described in ASTM
Standard D2512, Standard Test Method for Compatibility of Materials with
Liquid Oxygen 共Ref 16兲.
Some adaptations have been done to the test bench to ensure the purity of
LOX and to get reproducible result in aluminum ignition with hydrocarbon
coating.

Liquid Oxygen Shock Test Bench Mechanical Arrangement


The LOX shock test bench mechanical arrangement as described by ASTM
Standard D2512 has been modified for several purposes 共see Fig. 2兲.
• A chamber purged with GOX at the same purity as the LOX used to fill
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
56 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

E = 100 J
Upper guide Striker pin

LOX fill
GOX purge

Lower guide

Aluminum cup
Aluminum disks
O-ring

Liquid nitrogen bath Monel sleeve Monel disk

FIG. 2—LOX shock test bench sketch.

the test cup has been created to avoid air ingress and ensure the high
purity of the LOX. Not represented on the sketch is a relief plate, which
allows release of the pressure wave generated by the explosions.
• A liquid nitrogen bath associated to a hollow part surrounding the test
cup receptacle allows a perfect cooling of the test bench 共the fact that
very few bubbles can be observed in the receptacle filled with LOX
proves that the cooling is efficient兲.
• Lower and upper guides provide a precise vertical guiding of the striker
pin. This has been proven to be a very important parameter for the
reproducibility of results.
• A removable Monel disk provides a flat surface under the cup. It can be
replaced in case of damage by explosions. For the last test series, a
removable Monel sleeve has been added in the cup receptacle to protect
the side walls.
• Tightness between the LIN bath and the LOX receptacle is ensured by a
metallic O-ring.

Modified Striker Pin


In order to improve the sensitivity of the test bench, a modified striker pin
made of stainless steel has been used. The modified striker pin has a circular
void in the center of the impact zone 共see Fig. 3兲. A vent hole was pierced at the
top of the void to allow complete filling with LOX.
Several advantages have been found in using modified striker pins.
• The reduced impact surface 共approximately divided by a factor of two兲
makes the local energy applied on the disks higher and thus a higher
probability of aluminum combustion.
• An aluminum/hydrocarbons/LOX mixture is trapped under the void and
can react during a short period between the shock and the rebound of
the striker pin.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 57

Corrugated disk
Stacked aluminum
Flat disk disks coated with
hexadecane
Corrugated disk

Aluminum cup
sharp edges

FIG. 3—Modified striker pin and new stack tested.

• The vent hole collects the reaction products, which are projected on the
inner wall of the test cup.

Cup and New Stack Tested


Standard aluminum cups as specified in ASTM Standard D2512 have been used
for the tests but also for qualification purposes copper cups of identical dimen-
sions.
Not only aluminum but also copper disks have been used for the tests.
Their diameter was 18.5 mm. Corrugated disks had a corrugation length of 0.8
mm and a corrugation height of 0.4 mm for all the tests reported in this article.
During preliminary testing, it has been realized that the corrugated disks were
moving in the cup due to the LOX boiling during cooling down. In a few cases,
their rotation was large enough to align the corrugations and reduce the LOX
hold-up between the disks by fitting in each other. To avoid this undesirable
effect, it has been decided to insert a flat disk between the two corrugated disks
共see Fig. 3兲. This is one difference in comparison with the tests previously pub-
lished in Ref 6.

Oxygen Enriched Liquids Generation


In order to control properly the oxygen, argon, and nitrogen contents in the
liquid filling the cup, all the mixtures, including the 99.5 % oxygen purity one,
have been prepared using high purity cylinders and mass flow regulators. The
gaseous mixture was permanently checked by analyzers and liquefied through a
coil immersed in liquid nitrogen.

Test Protocol
• The corrugated 0.2 mm thick aluminum disks are coated with a hydro-
carbon layer of a controlled thickness 共by varying the hexadecane/
solvent proportion in the solution兲 as follows:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
58 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—LOX shock test bench model.

共1兲 Weighing of clean aluminum disks+ cup,


共2兲 Aluminum disks dipped in a solution solvent/hexadecane and stacked
in the cup,
共3兲 Drying of disks, and
共4兲 Weighing of polluted aluminum disks+ cup.
• The lower guide is placed on the aluminum cup, which is inserted in the
test bench receptacle and cooled down by the LOX flow.
• The upper guide and the modified striker pin, which have been pre-
cooled in LOX, are put in place.
• After at least 60 s, the 100 J mechanical shock is triggered.
The test results are very sensitive to a reproducible application of the test
protocol and to possible damages or deformations to parts of the test bench.
Accordingly, whatever the target conditions of the series of tests and every ten
shocks, one or two tests have been done with 99.5 % oxygen purity and about
30 g / m2 hexadecane deposit to check that positive results were still obtained
with a good reproducibility. It explains why a lot of tests have been done in
these conditions.

Liquid Oxygen Shock Test Bench Model


The objective of the LOX shock test bench used for aluminum ignition is to
reproduce LOX/aluminum/hydrocarbon ratios comparable to the ones that
could exist in actual brazed aluminum fins and may trigger aluminum ignition.
The large volume ratio of GOX, which usually exists in vaporization fins, is
taken into account in the shock test configuration by having very small corru-
gations on the disks and only LOX 共see Fig. 4兲. Typically, the LOX/aluminum/
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 59

hydrocarbon ratios are similar to the ones existing in vaporization fins where
the LOX volume fraction is about 15 %. A more detailed description of the
model of aluminum ignition by hydrocarbons can be found in Ref 4.

Shock Tests Interpretation


One objective of these LOX shock tests is to compare the probability of alumi-
num ignition for a given mass of hydrocarbon deposit in high purity LOX and
in standard purity LOX.
For a proper interpretation of the test results, negative and positive and
non-valid tests must be defined.
• Negative test: Hydrocarbon combustion but no aluminum ignition
• Positive test: Hydrocarbon combustion and aluminum ignition
• Non-valid test: No hydrocarbon combustion 共not taken into account for
the statistics兲
In order to support the test results interpretation and the subsequent con-
clusion, the following questions need to be addressed:
• Is the LOX purity in the test bench high enough?
• How can it be made sure that aluminum ignition took place in positive
tests?
• How can it be made sure that hydrocarbon combustion took place in
negative tests?

Liquid Oxygen Purity


In order to determine the actual LOX purity in the test cup during the tests, in
situ analysis of the oxygen purity has been performed.
For the purpose of the analysis, a striker pin has been modified to be con-
nected to a sampling line 共see Fig. 5兲.
Analysis sequences have been performed by installing the test cup with the
aluminum disks in the receptacle of the test bench, following the exact same
procedure and timing than for the real tests 共except for the shock itself兲.
It has been demonstrated that an appropriate purging time was sufficient
to ensure high purity of the LOX in the test cup. After 60 s, the purity is higher
than 99.95 % 共a figure which is believed to be conservative due to the purging
time of the sampling line兲.
A similar protocol has been used to show that the oxygen content in the cup
was close to the target for tests at 80 % O2 purity or below.
For all shock tests, a 60 s waiting period has been imposed by the proce-
dure between the moment the set up is ready for the shock 共polluted aluminum
disks, test cup, striker pin, and guides in place兲 and the shock.

Positive Tests Discussion


When performing the tests, some shocks result in a combustion mark inside the
test cup, in front of the vent hole of the striker pin.
These combustion marks are used to visually classify the tests as positive,
negative, or non-valid.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
60 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

2-42-4
Tube mm tubing
brasé

Axial
Trou axial4Ø mm
4 mm hole

Plugging
Bouchage à la of vent hole
brasure

FIG. 5—Modified striker pin for analysis.

In order to confirm that the visual criteria used to classify the tests do allow
to identify aluminum combustion, some qualification shock tests have been
performed with various test configurations.
Although all the test series have been performed with the aluminum cups
共as defined in ASTM Standard D2512兲 and with aluminum disks, the qualifica-
tions tests used copper test cups and copper disks. Scanning electron micro-
scope 共SEM兲 analysis 共energy dispersive spectroscopy method兲 has been per-
formed on the combustion marks, resulting from the various configurations of
the qualification tests. On the sample SEM picture shown on Fig. 6, the grey
area corresponds to alumina deposits as quantified by the bar chart. The white
spherical particles contain mainly iron, chrome, and a little of nickel.
From these analyses, combined with the visual observation of the corre-
sponding combustion marks, several conclusions have been drawn 共see Fig. 7兲.
Alumina was detected on the wall of the test cup with either aluminum cup
and/or aluminum disks. Therefore, some aluminum combustion can occur at
the test cup surface and on the disks.
The marks corresponding to tests with aluminum disks are wide black/grey
zone on the full height of the test cup with combustion residuals projections
共grainy surface containing alumina兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 61

60

50
Mass%

40

30

20 70 µm

10

0
C O Mg Al Cr Fe Cu

FIG. 6—Sample SEM analysis in the black/grey area of a test with copper cup.

The marks obtained with copper disks and copper cup are a small black
area with no residuals projections containing copper and maybe copper oxide.
Some residuals of the striker pin material are detected in the marks 共and
possibly residuals from a previous test trapped in the vent hole兲.

FIG. 7—SEM analysis of combustion marks.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
62 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Large black/grey zone


+ projections of residuals

FIG. 8—Typical positive result.

Based on the above analysis, the visual criteria used to identify positive
tests in the test series 共aluminum disks and aluminum cups兲 have been deter-
mined.

Positive Test Criteria 共See Fig. 8兲—


• Combustion marks in the cup in front of the vent orifice of the striker
pin
• Large black/grey zone+ projections of residuals
For positive tests, audible and visible explosion is generally observed dur-
ing the shock but not systematically.
In a few cases, there is a combustion mark, but it is not significant enough
to prove the aluminum combustion 共e.g., similar to the one observed on the
copper cup with copper disks兲. In such cases, for the stake of conservativeness,
the tests have been classified as non-valid and not taken into account in the
statistics.

Non-Valid Test Criteria—


• Limited combustion marks in the cup in front of the vent orifice of the
striker pin
• Small black/grey zone or no projections

Negative Tests Discussion


Negative tests are the ones where hydrocarbon combustion occurred but no
aluminum ignition.
To identify evidences of hydrocarbon combustion, some Blank Tests with-
out hydrocarbon contamination have been performed.
For Blank Tests, the following observations have been done 共see Fig. 9兲:
• No explosion sound or light emission observed during the shock,
• No traces of combustion, and
• Aluminum disks only flattened under the striker impact surface.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 63

Flattened zone
under striker
impact surface

FIG. 9—Aluminum disks after Blank Test.

When hydrocarbon pollution was present on the aluminum disks, their


appearance after the shock is different, they are almost always pierced under
the striker pin impact surface, and/or they have traces of heating. The compari-
son between aluminum disks appearance with and without hydrocarbon pollu-
tion is considered to be a sufficient proof that hydrocarbon combustion takes
place and that for negative tests it leads to local melting of aluminum.

Negative Test Criteria 共see Fig. 10兲—


• Absence of combustion marks in the cup in front of the vent orifice of
the striker pin
• Aluminum disks pierced under the striker impact surface or presenting
traces of heating

High Purity Shock Tests Results


In order to compare the probability of aluminum ignition for a given mass of
hydrocarbon deposit in standard oxygen purity and in high oxygen purity, LOX
shock tests have been performed in both conditions for various hydrocarbon
coating thicknesses.

Aluminum discs
pierced under the
striker impact
surface

FIG. 10—Aluminum disks typical negative result.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
64 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 11—Chart of high purity and standard purity LOX shock tests results.

The results of these LOX shock tests are given in Fig. 11. This chart shows
that a good reproducibility has been achieved with the test bench. For depos-
ited masses between 15 and 40 g / m2, the percentage of positive results is be-
tween 60 % and 70 %. For each deposited mass interval, a 90 % confidence
interval calculated using the binomial law4 as well as the number of tests per-
formed is indicated.
When comparing high purity with standard purity, one can notice that
within the error bars, they are close enough to show that the probability of
aluminum ignition for a given mass of hydrocarbon is similar in standard pu-
rity LOX 共99.5 %兲 and in high purity LOX 共⬎99.95 %兲.

Shock Tests Results as a Function of the Oxygen Content


A sufficient number of tests have been performed to plot the probability of
ignition of the aluminum samples as a function of the average thickness of
hexadecane coating and as a function of the oxygen content in the cryogenic
liquid varying from 50 % to 99.95 %. The remaining part in the cryogenic liquid
was nitrogen and 2.3 % argon except for one series of tests.
The results of the shock tests are given in Fig. 12. They are plotted as a

4
Bayesian calculation of the confidence interval using for the probability p to get a
positive result from one test in a given deposited mass interval an explicit prior distri-
bution uniform between zero and one. Knowing the test results 共k positive out of n tests兲,
the probabilities to have p lower than the upper bound and lower than the lower bound
are respectively 95 % and 5 %.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 65

FIG. 12—Chart of shock tests results as a function of the oxygen content.

function of the oxygen content for various hydrocarbon coating thicknesses.


The curves show that the probability of positive results 共aluminum ignition兲 for
a given mass of hydrocarbons is decreasing continuously between 100 % and
60 % oxygen content in the liquid. With 2.3 % argon content in the mixture and
below 70 % oxygen content, the observed percentage of positive results is below
10 %. It falls down to zero for oxygen content below or equal to 60 %.
In order to compare with the results published in Ref 6, a series of tests has
been performed with 30 g / m2 hexadecane deposit and 65 % oxygen content,
nitrogen only being the remaining part. It shows that the percentage of positive
results, which is 3 % with 2.3 % argon in the mixture, is 15 % in the absence of
argon. From these results a mitigating effect of argon on the probability of
aluminum ignition can be identified.

Conclusions for Air Separation Unit Safety


It is known that for ASU brazed aluminum bath type reboilers, the only mecha-
nism that may lead to major energy releases is ignition and combustion of the
aluminum fins of the vaporizer promoted by the combustion of flammable air
contaminants which had accumulated 共see Ref 3兲. The test results published in
this paper show that the probability of aluminum ignition for a given mass of
hydrocarbon is similar in standard purity LOX 共99.5 %兲 and in high purity LOX
共⬎99.95 %兲. As explained in the literature review, it is believed that once ig-
nited by hydrocarbons in LOX, aluminum combustion continues and leads to
VERs with the two purities. The level of energy released by such VERs cannot
be rated unambiguously as a function of oxygen purity. However, the data avail-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
66 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

able on the frequency 共or probability兲 of energy releases that is available for
‘‘standard purity’’ situations is also applicable to high purity applications. As a
consequence, the large experience of safe operation of ASUs, mostly gathered
when operating close to standard purity can be used to estimate the residual
frequency of energy releases in ASU brazed aluminum bath type reboilers op-
erating at high purity.
Another result concerns brazed aluminum vaporizers operated with low
oxygen purity. The fact that the observed percentage of positive results is below
10 % with 2.3 % argon content in the mixture and below 70 % oxygen content,
shows that aluminum ignition in nitrogen generators reboilers is very unlikely.
However, nitrogen generators operating with higher oxygen content 共around 80
% in the sump of the reboiler兲 keep an intermediate level of ignition probability.

References

关1兴 CGA G-4.8-2005, 2005, “Safe Use of Aluminum-Structured Packing for Oxygen
Distillation” 共Globally harmonized with EIGA-IGC Doc 701/04兲, Compressed Gas
Association, Inc., Chantilly, VA.
关2兴 CGA G-4.9-2004, 2004, “Safe Use of Brazed Aluminum Heat Exchangers for Pro-
ducing Pressurized Oxygen” 共Globally harmonized with EIGA-IGC Doc 145/08/E兲,
Compressed Gas Association, Inc., Chantilly, VA.
关3兴 CGA P-8.4-2006, 2006, “Safe Operation of Reboiler/Condensers in Air Separation
Units” 共Globally harmonized with EIGA-IGC Doc 65/06兲, Compressed Gas Associa-
tion, Inc., Chantilly, VA.
关4兴 Lehman, J.-Y., Wei, X. C., Hua, Q. X., and Delannoy, G., “Investigation of the
Fushun ASU Explosion in 1997,” J. Loss Prev. Process Ind., Vol. 16, 2003, pp.
209–221.
关5兴 Benning, M. A., Zabrenski, J. S., and Le, N. B., “The Flammability of Aluminum
Alloys and Aluminum Bronzes as Measured by Pressurized Oxygen Index,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Vol-
ume, ASTM STP 986, D. W. Schroll, Ed., ASTM International, West Conshohocken,
PA, 1988, pp. 54–71.
关6兴 Barthelemy, H., Roy, D., and Mazloumian, N., “Ignition of Aluminum by Impact in
LOX—Influence of Contaminants,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres: Ninth Volume, ASTM STP1395, Steinberg, T. A.,
Newton, B. E., and Beeson, H. D., Eds., ASTM International, West Conshohocken,
PA, 2000, 134–144.
关7兴 Yeh, C. L., Johnson, D. K., and Kuo, K. K., “Experimental Study of Flame-
Spreading Processes over Thin Aluminum Sheets,” Flammability and Sensitivity of
Materials in Oxygen-Enriched Atmospheres: Eighth Volume, ASTM STP 1319, W. T.
Royals, T. C. Chou, and T. A. Steinberg, Eds., ASTM International, West Consho-
hocken, PA, 1997, 283–296.
关8兴 Mench, M. M., Kuo, K. K., Sturges, J. H., Hansel, J. G., and Houghton, P., “Flame
Spreading and Violent Energy Release 共VER兲 Processes of Aluminum Tubing in
Liquid and Gaseous Oxygen Environments,” Flammability and Sensitivity of Mate-
rials in Oxygen-Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Stein-
berg, B. E. Newton, and H. D. Beeson, Eds., ASTM International, West Consho-
hocken, PA, 2000, 402–424.
关9兴 Mench, M. M., Haas, J. P., and Kuo, K. K., “Combustion and Flame Spreading of
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WERLEN ET AL., doi:10.1520/JAI102264 67

Aluminum Tubing in High Pressure Oxygen,” Combustion of Energetic Materials,


K. K. Kuo and L. T. De Luca, Eds., Begell House, Inc., New York, 2002, pp. 438–
452.
关10兴 Dunbobbin, B. R., Hansel, J. G., and Werley, B. L., “Oxygen Compatibility of High-
Surface Area Materials,” Flammability and Sensitivity of Materials in Oxygen En-
riched Atmospheres, ASTM STP 1111, Fifth Volume, Joel M. Stoltzfus and Kenneth
McIlroy, Eds., ASTM International, West Conshohocken, PA, 1991, pp. 338–353.
关11兴 Zawierucha, R., Million, J. F., Cooper, S. L., McIlroy, K., and Martin, J. R., “Com-
patibility of Aluminum Packing with Oxygen Environment Under Simulated Op-
erating Conditions,” Flammability and Sensitivity of Materials in Oxygen Enriched
Atmospheres, ASTM STP 1197, Sixth Volume, ASTM International, West Consho-
hocken, PA, 1993, pp. 255–275.
关12兴 Barthélémy, H. M., “Compatibility of Aluminum Packing with Oxygen, Test Re-
sults Under Simulated Operating Conditions,” Flammability and Sensitivity of Ma-
terials in Oxygen-Enriched Atmospheres: Sixth Volume, ASTM STP 1197, D. D. Jan-
off and J. M. Stoltzfus, Eds., ASTM International, West Conshohocken, PA, 1993,
pp. 276–290.
关13兴 Fano, E., Barthelemy, H., and Lehman, J.-Y., “Tests of Combustion of Aluminum
Packing and Trayed Columns,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. New-
ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000,
pp. 389–401.
关14兴 Meilinger, M., “Promoted Ignition-Combustion Tests with Structured Aluminum
Packings in Gaseous Oxygen with Argon or Nitrogen Dilution at 0.1 and 0.6 MPa,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Tenth
Volume, ASTM STP 1454, T. A. Steinberg, H. D. Beeson, and B. E. Newton, Eds.,
ASTM International, West Conshohocken, PA, 2003, pp. 137–163.
关15兴 Zawierucha, R. and Million, J. F., “Promoted Ignition-Combustion Tests of Brazed
Aluminum Heat Exchanger Samples in Gaseous and Liquid Oxygen Environ-
ments,” Flammability and Sensitivity of Materials in Oxygen Enriched Atmospheres:
Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. Newton, and H. D. Beeson,
Eds., ASTM International, West Conshohocken, PA, 2000, pp. 373–388.
关16兴 ASTM Standard D2512, 1995, “Standard Test Method for Compatibility of Materi-
als with Liquid Oxygen 共Impact Sensitivity Threshold and Pass-Fail Techniques兲,”
ASTM Standards related to Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, sponsored by ASTM Committee G-4 on Compatibility and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM International,
West Conshohocken, PA, pp. 1–11.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102240
Available online at www.astm.org/JAI

Katherine E. Robbins,1 Samuel Eddie Davis,2 and


Stephen D. Herald3

Verification of the ASTM G-124 Purge


Equation

ABSTRACT: ASTM G-124 seeks to evaluate combustion characteristics of


metals in high-purity 共⬎99 %兲 oxygen atmospheres. ASTM G-124 provides
the following equation to determine the minimum number of purges required
to reach this level of purity in a test chamber: n ⫽ ⫺4/log10共Pa/Ph兲, where n
is the total number of purge cycles required, Ph is the absolute pressure used
for the purge on each cycle, and Pa is the atmospheric pressure or the vent
pressure. The origin of this equation is not known and has been the source of
frequent questions as to its accuracy and reliability. This paper provides the
derivation of the G-124 purge equation, and experimentally explores the
equation to determine if it accurately predicts the number of cycles required.
KEYWORDS: purge equation, promoted combustion chamber, purge
cycles, oxygen concentration, oxygen percentage, Dalton, partial
pressures, high purity oxygen atmosphere

Introduction
ASTM G-124, “Standard Test Method for Determining the Combustion Behavior
of Metallic Materials in Oxygen-Enriched Atmospheres,” was first approved in
1994 and is employed internationally for comparisons of the combustion char-
acteristics of various metallic materials. The combustion characteristics that
can be evaluated include lowest burn pressure, highest no-burn pressure, and

Manuscript received November 20, 2008; accepted for publication August 19, 2009; pub-
lished online September 2009.
1
Student Researcher, Dept. of Mathematics, Vanderbilt Univ., 2201 West End Avenue,
Nashville, TN 37240.
2
Materials Engineer, Materials Combustion Research Facility, Materials Test Branch,
NASA George C. Marshall Space Flight Center, Huntsville, AL 35812.
3
Test Group Lead Engineer, InfoPro Corporation, Materials Test Branch, Materials
Combustion Research Facility, NASA George C. Marshall Space Flight Center, Hunts-
ville, AL 35812.
Cite as: Robbins, K. E., Davis, S. E. and Herald, S. D., ⬙Verification of the ASTM G-124
Purge Equation,⬙ J. ASTM Intl., Vol. 6, No. 10. doi:10.1520/JAI102240.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 68
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 69

regression rate 共apparent burn rate兲 of the sample. The intent of ASTM G-124 is
to evaluate these characteristics of metals in high-purity 共 ⬎ 99 % 兲 oxygen at-
mospheres. Obtaining high purity oxygen atmospheres in a laboratory can be
difficult, and the only method provided in G-124 for obtaining this purity is
what is sometimes called the “Purge Equation.” ASTM G-124 allows for no
greater than 0.01 % of the original atmosphere to remain, and leads to the
Purge Equation: n = −4 / log10共Pa / Ph兲, where n is the total number of purge
cycles required 共rounded up if not an integer兲, Ph is the absolute pressure used
for the purge on each cycle, and Pa is the atmospheric pressure or the vent
pressure. It is clearly useful to have such an equation because time and re-
sources can be saved by knowing the correct number of purges required to
reach the desired concentration of oxygen in a test chamber. However, the
validity of this equation has, in the past, been called into question. Applications
do exist where a difference of only a few tenths of a percent in the oxygen
concentration can yield different combustion behaviors. Therefore, it is impor-
tant that the accuracy and limitations of this equation are understood so that
the equation can serve as a useful tool when performing testing per ASTM
G-124.

Assumption Made
This derivation and research utilize the simplifying assumption that the two
gasses involved in this test, oxygen and nitrogen, behave ideally and compress
as ideal gasses. Existing literature supports this assumption.

Experimental

Test Apparatus
All testing was conducted in promoted combustion chambers that are used by
NASA’s Marshall Space Flight Center to perform ASTM G-124 testing. Two
chambers of different volumes were used to determine if chamber volume had
any effect on the purges. Chamber 1 is a 10 L 共0.35 ft3兲 promoted combustion
chamber, and Chamber 2 is a 17 L 共0.60 ft3兲 promoted combustion chamber.
Internal volumes of the chambers were calculated by measuring the inside
dimensions of the cylindrical chambers, then including separate calculations
for the internal hardware volume and deviations from perfect cylinders.
The test gas that was introduced into the chambers was commercial grade
oxygen, which was labeled as having purity greater than 99.6 % oxygen. Actual
evaluation of the oxygen gas inside the cylinders showed that the gas was ac-
tually close to 99.9 % oxygen. This oxygen grade is the one used when perform-
ing the ASTM G-124 testing, so it was also used in this evaluation.
Each of the two promoted chambers has an oxygen analyzer and separate
gauges for temperature, pressure, and humidity. A Servomex Corporation
model 572 oxygen analyzer is utilized on each chamber. Each analyzer is cali-
brated prior to every purge cycle by introducing a National Institute of Stan-
dards and Technology 共NIST兲 certified oxygen gas that is 99.999 % ± 0.0005 %
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
70 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

oxygen. Oxygen analyzers must read 100.0 % oxygen in order to be considered


accurate. Internal chamber pressure for each chamber is measured by a Viat-
ran Corporation model 240 pressure sensor leading to a Heise Corporation
model 710A digital gauge. The pressure gauges display the pressure from
0 to 10,000± 0.5 psig. Each component of the pressure system is calibrated
annually by an NIST accredited calibration facility.

Procedure
Chambers 1 and 2 utilized the same procedure throughout the verification.
Each chamber began the process by being sealed while in normal atmospheric
air at ambient temperature and pressure, Pa 共77° F 关25° C兴 and 14.5 psia
关100 kPa兴 at the test facility兲. First, an initial oxygen concentration reading was
taken from the air inside the chamber. Oxygen, as the purge gas, was then
allowed to enter the chamber through a series of opening and closings of a
valve until the chamber pressure was approximately the same as the desired
purge pressure, Ph. The chamber was stagnant for a sufficient length of time to
allow the temperature to return to near ambient, pressure to stabilize, and for
the gasses to become a homogeneous mixture. This process took an average of
3 min. The stabilized pressure was recorded, and the chamber was then vented
down to approximately 22.5 psia, or roughly 8 psia above atmospheric pres-
sure. Samples of the chamber gases were periodically withdrawn and analyzed
by an oxygen analyzer until the oxygen concentration stabilized. The chamber
was then vented to lower the pressure to Pa, completing the first purge. The
same process was repeated until the oxygen concentration either reached 100
% or reached the maximum value that the oxygen analyzer could detect 共i.e.,
when further purges did not increase the oxygen concentration reading on the
instrument兲. After each series of analyses was complete, the chamber was emp-
tied of gas, filled with atmospheric air, and a new purge cycle analysis was
begun with the next Ph.

Derivation

The following symbolism is used throughout this derivation:

Pi ⫽ The partial pressure of the chamber gas mixture resulting from


gas i;
Pa ⫽ The partial pressure of the chamber gas mixture due to the
initial chamber gas, normally air. This is also the lowest
pressure that will be witnessed by the test chamber environment
during the pressurization and venting cycles, to which ASTM
G-124 refers as purges. This is typically atmospheric pressure of
100 kPa, or the lowest chamber pressure witnessed;
Ph ⫽ The highest pressure that will be witnessed by the test chamber
environment during the pressurization and venting cycles. It is
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 71

the pressure resulting from the mixture of all of the gases


inside the chamber;
Xi ⫽ The molar ratio of the initial gas a of the chamber gas mixture
after the ith purge;
ni ⫽ The number of moles of the original chamber gas remaining
after the ith purge.

This derivation begins by assuming, even though the assumption does not
simplify the derivation, that the chamber is initially filled with ambient air at
atmospheric pressure. Purge cycles involve pressurizing and venting the
chamber using only oxygen gas 共thus the symbolism related to air and oxygen兲.
This derivation uses two elementary laws of chemistry and physics

Dalton ’ s law of partial pressures: Ph = Pa + PO2

and

Ideal gas law: PV = nRT

Where:
P ⫽ pressure,
V ⫽ volume,
n ⫽ number moles of gas,
R ⫽ gas constant and
T ⫽ temperature.

共Note: The Ideal Gas Law is assumed to be applicable, as noted earlier,


because ASTM G-124 is almost always conducted using oxygen and nitrogen,
which behave very similarly to ideal gases.兲
The impetus that leads to this equation in ASTM G-124 is a direct result of
the following sentence taken from the test method: “Pressurize and vent the
chamber a sufficient number of times to ensure that no more than 0.01 % of the
original atmosphere in the vessel remains.” This proportion refers to the mole
ratio of the original chamber gas versus the final chamber gases, i.e., the ratio
of na versus nh, or na / nh.
Given that PV = nRT, or equivalently, n = PV / RT, the ratio becomes

na RTa
冉 冊
P aV a
PaVaRTh

冉 冊
proportion of air remaining: = =
nh P hV h PhVhRTa
RTh
Since volume and temperature will not change, and since R is the gas
constant, then the equation simplifies to
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
72 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

na Pa
= = X1
nh Ph

This is convenient since the values of Pa and Ph can be measured with gauges on
the test chamber. Therefore, the mole ratio value can be easily obtained.

Initial Conditions
Start with n0 moles of air at atmospheric 共or ambient chamber兲 pressure, Pa;
Purges will be at a final, or highest, pressure of Ph using O2 gas.

First Purge
Pressurize—then vent to Pa;
Dalton’s Law of partial pressures: Ph = Pa + PO2;
Ideal gas law: PV = nRT;
Molar quantity of the original chamber gases after the first purge= X1
= 共Pa / Ph兲.
Since the chamber is vented to the original pressure, Pa, and both the
temperature and volume of the chamber have remained constant, the total
moles of the mixture of gases in the chamber are equal to n0.
Therefore, moles of original chamber air remaining= n1 = ntotalX1 = n0X1.

Second Purge
Pressurize—then vent to Pa;
The total pressure in the chamber is equal to the partial pressures of the
remaining air and the oxygen added during the current purge;
The molar quantity of the original chamber gas mixture after the second
purge

共 X 1P a兲
=X2 = = 共Pa/Ph兲2;
Ph
Again, the total number of moles of gas in the chamber is equal to n0;
Moles of original air remaining= n2 = ntotalX2 = n0X2;

n2
Proportion of original air remaining = = X2 = 共Pa/Ph兲2 .
n0

Nth Purge

nN
Proportion of original air remaining = = XN = 共Pa/Ph兲N 共1兲
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
n0
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 73

Solving for N

XN = 共Pa/Ph兲N

log共XN兲 = log关共Pa/Ph兲N兴 = N log共Pa/Ph兲

log共XN兲
N= 共2兲
log共Pa/Ph兲
When XN = 0.01 % = 0.0001, and log base 10 is used

log10共10−4兲 −4
N= = 共3兲
log10共Pa/Ph兲 log10共Pa/Ph兲
Equations 1 and 2 are themselves valuable since they are more general forms of
the Purge Equation and allow for a variety of conditions.

Data and Analysis

Step by Step Calculations


In order to look more closely at the data, it is useful to consider each purge
individually, especially because of the difficulty in pressurizing to exactly the
same Ph each cycle. Equation 3 was modified slightly for the actual experiment.
This modification was performed to correct for the inherent errors that would
creep into the equation because the initial chamber atmosphere is air, with 20.9
% oxygen, and the purge gas was oxygen. The equation, taking into consider-
ation oxygen in both the initial chamber air and purge gas, becomes

G = X N · G O + 共 1 − X N兲 · C 共4兲

Where:
G ⫽ the final oxygen concentration,
GO ⫽ initial oxygen concentration,
XN ⫽ the proportion of original atmosphere left in the chamber after
N purges, and C is the oxygen concentration of the purge gas.

From Eq 1 for a single purge, X1 = 共Pa / Ph兲, so after a single purge

G = 共Pa/Ph兲 · GO + 关1 − 共Pa/Ph兲兴 · C 共5兲


Equation 5 was used to generate the O2 % predicted by the purge equation after
each single purge. Using this method, a situation where errors would quickly
become compounded because of variations in Ph from purge to purge was
avoided.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
74 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Experimental Assessments

The true measure of theoretical calculations and analyses is how closely they
predict the values that will be produced in the laboratory. Validating the deri-
vation of the purge equation requires several tests to determine just how closely
the theoretical calculations and the measured values matched.
Table 1 provides the data for the purges that were conducted to validate the
calculations. The table provides, for each purge, the starting pressure Pa, final
pressure Ph, measured oxygen percentage O2 %, and the percentage oxygen
that is predicted by theoretical calculations. It should be noted that the oxygen
percentage at the starting point 共Purge #0兲 is the percentage that was measured
by an oxygen analyzer. The cylinders used to fill the chamber for the initial
measurement indicated the contents to be “air.” However, the air cylinders did
not exactly match the theoretical composition of air at standard temperature
and pressure, namely 20.95 % oxygen.

Graphs

The graphs below show a plot of the actual O2 % and the predicted O2 % versus
the purge number 共Fig. 1兲. The actual versus predicted points align very closely
together. The average difference between these points is reported as “avg. D” in
Table 2 for Chamber 1 and Table 3 for Chamber 2. One important point to note
is that in Fig. 2, the value for the oxygen concentration in Purge 3 appears to be
lower than for Purge 2 even after a purge cycle with oxygen. This anomaly is
due to the rounding issue with an oxygen analyzer that reads to only one deci-
mal point.

Results

Below is a summary of the data obtained for the two chambers. The average Ph
over the series of purges is shown for convenience. However, the average Ph
was not used for any calculations since the purges were analyzed individually,
step by step. An asterisk is present if the actual number of purges required to
reach the final O2 % was more than the number predicted by the purge equa-
tion. The column avg. D lists the average difference between the actual and the
predicted O2 % per purge.
The summary table above demonstrates that the purge equation typically
did not predict the number of purges exactly. However, the prediction was
almost always within one purge of the actual number of purges. One factor
affecting the measured values is that the available oxygen analyzer did not have
enough accuracy to allow XN = 0.01 %. The oxygen analyzer only displayed one
decimal place, so the smallest XN that could be detected was 0.1 % and this may
have changed smaller inaccuracies into larger ones in later purges. Also, the
average difference from the predicted oxygen percentage was not normally
within 0.1 %, but was typically within 0.2 %.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 75

TABLE 1—Purge data from Chambers 1 and 2.

Chamber 1 Chamber 2
Purge Pa Ph O2 Predicted Purge Pa Ph O2 Predicted
# 共psia兲 共psia兲 % O2 % # 共psia兲 共psia兲 % O2 %
0 20.5 20.5 0 20.8 20.8
1 14.7 82.7 88.7 85.86 1 14.82 89.82 86.8 86.85
2 14.7 94.7 98.1 98.24 2 14.82 89.82 97.7 97.74
3 14.7 85.7 99.5 99.67 3 14.82 95.82 99.5 99.56
4 14.7 88.7 99.7 99.92 4 14.82 91.82 99.8 99.84
5 14.7 128.7 99.8 99.97 5 14.82 98.82 99.9 99.89

0 20.5 20.5 0 20.8 20.8


1 14.76 167.76 93.7 93.01 1 14.82 163.82 93.0 92.74
2 14.76 159.76 99.4 99.42 2 14.82 173.82 99.1 99.31
3 14.76 164.76 99.9 99.95 3 14.82 177.82 99.7 99.83
4 14.76 158.76 100.0 99.99 4 14.82 162.82 99.8 99.88

0 20.5 20.5 0 20.8 20.8


1 14.76 280.76 96.3 95.82 1 14.82 268.82 95.6 95.54
2 14.76 270.76 99.8 99.80 2 14.82 266.82 99.5 99.66
3 14.76 274.76 100.0 99.99 3 14.82 269.82 99.7 99.88

0 20.5 20.5 0 20.7 20.7


1 14.76 441.76 97.7 97.34 1 14.79 441.79 97.2 97.06
2 14.76 451.76 99.9 99.92 2 14.79 408.79 99.7 99.80
3 14.76 445.76 100.0 100.00 3 14.79 416.79 99.8 99.89

0 20.5 20.5 0 20.7 20.7


1 14.76 628.76 99.1 98.13 1 14.79 608.79 98.0 97.98
2 14.76 640.76 99.8 99.98 2 14.79 604.79 99.7 99.85
3 14.76 620.76 99.9 100.00 3 14.79 608.79 99.6 99.90
4 14.79 608.79 99.7 99.89

0 20.5 20.5 0 20.6 20.6


1 14.76 809.76 98.7 98.55 1 14.8 820.8 98.4 98.47
2 14.76 807.76 99.9 99.98 2 14.8 929.8 99.7 99.88
3 14,8 809.8 99.8 99.90

0 20.5 20.5 0 20.6 20.6


1 14.76 1025.76 98.9 98.86 1 14.8 1016.8 98.8 98.75
2 14.76 1014.76 99.8 99.98 2 14.8 1024.8 99.8 99.88
3 14.76 1015.76 99.9 100.00

0 20.5 20.5 0 20.5 20.5


1 14.76 1513.76 99.3 99.22 1 14.76 1513.76 99.3 99.22
2 14.76 1515.76 99.7 99.99 2 14.76 1515.76 99.7 99.99
3 14.76 1511.76 99.8 100.00 3 14.76 1511.76 99.8 100.00
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
76 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1— 共Continued.兲

Chamber 1 Chamber 2
Purge Pa Ph O2 Predicted Purge Pa Ph O2 Predicted
# 共psia兲 共psia兲 % O2 % # 共psia兲 共psia兲 % O2 %
0 20.5 20.5 0 20.7 20.7
1 14.76 2027.76 99.5 99.42 1 14.8 2014.8 99.7 99.32
2 14.76 2026.76 99.9 100.00 2 14.8 2014.8 99.8 99.90
3 14.8 2014.8 99.9 99.90

0 20.5 20.5 0 20.6 20.6


1 14.8 5014.8 99.9 99.77 1 14.8 5014.8 99.5 99.67
2 14.8 5014.8 99.7 99.90
3 14.8 5014.8 99.8 99.90

Discussion of Uncertainties

Oxygen Analyzer and Purge Gas Purity


The largest contributors of uncertainty in this experiment were the inherent
inaccuracies of the oxygen analyzer and the true oxygen concentration of the

Chamber 1, 4 purges at 160 psia


101.0

100.0

99.0

98.0

97.0

96.0

95.0

94.0

93.0

92.0
0 1 2 3 4 5
Actual
Purge (N)
Predicted

FIG. 1—Predicted and actual oxygen concentrations by purge at ⬃160 psi.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 77

TABLE 2—Summary for Chamber 1.

P a, avg. Ph, Initial Final Purges avg. D, D ⬍ 0.1 % D ⬍ 0.2 %


psia psia Purges O2 % O2 % predicteda % ? ?
14.7 96.1 5 20.5 99.8 4* 0.7080b No No
14.76 162.76 4 20.5 100.0 4 0.1900 No Yes
14.76 275.43 3 20.5 100.0 4 0.1600 No Yes
14.76 446.43 3 20.5 100.0 3 0.1300 No Yes
14.76 630.09 3 20.5 99.9 2* 0.4167 No No
14.76 808.76 2 20.5 99.9 2 0.1150 No Yes
14.76 1018.76 3 20.5 99.9 2* 0.1067 No Yes
14.76 1513.76 3 20.5 99.8 2* 0.1900 No Yes
14.76 2027.26 2 20.5 99.9 2 0.0900 Yes Yes
14.8 5014.80 1 20.5 99.9 2 0.1300 No Yes
a
Purges predictions were obtained by rearranging Eq 4 to solve for XN, then substituting
into Eq 2.
b
The high value for D at low pressures in Chamber 1 may be a result of poor mixing, since
less turbulence has been observed in Chamber 1.

purge gas. The oxygen analyzer only displayed to one-tenth of 1 % oxygen


concentration. This created inaccuracies in that small changes of O2 % in the
chamber could not be determined. These inaccuracies compounded as purges
were repeated. Also, many of the purge cycles had readings that were above
99.0 % and only one significant digit could change. The purity of the purge gas
itself is also an issue because gas purity affects many calculations. The best
locally available gas supply was only labeled by the vendor as “⬎99.6 %.” For
the calculations made here, an assumption was made that the purity of the
purge gas was 100 %. An assumption of 100 % is required because running the
purge gas though the oxygen analyzer typically produced readings of
⬃99.9/ 100 %. However, it is possible that the purity was less at times. This

TABLE 3—Summary for Chamber 2.

P a, avg. Ph, Initial Final Purges avg. D, D ⬍ 0.1 % D ⬍ 0.2 %


psia psia Purges O2 % O2 % predicted % ? ?
14.82 93.22 5 20.8 99.9 4* 0.0400 Yes Yes
14.82 169.57 4 20.8 99.8 3* 0.1700 No Yes
14.82 268.49 3 20.8 99.7 2* 0.1333 No Yes
14.79 412.46 3 20.7 99.8 2* 0.1100 No Yes
14.79 607.79 4 20.7 99.7 2* 0.1650 No Yes
14.80 853.47 3 20.6 99.8 2* 0.1167 No Yes
14.80 1020.80 2 20.6 99.8 2 0.0650 Yes Yes
14.80 1514.80 3 20.5 99.6 2* 0.3733 No No
14.80 2014.80 3 20.7 99.9 2* 0.1600 No Yes
14.80 5014.80 3 20.6 99.8 2* 0.1567 No Yes
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
78 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Chamber 2, 4 purges at 608 psia


101.0

100.0

99.0

98.0

97.0

96.0

95.0

94.0

93.0

92.0
0 1 2 3 4 5
Purge (N) Actual
Predicted

FIG. 2—Predicted and actual oxygen concentrations by purge at ⬃600 psi.

would explain why purges in Chamber 1 could often reach 99.9 % or 100 %,
while purges in Chamber 2 could only reach 99.8 % or 99.9 %.

Multiple Purges
Utilizing chambers that are designed to purge at high pressures often makes it
difficult to stop the purge at an exact peak purge pressure, such as 80 psia.
Because of this, each purge was analyzed individually rather than assuming a
constant Ph over a series of purges. By using this method, deviations from the
purge equation could be determined without the introduction of additional
errors when the Ph values are not constant. This is a beneficial means of analy-
sis, but, in actual applications, these small, individual errors would be com-
pound though multiple purges. It should be noted that, in actual application,
more purges lead to greater overall errors.

Pressure and Temperature


The test chamber temperature could not always be held at a constant level
during the purge cycles, which led to other inherent compounded errors. Most
notably, whenever purge gas was added to a chamber, both the pressure and the
temperature would increase significantly. Even though additional time was
added to allow the temperature to drop back to normal prior to taking pressure
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROBBINS ET AL., doi:10.1520/JAI102240 79

readings, the temperature differential may have caused variations in the pres-
sure readings, especially for higher pressure purges in which the temperature
increases were more substantial.

Conclusions

The purge equation provides a good estimate for the minimum number of
purges needed to reach a given gas concentration. In this study, the actual
number of purges was within one purge of the number predicted by the purge
equation in 19 out of 20 independent tests. It is important to note that the
results obtained on the utilized test chambers are directly applicable to other
test chambers of similar volumes. Chambers with different volumes will
achieve oxygen concentrations at different rates than chambers of these vol-
umes, but the rate differences may be insignificant.
The purge equation itself holds up well with resulting deviations that are
small, typically within 0.2 % of the actual oxygen percentage for each purge. If
accuracy to within one-tenth of one-percent is desired, then this test method
requires using an accurate, calibrated oxygen analyzer to determine the oxygen
concentration at each step. However, the test method given by ASTM G-124
rarely requires this level of accuracy. Therefore, the purge equation is a valu-
able tool that can be used to determine the number of purge cycles that are
required to be performed prior to igniting a test sample in an ASTM G-124 test.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102351
Available online at www.astm.org/JAI

Kyle M. Sparks,1 Joel M. Stoltzfus,2 Theodore A. Steinberg,3 and


David Lynn4

Determination of Burn Criterion for Promoted


Combustion Testing

ABSTRACT: Promoted ignition testing 关1–3兴 is used to determine the relative


flammability of metal rods in oxygen-enriched atmospheres. In these tests, a
promoter is used to ignite each metal rod to start the sample burning. Ex-
periments were performed to better understand the promoted ignition test by
obtaining insight into the effect a burning promoter has on the preheating of
a test sample. Test samples of several metallic materials were prepared and
coupled to fast-responding thermocouples along their length. Various ignition
promoters were used to ignite the test samples. The thermocouple measure-
ments and test video were synchronized to determine temperature increase
with respect to time and length along each test sample. A recommended
length of test sample that must be consumed to be considered a flammable
material was determined based on the preheated zone measured from these
tests. This length was determined to be 30 mm 共1.18 in.兲. Validation of this
length and its rationale are presented.
KEYWORDS: Oxygen, ignition, temperature, burning, rods,
flammability, promoted combustion, heat affected, promoter, WSTF

Manuscript received January 27, 2009; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
Mechanical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces,
NM 88004.
2
Laboratories Office, NASA Johnson Space Center White Sands Test Facility, Las
Cruces, NM 88004.
3
Senior Lecturer, School of Engineering Systems, Queensland University of Technology
共QUT兲, Brisbane, QLD 4001, Australia.
4
Postgraduate Student, School of Engineering Systems, Queensland University of Tech-
nology 共QUT兲, Brisbane, QLD 4001, Australia.
Cite as: Sparks, K. M., Stoltzfus, J. M., Steinberg, T. A. and Lynn, D., ⬙Determination of
Burn Criterion for Promoted Combustion Testing,⬙ J. ASTM Intl., Vol. 6, No. 10.
doi:10.1520/JAI102351.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 80
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 81

Introduction
When metal burns, it does so in the liquid phase. In promoted ignition testing,
promoters are used to melt and initiate the burning of metal rod test samples.
When a promoter is ignited, it transfers energy as heat to the sample, and a
liquid molten drop forms at the bottom of the vertically mounted sample.
While a material is burning, the molten droplets continue to form and drip
away in a cyclic pattern as the rod is consumed. If conditions are not able to
sustain the burning, a drip will re-solidify and the sample will be extinguished.
When the promoter is ignited, a portion of the rod is heated beyond the loca-
tion of the promoter. This heating is termed “sample preheating.” Although
there have been earlier works researching promoter effects 关4兴, there are not
enough data to make conclusive determinations about the effect these promot-
ers have on the rod in terms of sample preheating. It has been shown in previ-
ous studies 关5–8兴 that metals are more flammable at elevated temperatures;
therefore, it is important to understand the preheating of the rods due to the
promoter when establishing burn criteria. Testing was proposed to determine
the distance that the promoter preheats the metal rod prior to molten drops
forming. The test sample configurations were chosen to replicate ASTM G124
关2兴 with the exception that thermocouples were inserted along the sample at
various depths and intervals. The distance heated by the promoter just prior to
promoter detachment or consumption is referred to as the promoter heat af-
fected zone 共HAZ兲.

Experimental

Test System
National Aeronautics and Space Administration 共NASA兲 Johnson Space Center
White Sands Test Facility 共WSTF兲 is equipped with a promoted combustion test
chamber in its Hazardous Fluids Test Area. This chamber has a maximum
allowable working pressure of 79.2 MPa 共11,500 psi兲 and is used primarily for
ASTM G124 standard promoted combustion testing. The gaseous oxygen used
in the chamber is sampled to ensure it conforms to the requirements of MIL-
PRF-27210G 关9兴, with a minimum oxygen concentration of 99.5 %. This system
and its use are described in several previous publications 关2,3兴.
For the current work, one of the chamber viewports was modified to incor-
porate a multi-thermocouple feedthrough. This feedthrough allowed for up to
twelve temperature readings at test pressures up to 34.5 MPa 共5,000 psi兲. Addi-
tionally, a high-speed camera was set up to capture the ignition event. Tempera-
ture, pressure, and high-speed video were recorded for each test. Temperature
and pressure data were recorded at a 50 or 100 Hz sample rate 共initial tests
were recorded at 50 Hz, the remainder at 100 Hz for better temperature data
resolution兲. Burning was observed via high-speed video, which was used to
establish the event duration between the ignition of the promoter and the time
the initial molten-metal drop separated from the sample or was consumed.
This event duration was then synchronized with sample temperature data for
each test.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
82 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1—Sample configuration and test matrix.

Sample Material Promoter Materiala Pressure MPa 共psi兲 Number of Tests


Copper Magnesium 0.7 共100兲 1
Monel®b400 Magnesium 0.7 共100兲 1
316 SS Magnesium 0.7 共100兲 1
Copper Pyrofuze®c 3.5 共500兲 1
Copper Magnesium 3.5 共500兲 4
Copper Aluminum 3.5 共500兲 4
Copper Magnesium 6.9 共1000兲 3
Copper Aluminum 6.9 共1000兲 3
Copper Magnesium 34.5 共5000兲 4
Copper Aluminum 34.5 共5000兲 3
a
Magnesium and aluminum promoters were initiated using Pyrofuze wire.
b
Monel® is a trademark of Inco Alloys International, Inc., Huntington, West Virginia.
c
Pyrofuze® is a registered trademark of Sigmund-Cohn Co., Mount Vernon, New York.

Sample Configuration and Test Matrix


Test samples were configured as 305 mm 共12 in.兲 long, 3.2 mm 共0.125 in.兲 di-
ameter rods. Three materials were tested: 316 stainless steel, Monel®5 400, and
commercially pure copper. Table 1 shows the test matrix for each material.
Copper was chosen as the material for the majority of the testing based on its
high thermal conductivity and resistance to burning at elevated pressures and
concentrations of oxygen.
All test samples were instrumented with Type-K bare-wire beaded-type
thermocouples. After the thermocouple holes were drilled, samples were
washed in a liquid detergent bath at 49– 66° C 共120– 150° F兲, immersed for 5
min, agitated for 1 min, and scrubbed with a nylon brush as required. They
were then agitated for 2 min with hot, deionized water until no visual evidence
of detergent solution was evident, and blown dry with gaseous nitrogen. Ther-
mocouple lead wires were 0.13 mm 共0.005 in.兲 in diameter. Typical samples had
ten thermocouples staked into 0.254–0.381 mm 共0.010–0.015 in.兲 diameter
holes with a depth of 0.254–0.508 mm 共0.010–0.020 in.兲. Thermocouples were
peened in place on the test sample rods as illustrated in Fig. 1. Spacing of
thermocouples was typically as shown in Fig. 2, with thermocouple placement
distances measured from the bottom of the rod. Some test samples varied in
that thermocouples holes were drilled at a depth of 1.27–1.78 mm 共0.050–0.070
in.兲. This was done to investigate the temperature deviation over the cross sec-
tion of the rod. The length of engagement of the promoter onto the bottom of
the rod varied. This was not by design, but rather by observation, after several
tests had been conducted. As this deviation could affect test results, promoter
engagement length was noted for each sample thereafter.
As shown in Fig. 3, dual-bore ceramic tubing and ceramic cement were
used on the lower seven thermocouples to protect them during ignition of the

5
Monel® is a trademark of Inco Alloys International, Inc.,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Huntington, West Virginia.
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 83

FIG. 1—Thermocouple sample interface.

promoter. Thermocouples TC-1 through TC-4 had a temperature range of


0 – 1 , 000° C 共32– 1 , 832° F兲. Thermocouples TC-5 through TC-7 had a tempera-
ture range of 0 – 500° C 共32– 932° F兲. Although this scheme worked well, it was
a result of thermocouple module 共isolation and amplification兲 availability and
was not a test requirement.

Test Procedure
Each sample was mounted to the sample support, and the thermocouple leads
terminated at the appropriate female thermocouple plug. These female plugs
were vertically mounted on the sample stand and numbered 1–10 to mate to
corresponding data channels. The sample or promoter was then wound with
eight wraps of Pyrofuze, inserted into the test chamber, and the thermocouple
and Pyrofuze wires were appropriately terminated. Nextel®6 ceramic fabric
sleeves were used to shield the thermocouple insulation from thermal damage
including igniting and burning. A brass plate was also used between the sample
and the shielded thermocouple wires for further protection. Thermocouple

6
Nextel® is a trademark of 3M Company, St. Paul, Minnesota.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
84 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Thermocouple spacing along test sample.

plug connections and shielding measures are shown in Fig. 4.


The chamber was closed and purged three times with test atmosphere at
3.5 MPa 共500 psi兲 to remove any residual gasses. The chamber was then pres-
surized to the desired test pressure. Temperature and pressure data were re-
corded along with high-speed video.

Test Results
High-speed video was used to determine the event duration, i.e., the duration
from the ignition of the Pyrofuze igniter wire to the first significant drip of
molten metal, promoter detachment, or consumption of the promoter 共if the
promoter did not drop off the sample兲. Since no appreciable energy was added
to the sample after the end of the event duration, it was assumed that the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 85

FIG. 3—Ceramic shielding on lower seven thermocouples.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
86 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Thermocouple plug connections and ignition shielding.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 87

TABLE 2—Material temperature comparison at 0.7 MPa.


Promoter Material
Pressure 共MPa兲 Magnesium 0.7
Sample Material 316 SS Monel® 400 Copper
Event Duration: Ignition to Drop 共ms兲 1,050 1,670 700
Thermocouple Placement 共mm兲 Temperatures at End of Event Duration 共°C兲
TC-1 共7.6兲 ¯ 937 732
TC-2 共11.7兲 124 247 408
TC-3 共16.3兲 78 108 248
TC-4 共21.1兲 60 71 107
TC-5 共26.2兲 52 58 67
TC-6 共31.8兲 50 54 54
TC-7 共43.8兲 43 43 42
TC-8 共94.0兲 41 35 35
TC-9 共144.8兲 31 32 31
TC-10 共195.6兲 37 33 30

sample would ignite and begin burning or quench and solidify after this event.
The event duration is thus considered to be the controlling time used when
assessing the temperature data from the thermocouples. To correctly plot event
duration on a time versus temperature graph, thermocouple response lag time
was considered and compensated for. Thermocouple response time was calcu-
lated based on the outside diameter of the bare-wire thermocouples 关14兴. Un-
less this thermocouple response time of 0.135 s is accounted for, the tempera-
ture along the rod at promoter detachment will be incorrect. When this
calculated response lag is applied to the data, its accuracy is verified in that the
ignition event shifts and lines up with the instant the thermocouple’s tempera-
ture rises occur. Heat losses due to thermocouple leads acting as heat sinks and
thermal effects of the ceramic shielding and cement were not addressed. Given
the large quantity of heat released by the burning metal test sample, the mag-
nitude of heat sink afforded by the thermocouples was judged to be negligible.
This conclusion was based on a parametric evaluation of the cross sectional
area and thermal conductivity of the materials. This evaluation indicated that
less than 5 % of the heat conducted from the burning promoter into the test
sample may be transferred into the seven thermocouples near the promoter
共see Fig. 3兲 and their ceramic insulators. In fact, if only the first three thermo-
couples are considered 共the maximum number of thermocouples measuring
near or above 260° C 共500° F兲 in the copper rod data recorded in Table 2兲 then
the loss by conduction is less than 2 %. Figures 5–7 show typical time vs.
temperature plots for copper samples with aluminum promoters, with the be-
ginning and end of the event duration shown as vertical lines. Test data in its
entirety is published in a WSTF Special Test Data Report 关10兴.
Initial tests exhibited some irregularities in the temperature data. Mainly,
TC-1 and TC-10 showed unrealistic oscillations and negative temperature read-
ings. In subsequent tests, these irregularities were corrected by electrically iso-
lating the samples from the sample stand.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
88 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—Copper sample with an aluminum promoter in gaseous oxygen at 3.5 MPa.

FIG. 6—Copper sample with aluminum promoter in gaseous oxygen at 6.9 MPa.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 89

FIG. 7—Copper sample with an aluminum promoter in gaseous oxygen at 34.5 MPa.

Discussion and Application of the Data


Although the majority of the test samples were copper rods, general observa-
tions can be made with regard to other test materials. As expected, due to their
lower thermal conductivity, 316 stainless steel and Monel 400 samples exhib-
ited a less severe promoter HAZ 共Table 2兲. It can be seen that the event dura-
tions for the stainless steel and Monel samples are larger than the copper
sample 共for a given pressure兲, giving the promoter more time to transfer heat to
the sample. However, the copper sample still exhibited higher temperatures
along TC-2 through TC-5. The high temperature for TC-1 of the Monel sample
is most likely due to the convective effects of the initial ignition of the promoter.
Theoretical heat transfer studies of the promoter HAZ have also been investi-
gated 关11兴 and, as expected, complement and agree with this experimental data.
It is also important to note that ASTM G124 recommends using the types of
promoters chosen in this experiment but does not require them to be used.
Where other types of promoters are chosen the promoter HAZ may vary and
additional tests may need to be done to verify severity.
As expected, it was also observed that promoters burn more slowly and
exhibit longer resident times when ignited at lower pressures 共Table 3兲. This
results in higher sample temperatures at the end of the event duration. Con-
versely, at elevated pressures promoters burn and detach from the sample
much more quickly and have less of a preheating effect.
To determine the promoter HAZ, it must be understood what amount of
sample preheating, in terms of temperature, is necessary for burning charac-
teristics to be affected. Data from high-temperature promoted ignition testing
关5兴 suggests that for several 300-series stainless steels, samples heated to 260° C
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
90 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 3—Average event duration with respect to test pressure.

Pressure MPa 共psia兲 Duration 共ms兲


3.5 共500兲 446
6.9 共1000兲 328
34.5 共5,000兲 140

共500° F兲 and below show little or no variation in flammability when compared


to ambient temperature testing. Changes in flammability would be noted by
changes in either the pressure where a metallic sample will support burning, or
the rate at which the sample melts while it is burning 共both are indicators of
metals flammability and are used historically to compare metals兲. Therefore,
260° C 共500° F兲 was chosen as a reasonable temperature to use in determining
what length of the sample was within the HAZ. The HAZ is considered the
length of the rod measuring temperatures greater than or equal to 260° C
共500° F兲 just as the ignition promoter is detaching from the test sample. Deter-
mining the location of this temperature condition allows accurate delineation
of the HAZ within the rod and the production of burn criteria beyond which
there are no further effects of the ignition promoter on the burning of the test
sample.
Temperatures at the end of the ignition events were used to determine the
location, from the bottom of the rod, where the samples reached 260° C
共500° F兲. For each test pressure series, this location was compiled and standard
deviations applied 共see Fig. 8兲. This shows that, including standard deviation
error bars, the promoter HAZ does not extend beyond 17.8 mm 共0.70 in.兲 above

FIG. 8—Distance along rod where 260° C was reached at promoter detachment at vari-
ous test pressures.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 91

the end of the rod. It should be noted that this distance includes the promoter
engagement distance, which is not included when reporting burn lengths as
instructed in ASTM G124. It was also observed that promoter material 共alumi-
num or magnesium兲 did not greatly affect the length of the HAZ produced
within the test sample 共for the conditions of these tests兲.
The mean value and standard deviation between the test pressure group
with the largest promoter HAZ was then found using the central limit theorem
共CLT兲, and a three sigma distribution was applied. Using the CLT assumes there
is a normal data distribution, and thus ⬃99.7 % of the data values will be
within three standard deviations of the mean. Using the most severe condition
in terms of heat generation/conduction 共3.5 MPa 共500 psi兲兲, and combining
both aluminum and magnesium promoter sample data, yields a mean HAZ
distance of 16 mm 共0.63 in.兲 with a standard deviation of 1 mm 共0.041 in.兲.
Applying three standard deviations results in a promoter HAZ of 19 mm 共0.75
in.兲 from the bottom of the rod 共Fig. 9兲.
Based on these data, which have also been validated by thermal modeling
关11兴, it was discussed and decided within NASA that samples which burn more
than 30 mm 共1.18 in.兲 above the promoter will be considered a burn 关1兴. To
come to this length of 30 mm 共1.18 in.兲 rather than the promoter-included
length of 19 mm 共0.75 in.兲, as determined in testing, roughly 50 % of the actual
HAZ was added as a conservative buffer. This ensures all metals that burn a
distance greater than 30 mm 共1.18 in.兲 are burning independently of any pro-
moter effects that preheated the test sample during the ignition process. For
metallic materials that have a lower thermal conductivity than copper 共that is,
most metals兲, burning a distance greater than 30 mm 共1.18 in.兲 is a large dis-
tance beyond the HAZ, and clearly the metals are burning.
The experimental observations and subsequent conclusion of the promoter
HAZ modifies the historical views of metals flammability in the specific appli-
cation of ASTM G124, as well as other tests that often require the test sample to
be consumed entirely or halfway to be considered flammable. This result is also
specific to the igniter materials tested. Copper was chosen as the major test
material based on several factors, most significantly its resistance to burning
and high thermal conductivity. A high thermally conductive material was de-
sired to ensure effective heat transfer from the promoters, leading to large
HAZs. Metals with lower thermal conductivities will have smaller HAZs. Addi-
tionally, since the promoter HAZ in this experiment is calculated using a 260° C
共500° F兲 temperature bias, it should be noted that, with further research into
the effects of metals flammability with respect to elevated temperature, the
promoter HAZ may need to be modified accordingly. For previously tested ma-
terials 共that use the same type of promoter兲 where burn lengths were recorded,
these burn criteria can be easily applied. However, since these burn criteria are
markedly different from the previous criteria, flammability of many materials
may no longer be defined. As the required number of tests to rank a material
are no longer met, additional testing may be needed. ASTM requires five no-
burn tests for a material to be considered nonflammable, NASA requires ten
关1,2兴. Additionally, after applying these burn criteria to existing material data,
many previously considered metals have not been tested at a nonflammable
condition. A small sample set 共see Table 4兲 shows some metals that are now
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
92 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 9—HAZ produced from ignition promoter after applying the CLT using a 260° C
共500° F兲 temperature limit.

considered flammable at pressures that were previously considered nonflam-


mable, and consequently there are insufficient data to establish a nonflam-
mable condition without further testing.

Conclusions
The goal of this testing was to characterize the heating effects of an ignition
promoter on a copper sample as is typically configured in promoted combus-
tion flammability testing. Furthermore, the goal was to apply these findings to
empirically establish better suited flammability criteria that incorporate the
heating effect of the ignition promoter on the test sample. It was determined
that the largest HAZ produced will likely be approximately 19 mm long 共or less兲
on a copper rod, and if a test sample burns a distance longer than this, it is
clearly burning independently of the ignition-promoter effect. The success of
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 4—Flammability of select metals using 30 mm burn criteria.

Nonflammability Pressure using Previous Cri- Flammability Pressures using 30 mm Criteria


Material teria MPa 共psia兲 关12兴 MPa 共psia兲 关13兴

Downloaded/printed by
Flammable at 20.7 共3000兲, no lower pressures
Haynes®a 188 20.7 共3,000兲 tested
Flammable at 17.2 共2500兲, no lower pressures
Hastelloy®a C22 20.7 共3,000兲 tested
Flammable at 10.3 共1500兲, no lower pressure
Elgiloy®b 10.3 共1,500兲 tested
Flammable at 3.5 共500兲, no lower pressure
Inconel®c 718 5.2 共750兲 tested
Flammable at 3.5 共500兲, no lower pressure
17-4 PH SS 3.5 共500兲 tested
a ® ®
Haynes and Hastelloy are registered trademarks of Haynes Stellite Co., Kokomo, Indiana.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
b
Elgiloy® is a registered trademark of Elgin National Watch Co., Elgin, Illinois.
c
Inconel® is a registered trademark of Inco Alloys International, Huntington, West Virginia.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


SPARKS ET AL., doi:10.1520/JAI102351 93
94 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

this testing allows scientific burn criteria to be established within NASA stan-
dards. In accomplishing this goal, there were many observations that aid in the
understanding of the intricacies of promoted combustion testing 共e.g., ignition
event duration with respect to test pressure, the severity of magnesium and
aluminum promoters on the test sample with respect to the gaseous oxygen
environment兲.

References

关1兴 NASA Technical Standard NASA-STD-共I兲-6001A, 2008, “Flammability, Offgassing,


and Compatibility Requirements and Test Procedures,” NASA Headquarters,
Washington, D.C.
关2兴 ASTM G124, 1995, “Standard Test Method for Determining the Combustion Be-
havior of Metallic Materials in Oxygen-Enriched Atmospheres, ASTM Standards
Related to Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres,” ASTM International, West Conshohocken, PA.
关3兴 ISO 14624-4:2003, 2003, “Determination of Upward Flammability of Materials in
Pressurized Gaseous Oxygen or Oxygen-enriched Environments,” Space Systems—
Safety and Compatibility of Materials, International Organization for Standardiza-
tion 共ISO兲, Geneva, Switzerland.
关4兴 De Wit, J. R., Steinberg, T. A., and Stoltzfus, J. M., “Igniter Effects on Metals
Combustion Testing,” Flammability and Sensitivity of Materials in Oxygen-Enriched
Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. Newton, and H.
D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关5兴 Slockers, M. J. and Robles-Culbreth, R., “Ignition of Metals at High Temperatures
in Oxygen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: 11th Volume, STP 1479, D. B. Hirsch, R. Zaweirucha, T. A. Steinberg, and
H. M. Barthelemy, Eds., ASTM International, West Conshohocken, PA, 2006, pp.
62–79.
关6兴 Engel, C. D., Herald, S., and Davis, E., “Promoted Metals Combustion at Ambient
and Elevated Temperatures,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: 11th Volume, STP 1479, D. B. Hirsch, R. Zawierucha, T. A.
Steinberg, and H. M. Barthelemy, Eds., ASTM International, West Conshohocken,
PA, 2006, pp. 51–61.
关7兴 Zaweirucha, R. and Million, J. F., “Promoted Ignition-Combustion Behavior of
Engineering Alloys at Elevated Temperatures and Pressures in Oxygen Gas Mix-
tures,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Ninth Volume, STP1395, T. A. Steinberg, B. E. Newton, and H. D. Beeson, Eds.,
ASTM International, West Conshohocken, PA, 2000.
关8兴 Sato, J., and Hirano, T., 1986, “Behavior of Fires Spreading Along High-
Temperature Mild Steel and Aluminum Cylinders in Oxygen,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Second Volume, STP 910,
M. A. Benning, Ed., ASTM International, West Conshohocken, PA, pp. 118–134.
关9兴 MIL-PRF-27210G, 1997, “Oxygen, Aviator’s Breathing, Liquid, and Gas,” Military
Performance Specification, U.S. Department of Defense, Washington, D.C.
关10兴 Stoltzfus, J. and Sparks, K., “Promoted Combustion Promoter Heat Affect Zone
Testing.” Special Test Data Report No. WSTF # 08-43113, NASA Johnson Space
Center White Sands Test Facility, Las Cruces, NM 共to be published兲.
关11兴 Steinberg, T., Lynn D., Sparks, K., Stoltzfus, J., and Smith, S., “Defining Self-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102351 95

Sustained Burning of Cylindrical Metal Rods through Characterization of the


Thermal Effects of the Ignition Promoter,” Flammability and Sensitivity of Materi-
als in Oxygen-Enriched Atmospheres: Twelfth Volume, ASTM STP 共to be published兲.
关12兴 Beeson, H., Stewart, W., and Woods, S., Safe Use of Oxygen and Oxygen Systems:
Handbook for Design, Operation, and Maintenance, ASTM International, West Con-
shohocken, PA, 2000.
关13兴 Beeson, H., Smith, S., and Stewart, W., Safe Use of Oxygen and Oxygen Systems:
Handbook for Design, Operation, and Maintenance, 2nd ed., ASTM International,
West Conshohocken, PA, 2007.
关14兴 Omega Engineering “Comparison of Time Constant vs. Overall Outside Diameter
of Bare Thermocouple Wires or Grounded Junction Thermocouples in Air,” Omega
Complete Temperature Measurement Handbook and Encyclopedia MMV Volume, 5th
ed., Omega Engineering Inc., Stamford, CT, 2004, pp. 51–52.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102300
Available online at www.astm.org/JAI

Greg A. Odom,1 Gwenael J. A. Chiffoleau,2 Barry E. Newton,3


and J. Ron Fielding4

Promoted Ignition Testing of Metallic Filters


in High-Pressure Oxygen

ABSTRACT: Currently, no test standard exists for evaluating the ignition tol-
erance and fault tolerance of metallic filters in high-pressure oxygen. Filters
are a critical component in oxygen systems to ensure system cleanliness
and mitigate ignitions by particle impact and contamination. However, filters
are at risk to these same ignition mechanisms and fires have occurred in
service. A new test method was developed using ASTM Standard G175
Phase 2 as a basis. The test subjects a pre-contaminated filter to a forced
ignition event using an ignition pill while the filter is maintained at elevated
pressure. Prior to testing, contaminant was applied to the filter element and
was also placed at the filter inlet. This additional contaminant was based on
contaminant that could potentially accumulate in a filter over time. This con-
taminant consisted of aluminum powder, iron particles, and perfluorinated
lubricant. An ignition pill, consistent with ASTM Standard G175 Phase 2, was
located on the upstream side of the filter. A back pressure was applied down-
stream of the ignition pill to ensure that the filter was pressurized during the
ignition and burning of the pill. Testing was performed on brass and stainless
steel filters of the same design using an oxygen shock to ignite the ignition
pill at a test pressure greater than the back pressure applied to the filter. The
brass filters safely contained the ignition event without breaching through the
filter element or body. For the stainless steel filters, the ignition event kindled

Manuscript received December 12, 2008; accepted for publication June 12, 2009; pub-
lished online August 2009.
1
Standard Test Manager, Wendell Hull & Associates, Inc., 5605 Dona Ana Rd., Las
Cruces, NM 88007.
2
Senior Scientist and Test Facility Manager, Wendell Hull & Associates, Inc., 5605 Dona
Ana Rd., Las Cruces, NM 88007.
3
VP R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM
88007.
4
President, Chase Filters and Components, LLC, 1003 48th St., Newport News, VA
23607.
Cite as: Odom, G. A., Chiffoleau, G. J. A., Newton, B. E. and Fielding, J. R., ⬙Promoted
Ignition Testing of Metallic Filters in High-Pressure Oxygen,⬙ J. ASTM Intl., Vol. 6, No. 8.
doi:10.1520/JAI102300.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 96
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ODOM ET AL., doi:10.1520/JAI102300 97

the filter element and burned through the filter body. This testing showed that
the ignition fault tolerance of the brass filters was far superior to that of the
stainless steel filters, which was consistent with the relative flammability of
these metallic materials, and therefore verified the test methodology.

Introduction
Within medical and industrial oxygen systems, it is crucial to ensure that all
components are compatible with oxygen. With careful selection of components
and materials, the risk of ignition within an oxygen system can be minimized.
When selecting a component for oxygen service, there are various ignition
mechanisms that should be considered. The ignition mechanisms include en-
ergy release from a second material 共promoted ignition兲, particle impact, flow
friction, resonance, electrical energy, and adiabatic compression 关1兴. In the case
of filters within oxygen systems, promoted ignition is a serious concern. Since
the primary function of filters is to collect particulate and contaminates from a
system, they are inherently susceptible to contaminant promoted ignition. Cur-
rently, there is no test standard for performing promoted ignition testing on
filters. ASTM G175 Phase 2 is a promoted ignition test for medical regulators.
The objective of this test is to determine the fault tolerance of oxygen regulators
when subjected to a promoted ignition event. A new test method for filters was
established using ASTM G175 Phase 2 as a guideline. The ASTM G175 Phase 2
procedure was modified to better simulate the actual service conditions of fil-
ters in service. The following paper presents the test design and procedure as
well as the results from testing on various filters.

Background
There is neither a test standard for performing promoted ignition testing on
filters nor is there any known previous data for this kind of testing. There have
been ignitions in the field involving filters in industrial gas pipelines 关2兴, oxygen
boosters, and breathing gas systems 共Fig. 1兲, in which the ignition event propa-
gated though the body of the filter and resulted in a catastrophic fire. These
ignition events can and have resulted in damage to components and systems.
The objective of this testing was to develop a method to evaluate the fault
tolerance of filters to promoted ignition.

Test Set-up
The test system design was consistent with the adiabatic compression test sys-
tem depicted in ISO 15001 Annex E 关3兴 and previously reported in ASTM STP
1319 关4兴. It consisted of an accumulator to supply heated 共⬃60° C兲, high-
pressure oxygen to a rapid opening valve 共impact valve兲. Prior to testing, cali-
bration cycles were performed on the test system to ensure that the required
pressure rise time of 15–20 ms was achieved at the end of a capped, 750-mm-
long, 14-mm inside-diameter tube. A test article was affixed to the test system at
an interface located a distance of 750 mm from the impact valve. A vent valve,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
98 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Condition of a breathing gas booster filter after an ignition in the field.

to relieve pressure in the test system, was located between the impact valve and
the test article. A high frequency response pressure transducer was provided to
ensure that the required test article pressurization rate was achieved for each
calibration cycle. Pressurization rates were monitored and recorded during
each pressure shock. The test was performed under computer control and data
was recorded digitally.
The filter test articles required preparation before testing. A standard
ASTM G175 ignition pill was installed immediately upstream of the filter inlet.
Additional contaminant was applied to the filter element and placed at the filter
inlet. This additional contaminant was based on contaminant potentially
present in a filter after a typical service life. A portion of the filter element for
each test article was contaminated with perfluorinated lubricant, aluminum,
and iron. A thin layer of lubricant was placed on the filter element and then
approximately one-half of the total aluminum and iron particles were spread
onto the filter element. Prior to the testing, the other one-half of the aluminum
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ODOM ET AL., doi:10.1520/JAI102300 99

and iron contaminant was poured into the filter inlet. A blank nylon disk 共ap-
proximately 82 mg兲 was placed just downstream of the ignition pill. This blank
nylon disk was used to isolate the standard ignition pill from the back pressure
that was applied within the filter before the impact event. The back pressure of
approximately 13.8 MPa 共2,000 psi兲 was applied to the filters to ensure that the
filters were pressurized during the time frame at which the fire from the igni-
tion pill entered the filters. A 1-mm 共0.040-in.兲 orifice was placed just down-
stream of the filter outlet. This orifice was used to simulate the flow rate of a
typical flow controlling component that would be installed downstream of
these filters in service. The filters were impacted with oxygen at the desired test
pressure. The oxygen pressure surge was applied to the filter inlet, thus igniting
the ignition pill.

Test Results
Table 1 provides further details regarding the test pressure, additional contami-
nant, and other specifics for each filter tested. Twelve 共12兲 filters were subjected
to positive ignition testing according to the requirements of a modified ASTM
G175 Phase 2 test procedure. Testing was performed on the following test ar-
ticles: five 共5兲 brass inline filters, five 共5兲 brass tee type filters, one 共1兲 stainless
steel inline filter and one stainless steel tee type filter. The filters were exposed
to the ignition and combustion of a standardized ignition pill, a typical 82-mg
nylon ignition pill blank and the additional contaminant applied to and added
to the filter.

Stainless Steel Filters


For the stainless steel inline and tee type filters, sustained promoted ignition of
the stainless steel filters bodies occurred causing a highly energetic fire release.
Pre- and post-test photographs of the stainless steel inline and tee type filters
are shown in Figs. 2–7. Based on the results observed during this testing, all of
the stainless steel filters 共inline and tee type兲 were judged to have failed the
requirements of the promoted ignition filter test because the filter body failed to
contain the ignition event.

Brass Filters
For the brass inline and tee type filters, sustained promoted ignition of the filter
bodies and internal components did not occur. Some heat deterioration and
melting on each test article was observed. These changes were due to the en-
ergy delivered by the pill and the additional contaminant that was consumed.
No external flame or breach of the pressurized filter was observed during the
testing. The filter element was not breached and therefore prevented slag and
particulate associated with the ignition event, from migrating downstream of
the filter. Post-test photographs of the brass inline and tee type filters are shown
in Figs. 8–14. Based on the results observed during this testing, all of the brass
filters 共inline and tee type兲 were judged to have successfully passed the require-
ments of the promoted ignition filter test. It is noteworthy that these brass
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Filter promoted ignition test data.

Qty. Contaminate Added to Test


Test Article Tested the Filter Element and Inlet Pressure Result

Downloaded/printed by
Brass inline filter 5 100 mg aluminum powder, 250 mg iron 34.5 MPa Ignition pill and contaminant ignited
particles inserted at the filter inlet. A thin 共5,000 psi兲 but did not kindle the filter element
layer of lubricant was smeared on the and did not propagate through the
filter element. filter body.
Brass tee filter 5 1 g aluminum powder, 2.5 g iron particles 34.5 MPa Ignition pill and contaminant ignited
inserted at the filter inlet. A thin layer of 共5,000 psi兲 but did not kindle the filter element
lubricant was smeared on the filter element. and did not propagate through the
filter body.
Stainless steel inline filter 1 100 mg aluminum powder, 250 mg iron 34.5 MPa Ignition pill and contaminant ignited
particles inserted at the filter inlet. A thin 共5,000 psi兲 and propagated through the filter body.
layer of lubricant was smeared on the

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
filter element.
Stainless steel tee filter 1 1 g aluminum powder, 2.5 g iron particles 20.7 MPa Ignition pill and contaminant ignited
inserted at the filter inlet. A thin layer of 共3,000 psi兲 and propagated through the filter body.
lubricant was smeared on the filter element.
Note:
article.共1兲 All tests were performed with a back pressure of 13.8 MPa 共2,000 psi兲 on the filter. 共2兲 A standard ASTM G175 ignition pill and

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


an 82-mg nylon blank were placed just upstream of the filters, pretest. 共3兲 There was a 1-mm 共0.04 in.兲 orifice downstream of the test
100 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
ODOM ET AL., doi:10.1520/JAI102300 101

FIG. 2—Pre-test photo of a stainless steel tee type filter.

filters also endured the ignition event without breaching the filter element,
which is above and beyond the requirement of this testing.

Conclusion
Promoted ignition testing was performed on brass and stainless steel filters.
The contaminant applied to the filters prior to testing was based on the amount

FIG. 3—Post-test condition of the stainless steel tee type filter 共View 1兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
102 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Post-test condition of the stainless steel tee type filter 共View 2兲.

and type of contaminant that could typically be found within filters that are out
in the field. The promoted ignition test results indicated that the brass filters
were able to contain the ignition event without breaching the filter element or
body. The stainless steel filters were not able to contain the ignition event dur-
ing the promoted ignition testing. The ignition event propagated the stainless
steel filter element as well as the stainless steel body, eventually burning
through the filter body. As previously determined, these findings are consistent

FIG. 5—Pre-test photo of a stainless steel inline filter.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ODOM ET AL., doi:10.1520/JAI102300 103

FIG. 6—Post-test condition of a stainless steel inline filter 共View 1兲.

FIG. 7—Post-test condition of a stainless steel inline filter 共View 2兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
104 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 8—Pre-test photo of a brass tee type filter.

FIG. 9—Post-test condition of a brass tee type filter.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ODOM ET AL., doi:10.1520/JAI102300 105

FIG. 10—Post-test condition of a brass tee type filter element 共View 1兲.

FIG. 11—Post-test condition of a brass tee type filter element 共View 2兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
106 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 12—Pre-test photo of a brass inline filter.

FIG. 13—Post-test condition of a brass inline filter 共View 1兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ODOM ET AL., doi:10.1520/JAI102300 107

FIG. 14—Post-test condition of a brass inline filter 共View 2兲.

with the difference in flammability between these two alloys 关5兴. Therefore, a
new test was developed that was capable of evaluating the ignition tolerance of
filters for high-pressure oxygen service.

Acknowledgments
An acknowledgment should be made to Chase Filters and Components for rec-
ognizing the need for a testing methodology for filters and for working with
Wendell Hull & Associates, Inc. to develop an applicable promoted ignition test
procedure. All brass filters that were referenced in this paper were designed and
manufactured by Chase Filters.

References

关1兴 Mcllroy, K., Zawierucha, R., and Drnevich, R. F., “Promoted Ignition Behavior of
Engineering Alloys in High Pressure Oxygen,” Flammability and Sensitivity of Ma-
terials in Oxygen-Enriched Atmospheres: Third Volume, ASTM STP 986, D. W.
Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp. 85–104.
关2兴 Colson, A. and Fano, E., “Filtration of Gaseous or Liquid Oxygen in Industrial
Applications and Associated Fire Risks,” J. ASTM Int., Vol. 3共4兲, 2006, pp. 301–313.
关3兴 ISO 15001, 2003, “Anesthetic and Respiratory Equipment-Compatibility with Oxy-
gen,” International Organization for Standardization 共ISO兲, Geneva, Switzerland.
关4兴 Newton, B., Porter, A., Hull, W. C., Stradling, J., and Miller, R., “A 6000 psig
Gaseous Oxygen Impact Test System for Materials and Components Compatibility
Evaluations,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997.
关5兴 Stoltzfus, J. M., Homa, J. M., Williams, R. E., and Benz, F. J., “ASTM Committee
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
108 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

G-4 Metals Flammability Test Program: Data and Discussion,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Volume, ASTM STP
986, D. W. Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp.
28–53.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102244
Available online at www.astm.org/JAI

Stephen D. Herald,1 Paul M. Frisby,2 and Samuel Eddie Davis3

Oxygen Compatibility of Brass-Filled PTFE


Compared to Commonly Used
Fluorinated Polymers for Oxygen
Systems

ABSTRACT: Safe and reliable seal materials for high-pressure oxygen sys-
tems sometimes appear to be extinct species when sought out by oxygen
systems designers. Materials that seal well are easy to find, but these ma-
terials are typically incompatible with oxygen, especially in cryogenic liquid
form. This incompatibility can result in seals that leak, or much worse, seals
that easily ignite and burn during use. Scientists at the Materials Combustion
Research Facility 共MCRF兲, part of NASA/Marshall Space Flight Center, are
constantly searching for better materials and processes to improve the safety
of oxygen systems. One focus of this effort is improving the characteristics of
polymers used in the presence of an oxygen enriched environment. Very few
systems can be built which contain no polymeric materials; therefore, mate-
rials which have good impact resistance, low heat of combustion, high auto-
ignition temperature, and those that maintain good mechanical properties
are essential. The scientists and engineers at the MCRF, in cooperation with
seal suppliers, are currently testing a new formulation of polytetrafluoroeth-
ylene 共PTFE兲 with brass filler. This brass-filled PTFE is showing great prom-
ise as a seal and seat material for high-pressure oxygen systems. Early
research has demonstrated very encouraging results, which could rank this

Manuscript received November 21, 2008; accepted for publication June 15, 2009; pub-
lished online August 2009.
1
Test Group Lead Engineer, InfoPro Corp., Materials Test Branch, Materials Combus-
tion Research Facility, NASA-George C. Marshall Space Flight Center, Huntsville, AL
35812.
2
Materials Engineer, InfoPro Corp., Materials Test Branch, Materials Combustion Re-
search Facility, NASA-George C. Marshall Space Flight Center, Huntsville, AL 35812.
3
Materials Engineer, Materials Test Branch, Materials Combustion Research Facility,
NASA-George C. Marshall Space Flight Center, Huntsville, AL 35812.
Cite as: Herald, S. D., Frisby, P. M. and Davis, S. E., ⬙Oxygen Compatibility of
Brass-Filled PTFE Compared to Commonly Used Fluorinated Polymers for Oxygen
Systems,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102244.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 109
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
110 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

material as one of the best fluorinated polymers ever tested. This paper will
compare the data obtained for brass-filled PTFE with other fluorinated poly-
mers, such as TFE-polytetrafluoroethylene 共PTFE兲, polychlorotrifluoroethyl-
ene 81, Viton® A, Viton® A-500, Fluorel®, and Algoflon®. A similar metal-
filled fluorinated polymer, Salox-M®, was tested in comparison to brass-filled
PTFE to demonstrate the importance of the metal chosen and relative per-
centage of filler. General conclusions on the oxygen compatibility of this
formulation are drawn, with an emphasis on comparing and contrasting the
materials performance to the performance of the current state-of-the-art oxy-
gen compatible polymers.
KEYWORDS: oxygen systems, polytetrafluoroethylene, fluorinated
polymers, MAPTIS

Introduction
Safe and reliable seal materials for high-pressure oxygen systems sometimes
appear to be extinct species when sought out by oxygen systems designers.
Materials that seal well are easy to find, but these materials are typically incom-
patible with oxygen, especially in cryogenic liquid form. This incompatibility
can result in seals that leak, or much worse, seals that easily ignite and burn
during use. Materials that are compatible with oxygen are easy to find, such as
the long list of compatible metals, but these metallic materials are limiting as
seal materials. A material that seals well and is oxygen compatible has been the
big game in the oxygen system designer’s safari.
Scientists at the Materials Combustion Research Facility 共MCRF兲, part of
NASA/Marshall Space Flight Center 共MSFC兲, are constantly searching for better
materials and processes to improve the safety of oxygen systems. One focus of
this effort is improving the characteristics of polymers used in the presence of
oxygen. Very few systems can be built which contain no polymeric materials;
therefore, materials which have good impact resistance, low heat of combus-
tion, high auto-ignition temperature, and those that maintain good mechanical
properties are essential.

The Search for a Better Polymer


Polymeric materials are widely used as seals and seat materials in oxygen sys-
tems. One of the most frequent causes of fires in oxygen systems using poly-
meric materials is ignition of the polymer, which promotes to surrounding
materials and results in a catastrophic failure. According to previous studies 关1兴
the most compatible polymers in oxygen systems are fluorinated polymeric
materials such as polytetrafluoroethylene 共PTFE兲 and PCTFE. This measure of
compatibility is determined by favorable characteristics for oxygen use such as
high oxygen index 共ignition resistance兲, low heat of combustion, and high auto-
ignition temperature. This knowledge has allowed the MCRF, in conjunction
with other groups within MSFC, to begin to experiment with different PTFE
and PCTFE based polymers. From both internal experience and previous stud-
ies, metal-filled polymers seem to be the key to improved properties among
oxygen compatible materials.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HERALD ET AL., doi:10.1520/JAI102244 111

In addition to exceptional performance in enriched oxygen environments,


metal-filled polymers obtain improved electrical and thermal conductivity,
wear, friction, and altered percolation threshold properties. Mamunya et al.
studied the effects of polyethylene and polyoxymethylene polymers filled with
dispersed iron and copper 关2兴. The polymer composites were produced by sev-
eral forming methods: extrusion of homopolymer 共polyethylene or polyoxym-
ethylene兲 filled with iron, extrusion of the blend polyethylene/polyoxymethylene
filled with iron, compression molding of PE filled with dispersed copper, and
extrusion of PE filled with copper. It was found that the extruded metal filler
with homopolymer allows for the spatial disposition of filler particles close to
matrix distribution.
This method resulted in low conductivity and a high value of the percola-
tion threshold. The extrusion of metal filler in a polymer blend of the polyeth-
ylene and polyoxymethylene 共one having a much higher viscosity兲 allowed for
localization of filler in the low viscosity polymer. This combination resulted in
an ordered spatial distribution of metal particles and a low value of the perco-
lation threshold. The hot pressing of a powder mixture of the polymer-metal
provides good contact between metal particles; therefore, a high electrical con-
ductivity and low value of percolation threshold.
The use of nickel, copper, and lead as a filler in PTFE was studied for its
wear and friction properties by Zhang et al. 关3兴. The powder metal-filled poly-
mers were tested for friction and wear by sliding against GCr15 bearing steel
and then examined by the use of high power microscopes. Scanning Electron
Microscopy and optical microscopy investigations of the rubbing surfaces show
that metal fillers of copper, lead, and nickel not only raise the load carrying
capacity of the PTFE composites, but also promote transfer of the PTFE com-
posites onto the counterfaces, so they greatly reduce the wear of the PTFE
composites.

Existing Data Library


The data used to compare brass-filled PTFE to other polymers was obtained
from other ASTM papers as well as NASA’s Materials and Processes Technical
Information System 共MAPTIS兲. The MAPTIS database is housed at the NASA
George C. Marshall Space Flight Center and is available to the public, with
minimal limitations, at the website http://maptis.nasa.gov.

Experimental

Test Apparatus
The data discussed in this paper was obtained by the use of both in-house
equipment and the use of preexisting documented results. The Brass-filled
PTFE samples discussed in this paper were evaluated using a Parr 6200 oxygen
bomb calorimeter to determine the heat of combustion. Three other testers
were also required in order to provide necessary data. These were the in-house
ambient-pressure mechanical impact, high-pressure mechanical impact, and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
112 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

autogeneous ignition temperature 共AIT兲 testers. Heat of combustion, AIT, am-


bient and high-pressure mechanical impact experiments were conducted in
accordance with ASTM G-72, ASTM G-86, ASTM D-4809, and NASA-STD-6001,
respectively.

Procedure

Heat of Combustion—The measurement of the heat of combustion was cal-


culated internally by the software programmed on the Parr 6200, which was
verified by comparison to a benzoic acid standard supplied by Sigma Aldrich.
The heat of combustion sample was placed inside a metallic combustion cup,
and an ignition wire was positioned above the sample as to touch the top of the
sample but not come in contact with any metal other than the terminal leads.
In order to assure that combustion will occur, a mineral oil spike was used, and
for which the calculations corrected. The combustion chamber was sealed and
pressurized with gaseous oxygen 共GOX兲, lowered into the water bath, and the
ignition wire leads were attached to the top of the combustion chamber. After
the test was over, the calorimeter displayed the gross heat content of the sample
in calories per gram weight of sample.

Autogeneous Ignition Test—The samples for AIT were prepared per the
standard and placed into a small glass test tube. The test tube was placed into
the sample holder and a thermocouple was positioned inside the test tube near
the sample. The chamber was purged several times with GOX to ensure a 100%
oxygen environment prior to pressurizing the chamber to the test pressure.
Once the test pressure was reached, the system was monitored in order to
assure that the chamber was properly sealed. The external heater was started
and allowed to increase the temperature of the sample at the rate of 5 ° F / min.
The indication of autogeneous ignition was the rapid increase of temperature
and pressure. If the system temperature rises above 425° C prior to ignition of
the sample, then the AIT is recorded as ⬎425° C.

Mechanical Impact Test—For ambient-pressure mechanical impact testing,


samples were positioned in the test chamber, which was cryogenically cooled
with nitrogen, and submerged in liquid oxygen 共LOX兲. The samples were im-
pacted with 98 J of energy 共72 ft· lb兲 while reactions were detected by the
operator in terms of a visual flash or audible noise. The high-pressure mechani-
cal impact test can be operated in a LOX or GOX mode, depending on the
parameters of the test. In order to achieve 100% oxygen inside the test chamber,
several lower pressure oxygen purges were required prior to the chamber being
pressurized to the full test pressure. Once the chamber was brought up to the
required test pressure, the plummet was allowed to impact the sample at 98 J of
energy 共72 ft· lb兲. The detection of a reaction was indicated by the illumination
of a photodiode, which was positioned outside of the chamber with an optical
viewing port. After several series of tests were conducted for both ambient and
high-pressure mechanical impact, the samples were analyzed for signs of com-
bustion that may not have been detected by the operator or equipment. The
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HERALD ET AL., doi:10.1520/JAI102244 113

TABLE 1—Heat of combustion results for fluorinated compounds.

Heat of Combustion Standard Deviation


Material 共kJ/g兲 共kJ/g兲
TFE-polytetrafluoroethylene 共PTFE兲a 6.35 0.102
polychlorotrifluoroethylene 81 共PCTFE兲a 10.70 0.072
Vinylidienefluoridehexafluropropylene A
共VDF/HFP兲a 16.71 0.075
Vinylidienefluoridehexafluropropylene A-500
共VDF/HFP兲a 16.47 0.046
Vinylidienefluoridehexafluropropylene
共VDF/HFP兲a 16.71 0.070
Brass-filled PTFEb 5.58 0.058
a
ASTM D328 Chou and Fiedorowicz.
b
NASA-STD-6001.

pass/fail criterion for a material, in both ambient and high-pressure mechani-


cal impact, is determined by the number of reactions recorded in a series of
tests. A sample is said to have passed the mechanical impact test at the given
pressure and temperature if no reactions are recorded out of 20 samples. If one
sample out of 20 reacts, then 40 additional samples must be tested without any
reactions in order for the material to pass the test.

Results and Discussion


Previous research conducted by Chou and Fiedorowicz examined several flu-
orinated polymeric materials that are of interest in the further characterization
of brass-filled PTFE including: TFE-polytetrafluoroethylene 共PTFE兲, polychlo-
rotrifluoroethylene 81 共PCTFE兲, Viton® A, Viton® A-500, and Fluorel®. In
Table 1 heat of combustion data for the fluorinated compounds is presented as
an average of three tests according the ASTM Standard D-3286 along with a
standard deviation calculated from the presented data. The heat of combustion
data for brass-filled PTFE presented in Table 1 was determined by the MCRF at
MSFC. The caloric content per gram of material for brass-filled PTFE was 12%
lower than the caloric content of the non-brass-filled PTFE determined by Chou
and Fiedorowicz. The reported values for the heat of combustion of TFE-
polytetrafluoroethylene 共PTFE兲 in the Safe use of oxygen and oxygen systems
vary substantially between the different test facilities with an average of 6.21
⫻ 103 共J / g兲 and standard deviation of 870.9 共J/g兲 关4兴. Although this value is
lower than that reported by Chou and Fiedorowicz, their value still falls within
the acceptable range of error.
The lower heat of combustion of the Brass-filled PTFE further exemplifies
the material’s compatibility in oxygen systems. The comparison of brass-filled
PTFE to the other fluorinated polymers studied by Chou and Fiedorowicz in
terms of heat of combustion, shows that brass-filled PTFE has a 48% lower
caloric content than polychlorotrifluoroethylene 81 共PCTFE兲 and a 66% lower
caloric content per gram than Viton® A, Viton® A-500, and Fluorel®. As de-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
114 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 2—AIT data of fluorinated compounds.

Initial Test Final Test


AIT Pressure Pressure
Material 共°C兲 共kPa兲 共kPa兲
TFE-polytetrafluoroethylene 共PTFE兲a ⬎425 1.034⫻ 104 1.517⫻ 104
polychlorotrifluoroethylene 81 共PCTFE兲a 404.3 1.034⫻ 104 1.517⫻ 104
Viton® Aa 239 1.034⫻ 104 1.379⫻ 104
Viton® A-500a 239 1.034⫻ 104 1.379⫻ 104
Fluorel® 296 1.034⫻ 104 1.448⫻ 104
Brass-filled PTFEb ⬎425 1.034⫻ 104 1.564⫻ 104
TFE-polytetrafluoroethylene 共PTFE兲b ⬎425 1.034⫻ 104 1.652⫻ 104
a
ASTM-AIT Chou and Fiedorowicz.
b
NASA-STD-6001 MSFC.

fined in the ASTM and NASA-STD-6001A standards for AIT testing, a sample
that ignites above a temperature of 425° C is recorded as having an AIT tem-
perature of ⬎425° C and the actual value is not specified. Due to this limitation,
the only comparison that can be made between the brass-filled PTFE and poly-
tetrafluoroethylene is that they both auto-ignite above 425° C and both materi-
als experience a pressure rise of roughly 4.83⫻ 103 kPa, which can be seen in
Table 2. When compared to the other materials, brass-filled PTFE has an AIT
value 44% greater than both Viton® A and Viton® A-500 and 30% greater than
Fluorel®. These percentages are only valid for the comparison of an AIT value
that is below the maximum test temperature of 425° C.
The determination that a material is oxygen compatible cannot be made
exclusively by AIT and heat of combustion data. A good example of the need for
several different parameters to be considered in the criteria for an oxygen com-
patible material is evident upon the comparison of the AIT data of brass-filled
PTFE and polychlorotrifluoroethylene 81 共PCTFE兲. The AIT for Brass-filled
PTFE is only 5% higher than for polychlorotrifluoroethylene 81 共PCTFE兲 based
on the maximum temperature threshold in the standard. As mentioned previ-
ously, the heat of combustion for polychlorotrifluoroethylene 81 共PCTFE兲 is
48% higher than that of the brass-filled PTFE, which indicates that upon igni-
tion the material can supply 48% more energy into the burning system.
Another test used to assess the viability of a material in a liquid or gaseous
oxygen environment is the mechanical impact test. The data presented by Chou
and Fiedorowicz is only an indication of whether or not the material passed the
test according to the ASTM standard. Their data indicates that only TFE-
polytetrafluoroethylene 共PTFE兲 passed the mechanical impact test in a LOX
environment at ambient-pressure. Figure 1 below, is a pretest picture of the
brass-filled PTFE, which is blue in color and came pre-cut into a circular
sample from the distributor. Although the name indicates that the material is
brass-filled, the small slivers cannot be seen on the surface in the picture. A
post-test picture of a brass-filled PTFE sample was taken in order to show how
the material behaves upon a non-reacting, high energy contact, which can be
seen in Fig. 2. A post-test picture of a reacting sample was not taken due to
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HERALD ET AL., doi:10.1520/JAI102244 115

FIG. 1—Pretest photo of the brass-filled PTFE.

complete combustion of the sample in a LOX environment. The post-test pic-


ture of a non-reactive sample shows that the material is extruded in an outward
radial direction when it is struck by the pin. Part of the sample was trapped
under the pin, generating a thin compressed layer. A small sliver of the brass
filling can be seen in the center of the compressed layer.
The mechanical impact data for the brass-filled PTFE, and several other
PTFE materials tested at MSFC, in both LOX and GOX environments, can be

FIG. 2—Mechanical impact 10,000 psi GOX post-test photo of the brass-filled PTFE.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
116 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

seen in Table 3. The data shows that brass-filled PTFE is highly compatible with
both LOX and GOX systems. It was impacted with 98 J at 6.89⫻ 104 kPa in
both LOX and GOX according the NASA-STD-6001 and passed in both envi-
ronments with no reactions out of 20 samples. In order to assure the accuracy
of the data, an additional 80 samples of the same batch of brass-filled PTFE
were tested in LOX and GOX. Each was impacted with 98 J at 6.89
⫻ 104 kPa in LOX and in GOX. The GOX tests showed no reactions out of the
80 impacts. Four of the additional 80 LOX samples reacted with no remaining
sample in the cup. A second batch of the brass-filled PTFE was tested under the
same conditions in order to determine the variability from batch to batch. It
can be seen in Table 3 that the second batch of brass-filled PTFE performed the
same as the first batch in a GOX environment and the reaction percentage
decreased in a LOX environment from 5% in Batch 1 to less than 2% in Batch
2. Two additional batches were tested at the same conditions and were shown
to produce zero reactions out of the samples tested. A determination could be
made through the batch mechanical impact screening that the compatibility of
brass-filled PTFE in high-pressure oxygen systems is not batch dependent.
In comparison to the brass-filled PTFE, a similar metal-filled PTFE com-
pound named Salox-M® uses bronze filler in PTFE instead of brass. The com-
position of Salox-M® is 40% bronze and 60% PTFE. The bronze component of
Salox-M® is made up of 90% copper, 8% tin, and 2% zinc. Although the per-
centage of brass in the brass-filled PTFE is not known, it can be determined
from Figs. 1 and 2 that it is far less than 40%. Since the exact chemical make up
of the brass component or its relative percentage in brass-filled PTFE is not
known, it can only be speculated that the main component of the brass is zinc
as opposed to the larger amount of tin in the bronze. The mechanical impact
results for Salox-M® impacted with 98, 55, and 40 J at 6.89⫻ 104 kPa in LOX
are shown in Table 3. The results demonstrate that the relative percentage of
bronze to PTFE has a strong influence on the sensitivity of Salox-M® in an
oxygen enriched environment.
The Salox-M® is much more reactive than the brass-filled PTFE. In fact,
reducing the pressure by 50% alone did not show the Salox-M® to be accept-
able in an oxygen enriched environment. The impact energy was reduced by
50% in order to meet the pass/fail acceptance criteria. The sensitivity of
Salox-M® in a high-pressure LOX environment has been shown to be far
higher than that of the brass-filled PTFE, which can be attributed to the high
percentage of bronze filler as well as the difference in the chemical composition
of the bronze as opposed to brass.
Two types of Algoflon®, another PTFE based material, were tested at the
same conditions in order to compare them to the brass-filled PTFE. Algoflon®
E2 is a PTFE virgin granular resin, which has been pre-sintered. This resin has
been treated thermally at, or above, the melting point of the resin at atmo-
spheric pressure. Algoflon® F7 is a PTFE virgin granular resin. It has been
designed as a molding resin for use as thin skived tape with excellent skivability
and good dielectric properties. As stated previously, two types of Algoflon®
were tested, using both a LOX and GOX environment. It was shown that the
pre-sintered E2 Algoflon® produced no reactions under the same conditions as
the brass-filled PTFE, whereas the F7 type did not perform well in a LOX
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 3—High-pressure mechanical impact data of fluorinated compounds.

System Pressure Average Sample Thickness


Material Impact Energy Test Conditions 共kPa兲 共cm兲 Reactions/Tests

Downloaded/printed by
Brass-filled PTFE 共Batch #1兲a 72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1728 0/20
72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1733 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1733 4/80
72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1733 0/80
Brass-filled PTFE 共Batch #2兲a 72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1576 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1576 1/60
Brass-filled PTFE 共Batch #3兲a 72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1576 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1576 0/20
Brass-filled PTFE 共Batch #4兲a 72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1576 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1576 0/20

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Salox-M®a 72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1579 2/2
55 ft· lb 共75 J兲 LOX/ −183° C 6.89⫻ 104 0.1589 2/4
40 ft· lb 共54 J兲 LOX/ −183° C 6.89⫻ 104 0.1589 2/25
50 ft· lb 共68 J兲 LOX/ −183° C 3.45⫻ 104 0.1589 2/18
36 ft· lb 共39 J兲 LOX/ −183° C 3.45⫻ 104 0.1589 0/20

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


Algoflon® F7a 72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.06217 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1589 2/8
72 ft· lb 共98 J兲 LOX/ −183° C 6.21⫻ 104 0.1589 2/3
72 ft· lb 共98 J兲 LOX/ −183° C 5.52⫻ 104 0.1589 2/11
72 ft· lb 共98 J兲 LOX/ −183° C 4.14⫻ 104 0.1589 2/4
72 ft· lb 共98 J兲 LOX/ −183° C 3.45⫻ 104 0.1589 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 103 0.1589 0/20
Algoflon® E2a 72 ft· lb 共98 J兲 GOX/ 24° C 6.89⫻ 104 0.1638 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1575 0/20
a
NASA-STD 6001 MSF.
HERALD ET AL., doi:10.1520/JAI102244 117
118 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

environment. The F7 required the pressure to be dropped to 3.45⫻ 104 kPa in


order for the material to pass.
Previous research conducted in 1983 by C. J. Bryan at the John F. Kennedy
Space Center, examined the oxygen compatibility of several different lubri-
cants, polymers, and metals at varying energy levels of mechanical impact. The
ambient-pressure mechanical impact data for some of their fluorinated poly-
meric materials, in a LOX environment, was obtained from the research con-
ducted by the Army Ballistic Missiles Agency 共ABMA兲. The mechanical impact
data is presented in Table 4 关5兴. The results show that the compounds were not
sensitive to mechanical shock at ambient-pressure. The same fluorinated com-
pounds were then tested in a high-pressure LOX environment. The results of
the mechanical impact tests under these conditions, as determined by data
from the Kennedy Space Center, can be seen in Table 5. polytetrafluoroethylene
and polychlorotrifluoroethylene 81 both passed the material acceptability test
up to a pressure of 1.03⫻ 104 kPa, whereas Fluorel® began reacting at pres-
sures above 3.45⫻ 103 kPa; therefore, failing the test. These same materials
were also tested in a GOX environment at ambient temperature and varying
pressures. Previous work conducted at MSFC in 1972 determined that materi-
als mechanically impacted were considerably more reactive in pressurized
GOX than in a pressurized LOX environment 关6兴. By comparison of the brass-
filled PTFE in Table 3 to TFE-polytetrafluoroethylene 共PTFE兲 in Table 5, it can
be seen that their performance in high-pressure LOX systems is very similar at
98 J and 0.1575 cm in thickness. The results presented in Table 5 demonstrate
the superior performance of the brass-filled PTFE over polychlorotrifluoroeth-
ylene 81 共PCTFE兲 and Fluorel®.
The data presented in Table 6 are high-pressure GOX impact results deter-
mined by the testers at White Sands Test Facility 共WSTF兲 and MSFC. In most
cases, the reactivity of materials increases from either a decrease in thickness
or an increase in pressure 关4兴. The data listed in Table 6 indicate that the reac-
tivity of TFE-polytetrafluoroethylene 共PTFE兲 tested at 98 J with a sample thick-
ness of 0.1575 cm increases by 25% as the pressure is increased from 60 psia to
2.24⫻ 104 kPa. The reactivity of the same material increases by 20% as the
thickness of the material is decreased by one-half. Comparing TFE-
polytetrafluoroethylene 共PTFE兲 to brass-filled PTFE, the data indicates that
under the same conditions, i.e., thickness, impact energy, and pressure, the
brass-filled PTFE outperforms the virgin PTFE. The data also shows that the
brass-filled PTFE significantly outperforms polychlorotrifluoroethylene 81
共PCTFE兲 and Viton® A. The data presented in Table 6 depicts Fluorel® as hav-
ing no reactions at varying pressures and sample thicknesses. However, it is not
known how the material would behave at double the maximum pressure pre-
sented in the table.

Conclusions
Polymeric materials, such as TFE-polytetrafluoroethylene 共PTFE兲, have proven
to be reliable and safe materials for use as seats and seals in oxygen systems.
The search for a polymer that performs as well as TFE-polytetrafluoroethylene
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 4—Ambient-pressure mechanical impact data of fluorinated compounds 关4兴.

Impact Energy Average Sample Thickness


Material 共ft· lb兲 共J兲 Test Conditions System Pressure 共cm兲 Reactions/Tests

Downloaded/printed by
TFE-polytetrafluoroethylene 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.005 0/20
共PTFE兲a 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.1575 0/100
72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.3175 0/100
polychlorotrifluoroethylene 81 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.1575 0/100
共PCTFE兲a 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.3175 0/100
Viton® Aa 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.1575 0/100
72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.3175 0/200
Fluorel®a 72 ft· lb 共98 J兲 LOX/ −183° C Ambient 0.1638 0/100
a
Determined by ABMA.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


HERALD ET AL., doi:10.1520/JAI102244 119
TABLE 5—High-pressure LOX mechanical impact data of fluorinated compounds 关6兴.

Impact Energy System Pressure Average Sample Thickness


Material 共ft· lb兲 共J兲 Test Conditions 共kPa兲 共cm兲 Reactions/Tests

Downloaded/printed by
Fluorel®a 72 ft· lb 共98 J兲 LOX/ −183° C 99.97 0.1589 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 689.5 0.1589 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 3.45⫻ 103 0.1589 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 103 0.1589 2/20
72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.1589 2/20
72 ft· lb 共98 J兲 LOX/ −183° C 99.97 0.3175 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 689.5 0.3175 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 3.45⫻ 103 0.3175 3/20
72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 103 0.3175 4/20
72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.3175 6/20

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
TFE-polytetrafluoroethylene 72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.1589 0/20
共PTFE兲a 72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.3175 0/20
TFE-polytetrafluoroethylene 72 ft· lb 共98 J兲 LOX/ −183° C 5.51⫻ 104 0.1589 2/3
共PTFE兲b 72 ft· lb 共98 J兲 LOX/ −183° C 6.89⫻ 104 0.1589 2/42
polychlorotrifluoroethylene 81 72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.1589 0/20

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


共PCTFE兲a,b 72 ft· lb 共98 J兲 LOX/ −183° C 1.03⫻ 104 0.3175 0/20
72 ft· lb 共98 J兲 LOX/ −183° C 1.38⫻ 104 0.1589 8/20
a
-MSFC and WSTF mechanical impact test results.
b
-Obtained from MAPTIS.
120 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 6—High-pressure GOX mechanical impact data of fluorinated compounds 关6兴.

Impact Energy System Pressure Average Sample Thickness


Material 共ft· lb兲 共J兲 Test Conditions 共kPa兲 共in.兲 Reactions/Tests

Downloaded/printed by
TFE-polytetrafluoroethylene 72 ft· lb 共98 J兲 GOX/ 127° C 413.7 0.1589 0/20
共PTFE兲a 72 ft· lb 共98 J兲 GOX/ 81° C 2.24⫻ 104 0.1589 5/20
20.3 ft· lb 共27.5 J兲 GOX/ 81° C 2.24⫻ 104 0.1589 0/20
72 ft· lb 共98 J兲 GOX/Ambient 3.45⫻ 104 0.3175 0/20
72 ft· lb 共98 J兲 GOX/Ambient 3.45⫻ 104 0.1589 4/20
TFE-polytetrafluoroethylene 72 ft· lb 共98 J兲 GOX/Ambient 4.83⫻ 104 0.1589 2/5
共PTFE兲b 72 ft· lb 共98 J兲 GOX/Ambient 5.52⫻ 104 0.1589 1/1
polychlorotrifluoroethylene 81 72 ft· lb 共98 J兲 GOX/Ambient 6.89⫻ 103 0.1589 0/20
共PCTFE兲a 50 ft· lb 共68 J兲 GOX/Ambient 1.03⫻ 104 0.1589 2/20
43 ft· lb 共59 J兲 GOX/Ambient 1.03⫻ 104 0.1589 0/20

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
40 ft· lb 共54 J兲 GOX/Ambient 1.38⫻ 104 0.1589 0/20
45 ft· lb 共61 J兲 GOX/Ambient 2.07⫻ 104 0.1589 0/20
55 ft· lb 共75 J兲 GOX/Ambient 3.45⫻ 104 0.1589 1/6
40 ft· lb 共54 J兲 GOX/Ambient 3.45⫻ 104 0.1589 0/20
72 ft· lb 共98 J兲 GOX/Ambient 6.89⫻ 103 0.3175 2/20

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


65 ft· lb 共88 J兲 GOX/Ambient 6.89⫻ 103 0.3175 0/20
72 ft· lb 共98 J兲 GOX/Ambient 1.03⫻ 104 0.3175 0/20
65 ft· lb 共88 J兲 GOX/Ambient 1.38⫻ 104 0.3175 0/20
55 ft· lb 共75 J兲 GOX/Ambient 2.07⫻ 104 0.3175 1/20
50 ft· lb 共68 J兲 GOX/Ambient 2.07⫻ 104 0.3175 0/20
72 ft· lb 共98 J兲 GOX/Ambient 3.45⫻ 103 0.3175 4/10
HERALD ET AL., doi:10.1520/JAI102244 121
TABLE 6— 共Continued.兲

Impact Energy System Pressure Average Sample Thickness


Material 共ft· lb兲 共J兲 Test Conditions 共kPa兲 共in.兲 Reactions/Tests

Downloaded/printed by
Viton® Aa 72 ft· lb 共98 J兲 GOX/Ambient 6.89⫻ 103 0.193 0/20
72 ft· lb 共98 J兲 GOX/Ambient 1.03⫻ 104 0.193 1/9
72 ft· lb 共98 J兲 GOX/Ambient 1.03⫻ 104 0.1589 1/20
72 ft· lb 共98 J兲 GOX/Ambient 1.38⫻ 104 0.1589 1/17
30 ft· lb 共41 J兲 GOX/Ambient 3.45⫻ 104 0.1589 0/20
Fluorel®a 72 ft· lb 共98 J兲 GOX/Ambient 1.72⫻ 103 0.1589 0/20
51 ft· lb 共68 J兲 GOX/Ambient 3.45⫻ 103 0.1589 0/20
30 ft· lb 共41 J兲 GOX/Ambient 1.03⫻ 104 0.1589 0/20
25 ft· lb 共34 J兲 GOX/Ambient 3.45⫻ 104 0.1589 0/20
72 ft· lb 共98 J兲 GOX/Ambient 1.72⫻ 103 0.3175 0/20

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
51 ft· lb 共68 J兲 GOX/Ambient 3.45⫻ 103 0.3175 0/20
30 ft· lb 共41 J兲 GOX/Ambient 1.03⫻ 104 0.3175 0/20
25 ft· lb 共34 J兲 GOX/Ambient 3.45⫻ 104 0.3175 0/20
a
MSFC and WSTF mechanical impact test results.
b
Obtained from MAPTIS.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


122 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
HERALD ET AL., doi:10.1520/JAI102244 123

共PTFE兲, in terms of oxygen compatibility, and also maintains its physical integ-
rity at high-pressures and cryogenic temperatures has become an ongoing ex-
pedition. Through research at the MCRF at MSFC a brass-filled PTFE has been
uncovered that exemplifies all of the positive characteristics that are valued in
an oxygen compatible material. The low heat of combustion, elevated AIT, and
very high resistance to mechanical shock in pressurized LOX and GOX envi-
ronments further exemplify the suitability of the brass-filled PTFE for oxygen
systems. The data presented clearly shows that the choice of the filler material
and the percentage of filler relative to PTFE can make all the difference in a
high-performing oxygen compatible material. Even though the brass-filled
PTFE performs similarly to TFE-polytetrafluoroethylene in a high-pressure
LOX environment, its performance far exceeds that of TFE-
polytetrafluoroethylene in a high-pressure GOX environment, which makes
brass-filled PTFE the best choice for oxygen systems.

References

关1兴 Chou, T. C. and Fiedorowicz, A., “Oxygen Compatibility of Polymers Including


TFE-Teflon®, Kel-F® 81, Vespel® SP-21, Viton® A, Viton® A-500, Fluorel®, Neo-
preneg, EPDM, Buna-N, and Nylon 6,6,” Flammability and Sensitivity of Materials
in Oxygen-Enriched Atmospheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T.
C. Chou, and T. A. Steinberg, Eds., ASTM International, West Conshohocken, PA,
1997.
关2兴 Mamunya, Y.-P., Muzychenko, Y.-V., Pissis, P., Lebedev, E. V., and Shut, M. I.,
“Processing, Structure, and Electrical Properties of Metal-Filled Polymers,” J. Mac-
romol. Sci., Phys., Vol. B40共3–4兲, 2001, 591–602.
关3兴 Zhang, Z. Z., Xue, Q. J., Liu, W. M., and Shen, W. C., “Friction and Wear Properties
of Metal Powder Filled PTFE Composites Under Oil Lubricated Conditions,” Wear,
Vol. 210, 1997, pp. 151–156.
关4兴 Beeson, H. D., Smith, S. R., and Stewart, W. F., Safe Use of Oxygen and Oxygen
Systems: Handbook for Design, Operation, and Maintenance, 2nd ed., ASTM Inter-
national, West Conshohocken, PA, 2007.
关5兴 Bryan, C. J., “NASA Mechanical Impact Testing in High-Pressure oxygen,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP
812, B. L. Werley, Ed., ASTM International, West Conshohocken, PA, 1983, pp.
9–42.
关6兴 Schwinghamer, R. J., “Impact Sensitivity of Materials in Contact with Liquid and
Gaseous Oxygen at High Pressure,” Rep. No. TMX-64634, National Aeronautics
and Space Administration, Washington, D.C., January 1972.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102262
Available online at www.astm.org/JAI

Chr. Binder,1 T. Brock,1 O. Hesse,1 S. Lehné,1 and T. Tillack1

Ignition Sensitivity of Nonmetallic Materials


in Oxygen-Enriched Air „NITROX…: A
Never Ending Story in SCUBA Diving?

ABSTRACT: Self-contained underwater breathing apparatus 共SCUBA兲 divers


use different mixtures of nitrogen and oxygen as breathing gas. The so-
called NITROX mixture often contains more oxygen than is contained in air
and may increase the fire hazard in diving equipment. The SCUBA diving
community, however, still differentiates between NITROX mixtures that con-
tain more than 40 % oxygen or less. They consider a mixture with up to 40 %
oxygen often as regular air. In filling operations of SCUBA cylinders, gas
flows from high pressure to low pressure. Because of compressive heating, a
sudden temperature rise in the gas occurs and nonmetallic materials, e.g.,
seals may ignite. BAM has used the pneumatic impact test to investigate the
ignition sensitivity of ethylene propylene diene monomer 共EPDM兲, fluorelas-
tomer 共FPM兲, polytetrafluorethylene 共PTFE兲, polyetheretherketone 共PEEK兲,
and nylon 共PA 6.6兲 to gaseous impacts in synthetic air, in various NITROX
mixtures, and in pure oxygen. The test results clearly show that for nonme-
tallic materials, the maximum pressure of nonreaction in NITROX mixtures
decreases at a content of 29 % oxygen in comparison to those in air. In
addition, autoignition temperatures of the nonmetallic materials were also
determined. The findings of this investigation support very well the results of
other publications on oxygen enrichment. As a consequence of this study, in
SCUBA diving, the same safety requirements for NITROX mixtures with
more than 21 % oxygen should be applied as for pure oxygen in the industry.

Manuscript received December 5, 2008; accepted for publication June 12, 2009; pub-
lished online July 2009.
1
Specialist of Working Group “Safe Handling of Oxygen,” Division II.1 “Gases, Gas
Plants,” BAM Federal Institute for Materials Research and Testing, Berlin 12205, Ger-
many.
Cite as: Binder, Chr., Brock, T., Hesse, O., Lehné, S. and Tillack, T., ⬙Ignition Sensitivity
of Nonmetallic Materials in Oxygen-Enriched Air 共NITROX兲: A Never Ending Story in
SCUBA Diving?,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102262.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 124
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102262 125

KEYWORDS: NITROX, SCUBA diving, filling process, pneumatic


impact, autoignition temperature, flammability, nonmetallic materials,
oxygen mixtures, fire hazard, oxygen enrichment

Introduction
In self-contained underwater breathing apparatus 共SCUBA兲 diving, nitrogen
may accumulate in the human body depending on the diving depth and on the
duration of diving time. This means for divers that they have to dispose of
nitrogen before reaching the water surface by undertaking several decompres-
sion stops. If not, there is a high risk to the divers’ health. The advantage in
minimizing the nitrogen and increasing the oxygen content reduces the risk of
nitrogen enrichment in the human body.
Therefore, SCUBA divers use different mixtures of nitrogen and oxygen as
breathing gas. The so-called NITROX mixtures mostly contain more oxygen
than air and help reduce the risk of decompression sickness. Several NITROX
mixtures, partially with misleading names are used in SCUBA diving. Table 1
contains nitrogen/oxygen mixtures that are common in the diving community.
For example, the so-called SaveAir® 50 suggests a safe use of this gas,
comparable to normal air, but this mixture actually contains 50 % oxygen. Even
NITROX 100, e.g., suggests being a nonhazardous mixture with nitrogen, but
NITROX 100 is nothing else than pure oxygen!
In industry, it is well known that higher oxygen concentrations than in air
increase the fire hazards in a mixture of nitrogen and oxygen. Parts of the
SCUBA diving community, however, still differentiate between NITROX mix-
tures that contain more than 40 % oxygen or less. A mixture with up to 40 %
oxygen is often considered as regular air. In this context, many people forget
that besides the pressure, compressed air exhibits additional risks. It could be
shown that nonmetallic materials that are considered as nonflammable in air at
ambient pressures may ignite and burn in air at pressures greater than 20.7
MPa 关2兴.
Typical nonmetallic materials in valves, fittings, and other parts are sealing
materials, lubricants, residues of abrasion, detergents, oils, etc. Figures 1 and 2
give examples of what flammable nonmetallic materials may be present in div-
ing components.

TABLE 1—NITROX mixtures in SCUBA diving 关1兴.

Content of Nitrogen/Oxygen
Name of NITROX Mixture in Volume %
NITROX 0 71/29
NOAA-1 68/32
NITROX-D 67,5/32,5
NOAA-II 64/36
NITROX-C 60/40
SaveAir® 50 50/50
NITROX-B 40/60
NITROX
Copyright 100
by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014 0/100
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
126 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Sealing tape on pressure gauge connector.

It is widely spread that diving equipment for such mixtures does not re-
quire special attention to fire hazards. Furthermore, it is not regarded as essen-
tial to have this equipment cleaned as usually required for oxygen service in
industry or for medical purposes.
That this attitude may result in incidents has already been published more
than 10 years ago 关3兴. In addition, it has already been shown that an increase in
oxygen content results in decreasing autoignition temperatures 共AIT兲 and me-
chanical impact energy thresholds for most of the nonmetallic materials 关4,5兴.
Filling operations of SCUBA cylinders by the gas blending technique or by
premix decanting are common in diving centers. During these processes, gas
flows from high pressure to low pressure. Because of compressive heating, a
sudden temperature rise in the gas occurs and nonmetallic materials may ignite
and cause a burn-out of diving components. Figure 3 shows the burn-out of an
oxygen pressure reducer. Even the metallic housing burned. Figure 4 is a pic-
ture of a metallic oxygen filling hose with an organic liner that burned out
because of compressive heating.
The incidents shown in the above-mentioned figures can be understood if
the theoretical maximum temperature is calculated that can be developed in
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102262 127

FIG. 2—Residue of sealing tape on valve connector.

rapid pressurization of oxygen from atmospheric pressure at 20° C to 100 MPa,


yielding a temperature of 815° C 关6兴.
So far, only few publicly available data exist on the ignition sensitivity of
nonmetallic materials to gaseous oxygen impacts in NITROX mixtures. BAM,
The Federal Institute for Materials Research and Testing in Germany, decided
to perform tests on some sealing materials that may typically found in SCUBA
diving equipment such as EPDM, FKM/FPM, PA 6.6, PEEK, and also PTFE.
Additionally, AIT determinations on these materials were performed to find
out whether former findings in oxygen enriched mixtures 关4,5兴 can be sup-
ported with BAM’s unique inductive heating test method 关7兴.

Experimental
A standardized test method that simulates compressive heating or pressure
shocks that may occur during filling operations of cylinders is the pneumatic
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
128 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—Burned oxygen pressure reducer.

FIG. 4—Oxygen filling hose.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102262 129

TABLE 2—List of NITROX mixtures used in this study.

Content of Nitrogen/Oxygen
Gas Mixture in vol %
Air 79/21
NITROX-0 71/29
NITROX-C 60/40
NITROX-B 40/60
Pure oxygen 0/100

impact test according to ISO 21010, Annex C 关8兴. This test method utilizes a 750
mm long connecting tube with an internal diameter of 14 mm between the
quick opening valve and the sample holder. Instead of pure oxygen alone, this
test was carried out with various NITROX mixtures and with synthetic air. Air
Liquide Deutschland GmbH supplied all the gases with a purity of 99.2 %.
Table 2 shows what NITROX mixtures were used in this study.
The standardized pressure shock test was performed with these different
NITROX mixtures at 60° C. By opening of a quick-opening valve, the samples
are exposed to gaseous oxygen impacts from atmospheric pressure to the final
test pressure. The pressure rise time is 15–20 ms.
In addition, AIT determinations on the materials with the above-mentioned
nitrogen/oxygen mixtures were performed using BAM’s AIT tester, which is de-
scribed elsewhere 关7兴. The sample holder, an autoclave, is pressurized to the
desired pressure at the beginning of the test and a low-frequency heater induc-
tively heats the autoclave in an almost linear way at a rate of 110 K/min. Pres-
sure and temperature are recorded so that the ignition of the sample can be
recognized by a sudden rise in temperature and pressure.
The nonmetallic materials were taken from the stock and prepared for the
tests according to the above-mentioned standard.

Results and Discussion


Table 3 shows the results that were obtained in oxygen pressure shock testing.

TABLE 3—Maximum pressures of nonreaction in bar at 60° C as a function of oxygen


content.

Oxygen Content in vol %

Material 21 Air 29 NITROX-0 40 NITROX-C 60 NITROX-B 100 Pure Oxygen


PTFE 90 75 45 35 20
PA 6.6 55 40 35 25 20
PEEK 50 50 30 25 20
FKM/FPM 90 55 55 40 30
EPDM 30 25 20 15 10
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
130 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—Influence of oxygen content on the maximum pressure of nonreaction in pres-


sure shock testing.

It contains the maximum pressure of nonreaction for five materials depending


on the oxygen content in the NITROX mixture.
Generally, the table shows that for a given test temperature, the maximum
pressures of nonreaction decrease with increasing oxygen content. The trend
curves in Fig. 5 depict that this mainly occurs between an oxygen content of 21
% as in air and 29 % in NITROX-0.
These results show that nitrogen/oxygen mixtures with slightly greater oxy-
gen content than in air already raise the risk of a fire incident.
Table 4 contains the results of the AIT determinations at 50 bar for the
above-mentioned five materials depending on the oxygen content.
For the given test pressure, the AIT of the nonmetallic materials decreases
with increasing oxygen content. The trend curves are shown in Fig. 6. The AIT
is almost constant only for PTFE. For this material, no distinctive influence of
the oxygen content on its AIT is detectable. This result confirms former findings
on AIT determinations in oxygen enriched environments 关5兴, where the same
method of investigation but a different test apparatus was used at 34 and 69
bar.

TABLE 4—AITs in °C at 50 bar as a function of oxygen content.

Oxygen Content in vol %

Material 21 Air 29 NITROX-0 40 NITROX-C 60 NITROX-B 100 Pure Oxygen


PTFE 475 473 471 471 466
PA 6.6 327 297 269 250 136
PEEK ⬎500 413 390 436 407
FKM/FPM 363 350 338 334 306
EPDM 249 231 213 201 183
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102262 131

FIG. 6—Influence of oxygen content on the AIT of nonmetallic materials.

Also, the results of the AIT determinations depict that NITROX mixtures
with an oxygen content higher than in air may increase the chance of having a
fire incident.

Summary and Conclusions


This investigation with different NITROX mixtures used for SCUBA diving
shows that already a slight increase in oxygen content in nitrogen/oxygen mix-
ture increases the fire hazard of nonmetallic materials in comparison to air. The
findings of this work support very well the results of other papers on oxygen
enrichment and leave no more any doubt that for safety-related reasons:
• NITROX mixtures with an oxygen content of more than 21 % should be
handled with care like pure oxygen.
• Valves, fitting, hoses, and other parts used for NITROX mixtures should
be tested for burn-out safety according to National and International
standards for oxygen components.
• Regulations for pure oxygen should be also adopted for NITROX mix-
tures.
Regarding the question in the title of this paper “NITROX–A Never Ending
Story in SCUBA diving?” the answer today must be: “Yes, the story is over!”
It should be clear to everyone in the diving community that even NITROX
mixtures with oxygen contents between 21 % and 40 % need to be handled with
care.

References

关1兴 http://www.rebreather.de/Nitrox/tauchen.html 共Last accessed 11 Nov. 2008兲.


关2兴 Hirsch, D. B. and Bunker, R. L., “Effects of Diluents on Flammability of Nonmet-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
132 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

als in High-Pressure Oxygen Mixtures,” Flammability and Sensitivity of Materials in


Oxygen-Enriched Atmospheres: Sixth Volume, ASTM STP 1197, D. D. Janoff and J.
M. Stoltzfus, Eds., ASTM International, West Conshohocken, PA, 1993.
关3兴 Gabel, H. and Janoff, D., “Use of Oxygen-Enriched Mixtures in Recreational
SCUBA Diving–Is the Public Being Informed of the Risks?,” Flammability and Sen-
sitivity of Materials in Oxygen-Enriched Atmospheres: Eighth Volume, ASTM STP
1319, W. T. Royals, T. C. Chou, and T. A. Steinberg, Eds., ASTM International, West
Conshohocken, PA, 1997.
关4兴 Hirsch, D., Hshieh, F., Beeson, H., and Bryan, C., “Ignitability in Air, Gaseous
Oxygen, and Oxygen-Enriched Environments of Polymers Used in Breathing-Air
Devices,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997.
关5兴 Bryan, C. J., Hirsch, D. B., Haas, J., and Beeson, H., “Ignitability in Air, Gaseous
Oxygen, and Oxygen-Enriched Environments of Polymers Used in Breathing-Air
Devices, Final Report,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. New-
ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关6兴 ASTM G63-99, 2007, “Standard Guide for Evaluating Nonmetallic Materials for
Oxygen,” Annual Book of ASTM Standards, Vol. 14.04, ASTM International, West
Conshohocken, PA.
关7兴 Wegener, W., Binder, C., Hengstenberg, P., Herrmann, K.-P., and Weinert, D.,
“Tests to Evaluate the Suitability of Materials for Oxygen Service,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Volume, ASTM
STP 986, D. W. Schroll, Ed., ASTM International, West Conshohocken, PA, 1988.
关8兴 ISO 21010, 2004, “Cryogenic Vessels—Gas/Materials Compatibility,” Beuth Verlag
GmbH, Berlin.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102309
Available online at www.astm.org/JAI

Chr. Binder,1 K. Arlt,1 T. Brock,1 P. Hartwig,1 O. Hesse,1 and


T. Tillack1

The Importance of Quality Assurance and


Batch Testing on Nonmetallic Materials
Used for Oxygen Service

ABSTRACT: In oxygen components, even for high pressure oxygen service,


it is possible to use organic seals, lubricants, or filling liquids, provided their
oxygen compatibility has been checked. However, fire incidents in oxygen
systems still occur because these materials ignite and burn. There are many
reasons, such as incorrect design, contamination, faulty operation, unsuit-
able materials, etc., why this happens. Another cause that is overseen very
often is proper maintenance on the user’s side. It is very important to replace
in a component a worn out seal by the same one with identical oxygen
compatibility properties. On the part of the producer or distributor of materi-
als, batch testing and also a quality assurance system play a key role in the
safety of an oxygen component. Any change in the manufacturing process of
a material, or in its composition, and even its further processing may have an
impact on its oxygen compatibility and finally on the component in which it is
used. Numerous investigations by BAM over decades reveal the influence of
minor constituents and fillers on a material’s oxygen compatibility. The test
results in this paper show how important it is to regularly perform batch
testing on nonmetallic materials used for oxygen service and to have a qual-
ity assurance system that helps minimize incidents where unsuitable mate-
rials are chosen by accident.
KEYWORDS: oxygen compatibility, nonmetallic materials, pneumatic
impact, flammability, components, maintenance, batch testing, quality
assurance

Manuscript received December 19, 2008; accepted for publication June 12, 2009; pub-
lished online July 2009.
1
Specialists of Working Group “Safe Handling of Oxygen,” Division II.1 “Gases, Gas
Plants,” BAM Federal Institute for Materials Research and Testing, Berlin 12205, Ger-
many
Cite as: Binder, Chr., Arlt, K., Brock, T., Hartwig, P., Hesse, O. and Tillack, T., ⬙The
Importance of Quality Assurance and Batch Testing on Nonmetallic Materials Used for
Oxygen Service,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102309.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 133
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
134 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Introduction

Most of the components for oxygen service contain nonmetallic materials, such
as seals, lubricants, filling liquids, etc. Their oxygen compatibility and the de-
sign of a component help ensure safe handling.
Very often, for low pressure oxygen applications, only the nonmetallic seal
or the lubricant of a component needs to be tested and evaluated for oxygen
compatibility. For operating pressures greater than 100 bar, mostly the oxygen
component with its nonmetallic seals inside is tested for burn-out safety.
National, European, and International standards describe how to deter-
mine either the oxygen compatibility of a material or how to test the burn-out
safety of a component. The reference list at the end of this paper gives some
examples 关1–14兴.
The user should not encounter any fire incidents, provided the operating
conditions and the cleanliness requirements of a component are met. However,
the reality is different. Occasionally, fires still happen and it is often very diffi-
cult to find out the cause of such an incident.
BAM, The Federal Institute for Materials Research and Testing in Berlin,
has been testing and evaluating materials and components for oxygen service
for many decades. This paper intends to campaign for batch testing and quality
assurance procedures on nonmetallic materials for oxygen service on the
manufacturer’s side as well as on user’s side. In this context, a user can also be
a valve producer, who employs lubricants, filling liquids, and seals for his prod-
ucts or the end-user of an oxygen component, who does maintenance.
Unfortunately, according to BAM’s experience, only some manufacturers
regularly perform batch testing mostly on lubricants, filling liquids, and occa-
sionally on seals. This may either be related to a lack of understanding or to the
test cost.
At BAM, the test of first choice for many decades has been the pressure
shock or pneumatic impact tester according to ISO 21010 关1兴 or EN 1797 关2兴 for
nonmetallic materials with operating temperatures up to 60° C. This test inves-
tigates the ignition sensitivity of materials to gaseous oxygen impacts. It simu-
lates a scenario that happens in real life in a component or in piping when
oxygen flows from high pressure to low pressure and compressive heating oc-
curs. In the laboratory, a heatable steel tube contains the nonmetallic material
under test, usually at an initial pressure of 1 bar 共see Fig. 1兲. By opening of a
high-speed valve, the sample is then exposed to gaseous oxygen impacts accord-
ing to a given test scheme. A distinct temperature rise in the sample tube indi-
cates a positive reaction.
Some papers in the literature doubt that this test method may be applied to
batch testing 关15–17兴. They conclude from their findings that this method is not
sensitive enough and that it has a poor repeatability. The authors of this paper,
however, cannot confirm the variability in the data that has been reported. This
may be attributed to a different test system and a different sample preparation.
At BAM, this method has been used to determine the maximum working pres-
sure of nonmetallic materials for oxygen service, for more than 25 years.
Figure 2 illustrates what happened when BAM rechecked a graphite seal in
2007 that qualified for oxygen service with a maximum use pressure of 300 bar
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102309 135

FIG. 1—Sample holder on pneumatic impact tester at BAM, pre-test.

FIG. 2—Sample holder on pneumatic impact tester at BAM, post-test.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
136 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

in 1997. Testing at 220 bar oxygen pressure of the so-called “same material”
caused a violent reaction of the material and resulted in a complete demolition
of the sample holder and the heater of the test apparatus.
The reason for this incident was not an altered production process of the
seal or a different run of the test. It turned out that the manufacturer had
simply provided the wrong material, modified graphite, for testing, possibly
due to a faulty or missing quality assurance system. In this case, only damaged
test equipment was the result. However, for end-users, this could have resulted
in injured operators, loss of production, and repair cost, all of which were
preventable.

Experimental

Test Method
The test results in this paper are based on BAM’s standardized pneumatic im-
pact tester for determining the ignition sensitivity of nonmetallic materials to
gaseous oxygen impacts. The pneumatically actuated high-speed valve of the
tester was designed and built by BAM and is similar to the types described in
previous papers 关18,19兴. It guarantees a pressure rise from atmospheric pres-
sure to the test pressure within 15–20 ms in the sample tube.
According to the above-mentioned standards, solid samples are finely di-
vided into a minimum of six pieces. Liquids and paste-like products are tested
on fibrous ceramics soaked by the sample to be tested. A possible reaction of
the sample with oxygen is indicated by a steep temperature rise of at least 20° C
superimposed on the temperature rise in a blank test. Usually, a pressure incre-
ment of 10 bar is used to determine the maximum pressure of non-reaction.
To check the repeatability of the test system, BAM is using a particular
virgin polytetrafluoroethylene 共PTFE兲 as a kind of reference material. Figure 3
depicts the test results of recent years. Each triangle reflects two test series at
30 bar, each consisting of five consecutive pressure impacts on the reference
specimen, at which no reactions with oxygen were detected. This pressure is
called the maximum pressure of non-reaction as referred to throughout this
entire paper. The symbols above each triangle, at a 10 bar higher pressure,
show the number of impact on what the specimen did react.
Figure 3 is an example for the repeatability of the test method on PTFE.
This is representative for most of the tests that have been performed on non-
metallic materials over the past years.

Materials
The materials and lubricants were provided by industry for safety evaluation by
BAM. The particular test results in this paper mostly refer to a particular test
application. Because of this fact, the manufacturer or distributor and the prod-
ucts had to be made anonymous. All the materials and lubricants were pre-
pared for testing as required in the standard 关1,2兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102309 137

FIG. 3—Triangles depict maximum pressures of non-reaction at 60° C of virgin PTFE


as a reference material for quality control checks on the pneumatic impact tester. Other
symbols show reactions at a 10 bar higher oxygen pressure.

Materials and Lubricants—Graphite was chosen as an example because of


its high autoignition temperature 共AIT兲 and ignition resistance to pneumatic
impacts in oxygen. PTFE, a thermoplastic material, is frequently used because
of its high AIT in oxygen, and nitrile butadiene rubber 共NBR兲 is a typical rep-
resentative for elastomers found in many components, although it exhibits a
relatively low AIT. In addition, findings of a liquid and a paste-like lubricant,
both based on perfluoroalkyl polyether oil, of two different manufacturers were
selected for this paper.

Results and Discussion


The figures in this section present the maximum pressures of non-reaction. For
readability, in contrast to Fig. 3, all the data at a 10 bar higher pressure at
which reactions regularly occurred has not been added to the figures.
First, the test results on graphite, obtained over a 25 year period of time.

Graphite
Figure 4 depicts the maximum pressures of non-reaction at 60° C of different
graphite-types of different producers. Four of the six data points show that the
expected pressure range for graphite is usually between 250 and 300 bar
共shaded area兲. The deviations in the years 1985 and 2005 explain why the
above-mentioned incident happened. The wrong delivered material was a com-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
138 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Maximum pressures of non-reaction at 60° C of different graphite-types of dif-


ferent producers. The expected range is shaded. There are distinct exceptions in the years
1985 and 2005.

pound of graphite and PTFE that already reacted at a very low oxygen pressure
of 30 bar. The pure graphite tested in 1997 did not react up to 300 bar oxygen
pressure.

PTFE
Figure 5 shows the maximum oxygen pressures of non-reaction at 60° C of
different PTFE-types and different producers in a 25 year period of time. Ac-
cording to BAM’s experience, virgin PTFE normally exhibits a maximum pres-
sure of non-reaction between 30 and 40 bar 共shaded area兲. This threshold can
be raised if PTFE is combined with a filler material such as carbon, graphite,
glass, etc. The grey-colored triangles in Fig. 5 depict a higher maximum pres-
sure of non-reaction. However, impregnations by PTFE dispersion or expanded
PTFE usually lower the threshold pressure of non-reaction. The dark-colored
triangles in Fig. 5 are examples of such types.
Figure 6 depicts a comparison of different batches of the same PTFE-type
without filler of one producer. The differences in the maximum pressures of
non-reaction are not as big as between different PTFE-types of different manu-
facturers. Here, molding pressure and sintering temperature play a key role in
processing of the batch.
For example, for batch number 2, different molding pressures were used
marked in Fig. 6 by different squares 共black: 150 bar molding pressure, non-
filled: 250 bar molding pressure, and grey: 400 bar molding pressure兲. In this
case, an increasing molding pressure yielded a higher maximum pressure of
non-reaction.
Hence, for safety reasons, it is always important to know exactly what
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102309 139

FIG. 5—Maximum pressures of non-reaction at 60° C of different PTFE-types of differ-


ent manufacturers. The expected range for virgin PTFE is shaded. Grey-colored and
dark-colored triangles show distinct influence on performance of PTFE by filler or
impregnation.

FIG. 6—Maximum pressures of non-reaction at 60° C of different batches of one par-


ticular PTFE-type of the same manufacturer. The expected range is shaded. Batch num-
ber 2: Processing by different molding pressures, increasing from the bottom up.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
140 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Maximum pressures of non-reaction at 60° C of NBR-types of different manu-


facturers. The expected range is shaded.

PTFE-batch of the manufacturer may be used for a particular application in


oxygen service, especially for operating pressures above 50 bar.

NBR
Figure 7 shows the maximum pressures of non-reaction of different NBR-types
supplied by different manufacturers. Similar to virgin PTFE, the maximum
pressure varies slightly; in this case between 20 and 30 bar. Also in this case, for
safety reasons, it is absolutely necessary to replace an NBR seal in the oxygen
wetted area of a component by an NBR seal with identical or superior oxygen
compatibility properties, if the maximum use pressure is above 30 bar.

Paste-like Lubricant
Figure 8 presents the maximum pressures of non-reaction of different batches
of one particular paste-like lubricant of one particular manufacturer. The graph
shows that most of the pressures of non-reaction can be found at the expected
pressure of 250 bar. But there are exceptions between 160 and 310 bar, depend-
ing on the particular batch of production.

Liquid Lubricant
Figure 9 depicts the maximum pressures of non-reaction of different batches of
a particular liquid lubricant of one particular producer. Again, this result is
based on a perfluoroalkyl polyether but it contains no fillers as the above-
mentioned paste-like lubricant. In comparison to the latter one, the maximum
pressure of non-reaction varies “only” slightly between 250 and 270 bar. The
shaded area in Fig. 9 shows this small range.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102309 141

FIG. 8—Maximum pressures of non-reaction at 60° C of different batches of one par-


ticular type of paste-like lubricant of one manufacturer. The expected pressure range is
shaded.

Summary
This paper presents test results of only a small section of numerous nonmetallic
materials and lubricants that BAM has tested and evaluated for oxygen service

FIG. 9—Maximum pressures of non-reaction at 60° C of different batches of one par-


ticular type of liquid lubricant of one manufacturer. The expected pressure range is
shaded.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
142 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

over a couple of decades. It does not intend to cover any statistic aspects on the
applied test method or on the test results obtained nor does it address all com-
monly used nonmetallic materials or lubricants.
This paper just wants to alert manufacturers, distributors, and users of
nonmetallic materials that these need to be carefully selected for any oxygen
component regarding its particular operating conditions.
For users, it is necessary to know that equally named nonmetallic materials
may have a different performance in oxygen compatibility. A help to the user
may be the test certificate on a particular batch of the material. It should show
that the candidate material has been tested for oxygen compatibility according
to a standardized test method.
A cooperation between manufactures and users may help to further im-
prove the safety of oxygen components and help avoid incidents. An example is
the manufacturing of certified products under surveillance by an accredited
test laboratory. In this way, it is continually demonstrated by batch or check
tests that the safety-related parameters remain acceptable.
A quality assurance system on the manufacturer’s and the user’s side helps
minimize mix-ups of nonmetallic materials for oxygen components. This may
be also a way to prevent incidents like the one described in the introduction to
this paper.

References

关1兴 ISO Standard 21010, 2004, “Cryogenic Vessels—Gas/Materials Compatibility,”


Beuth Verlag GmbH, 10772 Berlin.
关2兴 EN Standard 1797, 2001, “Cryogenic Vessels—Gas/Material Compatibility,” Ger-
man version EN 1797, Beuth Verlag GmbH, 10772 Berlin.
关3兴 ISO/DIS Standard 11114–3, 2008, “Transportable Gas Cylinders—Compatibility of
Cylinder and Valve Materials with Gas Contents-Part 3: Autogenous Ignition Test
for Non-Metallic Materials in Oxygen Atmosphere,” Beuth Verlag GmbH, 10772
Berlin.
关4兴 DIN EN ISO Standard 14113 1–3, 2007 “Gas Welding Equipment—Rubber and
Plastics Hose and Hose Assemblies for Use with Industrial Gases up to 450 bar 共45
MPa兲 共ISO 1413兲,” Beuth Verlag GmbH, 10772 Berlin.
关5兴 ISO Standard 2503, 1998, “Gas Welding Equipment—Pressure Regulators for Gas
Cylinders Used in Welding, Cutting and Allied Processes up to 300 bar 共30 MPa兲
共ISO 2503兲,” Beuth Verlag GmbH, 10772 Berlin.
关6兴 ISO Standard 10297, 2006, “Transportable Gas Cylinders—Cylinder Valves-
Specification and Type Testing,” Beuth Verlag GmbH, 10772 Berlin.
关7兴 ISO/DIS Standard 15001, 2008, “Anesthetic and Respiratory Equipment—
Compatibility with Oxygen,” Beuth Verlag GmbH, 10772 Berlin.
关8兴 ASTM Standard D2512–95, 2008, “Standard Test Method for Compatibility of Ma-
terials with Liquid Oxygen 共Impact Sensitivity Threshold and Pass-Fail Tech-
niques兲,” Annual Book of ASTM Standards, Vol. 14.04, ASTM International, West
Conshohocken, PA.
关9兴 ASTM Standard G72–01, 2001, “Standard Test Method for Autogenous Ignition
Temperature of Liquids and Solids in a High-Pressure Oxygen-Enriched Environ-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
BINDER ET AL., doi:10.1520/JAI102309 143

ment,” Annual Book of ASTM Standards, Vol. 14.04, ASTM International, West
Conshohocken, PA.
关10兴 ASTM Standard G74–08, 2008, “Standard Test Method for Ignition Sensitivity of
Materials to Gaseous Fluid Impact2,” Annual Book of ASTM Standards, Vol. 14.04,
ASTM International, West Conshohocken, PA.
关11兴 ASTM Standard G86–98a, 2005, “Standard Test Method for Determining Ignition
Sensitivity of Materials to Mechanical Impact in Ambient Liquid Oxygen and Pres-
surized Liquid and Gaseous Oxygen Environments,” Annual Book of ASTM Stan-
dards, Vol. 14.04, ASTM International, West Conshohocken, PA.
关12兴 ASTM Standard G124–95, 2003, “Standard Test Method for Determining the Com-
bustion Behavior of Metallic Materials in Oxygen-Enriched Atmospheres,” Annual
Book of ASTM Standards, Vol. 14.04, ASTM International, West Conshohocken, PA.
关13兴 ASTM Standard G175–03, 2003, “Standard Test Method for Evaluating the Igni-
tion Sensitivity and Fault Tolerance of Oxygen Regulators Used for Medical and
Emergency Applications,” Annual Book of ASTM Standards, Vol. 14.04, ASTM In-
ternational, West Conshohocken, PA.
关14兴 ASTM Standard G63–99, 2007, “Standard Guide for Evaluating Nonmetallic Ma-
terials for Oxygen Service,” Annual Book of ASTM Standards, Vol. 14.04, ASTM
International, West Conshohocken, PA.
关15兴 Moffet, G. E., Pedley, M. D., Schmidt, N., Williams, R. E., Hirsch, D., and Benz, F.
J., “Ignition of Nonmetallic Materials by Impact of High-Pressure Gaseous Oxy-
gen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Third Volume, ASTM STP 986, D. W. Schroll, Ed., ASTM International, West Con-
shohocken, PA, 1988.
关16兴 Schmidt, N., Moffet, G. E., Pedley, M. D., and Linley, L. J., “Ignition of Nonmetal-
lic Materials by Impact of High-Pressure Oxygen II: Evaluation of Repeatability of
Pneumatic Impact Test,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Fourth Volume, ASTM STP 1040, J. M. Stoltzfus, F. J. Benz,
and J. S. Stradling, Eds., ASTM International, West Conshohocken, PA, 1989.
关17兴 Janoff, D., Pedley, M. D., and Bamford, L. J., “Ignition of Nonmetallic Materials by
Impact of High-Pressure Oxygen III: New Method Development,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Fifth Volume, ASTM
STP 1111, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West Consho-
hocken, PA, 1991.
关18兴 Wegener, W., Binder, C., Hengstenberg, P., Herrmann, K.-P., and Weinert, D.,
“Tests to Evaluate the Suitability of Materials for Oxygen Service,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Volume, ASTM
STP 986, D. W. Schroll, Ed., ASTM International, West Conshohocken, PA, 1988.
关19兴 Binder, C., Kieper, G., and Herrmann, P., “A 500 bar Gaseous Oxygen Impact Test
Apparatus for Burn-Out Testing of Oxygen Equipment,” Flammability and Sensi-
tivity of Materials in Oxygen-Enriched Atmospheres: Seventh Volume, ASTM STP
1267, D. D. Janoff, W. T. Royals, and M. V. Gunaji, Eds., ASTM International, West
Conshohocken, PA, 1995.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102293
Available online at www.astm.org/JAI

M. Carré,1 H. Barthélémy,2 J. Bruat,1 S. Lombard,1 O. Longuet,3


and J. P. Schaaff3

Identification and Quantification of


Combustion Products Released by
Non-Metallic Materials Used for Medical
Oxygen Equipment

ABSTRACT: Non-metallic materials 共plastics, elastomers, and lubricants兲 are


known to be the most critical ones in terms of safety in high pressure oxygen
distribution systems. When proper oxygen system design guidelines are not
followed, the risk of accidental ignition in service arises and has been known
for many years. It has been observed that if sufficient mass was available
when ignited, concentrations of species could arise in a level that could be
considered toxic. This is a concern in particular for those non-metallic parts
containing halogens. This risk shall especially be addressed for medical
equipment and for any high pressure breathing gas system. International
standards 共ISO 15001 or EIGA 73/08兲 recommend how to address this tox-
icity risk. They give specific recommendations such as minimising non-
metallic materials in design of equipment and use of non-metallic materials
whose thermal decomposition and/or combustion in oxygen produce less
toxic products 共with lethal concentration LC50⬎5,000 ppm兲. They also
present a list of theoretical potential combustion products but no quantitative
values of real combustion by-products from which a toxicity risk assessment
can be achieved. Then, in order to evaluate the toxicity risk for medical
oxygen and breathing gas applications, it is necessary not only to identify but
to quantify compounds generated by combustion. In this paper, an analytical
procedure used to identify and quantify the released products will be de-

Manuscript received December 12, 2008; accepted for publication August 20, 2009; pub-
lished online October 2009.
1
Air Liquide, Centre de Recherche Claude-Delorme, 78353 Les Loges en Josas, France.
2
AL Fellow, Materials Gas Cylinders and Pressure Equipment, Air Liquide Group 75321,
Paris, France.
3
Air Liquide, Centre de Technologie et d’Expertise, 93155 Blanc-Mesnil, France.
Cite as: Carré, M., Barthélémy, H., Bruat, J., Lombard, S., Longuet, O. and Schaaff, J. P.,
⬙Identification and Quantification of Combustion Products Released by Non-Metallic
Materials Used for Medical Oxygen Equipment,⬙ J. ASTM Intl., Vol. 6, No. 10.
doi:10.1520/JAI102293.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 144
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 145

scribed. The target is to analyze these compounds with the shortest delay
after combustion promoted by adiabatic compression with oxygen at 200
bars 共20,000 kPa兲. The analysis of combustion by-products must be done in
the few minutes following the combustion. Therefore direct analytical tech-
niques, like Fourier transformed infrared spectrometry and mass spectrom-
etry, were used to identify and quantify the generated species and sample
extraction was taken on the closest emission point to the combustion on the
installation. All such precautions were intended for avoiding modification or
subsequent reaction of compounds due to sample preparation steps or de-
lays in analysis. Comparisons were made for several plastics materials and
elastomers. The list of quantified combustion by-products will be presented
for polytetrafluoroethylene and polychlorotrifluoroethylene. The quantification
data obtained during this study is proposed for use in a toxicity risk assess-
ment. Then a global risk evaluation can be done including this toxicity risk
assessment and an ignition risk assessment. Based on this global risk evalu-
ation, practical recommendations could be provided.
KEYWORDS: combustion, non-metallic materials, oxygen, gas
analysis, toxicity risk, PTFE, PCTFE

Introduction
Non-metallic materials 共plastics, elastomers, and lubricants兲 are known to be at
the origin of most of the safety and reliability problems in high pressure oxygen
applications. When oxygen design guidelines and component testing are not
strictly followed, ignition sources can remain that may cause ignition of such
non-metallic materials in service. These potential ignition sources can be adia-
batic compression subsequent to a fast acting valve opening, impact of en-
trained particles by local high velocity gas flow. In these cases, it has been
observed that if sufficient mass of non-metallic material was available when
ignited, concentrations of gas species could be released by combustion in a
level that could be considered toxic 共with lethal concentration LC50
⬍ 5 , 000 ppm兲. The level of toxicity for gases and gas mixtures, defined in In-
ternational Organization for Standardization 共ISO兲 standard 10298 关1兴, is di-
vided into three groups
共1兲 Non-toxic 关when LC50 ⬎ 5 , 000 ppm 共V / V兲兴;
共2兲 Toxic 关when 200 ppm 共V / V兲 ⬍ LC50 ⱕ 5 , 000 ppm 共V / V兲兴; and
共3兲 Very toxic 关when LC50 ⱕ 200 ppm 共V / V兲兴.
Where LC50 values correspond to one hour exposure to gas and ppm 共V/V兲
indicates part per million, by volume. In our study we will focus on the species
with LC50 ⬍ 5 , 000 ppm 共V / V兲, which can be called toxic.
Because of this risk, the concentration and therefore toxicity of combustion
products needs to be considered for medical devices handling high pressure
oxygen, such as breathing and anesthetics circuits. International standards
共ISO 15001 关2兴 or EIGA 73/08 关3兴兲 recommend how to address this toxicity risk.
They give specific recommendations such as minimising non-metallic materials
in design of equipment and use of non-metallic materials whose thermal de-
composition and/or combustion in oxygen produce less toxic products 共with
lethal concentration LC50 ⬎
Copyright by ASTM Int'l (all rights reserved); Tue May
5 , 000 ppm兲. They also present a list of potential
6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
146 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

combustion products but no quantitative values of real combustion by-


products are given and a toxicity risk assessment cannot be achieved.
In order to support a toxicity risk assessment for medical oxygen and
breathing gas applications, it is necessary not only to identify and but also to
quantify compounds with low concentration of LC50 generated by such fortu-
itous combustion. The quantification of these compound concentrations pro-
duced with respect to expected amount of material burned can assist in a risk
assessment that would compare these estimated values to the LC50 of each
compound and determine their toxicity risk. It can be noticed that real quanti-
fication data concerning the released compounds after combustion are limited.
Most of the data came from theoretical calculation and not from experimental
measurement.
Due to reported accidents in 1993 in France and Canada for which the
official cause was attributed to the combustion of a PCTFE seat in a check
valve, a study was undertaken to identify the actual toxic compounds generated
by such combustion. In laboratory, tests ignition was promoted by adiabatic
compression following the standard test method described in annex of ISO
15001 关2兴. A publication 关4兴 was done in 2003 in which experiments were man-
aged for quantifying the gaseous by-products released by combustion of non-
metallic materials. However, the analytical setup and procedure at that time
had some limitations in determining the complete nomenclature of generated
products and in quantifying them accurately.
In an attempt to complete these results, this new study was done. Its target
was to identify and quantify all of the released compounds with the shortest
delay after combustion promoted by adiabatic compression with oxygen at 200
bars 共20,000 kPa兲. The analysis of combustion by-products must be done in a
few minutes following the combustion. Therefore direct fast response analyti-
cal techniques, like Fourier transformed infrared spectrometry 共FTIR兲 and
mass spectrometry 共MS兲, were used. The first objective of this paper is to de-
scribe the analytical system designed for use on adiabatic compression test
bench and the analytical procedures developed to obtain quantitative results
for several materials. Then, comparisons are made for several plastics materials
and elastomers. The list of quantified combustion by-products is presented for
several materials and a mass balance calculation is estimated for polytetrafluo-
roethylene 共PTFE兲 and polychlorotrifluoroethylene 共PCTFE兲.
The quantification data obtained during this study is proposed for use in a
toxicity risk assessment. Then a global risk evaluation can be done including
this toxicity risk assessment and an ignition risk assessment. Based on this
global risk evaluation, practical recommendations could be provided.

Standards and Regulations

Principal recommendations are described in international standards and regu-


lations. The European Industrial Gas Association 共EIGA兲 and the Compressed
Gas Association published a joined document to address design considerations
in order to help mitigate ignitions and the associated potential toxicity risk
关3,5兴. This document includes considerations in the areas of system design,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 147

system cleanliness, material susceptibility to combustion or decomposition,


and toxicity risks associated with non-metallic material decomposition. More-
over, the Standard ISO 15001 was also published concerning the use of anes-
thetic and respiratory equipment 关2兴. This international standard includes the
recommendation to perform risk analysis via ISO 14971 关6兴, directs the consid-
eration of the hazards associated with potentially toxic products of combustion
and decomposition, and details a test method for analyzing these potentially
toxic gases.
Furthermore, in France, by regulation dated Jan. 18, 2000, it is forbidden to
use valves containing seats and sealing parts made of halogenated polymers in
high pressure 200 bars 共20,000 kPa兲 oxygen systems for medical application
bundles.
In a conflicting requirement, halogenated materials have been used histori-
cally in oxygen systems due to their superior performance in resisting ignition.
The risk of moving away from halogenated materials is the introduction of less
ignition resistant materials and therefore an increased risk of fires. Oxygen
system design guidelines such as ASTM G88-90 关7兴, ISO 15001:2004 关2兴, and
component acceptance tests ISO 10297:2006 关8兴 have been established to help
designers to build systems and components that, because of their proper de-
sign, inherently prevent ignition of non-metallic parts from occurring in their
oxygen systems and hence simultaneously reduce the risk of potential exposure
to combustion products. Risk management for medical devices is proposed in
ISO 14971:2000 关6兴 and could be used for evaluating these two opposing risks.
In summary, the selection of the right non-metallic materials must in gen-
eral take into account two conflicting requirements 关15兴
共1兲 Use a less ignitable material with a lower heat of combustion and a
higher potential toxicity of burning products;
共2兲 Use a more easily ignitable material with a higher heat of combustion
and a lower toxicity potential.
In order to make a complete risk assessment, it is necessary to have sufficient
data for doing a toxic risk analysis. It is then necessary to identify and quantify
combustion products that may be toxic if generated in sufficient concentra-
tions, to support the global risk evaluation for medical oxygen and breathing
gas applications.

Previous Work in Theoretical Estimates and Identification of


Non-Metallic Combustion Products

The majority of polymers contain only carbon and hydrogen and their com-
plete combustion yields carbon dioxide and water. From the literature review
关9–13兴, it was noticed that several plastic materials with high auto-ignition tem-
peratures 共AITs兲 and commonly used in oxygen-compatible equipment contain
fluorine or chlorine. This is in particular the case of PTFE 共–共CF2 – CF2–兲n兲,
Viton 共–共–CH2 – CF2–兲n兲 . . . 共–CF2 – CF兲n, –CF3, and PCTFE 共–共CF2 – CFCl–兲n兲.
For the ones containing fluorine, chlorine or sulfur, compounds of especially
low LC50 and therefore potential toxicity could be generated. The Table 1,
found in ISO 15001:2004 and in ASTM STP 1454 关2,4兴 give a list of theoretical
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Generation of toxic gases—theoretical assumption from ISO 15001 关2兴.
b
LC50 TLVc

Downloaded/printed by
Material and Constituent Elements Potential Combustion Productsa mL/ m3 mL/ m3 mg/ m3
d
Polytetrafluoroethylene 共PTFE兲 C2Cl4O2 diphosgene 2
COCl2 phosgene 5 0,05 0,2
Polychlorotrichloroethylene 共PCTFE兲 F2O oxygen difluoride 2,6 0,05 0,1
Polyvinylchloride 共PVC兲
F2 fluorine 185 1 2
Cl2 chlorine 293 1 3
COF2 carbonyl fluoride 360
HF hydrogen fluoride 1,276 3 2,5

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
HCl hydrogen chloride 3,120 5 7,5
ClF5 chlorine pentafluoride 122
ClF3 chlorine trifluoride 299 0,1 0,4
e
Polyamide 共PA兲 NO nitric oxide 115 25 30

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


Polyurethane 共PU兲
NO2 nitrogen dioxide 115 5 9
N2O3 nitrogen trioxide 115
HCN 140 10 10
共CN兲2 cyanogens 350 10
f h
Polyethylene 共PE兲 CO carbon monoxide 3,760 50 55
148 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 1— 共Continued.兲
b
LC50 TLVc

Downloaded/printed by
Material and Constituent Elements Potential Combustion Productsa mL/ m3 mL/ m3 mg/ m3
Polypropylene sulfide 共PPS兲g H2S hydrogen sulfide 712 10 15
CH3SH methyl mercaptan 1,350
COS carbonyl sulfide 1,700
SO2 sulfur dioxide 2,520 5 13

Polyetheretherketone 共PEEK兲f CO carbon monoxideh 3,760 50 55


f h
Ethylene-propylene 共EPDM兲 CO carbon monoxide 3,760 50 55
i
SO2 sulfur dioxide 2,520 5 13

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
a
Assumes complete combustion.
b
Lethal concentration according to ISO 10298.
c
Threshold limit value: This is the maximum permanent concentration that is regulatory in a working ambient. See legal definitions for
further details.
d
The materials in this group contain: Carbon; oxygen; hydrogen; chlorine; fluorine.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


e
The materials in this group contain: Carbon; oxygen; hydrogen; chlorine; nitrogen.
f
Contains carbon; oxygen; hydrogen.
g
Contains carbon; oxygen; hydrogen; sulfur.
h
Normally transformed into CO2 because of oxygen excess.
i
Possibility of sulfur traces depending on the manufacturing process.
CARRÉ ET AL., doi:10.1520/JAI102293 149
150 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

low LC50 gases, which could be formed based solely on the chemical elements
in the chemical formula of the polymers and not on chemical reactions kinetics
or stability of theoretical products. This table was used as a starting point from
which to perform analytical experiments and determine what compounds
might realistically be found as a combustion product of these polymers 共see
Table 1兲. It should be noted however, that toxicology data is always being up-
dated as new information comes to light and that some of the values published
in ISO 15001 have changed. Nonetheless, these numbers give us a guide by
which to compare the toxicity of the compounds presented.
In this table, the gases are grouped into categories, the ones containing
fluorine and/or chlorine in the chemical formula 共they are the ones susceptible
to generate the lowest LC50 gases兲, the ones containing nitrogen, 共these are
susceptible to generate low LC50 nitrogen-bearing gaseous compounds兲, the
ones containing sulfur 共risk of generation of low LC50 sulfur-bearing gaseous
compounds兲 and the lower risk gases containing only carbon and hydrogen. In
this latter case, carbon monoxide might be generated but in fact its formation
is very unlikely in the presence of high pressure oxygen.
In an effort to identify what products are realistically formed during poly-
mer combustion analytical work was performed in a previous study 关4兴. At that
time, each plastic was submitted to adiabatic compression test and two types of
gas samples were then extracted just after completing the test
共1兲 One sample was passed directly through the Dräger tubes or the chemi-
luminescence apparatus 共for NO2 analysis兲 connected in series or in
derivation to the outlet of the adiabatic compression tube. This type of
sample allows an initial, approximate analysis to be made. It can detect
gases that are present in high concentrations and is a good method to
try to approach detection of unstable gases that are prone to react
further. These gases may not be seen if the analysis was performed at a
later time. Nonetheless, it is not a highly accurate method and prone to
interferences.
共2兲 The second type of analysis consisted of collecting a gas sample at the
outlet of the adiabatic compression tube using a cylinder or a bubbler
to be kept for off-site deferred-time analysis. The hydrolyzed products
in the bubblers permitted chemical analysis to be made 共looking in
particular for hydrogen cyanide 共HCN兲 or sulfurated products兲 and the
cylinder samples were used to carry out infrared 共FTIR兲, ultraviolet
共UV兲, and MS analysis.
The products identified and quantified are summarized in Table 2.
These results give useful information concerning the released compounds.
Measurement of the major compounds of concern in Table 2 could not be made
immediately following the combustion due to the necessary transfer of com-
bustion products to analytical facility for FTIR, UV, or MS analysis. Captured
species could continue to react in the gaseous state or with the container sur-
faces in the interval between the sampling and the analysis, leading to the
potential misrepresentation of the identified products. Due to lack of reaction
kinetic information, it is unknown if the mixture stabilizes in a matter of mil-
liseconds or hours. However, it can be assumed that the worst case scenario for
the major compounds of concern is immediately after ignition as these com-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—Generation of combustion gases—analytical results from 关4兴.

With Dräger
Products Tubes Chemiluminescence FTIR UV MS Bubbler

Downloaded/printed by
Kel-f Not done Not done CO2, CF4, COCl2 CO2, CF4 Not done
CH4, CoCl2, COF2, HCl
COF2, HF⬎ few 100 ppm COCl2
⬎ few 100 ppm
Vespel NOX ⬍ 5 ppm NO2 ⬇ 1 ppm H2O, CO2 Not done H 2O , N 2 no HCN
HCN measure O2, CO2
contaminated by
NOX
Vulkollan NOX ⬍ 5 ppm Not done H2O, CO2 Not done H 2O , N 2 No HCN
HCN not measurable CH4 共traces兲 O2, CO2
CO 共traces兲

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Ultrason PES no CS2, ¯ CH4 共10 ppm兲 SO2 共⬍1 ppm兲
H2SO4, SO2 CO2 共1.7 %兲
Note: The major toxic products 共with lower LC50兲 are shown in bold letters.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


CARRÉ ET AL., doi:10.1520/JAI102293 151
152 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Description of test bench used in ISO 15001 关2兴 and in previous publication
关4兴.

pounds are highly reactive and will have limited time to react with other spe-
cies to potentially recombine into other more stable compounds with higher
LC50 if they have not already done so. Moreover the quantification of all com-
pounds was not possible due to the lack of standard mixtures for calibration.
For a complete toxicity risk analysis it is necessary to quantify combustion
products in order to determine dose and its comparison to LC50 values. In
order to objectively compare various polymers and their gaseous combustion
by-products toxicity a new study was undertaken to identify and then quantify
accurately all such by-products of combustion.

Identification of Gaseous Combustion Products by Fourier Transformed


Infrared Spectrometry Analysis Using 26 L Adiabatic Combustion Test
Bench
The target of this study is to determine the combustion compounds immedi-
ately following the promoted burning event in oxygen. The adiabatic compres-
sion test procedure, described in our previous publication 关4兴 and recom-
mended in ISO 15 001 关2兴, was modified in the purpose of allowing direct
analysis.

Test Apparatus
The test apparatus 共see Fig. 1兲 consisted of a high pressure oxygen supply con-
nected, via an inlet valve, to a vessel equipped with a preheating device which
heats the oxygen up to 60° C.
Oxygen at 60° C and high pressure is made to suddenly impact the material
to be tested via a 14-mm diameter, 750-mm-long Monel tube, the inlet of which
is opened and closed by a quick-opening valve.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 153

Some modifications were applied to this standard ISO setup and procedure
to comply with the specific objectives of our study and integrate the analytical
constraints
共1兲 Purging Cylinder 10 and Purging Valve 9 were removed to avoid dead
end and Oxygen Supply 1 was used to purge the system;
共2兲 No sampling Cylinder 12 and Cylinder Valve 13 was used because the
line after the Shutoff Valve 11 was directly connected to FTIR and/or
mass spectrometry 共MS兲 analyzer;
共3兲 A Shut-off Valve 6 and a Check Valve 5 were provided to isolate the
sample tested and the combustion products released. Considering the
difficulty to avoid ignition at these locations and to prevent contamina-
tion of the combustion gases, these two valves were not used to per-
form the tests;
共4兲 The diameter of the connection tube is 14 mm with a length of 750
mm;
共5兲 The sampling line between combustion chamber and FTIR is 3 m
length with a diameter of 2 mm; and
共6兲 To reduce the volume of gas to be analyzed, we performed the adiabatic
compression at 200 bars 共20,000 kPa兲 instead of 240 bars 共24,000 kPa兲.
In these conditions the total volume of oxygen in the test bench is 26 l
共around 1.1 moles of O2兲. Provided the combustion of tested material
was complete, it was not noticed that there were any difference be-
tween identification of combustion products produced during this test-
ing at 200 bars 共20 000 kPa兲 or previous testing performed at 240 bars
共24,000 kPa兲 关4兴.

Sample Holder
The material sample to be tested was set up within a sample holder in brass at
the downstream end of the connector tube. It was important to ensure that any
material used to tightly seal the sample holder 共e.g., PTFE tape兲 was protected
again the risk of ignition or decomposition during the adiabatic compression,
so as not to interfere with results interpretation.

Protocol for Gaseous Analysis of Combustion Products with Fourier Trans-


formed Infrared Spectrometry
In this experiment, FTIR was used to identify the generated species. The analy-
sis of combustion by-products, as mentioned previously, was done as soon as
possible after the combustion, by direct extraction on the installation avoiding
modifications or reactions of compounds due to prolonged time for sample
preparation. The sampling system was arranged for direct connection to the
FTIR cell 共see Fig. 2兲. In this experiment, no passivation procedure was done on
the tubing between combustion chamber and FTIR.
The operating conditions used for FTIR analysis were as follows:
共1兲 Type of equipment: FTIR spectrometer Nicolet 380;
共2兲 Optical path of cell: 10 m;
共3兲 Volume of cell: 2 L;
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
154 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Sampling system for direct extraction and introduction into FTIR cell.

共4兲 Spectral range measured: 650– 4 , 200 cm−1;


共5兲 Number of scans: 64;
共6兲 Total time for one FTIR measurement 共for 2 L兲: 2 min, 20 s;
共7兲 Time for purge between each FTIR measurement: Around 3 min; and
共8兲 Total time for the analysis of the 26 L: Around 60 min 共including
purge兲.
The tube after Valve 11 is in stainless steel with dimension of 3 m in length and
2 mm in diameter. The filter and the valves are also in stainless steel.
At least three test samples were subjected to experiment to achieve accept-
able reproducibility of combustion product analysis. Samples of weight be-
tween 150 and 300 mg 共from 0.6 to 3 millimoles of plastic兲 were used, in order
to obtain complete combustion and to avoid excessive concentration of com-
bustion products to be analyzed. These masses could be larger than the ones in
use for oxygen components that are in the 30–100 mg range. Larger quantities
of non-metallic products used in this experiment aid identification and quanti-
fication of by-products but concentrations detected need to be scaled in order
to be comparative to real life scenarios.
The adiabatic compression cycle was performed as follows:
共1兲 The test sample was set up in the sample holder 共after recording the
exact weight of sample兲;
共2兲 The test and sampling systems were purged with oxygen through the
FTIR cell;
共3兲 Valve 11 was closed to isolate the sampling line;
共4兲 The sampling line was pumped down to 0.02 bar 共2 kPa兲. No passiva-
tion was applied in this experiment;
共5兲 The gas flow was controlled by the quick-opening Valve 4 to increase in
a very short time 共20+ 0 – 5 ms兲 the pressure of the test sample from
atmospheric pressure to maximum test pressure 200 bars 共20,000 kPa兲.
After the test, Valve 4 was rapidly closed 共less than 30 ms兲 to prevent
any backflow of the combustion products. The ignition of the sample
could be verified by monitoring the increase in temperature of sample
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 155

holder. The total volume of oxygen in the adiabatic compression bench


was around 26 L;
共6兲 Valve 11 was opened to fill the sampling line and to flush the gaseous
combustion products off the connection tube and sample holder into
the FTIR cell 共with outlet valve closed兲. The FTIR cell was filled up to
1.2 bar 共120 kPa兲 using for control the manometer at the outlet of the
cell and by closing the valve at the inlet of the cell. The volume of the
cell is 2 L, and then each sampling volume is around 2.4 Nliters;
共7兲 The analysis of the gas was performed in the cell in static mode;
共8兲 The cell was purged by opening the outlet valve and pump down to 0.02
bar 共2 kPa兲;
共9兲 The outlet valve was closed and the inlet valve was opened to refill the
cell up to 1.2 bar 共120 kPa兲. The volume of FTIR cell is 2 L at atmo-
spheric pressure. Steps 6–9 were repeated until purging all the oxygen
and the combustion products in the adiabatic compression bench. This
resulted in eleven to twelve successive FTIR measurements per test
sample;
共10兲 Valve 11 was closed to isolate the sampling system;
共11兲 The sampling line was purged through the FTIR cell and then
pumped down to 0.02 bar 共2 kPa兲;
共12兲 The sample holder was disassembled, and rinsed with an appropriate
solvent, e.g., water;
共13兲 The sample holder was dried and then another sample was set up and
connected to the apparatus; and
共14兲 Valve was re-opened, the system was purged and the process was
started again from Step 2.

Materials Tested
共1兲 PCTFE 共kel F®兲;
共2兲 PTFE;
共3兲 FKM 共Viton®兲;
共4兲 Polyamide PA 共Tecamid 66®兲; and
共5兲 EPDM 共ethylene-propylene兲.
We selected three fluorinated products 共PCTFE, PTFE, and Viton®兲, one prod-
uct containing nitrogen 共polyamide兲 and one composed in part of C, O, and H
but that could possibly contain sulfur depending of the manufacturing process
共EPDM兲. Each sample was submitted to adiabatic compression 共as indicated
before兲 and the combustion products were identified and appreciatively quan-
tified by FTIR.

Analytical Results with Fourier Transformed Infrared Spectrometry Using 26 L


Test Bench
Following our procedure, eleven to twelve successive FTIR analyses were
achieved per test sample. Due to the fact that our sampling port location was
close to the sample holder, the first analysis gave the highest concentration of
combustion products. Then the following samples show decreasing values 共see
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
156 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—FTIR spectra obtained for PTFE. In red-measurement 1: 4 min after combus-
tion. From 2 to 8: 5.5 min between each measurement. In blue-measurement 8: 42.5
min after combustion.

Fig. 3兲. These decreasing values seem to show that gas diffusion is limited
between combustion products and oxygen. Indeed, even after 45 min the con-
centration of combustion products in oxygen is not constant. The lower con-
centration is always observed at the last analysis.
According to the FTIR spectra obtained for each material we were able to
identify the major components released. The identification was done by com-
parison of these spectra with the spectra of each pure compound
共CO2 , CF4 , CH4. . .兲 at concentrations between 1,000 ppm and 1 %. The spec-
trum of each pure compound was obtained by analysis of mixtures in cylinders.
In the case of COF2, no mixture is available in cylinders and the spectrum of
the product at 6,000 ppm, obtained in infra-red 共IR兲 spectra database, was used
for comparison. Some overlapping in wavelength was observed for CF4 and
COF2 but the identification was possible based on other absorption wave-
lengths. It was also possible to prove that some products listed in Table 1 were
not generated 共or below the detection limit of FTIR analyzer兲. Indeed, in Fig. 4,
the spectrum of COCl2 at 240 ppm is compared to the one of combustion
by-product of Kel-F. The specific wavelength of COCl2 cannot be seen in the
released products from Kel-F combustion.
The gases identified by FTIR and their range of concentrations are indi-
cated in Table 3.
Data confirmed that some of PCTFE’s main products of combustion are low
LC50 compounds of concern 共LC50 below 5,000 ppm兲. In reference to Table 1,
the generation of COF2, and HF was evidenced. On the other hand, C2Cl4O2,
COCl2, F2O, HCl, ClF5, and ClF3 were not detected even after further checking
on all the spectra obtained for each sample 共eleven spectra were measured for
each plastic sample兲. It can be concluded that lowest LC50 gas released by
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 157

FIG. 4—Spectra comparison between COCl2, COF2, CF4, and CO2 mixtures with com-
bustion products of Kel-F.

PCTFE is COF2 as phosgene was not detected a short time after combustion.
FTIR is of course not able to detect homo-nuclear molecules F2 and Cl2, and
additional measurements by MS need to be done to monitor their concentra-
tions. CF4 and CO2 were also easily detected as they were generated as main
products of combustion.
The main generated compounds are the same for PTFE except the chlori-
nated ones: COF2, CO2, and CF4. Again for analysis of F2, it will be necessary to
use MS.

TABLE 3—Identification of gaseous combustion by-products by FTIR.

Polyamide
Material PCTFE PTFE 共Nylon兲 EPDM FKM
Major COF2, CO2, COF2, CO2, CO2 CO2 CO2
generated CF4, CF4
products
⬎1000 ppm

Other products HF, SiF4 HF NO2 CO HF, CO,


⬎30 ppm and SO2, SiF4
⬍100 ppm
Traces of C 2H 4, C 2H 6, COF2, CF4,
products C4H10 HCL
⬍30 ppm
Non-detected COCl2, OF2, OF2 N2O3, NO, SO2
products C2Cl4O2, HCN, 共CN兲2
HCL, ClF5,
ClF3
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
158 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Only CO2 was released at high level by polyamide. NO2 was detected but at
lower values 共below 50 ppm for 150 mg of polyamide兲, even lower than the ones
estimated by mass balance from the chemical formula of polyamide. This gen-
erated product must be considered because its LC50 concentration is at 115
ppm. It should also be noted that no HCN nor 共CN兲2 were detected. Such toxic
gases had been observed upon burning in air of a variety of polymers, especially
the nitrogen-containing materials 关14兴. In high pressure oxygen, HCN cannot
be formed because the increase of oxygen supply to reaction reduces the quan-
tity of HCN formed 关14兴.
For EPDM, no SO2 was detected and only CO2 and CO were found in larger
quantities 共above 100 ppm兲. FKM releases CO2 共main compound兲 and also in
concentration below 50 ppm HF, CO, SO2, and SiF4. SiF4 could come from the
additives used to manufacture Viton®.
Table 3 clearly stresses the great difference between the types of com-
pounds and their general levels of production that could be formed by the
various polymers. It can be seen that halogenated ones have a tendency to form
greater quantities of low LC50 molecules and therefore have a greater potential
for toxicity than those containing nitrogen or sulfur as they were mentioned in
publications in Refs 2–4.

Quantification of Gaseous Combustion Products from Polytetrafluoroet-


hylene and Polychlorotrifluoroethylene by Fourier Transformed Infrared
Spectrometry and Mass Spectrometry Analysis Using 10 L Adiabatic
Combustion Test Bench
From the work presented earlier, it has been shown that PTFE and PCTFE are
the materials which can generate the largest quantity of low LC50 compounds.
In order to precisely determine the amount of such products released by both
materials, a specific campaign of analysis was performed including the use of
FTIR and MS. The target was to quantify the number of moles of combustion
product generated overall per gram of material burned. As discussed in the
previous section, the adiabatic compression test bench has a total of 26 L of gas
available for extractive analysis which requires a minimum of eleven to twelve
analyses for each tested material. In order to obtain the global quantity of
generated combustion by-products for each plastic sample, it is necessary to
add the quantification result of the eleven of twelve measurements. Hence, if
we cumulate twelve analytical results, the global uncertainty of the total
amount is higher than the uncertainty for each measurement. To improve the
accuracy and precision of the analytical methodology it was necessary to re-
duce the number of measurement per plastic sample. Then it was decided to
use another type of adiabatic compression test bench that features a capacity of
only 10 L of gas. For analyzing this volume of gas a minimum of four analyses
must be made. Moreover this installation can be used for reactive gases like
fluorine and chlorine.

Test Apparatus
The combustion in pure oxygen is promoted by adiabatic compression at 200
bars 共20,000 kPa兲 with 10 L of oxygen and at room temperature 共20° C兲. These
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 159

FIG. 5—Adiabatic compression test bench for quantification of toxic compounds.

parameters are equal to the ones used for the 26-L test bench with the excep-
tion of a more limited volume of oxygen available for combustion. The target
here again is not to assess whether the material is prone to burning in stan-
dardized conditions but to find conditions where the material totally burns and
to analyze more precisely the gases generated by this combustion. In these
conditions the amount of oxygen is about 420 millimoles and the maximum
size of plastic sample is about 300 mg, which corresponds to a maximum of 3
millimoles of plastic monomer. The quantity of oxygen should be enough for
complete combustion.
This test is therefore done with the bench design presented in Fig. 5
With this test setup, the adiabatic compression is triggered by the bursting
of a rupture disk 13兲 calibrated at 200 bars 共20,000 kPa兲.
In a first step: Oxygen was introduced into the O2 cavity 4兲 at 50 bar and
into the connection tube at 1 bar. Then the rapid-opening valve 2兲 was opened
to introduce N2 at 300 bar, having the effect of rapidly pushing the piston
forward. The piston then compressed the oxygen in the cavity up to 200 bar
共set-point rupture pressure of the disk兲.
Once the rupture disk bursts, the oxygen from the cavity now at 200 bar
adiabatically compressed the downstream volume of oxygen previously at at-
mospheric pressure thus applying the gaseous impact to the sample holder 12兲.

Sample Holder
The sample holder 12兲 is made of brass, and is the same as the one used in the
26 L adiabatic compression testing. The cavity made of Monel had dimensions
of 14-mm diameter and 750-mm length. All the tubes and valves that were used
are made of Monel with the exception of the sampling line and Valve 2 which
are made of stainless steel as in the previous experiments.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
160 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Sampling system for FTIR and mass spectrometer.

Protocol for Gaseous Analysis of Combustion Products with Fourier Trans-


formed Infrared Spectrometry and Mass Spectrometry
For quantification of released compounds from PTFE and PCTFE, it was nec-
essary to use both FTIR and MS. They allowed quantifying all of the generated
species including F2 and Cl2. For the same reasons that in the previous experi-
ment, the analysis of combustion by-products was done in very short deferred-
time after the combustion, again with direct extraction. In this phase, the sam-
pling system was adapted to add MS in addition to FTIR for direct sampling
共see Fig. 6兲.
In order to limit the potential loss of F2 in the sampling lines and obtain
accurate values for F2 quantification, all the sampling lines were passivated
with a mixture of F2 0.5 % in nitrogen.
At least three experiments on test samples were carried out to achieve ac-
ceptable reproducibility of combustion product analysis. Sample of weight be-
tween 150 and 250 mg were again used in the purpose of achieving complete
combustion and limiting the concentration of combustion product to be ana-
lyzed. These quantities are larger than the average one used for oxygen compo-
nents that are in the 30–60 mg range. The concentrations detected here need to
be scaled in order to be comparative to real life scenarios in risk analysis. The
test was performed as follows:
共1兲 The test sample was set up in the sample holder 共record the exact
weight of sample before兲;
共2兲 The test and sampling systems were purged with oxygen through
Valves 5 and 11;
共3兲 The valve 共10兲 was closed to isolate the sampling line;
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 161

共4兲 The sampling line was pumped down to 0.02 bar 共the sampling line was
passivated with 0.5 % F2 in nitrogen two times during 20 min and
pumped down to 0.02 bar兲;
共5兲 The O2 cavity 共4兲 was filled with 50 bar of oxygen and then the valve 共5兲
was closed 共total volume of O2 is 10 L兲 ;
共6兲 The connecting tube section 共14兲 and the sample holder 共12兲 were filled
with oxygen up to 1 bar and then the valve 共11兲 was closed;
共7兲 The rapid valve 共2兲 was opened to make oxygen from high pressure
supply push forward the piston 共3兲 which in-turn compressed the oxy-
gen in cavity up to the rupture disk set-point pressure 共200 bar兲. The
rupture disk 共13兲 breaks and generated the adiabatic compression of
the oxygen within the connector 共14兲. The actual ignition of the sample
was followed by measuring the increase in temperature of sample
holder;
共8兲 Valves 10 and 11 were opened to fill the sampling line and to flush the
gaseous combustion products from the connection tube and sample
holder into the FTIR cell 共with outlet valve closed兲. The FTIR cell was
filled up to 1.2 bar using the manometer at the outlet of the cell to
control pressure and the valve at the inlet of the cell was closed. The
volume of FTIR cell is 2 L at atmospheric pressure;
共9兲 The analysis of the gas in the FTIR cell was performed in static mode;
共10兲 The cell was purged by opening the outlet valve and directing the gas
to the mass spectrometer for analysis of F2 and Cl2;
共11兲 The FTIR cell was pumped down to 0.02 bar;
共12兲 The outlet valve was closed and the inlet valve was opened to refill the
cell up to 1.2 bar.
Due to the 2 L atmospheric pressure FTIR volume and the 10 L of
combustion gases to analyze, the process was repeated from Steps
8–12 until purging all the oxygen and the combustion products in the
adiabatic compression bench 共repeat four or five times兲.
共13兲 Valve 10 was closed to isolate the sampling system;
共14兲 The sampling line was purged through the FTIR cell and then was
pumped down to 0.02 bar;
共15兲 The sample holder was disassembled and rinsed with an appropriate
solvent, e.g., water;
共16兲 The sample holder was dried, and the second material sample was set
up and connected to the apparatus; and
共17兲 Re-open. The valve was re-opened, the system was purged and the
process was repeated from Step 2.

Materials Tested
共1兲 PCTFE 共Kel F®兲; and
共2兲 PTFE
Only the two fluorinated polymers were selected 共PCTFE, PTFE兲 because
they were the ones generating the largest quantity of the low LC50 compounds
of concern. Each sample was submitted to adiabatic compression 共as described
before兲 and the combustion products were quantified by both FTIR and MS.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
162 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Quantification Results with Fourier Transformed Infrared Spectrometry and


Mass Spectrometry Using 10 L Test Bench and Reaction Modelling: PTFE
Following the protocol described before, four or five successive FTIR and MS
measurements were achieved per test sample. The time for one measurement is
around 24 min 共2.5 min for FTIR+ 20 min for MS+ 1.5 min of purge兲 and the
total time of analysis for one sample is between 96 and 120 min for four or five
measurements.
The calibration of FTIR was made with COF2, CO2, and CF4 standard mix-
tures before the experiment. The uncertainty for each measurement is calcu-
lated by taking in account the major sources of uncertainty
共1兲 The repeatability of FTIR measurement: 0.6 % relative to the measure-
ment;
共2兲 The uncertainty of standard mixtures 共from 2 to 20 % relative depend-
ing of the molecule兲; and
共3兲 The uncertainty of the sampling: between 5 to 30 % relative taking into
account the kind of molecules and the precision of the manometer used
to measure cell pressure.
The total amount of generated moles for each sample is calculated using the
analytical concentration of each measurement. Knowing the volume of gas
sampled for each measurement 共volume of FTIR cell is known and the pressure
is measured兲 it is possible to calculate the amount of moles in each measure-
ment. The total amount of generated moles is given by the sum of each molar
amount obtained for each measurement 共in this experiment the sum of the four
or five measurements兲.
The global uncertainty of this total amount of generated moles is 15 %
relative for CO2, 15 % relative for CF4, and 30 % relative for COF2. These
uncertainties are estimated for a risk of 33 %. All the results with FTIR are
summarized in Table 4.
The calibration of MS was made before the experiment with F2 and Cl2
standard mixtures at 5000 ppm in nitrogen. The uncertainty for each measure-
ment is calculated by taking in account the major sources of uncertainty
共1兲 The repeatability of MS measurement: 2 % relative to the measure-
ment;
共2兲 The uncertainty of standard mixtures 共5 % relative兲; and
共3兲 The uncertainty of the sampling: 15 % for F2 to 30 % for Cl2 taking into
account the possible loss of these reactive molecules.
The total amount of generated moles for each sample is calculated using the
analytical concentration of each measurement. All the gas sampled in the FTIR
is sent to MS, then, knowing the volume of gas sampled for each measurement
共volume of FTIR cell is known and the pressure is measured兲, it is possible to
calculate the amount of generated moles in each measurement. Like for FTIR,
the total amount of generated moles is given by the sum of each molar amount
obtained for each measurement 共in this experiment the sum of the four or five
measurements兲.
The global uncertainty of this total amount of generated moles is 15 %
relative for F2 and 30 % relative for Cl2 due to possible loss during sampling.
These uncertainty are estimated for a risk of 33 %. All the results with MS are
summarized in the Table 4.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 4—Number of moles generated per PTFE samples analyzed by FTIR and MS.
Total Moles Total Moles Total Moles Total Moles
of COF2 Uncertainty of CF4 Uncertainty of CO2 Uncertainty of F2 Uncertainty

Downloaded/printed by
Mass of Generated for COF2 Generated for CF4 Generated for CO2 Generated for F2
sample in per 共30 % per 共15 % per 共15 % per 共15 %
Milligrams Sample Relative兲 Sample Relative兲 Sample Relative兲 Sample Relative兲
272.08 0.000524 +/− 0.000157 0.001805 +/− 0.000271 0.002182 +/− 0.000327 Not done +/−
284.73 0.000810 +/− 0.000243 0.002279 +/− 0.000342 0.002030 +/− 0.000305 Not done +/−
166.53 0.000561 +/− 0.000168 0.001499 +/− 0.000225 0.001440 +/− 0.000216 Not done +/−
158.21 0.000349 +/− 0.000105 0.001094 +/− 0.000164 0.001354 +/− 0.000203 0.000274 +/− 0.00041
157 0.000400 +/− 0.000120 0.001104 +/− 0.000166 0.001301 +/− 0.000195 0.000238 +/− 0.00036

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


CARRÉ ET AL., doi:10.1520/JAI102293 163
164 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Relationship between mass of PTFE and generated compounds.

As a countercheck of our results it was tried to model the chemical reaction


involved in combustion with O2. Knowing the major compounds generated
共CF4 , CO2 , COF2 , F2兲 it can be assumed the following reaction scheme:
共CF2 – CF2兲“monomer” + O2 → nCOF2COF2 + nCF4CF4 + nCO2CO2 + nF2F2
where:
nspecies = coefficients to be determined.
Since the mass of material used for each test is known, we can calculate the
number of moles of PTFE “monomer” units in our sample 共assuming a molar
weight of 100.01 g for PTFE兲. There should be proportionality between mass of
PTFE and number of moles of generated products. Indeed, using our analytical
results we were able to establish such linear relationship 共see Fig. 7兲
The slopes of the regressions are representative of the coefficients nspecies of
the chemical reaction. Based on these values, we obtained an estimation of the
coefficients that is presumably correct although we do not have enough
samples to calculate the regression coefficients with more confidence. Then we
apply correction to these values to obtain elemental balance 共for O, C, F兲 for the
reaction scheme
共CF2 – CF2兲“monomer” + O2 → 0.3COF2 + 0.775CF4 + 0.925CO2 + 0.15F2
By assuming this reaction, we can evaluate how well the analytical results fit
our reaction model by comparison between calculated and analyzed numbers
of moles of combustion products. The Table 5 shows the recovery rate obtained
by the FTIR and MS quantitative analysis compared to the reaction mass bal-
ance model.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 5—Comparison between quantitative FTIR and MS analyzed data and mass balance reaction model moles of combustion products
of PTFE.

Milligrams of PTFE Sample Moles of CO2 Moles of COF2 Moles of CF4 Moles of F2

Downloaded/printed by
Total moles analyzed 272.08 0.00218 0.00052 0.00180 Not done
Total moles calculated 272.08 0.00252 0.00082 0.00211 Not done
% of recovery 87 % 64 % 86 %
Total moles analyzed 284.73 0.00203 0.00081 0.00228 Not done
Total moles calculated 284.73 0.00263 0.00085 0.00221 Not done
% of recovery 77 % 95 % 103 %
Total moles analyzed 166.53 0.00144 0.00056 0.00150 Not done
Total moles calculated 166.53 0.00154 0.00050 0.00129 Not done
% of recovery 94 % 112 % 116 %
Total moles analyzed 158.21 0.00135 0.00035 0.00109 0.00027396

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Total moles calculated 158.21 0.00146 0.00047 0.00123 0.00023728
% of recovery 93 % 74 % 89 % 115 %
Total moles analyzed 157 0.00130 0.00040 0.00110 0.00023849
Total moles calculated 157 0.00145 0.00047 0.00122 0.00023547
% of recovery 90 % 85 % 91 % 101 %

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


CARRÉ ET AL., doi:10.1520/JAI102293 165
166 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

One can see that for compounds analyzed by FTIR and MS, closing of the
mass balance is fairly good considering the aforementioned uncertainties and
approximations.

Quantification Results for Polychlorotrifluoroethylene with Fourier Transformed


Infrared Spectrometry and Mass Spectrometry Using 10 L Test Bench and Reac-
tion Modelling
The quantification of released compounds was made by five successive FTIR
and MS measurements per test sample. The time for one measurement is
around 13 min 共2.5 min for FTIR+ 9 min for MS+ 1.5 min of purge兲 and the
total time of analysis for one sample is between 52 and 65 min for four or five
measurements.
The calibration of FTIR was made with COF2, CO2, and CF4 standard mix-
tures before the experiment. The uncertainty for each measurement is calcu-
lated by taking in account the major sources of uncertainty, as explained in
previous experiment. The total amount of generated moles for each sample is
calculated using the analytical concentration of each measurement. All the gas
sampled in the FTIR is sent to MS, then, knowing the volume of gas sampled
for each measurement 共volume of FTIR cell is known and the pressure is mea-
sured兲, it is possible to calculate the amount of generated moles in each mea-
surement. Like for FTIR, the total amount of generated moles is given by the
sum of each molar amount obtained for each measurement 共in this experiment
the sum of the four or five measurements兲.
The global uncertainty of this total amount of generated moles is 15 %
relative for F2, CO2, and CF4 and 30 % relative for Cl2 and COF2 due to possible
loss during sampling. These uncertainty are estimated for a risk of 33 % 共k
= 1兲. All the results with MS and FTIR are summarized in the Table 6.
Knowing the major compounds generated 共CF4, CO2, COF2, F2, and Cl2兲
we can assume the following reaction with O2:
共CF2 – CFCl兲“monomer” + O2 → aCOF2COF2 + aCF4CF4 + aCO2CO2 + aCl2Cl2 + aF2F2
COFCl and COCl2 products were not detected by FTIR nor by MS and are not
included in the chemical balance.
The same calculation of number of moles of PCTFE 共here again in the
sense of moles of the monomer unit兲 can be done with a molar weight of the
monomer of 116.5 g. There is a linear relationship between moles of PCTFE
and moles of generated products. Using the analytical results it was possible to
evaluate this linear relationship 共see Fig. 8兲
Compared to the previous case of PTFE, here the regression coefficients
can be obtained with better confidence, except for COF2 and Cl2 that poses
specific analytical problems. Again, coefficients are corrected to provide el-
emental balancing of the reaction scheme
共CF2 – CFCl兲“monomer” + O2 → 0.25COF2 + 0.55CF4 + 1.2CO2 + 0.5Cl2 + 0.15F2
Assuming this reaction it is possible to evaluate how well the analytical results
fit our reaction model by comparison between calculated and analyzed num-
bers of moles of combustion products. Table 7 shows the recovery rate obtained
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 6—Number of moles generated per PCTFE samples analyzed by FTIR and MS.
Total moles
Total Moles Uncertainty of CF4 Uncertainty Total Moles Uncertainty Total Moles Uncertainty Total Moles Uncertainty
Mass of of COF2 for COF2 Generated for CT4 of CO2 for CO2 of F2 for F2 of Cl2 for Cl2

Downloaded/printed by
PCTFE in Generated 共30 % per 共15 % Generated 共15 % Generated 共15 % Generated 共30 %
Milligrams per Sample Relative兲 Sample Relative兲 per sample Relative兲 per Sample Relative兲 per Sample Relative兲
134.6 0.000405 +/− 0.000122 0.000637 +/− 0.000096 0.001312 +/− 0.000197 Not done +/− Not done +/−
143.37 0.000268 +/− 0.000080 0.000583 +/− 0.000087 0.001075 +/− 0.000161 Not done +/− Not done +/−
135.73 0.000354 +/− 0.000106 0.000664 +/− 0.000100 0.001233 +/− 0.000185 Not done +/− Not done +/−
150.53 0.000299 +/− 0.000090 0.000732 +/− 0.000110 0.001380 +/− 0.000207 Not done +/− Not done +/−
158.36 0.000180 +/− 0.000054 0.000695 +/− 0.000104 0.001546 +/− 0.000232 0.00028 +/− 0.000042 0.00056 +/− 0.000167
141.92 0.000212 +/− 0.000063 0.00064897 +/− 0.000097 0.00132521 +/− 0.000199 0.00018 +/− 0.000027 0.00053 +/− 0.000159
176.72 0.000271 +/− 0.000081 0.00077696 +/− 0.000117 0.00158468 +/− 0.000238 0.00028 +/− 0.000042 0.00057 +/− 0.000171

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


CARRÉ ET AL., doi:10.1520/JAI102293 167
168 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 8—Relationship between mass of PCTFE and generated compounds.

by the quantitative FTIR and MS analysis compared to the mass balance


model.
Again closing of mass balance may be considered as satisfactory, the largest
discrepancies being for COF2. It could be noticed that the results for F2 are
generally above the calculated value. On the other hand the results for Cl2 are
below the calculated value. This could be related to loss during the sampling of
the gas sample after the combustion. Indeed, the sampling tube which was in
stainless steel was passivated with F2 mixture only.

Toxicity Risk Discussion for Polytetrafluoroethylene and Polychlorotrifluoroeth-


ylene
The analytical results obtained for PTFE and PCTFE show that these two ma-
terials are the ones able to generate the lowest LC50 value compounds. In order
to evaluate the real toxicity risk concerning these two materials, it is possible to
use the equation mass balance model explained previously. Indeed it was
shown that there is proportionality between mass of material and number of
moles of generated gases.
The following tables present the quantity of the gases with lowest LC50
generated by combustion of different masses of PTFE and PCTFE 共from 20 to
200 mg兲 using the proposed model.
Using the equation mass balance model it is possible to approximate the
number of mole of COF2 and F2 generated by combustion of masses of PTFE
from 20 to 200 mg 共Table 8兲. The same calculation can be done for COF2, F2,
and Cl2 with PCTFE 共Table 9兲.
Knowing this number of moles, it is possible to approximate the concen-
tration of COF2, F2 and Cl2 in a known volume of oxygen.
This discussion exemplifies the complexity of understanding toxicity con-
cerns due to impact of the scenario on the products encountered. Though it is
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 7—Comparison between quantitative FTIR and MS analyzed data and mass balance reaction model moles of combustion products
of PCTFE.

Milligrams of sample Moles of CO2 Moles of COF2 Moles of CF4 Moles of F2 Moles of Cl2

Downloaded/printed by
Total of moles analysed 134.6 0.00131 0.00041 0.00064 Not done Not done
Total of moles calculated 0.00139 0.00029 0.00064 Not done Not done
% recovery 95 % 140 % 100 %
Total of moles analysed 143.37 0.00108 0.00027 0.00058 Not done Not done
Total of moles calculated 0.00148 0.00031 0.00068 Not done Not done
% recovery 73 % 87 % 86 %
Total of moles analysed 135.73 0.00123 0.00035 0.00066 Not done Not done
Total of moles calculated 0.00140 0.00029 0.00064 Not done Not done
% recovery 88 % 121 % 104 %
Total of moles analysed 150.53 0.00138 0.00030 0.00073 Not done Not done

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Total of moles calculated 0.00155 0.00032 0.00071 Not done Not done
% recovery 89 % 92 % 103 %
Total of moles analysed 158.36 0.00155 0.00018 0.00070 0.00028 0.00056
Total of moles calculated 0.00163 0.00034 0.00075 0.00020 0.00068
% recovery 95 % 53 % 93 % 136 % 82 %

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


Total of moles analysed 141.92 0.00133 0.00021 0.00065 0.00018 0.00053
Total of moles calculated 0.00146 0.00030 0.00067 0.00018 0.00061
% recovery 91 % 69 % 97 % 100 % 87 %
Total of moles analysed 176.72 0.00158 0.00027 0.00078 0.00028 0.00057
Total of moles calculated 0.00182 0.00038 0.00083 0.00023 0.00076
% recovery 87 % 72 % 93 % 123 % 75 %
CARRÉ ET AL., doi:10.1520/JAI102293 169
170 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 8—Combustion model approximation for concentration of released gases with low
LC50 in function of the mass of PTFE.

With With With With With


200 mg 150 mg 100 mg 50 mg 20 mg
of PTFE of PTFE of PTFE of PTFE of PTFE
Mole of COF2 0.000600 0.000450 0.000300 0.000150 0.000060
generated

Equivalent 4,799 3,599 2,400 1,200 480


concentration of
COF2 in 3 L of O2
共in ppm volume兲

Equivalent 28,796 21,597 14,398 7,199 2,880


concentration of
COF2 in 0.5 L of O2
共in ppm volume兲

LC50 of COF2 360 360 360 360 360


共in ppm volume兲

Mole of F2 0.000300 0.000225 0.000150 0.000075 0.000030


generated

Equivalent 2400 1,800 1,200 600 240


concentration of F2
in 3 L of O2
共in ppm volume兲

Equivalent 14,398 10,798 7,199 3,599 1,440


concentration of
F2 in 0.5 L of O2
共in ppm volume兲

LC50 of F2 185 185 185 185 185


共in ppm volume兲

clear that the selecting a modelling scenario is complicated, for simplification


in this discussion, we considered a combination the two scenarios presented
here. We assume both a low dilution gas as well as the high initial concentra-
tions of generated products found if immediate contact post-combustion where
to occur. In order to present some concentration values, assumptions were
done on the volume of oxygen in which the by-products can be released. Com-
bustion of all or part of a non-metallic component in medical devices might not
be immediately apparent and the products of combustion might remain en-
tirely confined within a small part of the equipment. In this case, subsequently
these products of combustion might be delivered to patient as a bolus of high
concentration. Knowing that standard flow rate of oxygen used in oxygen de-
livery systems are 0.5 liter/min for somebody in good shape and around 3 liter/
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 171

TABLE 9—Concentration of released toxic gases in function of the mass of PCTFE.

With With With With With


200 mg 150 mg 100 mg 50 mg 20 mg
of of of of of
PCTFE PCTFE PCTFE PCTFE PCTFE
Mole of COF2 0.000429 0.000322 0.000215 0.000107 0.000043
generated

Equivalent 3,434 2,576 1,717 859 343


concentration of
COF2
in 3 L of O2
共in ppm volume兲

Equivalent 20,606 15,455 10,303 5,152 2,061


concentration of
COF2
in 0.5 L of O2
共in ppm volume兲

LC50 of COF2 360 360 360 360 360


共in ppm volume兲

Mole of F2 0.000258 0.000193 0.000129 0.000064 0.000026


generated

Equivalent 2,061 1,545 1,030 515 206


concentration of
F2
in 3 L of O2
共in ppm volume兲

Equivalent 12,364 9,273 6,182 3,091 1,236


concentration of
F2
in 0.5 L of O2
共in ppm volume兲

LC50 of F2 185 185 185 185 185


共in ppm volume兲

Mole of Cl2 0.000859 0.000644 0.000429 0.000215 0.000086


generated

Equivalent 6,869 5,152 3,434 1,717 687


concentration of
Cl2
in 3 L of O2
共in ppm volume兲

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
172 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 9— 共Continued.兲

With With With With With


200 mg 150 mg 100 mg 50 mg 20 mg
of of of of of
PCTFE PCTFE PCTFE PCTFE PCTFE
Equivalent 41,213 30,910 20,606 10,303 4,121
concentration of
Cl2
in 0.5 L of O2
共in ppm volume兲

LC50 of Cl2 293 293 293 293 293


共in ppm volume兲

min for people with respiratory disease 关16兴, these two values were selected for
our calculation. Using these assumptions, the patient will be in contact of these
high concentrations during one minute.
In both cases, 3 and 0.5 L/min, the concentrations of F2, COF2, and Cl2 are
above the LC50 concentration of each gas even with 20 mg of plastic. To truly
evaluate the toxicity risk, the duration of exposure from the combustion prod-
ucts needs to be determined. The LC50 values given in Table 1 are for 1 hour
duration and would need to be converted to the same duration of exposure for
comparison using toxicological methods or obtaining LC50 data for the same
duration of exposure. Determining the duration of exposure from the combus-
tion products is a challenging task because it is configuration dependent and
variables such as pressure, volumes, flow rates and non-metal mass can all
influence the final concentration and, therefore the toxicity risk.

Conclusions
This paper presented a new analytical setup and procedure to identify and
quantify the released products obtained after promoted combustion of non
metallic materials with oxygen. The target of this study was to determine these
compounds in a shortest delay after combustion induced by adiabatic compres-
sion with oxygen at 200 bars 共20,000 kPa兲 and 10 or 26 L of oxygen. Real time
analysis can be done and the quantification of combustion by-products was
done with better accuracy than in previous study off-site. Indeed, very signifi-
cant improvement resulted from minimization of the time slot between the end
of the combustion event and introduction of sampled gases into the analytical
instrument 共FTIR or MS兲.
It was possible to determine the actual quantitative nomenclature of gas-
eous products released from the combustion event itself. The presence of prod-
ucts of low LC50 values could unambiguously evidenced. Those, whose pres-
ence can be definitely ruled out, such as phosgene, could be also specified. The
table presented in the ISO 15001, the content of which was mainly speculative,
can be amended accordingly.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
CARRÉ ET AL., doi:10.1520/JAI102293 173

The list of quantified combustion by-products could be obtained for a range


of materials to be considered for medical oxygen and breathing systems. More-
over, mass balance reaction models methodology was demonstrated for, in par-
ticular for PTFE and PCTFE. Once sufficient data are obtained to generate a
high confidence model, a simple calculation could be proposed to approximate
the number of moles of each of the toxic compound generated by combustion
of a known mass of a given polymer material. The quantification of these com-
pound concentrations produced in respect to expected amount of material
burnt can assist a toxicologist in a risk assessment that would compare these
estimated values to the LC50 and other establish limits of each compound for
determining their toxicity risk. Although uncertainties remain, this approxima-
tion done for 3 and 0.5 liters of oxygen per minute flow rates predicts that
PTFE and PCTFE generate COF2, F2, and Cl2 at higher concentrations, assum-
ing one minute exposure, than the 1-hour LC50 of these gases. However, a
toxicologist would need to evaluate concentrations over the same time period
for a toxicity risk analysis.

References

关1兴 ISO 10298, 1995, “Gas Cylinder—Determination of Toxicity of a Gas or Gas Mix-
ture,” ISO International, Geneva, Switzerland.
关2兴 EIGA Doc. 73/08, 2008, “High Pressure Breathing Gas Systems Toxicity Risks of
Using Non-metallic Materials,” Ed. Eiga, Av. des Arts, Brussels, Belgium.
关3兴 ISO 15001, 2003, “Anaesthetic and Respiratory Equipment—Compatiblity with
Oxygen,” ISO International, Geneva, Switzerland.
关4兴 Barthélémy, H., 2003, “Combustion Products Toxicity of Non-Metallic Materials
Used for Medical Oxygen Equipment,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres, ASTM STP 1454, T. A. Steinberg, H. D. Beeson, and
B. E. Newton, Eds., ASTM International, West Conshohocken, PA, pp. 9–20.
关5兴 CGA G-4, 2008, “Design Considerations to Mitigate the Potential Risks of Toxicity
When Using Non-metallic Materials in High Pressure Oxygen Breathing Gas Sys-
tems 共EIGA Doc 73/08兲,” ISO International, Geneva, Switzerland.
关6兴 ISO 14971, 2007, “Medical Devices. Application of Risk Management to Medical
Devices,” ISO International, Geneva, Switzerland.
关7兴 ASTM G88-90, 1997, “Standard Guide for Designing Systems for Oxygen Service,”
Annual Book of ASTM Standards, Vol. 14, ASTM International, West Consho-
hocken, PA.
关8兴 ISO 10297, 2006, “Transportable Gas Cylinders. Cylinder Valves—Specification
and Type Testing,” ISO International, Geneva, Switzerland.
关9兴 Gas Encyclopedia, L’Air Liquide, France, 1976.
关10兴 Ullmann’s, Encyclopedia of Industrial Chemistry, Vol. A21, Wiley, 1990.
关11兴 Biron, M. and Biron, C., Eds., Les Thermoplastiques, 1992.
关12兴 Mark, H. F., Bikales, N. M., Overberger, C. G., and Menges, H., Encyclopedia of
Polymer Science and Engineering, 2nd ed., Vol. 3, Wiley, 1985.
关13兴 Sax, N. I., Dangerous Properties of Industrial Material, 6th ed., Wiley, 1984.
关14兴 Tsuchiya, Y., “Significance of HCN Generation in Fire Gas Toxicity,” Journal of
Combustion Toxicology, Vol. 4, 1977, pp. 271–282.
关15兴 Barthélémy, H., Delode, G., and Vagnard, G., “Ignition of Materials in Oxygen
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
174 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Atmospheres: Comparison of Different Testing Methods for Ranking Materials.”


Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 1111, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West Consho-
hocken, PA, pp. 506–515.
关16兴 Flenley, D. C., “Clinical Hypoxia: Causes, Consequences and Correction,” The Lan-
cet, March 11, 1978, p. 542.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102306
Available online at www.astm.org/JAI

Susana Harper,1 David Hirsch,2 and Sarah Smith3

Determination of Time Required for


Materials Exposed to Oxygen to Return
to Reduced Flammability

ABSTRACT: Increased material flammability due to exposure to high oxygen


concentrations is a concern from both safety and operational perspectives.
Localized high oxygen concentrations can occur when exiting a higher oxy-
gen concentration environment due to material saturation as well as oxygen
entrapment between barrier materials. Understanding of oxygen diffusion
and its correlation to flammability risks can reduce the likelihood of fires while
improving procedures as NASA moves to longer missions with increased
extravehicular activities in both spacecraft and off-Earth habitats. This paper
examines the time required for common spacecraft materials exposed to
oxygen to return to reduced flammability after removal from the increased
oxygen concentration environment. Specifically, NASA-STD-6001A maxi-
mum oxygen concentration testing and ASTM Standard F1927 diffusion test-
ing were performed on Nomex®4 HT90-40, Tiburon®5 surgical drape, cotton,
Extravehicular Mobility Unit 共EMU兲 Liquid-Cooled Ventilation Garment, EMU
Thermal Comfort Undergarment, EMU Mosite foam with spandex covering,
Advanced Crew Escape Suit 共ACES兲 outer cross-section, ACES Liquid-
Cooled Garment, ACES O2 hose material, Minicel®6 polyethylene foam,

Manuscript received December 29, 2008; accepted for publication August 19, 2009; pub-
lished online October 2009.
1
Engineer, NASA Johnson Space Center White Sands Test Facility, P.O. Box 20, Las
Cruces, NM 88004.
2
Engineer, Jacobs Technology, Inc., NASA Johnson Space Center White Sands Test Fa-
cility, P.O. Box 20, Las Cruces, NM 88004.
3
Mechanical Engineer, NASA Johnson Space Center White Sands Test Facility, P.O. Box
20, Las Cruces, NM 88004.
Cite as: Harper, S., Hirsch, D. and Smith, S., ⬙Determination of Time Required for
Materials Exposed to Oxygen to Return to Reduced Flammability,⬙ J. ASTM Intl., Vol. 6,
No. 10. doi:10.1520/JAI102306.
4
Nomex® is a registered trademark of E. I. Du Pont de Nemours and Co., Wilmington,
DE.
5
Tiburon® is a registered trademark of Allegiance Corporation, McGraw Park, IL.
6
Minicel® is a registered trademark of Sekisui Voltek, Lawrence, MA.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 175
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
176 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Minicel® polyethylene foam with Nomex® covering, Pyrell polyurethane


foam, and Zotek®7 F-30 foam.
KEYWORDS: flammability, nonmetals, permeation, diffusion,
diffusion, maximum oxygen concentration 共MOC兲, gaseous oxygen
enrichment, textiles, Extravehicular Mobility Unit, crew escape suit

Introduction

Catastrophic fires have occurred as a result of gaseous oxygen enrichment even


in low pressure environments. One contributing factor is that textile materials
become more flammable and easier to ignite after exposure to gaseous oxygen
enrichment or saturation. In addition, these materials can serve as barriers by
trapping localized oxygen-enriched environments.
When materials are moved from a higher oxygen concentration environ-
ment to an environment with lower oxygen concentration, the corresponding
flammability and ignitability risks are difficult to characterize. An industrial
example of such a scenario would occur when a person performs liquid oxygen
filling operations and is exposed to a high amount of oxygen vapor. An aero-
space example of such a scenario would occur when an astronaut completes an
extravehicular activity 共EVA兲 performed in 100 % oxygen and then moves into a
spacecraft with a lower oxygen concentration 共such as 34 % oxygen兲. In each of
these scenarios, the duration of time for the person’s garments to return to the
flammability and ignitability expected in the lower oxygen concentration is
uncertain. The generally accepted rule of thumb has historically been to allow
30 min for materials to return to their flammability and ignitability in the lower
oxygen concentration 共and essentially to avoid any potential ignition sources
during that 30-min time frame兲 关1兴. However, this rule of thumb was not based
on data and was established as a presumably conservative estimate. This paper
is a first step in approximating how well this rule of thumb lines up with reality.
Understanding the true times needed for appropriate diffusion will help to en-
sure that oxygen enrichment flammability concerns are not neglected while
limiting significant time lost particularly in the case of an astronaut moving
back and forth between different environments.
NASA Johnson Space Center 共JSC兲 White Sands Test Facility developed a
test and analysis methodology to relate oxygen diffusion rates with flammabil-
ity of materials that have been exposed to oxygen-enriched environments. This
report examines two oxygen enrichment scenarios, the scenario of saturation of
materials exposure to oxygen-enriched environments and the scenario of en-
trapment when materials function as a potential barrier to create localized high
oxygen concentrations. The entrapment scenario is particularly focused on
simulating oxygen trapped between a material and a person’s body. These two
scenarios are depicted in Fig. 1.

7
Zotek® is a registered trademark of Zotefoams plc, Walton,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
KY.
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HARPER ET AL., doi:10.1520/JAI102306 177

FIG. 1—Oxygen saturation and entrapment scenarios for oxygen concentration decay
diffusion analysis.

Test Methodology
The goal in developing the test methodology was to characterize flammability
as a function of time, thereby relating the flammability to the diffusion of oxy-
gen out of concentrated areas and materials in entrapment and saturation sce-
narios. It was determined that a two-phase approach could be used and data
correlated. The two phases of the methodology were diffusion calculations and
testing as well as flammability testing. In the diffusion phase mass transport
calculations were performed to theoretically correlate how the concentration of
oxygen would vary with time for a specified configuration. Material diffusion
testing was performed in this phase to obtain critical diffusion values required
for mass transport calculations. In the flammability phase, material flammabil-
ity testing was performed to determine what concentration would be consid-
ered a reduced flammability environment for each material. The approximated
time for a material exposed to 100 % oxygen to reach this reduced flammability
concentration is then the recommended diffusion time for personnel or mate-
rials to wait before resuming normal operations. A general flow chart of meth-
odology is given in Fig. 2 for general understanding; a more detailed explana-
tion of each phase and its test methods will be given in the following sections.

Diffusion Phase of Methodology


When a material is moved from a high oxygen concentration to a low oxygen
concentration, the oxygen concentration in the material decreases over time
through diffusion. Mass transfer is mass in transit as the result of a species
concentration difference in a mixture as a driving potential for transport. This
transport process is called diffusion. The oxygen transmission rate is the flux of
a species through a specific medium due to species concentration differences
on either side normalized to the pressure gradient. For simplification, only
oxygen transport was considered in our models. Nitrogen transport was as-
sumed to be reasonably negligible due to a significantly smaller concentration
gradient that would drive transport and transport rates that tend to be two to
three times slower than those seen for oxygen.
The relationship between the oxygen transmission rate 共absolute flux兲 and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
178 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Approximation of time required for materials exposed to oxygen to return to


reduced flammability methodology.

the concentration gradient is described by Fick’s First Law of Diffusion when


describing a steady-state stationary medium as shown in Eq 1. Calculations
were performed using this stationary medium model as it is worst case situa-
tion from the perspective of maintaining localized pockets of high oxygen con-
centrations. The molar diffusion coefficient shown in this equation must be
determined experimentally and pertains to a specific permeant and media com-
bination

⳵C
J=−D 共1兲
⳵x
Equation 1 is Fick’s First Law for a steady-state stationary medium, where J is
the absolute flux per unit area of permeant through a medium, D is the diffu-
sion coefficient, and ␦C / ␦x is the concentration gradient of the permeant across
a thickness ␦x.
To obtain diffusion coefficients, testing was performed. Testing generated
oxygen transmission rate data 共absolute flux兲. Rearrangement of Eq 1, with
known concentrations and material thicknesses, allowed the derivation of ma-
terial and gas specific diffusion coefficients. Testing was performed in accor-
dance with ASTM Standard F1927 关2兴, Standard Test Method for the Determi-
nation of Oxygen Gas Transmission Rate, Diffusion and Permeance at
Controlled Relative Humidity Through Barrier Materials Using a Coulometric
Detector. This test method determines the rate of transmission of oxygen gas at
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HARPER ET AL., doi:10.1520/JAI102306 179

steady-state conditions at a given temperature and percent relative humidity 共%


RH兲 through films, sheeting, laminates, coextrusions, or plastic-coated papers
and fabrics.
Oxygen transmission rate or absolute flux can further be correlated with
total amount of permeant using Eq 2 关3兴

Q
J= 共2兲
At
Equation 2 is correlation of absolute flux 共J兲 to amount of permeant 共Q兲 that
has passed through area 共A兲 during time 共t兲.
The scenarios in question are not steady state due to their dependence on
time for atoms to accumulate in a region and correspondingly deplete from
another a region. Nonetheless, the two absolute flux equations, Eqs 1 and 2, can
be equaled and manipulated to determine the amount of oxygen diffused as a
function of time and therefore the oxygen concentration decay. The equations
shown in Fig. 3 were evaluated at discrete time steps, which can be considered
steady state within each time step. By using this method, the time needed for
materials to return to reduced oxygen concentrations was determined 关4兴.
For entrapment scenarios and for porous saturation scenarios, it was as-
sumed that the full volume of the defined space was filled with 100 % oxygen. In
the entrapment scenario, the model simulates oxygen trapped between an im-
permeable surface 共human body兲 and that of a barrier material, the gap be-
tween these two being 1 cm deep and filled with oxygen. Diffusion was assumed
to take place only through the main face area and not through the sides. For
saturation scenarios, bulk materials were modeled as a cube with a side length
of 0.5 m. Diffusion was assumed to take place through all six faces of the cube.
For porous materials it was amused that the full volume of the modeled cube
was filled with 100 % oxygen.
For non-porous materials the amount of oxygen that is actually in the cube
volume is limited by the extent to which a material can absorb a gas. If treated
as a uniform substance, the concentration of the gas in the solid may be ob-
tained through the use of a property known as the solubility. It is defined by the
expression below 关3兴. For the value of solubility it is necessary to rely exclu-
sively on experimental measurements. The mechanism of diffusion of gases,
liquids, and solids in solids is extremely complicated, and generalized theories
are not available. Furthermore, only limited experimental results are available
in the literature. Due to limited solubility information, materials were assumed
to have solubility equal to that of the published solubility of oxygen in rubber at
298 K 关3兴. If actual material solubility is smaller, then diffusion times would
also be due to a decreased amount of initial oxygen absorbed into the material

C A共 0 兲 = S · p A 共3兲
Equation 3 is the relationship between molar concentration 共CA共0兲兲 of A of a
gas in a solid at the interface, solubility 共S兲 in units of mol/ m3 · Pa, and the
partial pressure of that gas adjoining the interface 共pA兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
180 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—Oxygen decay diffusion time-dependent calculations, where D is the diffusion


coefficient, C is concentration, Q is amount of permeant that has passed through area
A, J is absolute flux, mol is quantity of moles of permeant, t is time, Volume is volume
of entrapped or saturated area, and O2 % is oxygen concentration remaining.

The volumes and solubility relationships discussed above were used to de-
fine initial concentration and initial mol quantities for diffusion calculations.
All approximations were made assuming a well-mixed external atmospheric
environment of 20.9 %.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HARPER ET AL., doi:10.1520/JAI102306 181

Flammability Phase of Methodology


Flammability testing was conducted to determine the maximum oxygen con-
centration 共MOC兲 at which a material will be acceptable for use without restric-
tions per NASA-STD-6001A requirements for NASA vehicles. NASA requires
that all non-metallic materials pass NASA-STD-6001A Test 1 flammability test-
ing. Materials are required to self-extinguish before a burn length of 6 in. is
reached with no dripped burning material that propagates fire to a K-10 paper
set below the sample in a static environment 关5兴. The test setup was a modified
NASA-STD-6001A Test 1 configuration, as shown in Fig. 4. Non-edge ignition
was performed to better simulate a realistic ignition scenario. To achieve this
scenario, materials were angled at 15° from vertical, with the igniter positioned
2 in. above the bottom of the sample and 1/4 in. below the surface at 101 kPa.
To determine the MOC threshold the NASA-STD-6001A Test 1 Flammabil-
ity Test is performed at incrementally higher concentrations until a threshold is
established between a concentration in which the material will fail, burning
longer than 6 in. and one in which five samples will pass, self-extinguishing
prior to burning 6 in. 关5兴 The MOC threshold established for each material can
be considered an upper boundary, below which materials exhibit reduced flam-
mability with self-extinguishment characteristics. In this paper and for NASA
programs, the MOC threshold will be the concentration at which we will con-
sider the material to have established reduced flammability and at which stan-
dard operations may proceed.

Test Materials
The NASA groups that encounter potentially oxygen-enriched entrapment and
saturation scenarios collaborated to develop a list of realistic and applicable
materials for testing. Groups involved in material selection were the JSC Ma-
terial and Process Branch, the EVA Office, and the Crew Escape Suit and Sys-
tems Group. Some materials not used in NASA scenarios were chosen for their
comparison with past data 关6兴. The materials chosen were grouped into sepa-
rate categories by function and tested in their use thickness. Table 1 describes
the 13 materials and layups that were tested.

Test Results and Discussion


Results from the two phases of testing are discussed in the following sections,
as well as the analysis and correlation of test data.

Diffusion Phase Testing and Derivations


ASTM Standard F1927 was conducted at the specified temperature and % RH
expected during material usage to determine absolute flux and diffusion coef-
ficients for non-porous materials. Results of these tests are shown in Table 2
and Fig. 5. Materials that cannot contain a gas, such as common clothing ma-
terials, are considered porous and exhibit free diffusion. The Advanced Crew
Escape Suit 共ACES兲 layup was questionable and was tested using ASTM Stan-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
182 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—NASA-STD-6001A Test 1 Flammability Test configuration on ACES suit layup


modified for non-edge ignition with 15° from vertical position.

dard F1927. Testing showed that the layup was indeed porous. Materials that
were determined to be porous exhibit free diffusion for which a published value
for oxygen through air was used 关3兴. Porous materials and free diffusion value
are presented in Table 3.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HARPER ET AL., doi:10.1520/JAI102306 183

TABLE 1—Test materials.

Material Description
Common/Comparison Fabrics
Aramid fabric with high-performance heat-
Nomex® HT90-40 and flame-resistant properties; L/N 7254
Microfiber composite consisting of three
layers: An absorbent fluid-control layer made
of microfiber fabric, an impermeable
cast-extruded polyethylene membrane
laminated to the non-woven components, and
Tiburon surgical drape a patient comfort layer
Cellulose fabric in 100 % cotton Hanes Beefy
Cotton fabric T-shirt
Extravehicular Mobility Unit Materials
EMU Liquid-Cooled Ventilation
Garment 共LCVG兲 Polyamide material nylon tricot ST11N791-01
Thermal Comfort Undergarment Polyester-based material TCU bottom, P/N
共TCU兲 SKD38114488-01; 100 % polyester
Polyurethane/polyethylene glycol elastomeric
spandex ST11N117-07 Covered Viton 共Mosite兲
fluoroelastomer closed-cell foam
ST66V2590-01 is used in a variety of pads
available for use in the EMU; these pads were
designed to reduce hot-spots created by suit
contact with the shoulders, elbows, ribs, or
knees; the pads are inserted into spandex
pockets that are form-fitted to each pad, and
Spandex-covered Viton 共Mosite兲 foam these are whip-stitched to the LCVGa
Advanced Crew Escape Suit
共ACES兲/Equipment
Layup consisting of two outer layers of ACES,
the outer material being composed of the
aramid fiber-based Nomex® and the inner
material composed of a
polytetrafluoroethylene based Gore-Tex® for
ACES suit layup use as a bladder
An assembly composed of thick
polypropylene-based undergarment with
plastic tubing stitched in; this garment is
worn under the outer ACES garment for
ACES Liquid-Cooled Garment 共LCG兲 temperature controlb
Cabin Environment Materials

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
184 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1— 共Continued.兲

Material Description
®
L-200 Minicel polyethylene foam; this foam
is extremely fine-celled, chemically
cross-linked closed-cell foam commonly used
in the various space vehicles; this foam is
commonly covered an aramid 共Nomex兲 to
Minicel® polyethylene foam mitigate fire risk
HT90-40 Nomex® Covered L-200 Minicel®
polyethylene foam; this foam is closed-cell
Minicel® polyethylene foam with foam commonly used in the various space
Nomex® covering vehicles
Pyrell polyurethane foam; this foam is
Pyrell polyurethane foam open-celled foam
Polyvinylidene fluoride, highly non-reactive
and pure thermoplastic fluoropolymer
closed-cell foam; a foam being considered for
Zotek® F-30 Foam extensive use in future NASA vehicles
a
Reference 7.
b
Reference 8.

Flammability Phase Testing


NASA-STD-6001A MOC flammability testing was conducted at specified envi-
ronmental pressure, while the oxygen concentration was varied to determine
the flammability threshold. Though the actual materials are used in a variety of
pressure and concentration combinations, a single worst case pressure of 101
kPa 共14.7 psia兲 was chosen for consistency and ease of data comparison 关9兴.
MOC threshold results are shown in Tables 4 and 5 for non-porous and porous
materials, respectively.

Correlation of Diffusion and Flammability Data for Oxygen Decay Time Ap-
proximations
Once oxygen transmission testing was performed and diffusion coefficients are
calculated, diffusion calculations were used to determine oxygen concentration
dependence with time for each material in its specific configuration for either
entrapment or saturation modeling. Thicknesses for oxygen transmission rate
testing, diffusion calculations, and flammability testing were chosen to repre-
sent in-use thicknesses of materials. Nonetheless it should be noted that the
thicknesses of each material have an effect on the time needed for diffusion to
occur and that time will vary if thickness is changed.
Figures 6 and 7 are example plots showing the decay of the oxygen concen-
tration in the trapped area as diffusion progresses with time.
Once oxygen decay rates were established for the materials in their specific
configuration, it was possible to determine the time needed to reach the flam-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—Non-porous materials and their calculated diffusion coefficients.

Analysis Conditions

Downloaded/printed by
Temp Test Gas 共O2兲 Carrier Gas 共N2兲 Diffusion Coefficient
Material 共°F兲 Humidity 共RH %兲 Humidity 共RH %兲 共 mol· m / m2 · S · Pa 兲
Tiburon surgical drape 75 35 35 1.4018E-14
Spandex-covered Viton 共Mosite兲 foam 72 47 35 4.1486E-15
Minicel polyethylene foam 75 35 35 1.3362E-13
Minicel polyethylene foam with Nomex
covering 75 35 35 1.3543E-13
Zotek F-30 foam 75 35 35 3.4117E-13

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


HARPER ET AL., doi:10.1520/JAI102306 185
186 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—Diffusion coefficients for non-porous materials.

TABLE 3—Porous materials and their diffusion coefficients.

Natural Oxygen Gas Diffusion Coefficient Air


Material @ 278 K and 1 atm 共m2 / s兲
Nomex HT90-40 2.1E-05
Cotton fabric 2.1E-05
EMU Liquid-Cooled Ventilation Garment
共LCVG兲 2.1E-05
Thermal Comfort Undergarment 共TCU兲 2.1E-05
ACES suit layup 2.1E-05
ACES Liquid-Cooled Garment 共LCG兲 2.1E-05
Pyrell polyurethane foam 2.1E-05

TABLE 4—Non-porous materials MOCs derived from NASA-STD-6001A Test 1 Flammabil-


ity Test modified for non-edge ignition with 15° from vertical position at 101 kPa.

Maximum Oxygen
Material Test Pressure 共Pa兲 Concentration 共%兲
Tiburon surgical drape 101 325 20
Spandex-covered Viton 共Mosite兲 foam 101 325 18
Minicel polyethylene foam 101 325 20
Minicel polyethylene foam with Nomex
covering 101 325 28
Zotek F-30 foam 101 325 36
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 5—Porous materials MOCs derived from NASA-STD-6001A Test 1 Flammability Test modified for non-edge ignition with 15° from
vertical position at 101 kPa.

Maximum Oxygen

Downloaded/printed by
Material Test Pressure 共Pa兲 Concentration 共%兲
Nomex HT90-40 101 325 24
Cotton fabric 101 325 13
EMU Liquid-Cooled Ventilation Garment
共LCVG兲 101 325 23
Thermal Comfort Undergarment 共TCU兲 101 325 21
ACES suit layup 101 325 34
ACES Liquid-Cooled Garment 共LCG兲 101 325 18.1
Pyrell polyurethane foam 101 325 19

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


HARPER ET AL., doi:10.1520/JAI102306 187
188 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Non-porous Tiburon® surgical drape 1 cm gap entrapment, oxygen concentra-


tion decay versus time.

mability MOC threshold. Below this threshold we can assume decreased flam-
mability and return to normal operations for a given material. This correlation
was depicted earlier in this paper and can be found in Fig. 2. The entrapment
scenario represents oxygen found between clothing and a person’s body. Tables

FIG. 7—Porous Pyrell polyurethane foam 共thickness: 0.0508 m兲 1 cm gap entrapment,


oxygen decay versus time.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HARPER ET AL., doi:10.1520/JAI102306 189

6 and 7 show the amount of time needed for the entrapped area to reach the
MOC threshold or to reach 20.9 % oxygen if the MOC threshold was deter-
mined to be below 20.9 % oxygen.
Because free oxygen diffusion takes place through all porous materials, the
only variation in diffusion time for these materials is their dependence on
thickness as shown in equations from Fig. 3. Table 7 presents the thickest and
thinnest porous materials tested to give a range of time for decreased flamma-
bility.
The saturation scenario models the case of a bulk material saturated in
oxygen. Saturation is a concern for larger and bulkier materials as they can
hold the largest quantity of oxygen when saturated. Tables 8 and 9 below give
the approximated time for a saturated material to permeate out enriched oxy-
gen and equilibrate to reach the MOC threshold or 20.9 % oxygen if the MOC
was determined to be below 20.9 % oxygen. The data presented here is for the
time to reach the corresponding environmental concentration, derived from the
partial pressure, pertaining to the internal oxygen concentration absorbed
within the material using Eq 3. The concentration that is found within the
material is also listed. There is little understanding as to how absorbed oxygen
contributes to the increased flammability of the material. There is a possibility
that this internal concentration greatly contributes to the speed of combustion,
or it is possible that this quantity is too small compared to the quantity of the
solid material to support combustion. This is one of the primary questions that
will be investigated in the follow-on time-dependent flammability testing.

Conclusions and Future Work

The two-phase methodology consisting of NASA-STD-6001A MOC flammability


testing and ASTM Standard F1927 diffusion testing and analysis was successful
in approximating time required for materials exposed to oxygen to return to
reduced flammability. This methodology is recommended for future use in ap-
proximating flammability risks for localized enriched-oxygen environments.
From these theoretical approximations, it is clear that oxygen entrapment and
saturation can be a real concern and need to be further investigated, especially
when dealing with non-porous materials.
For porous materials, the 30-min rule of thumb may be overly conservative.
From approximations, it can be expected that oxygen concentration will reach
their reduced flammability conditions in a maximum of 2 min for a 0.5 m
length bulk material. In the case of clothing, even when entrapment of enriched
oxygen occurs, diffusion to reduced flammability conditions should occur in
less than 1 s. Though it is evident that diffusion is relatively quick out of porous
materials, precaution should still be taken as there are still other associated
risks and configuration concerns to take into consideration. Configurationally a
concern would be the exposure of large porous materials, such as mattress, to
an enriched environment. Here the volume for saturation can be very large, yet
surface area for diffusion is limited. Approximations in this paper assumed
diffusion from all six faces in a bulk sample, where in the case of a mattress at
least one of the large faces for diffusion would likely be blocked. This would
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 6—Non-porous materials time for oxygen concentration decay for 1 cm gap entrapment scenario.

Material Thickness 共m兲 MOC 共%兲 Time 共t兲 for MOC 共h:min:s 共O2 %兲兲

Downloaded/printed by
Tiburon surgical drape 0.000 22 20 05:10:01 共20.9兲
Spandex-covered Viton 共Mosite兲 foam 0.007 94 18 447:18:21 共20.9兲
Minicel polyethylene foam 0.051 00 20 36:55:01 共20.9兲
Minicel polyethylene foam with Nomex
covering 0.051 69 28 13:34:59
Zotek F-30 foam 0.0254 36 04:49:59

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


190 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 7—Porous materials time for oxygen concentration decay for 1 cm gap entrapment scenario.

Material Thickness 共m兲 MOC 共%兲 Time 共t兲 for MOC h:min:s 共O2 %兲

Downloaded/printed by
EMU Liquid-Cooled Ventilation Garment
共LCVG兲 0.000 22 23 00:00:026
Pyrell polyurethane foam 0.0508 19 共20.9兲 00:00:10

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


HARPER ET AL., doi:10.1520/JAI102306 191
TABLE 8—Non-porous materials time for oxygen concentration decay for 0.5 m length cube saturation scenario and O2 in solid rubber
solubility.

Thickness MOC Time 共t兲 for Internal

Downloaded/printed by
Material 共m兲 共%兲 Corresponding MOC 共h:min:s 共O2 %兲兲 Material O2%
Spandex-covered Viton 共Mosite兲 foam 0.50 18 10,000:00:00 共20.9兲 2.47
Minicel polyethylene foam 0.50 20 3050:00:00 共20.9兲 1.48
Minicel polyethylene foam with Nomex
covering 0.50 28 350:00:00 1.98
Zotek F-30 foam 0.50 36 121:00:00 2.54

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


192 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
HARPER ET AL., doi:10.1520/JAI102306 193

TABLE 9—Porous materials time for oxygen concentration decay for 0.5 m length cube
saturation scenario.

Time
Thickness MOC 共t兲 for MOC
Material 共m兲 共%兲 共h:min:s 共O2 %兲兲
Pyrell polyurethane foam 0.50 19 00:02:10 共20.9兲

increase the time for achieving reduced flammability via diffusion. It should
also be noted that there is still a risk of oxygen accumulation due to density that
may cause oxygen to act as a mass with minimized faces for free diffusion.
For non-porous materials, theoretical approximations suggest that they are
excellent at trapping and retaining oxygen. Though more investigation is
needed to fully understand the saturation phenomenon for non-porous materi-
als, entrapment approximations give us some meaningful data. Closed-cell
foams excel as superior barriers in preventing diffusion. This is especially clear
with Mosite foam with a required 447 h to allow 1 cm depth of oxygen behind
it to diffuse. Localized enriched-oxygen concentrations should be considered in
operational planning, especially as there is a move toward a greater use of
closed-cell foams.
This is particularly alarming when we consider that Mosite foam is used in
the astronaut’s Liquid-Cooled Ventilation Garment 共LCVG兲 that is worn below
the EVA suit and can be worn after a mission is complete and suit is removed.
Here the astronaut would retain enriched-oxygen concentrations, trapped in
contact with his body for extended periods of time. Another scenario for en-
trapment is of a NASA vehicle changing from a more enriched environment to
a decreased one, as is the case in the air lock. Non-porous materials in cabin
walls or in other uses could trap oxygen behind them and maintain enriched
concentrations in those pockets from 5 to 447 h based on configurations ana-
lyzed here. In the case of non-porous material entrapment, the 30-min rule of
thumb may not be sufficient.
With non-porous materials, the risks must be weighed to not overly restrict
operations nor to ignore the risk for entrapment and increased flammability.
Testing and approximations here were done in a quiescent environment as this
is the worst case scenario for minimizing diffusion. It is recommended that
operational awareness to introduce bulk flow into the environment be consid-
ered as it would greatly increase the speed at which diffusion takes place.
Real life flammability validation testing will be performed to examine how
well modeled scenarios agree with real life phenomenon, but these were not
completed in time for this publication. Validation testing will consist of burn-
ing materials at time intervals after being removed from 100 % oxygen environ-
ments. Their burn lengths and burn rates will be then compared flammability
burn rates for materials at specified concentrations.

References

关1兴 Beeson, H. D., Smith, S. R., and Stewart, W. F., Safe Use of Oxygen and Oxygen
Systems: Handbook for Design, Operation, and Maintenance, 2nd ed., ASTM Inter-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
194 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

national, West Conshohocken, PA, 2007.


关2兴 ASTM Standard F1927, 1999, “Standard Test Method for the Determination of
Oxygen Gas Transmission Rate, Diffusion and Permeance at Controlled Relative
Humidity Through Barrier Materials Using a Coulometric Detector,” Annual Book
of Standards, ASTM International, West Conshohocken, PA.
关3兴 Incropera, F. and DeWitt, D., Fundamentals of Heat and Mass Transfer, 4th ed.,
John Wiley and Sons, New York, 1996.
关4兴 Harper, S., “Detailed Calculations for Correlating Diffusion and Diffusion Coeffi-
cient to Time Required for Materials Exposed to Oxygen to Return to Reduced
Ignitability,” White Paper WP-08-0010, NASA Johnson Space Center White Sands
Test Facility, Las Cruces, NM, 2009.
关5兴 NASA-STD-6001A, April 21, 2008, “Flammability, Offgassing, and Compatibility
Requirements and Test Procedures for Materials in Environments that Support
Combustion,” Interim NASA Technical Standard NASA-STD-共I兲-6001A, Test 1,
“Upward Flame Propagation,” NASA Headquarters, Washington, D.C.
关6兴 Sepulveda, R. and Smith, S., “Flammability of Textiles in Air and in Gaseous Oxy-
gen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
11th Volume, ASTM STP 1479, D. B. Hirsch, R. Zawierucha, T. A. Steinberg, and H.
Barthelemy, Eds., ASTM International, West Conshohocken, PA, 2006.
关7兴 Williams, D. R. and Johnson, B. J., “EMU Shoulder Injury Tiger Team Report,”
Report No. NASA/TM-2003–212058, NASA Johnson Space Center, Houston, TX,
September 2003.
关8兴 Stuart, M. C., Lee, M. S., McDaniel, A., Jacobs, T., and Schneider, S. M., “Perfor-
mance of the Liquid-Cooling Garment with the Advanced Crew Escape Suit in
Elevated Cabin Temperatures,” Report No. NASA/TP—2004–212074, NASA
Johnson Space Center, Houston, TX, 2004.
关9兴 Campbell, P. D. and Henninger, D. L., “Recommendations for Exploration Space-
craft Internal Atmospheres, The Final Report of the NASA Exploration Atmo-
spheres Working Group,” Report No. JSC 63309 EAWG, NASA Johnson Space Cen-
ter, Houston, TX, 2006.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 7
doi:10.1520/JAI102285
Available online at www.astm.org/JAI

Sarah Smith1

Effect of Oxygen Concentration on


Autogenous Ignition Temperature and
Pneumatic Impact Ignitability of
Nonmetallic Materials

ABSTRACT: Extensive test data exist on the ignitability of nonmetallic mate-


rials in pure oxygen, but these characteristics are not as well understood for
lesser oxygen concentrations. In this study, autogenous ignition temperature
testing and pneumatic impact testing were used to better understand the
effects of oxygen concentration on ignition of nonmetallic materials. Tests
were performed using oxygen concentrations of 21, 34, 45, and 100 %. The
following materials were tested: PTFE Teflon®, Buna-N, Silicone, Zytel® 42,
Viton® A, and Vespel® SP-21.
KEYWORDS: autogenous ignition temperature, pneumatic impact,
oxygen concentration, nonmetallic materials

Introduction

Increased fire hazards are inherent in oxygen-enriched environments. Exten-


sive data have been generated to study the flammability and ignitability of
materials in pure oxygen. However, there are limited data for environments
using less than 100 % oxygen but greater than 21 % oxygen 共as is found in air兲.
The purpose of this study was to determine the effects of oxygen concentration
on the ignitability of nonmetallic materials. Specifically, autogenous ignition
temperature 共AIT兲 and pneumatic impact tests were performed in several dif-
ferent oxygen concentrations.

Manuscript received December 11, 2008; accepted for publication May 19, 2009; pub-
lished online June 2009.
1
NASA Project Manager, White Sands Test Facility, National Aeronautics and Space
Administration 共NASA兲, P.O. Box 20, Las Cruces, NM 88004.
Cite as: Smith, S., ⬙Effect of Oxygen Concentration on Autogenous Ignition Temperature
and Pneumatic Impact Ignitability of Nonmetallic Materials,⬙ J. ASTM Intl., Vol. 6, No.
7. doi:10.1520/JAI102285.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 195
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
196 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Test Materials
Tests were performed on several elastomers and plastic materials, as described
in Table 1. According to ASTM G63 关1兴, materials preferred for oxygen service
are those that have AITs of 400° C or greater, heats of combustion of 10,500 J/g
or less, and oxygen indices greater than 55. The materials selected for this study
exhibit varying degrees of oxygen compatibility, as shown in Table 1.

Test Environments
Tests were performed in four different oxygen concentrations 共21, 34, 45, and
100 %兲, with the balance being nitrogen. The intermediate oxygen concentra-
tions of 34 and 45 % were chosen because they are environments currently
being used within the National Aeronautics and Space Administration 共NASA兲.
The upper limit for the cabin atmosphere for NASA’s Constellation Program is
34 % oxygen, and the environment used by the scuba divers at NASA’s Neutral
Buoyancy Laboratory is 45 % oxygen.

Test Methods
The AIT test measures the minimum temperature where a material will spon-
taneously ignite when heated in an oxygen or oxygen-enriched atmosphere. The
test apparatus is depicted in Fig. 1, and the procedure used was ASTM G72 关2兴.
The general test procedure was as follows. A sample was placed in the sample
holder and the sample holder was placed inside the reaction vessel. The vessel
was then pressurized with the test media to 10.3 MPa 共1,500 psi兲. The heating
jacket was then used to heat the vessel at a rate of 5 ± 1 ° C / min from 60 to
260° C, and ⬎3 ° C for temperatures greater than 260° C. During the test, the
reaction vessel was not vented, and therefore the pressure in the vessel in-
creased along with the temperature. The ignition of the sample was indicated
by a rapid temperature rise of at least 20° C as measured by the thermocouple.
The pneumatic impact test measures the relative ignitability of materials by
heat of compression. The test apparatus and system are depicted in Figs. 2 and
3, respectively. The procedure used was ASTM G74 关3兴. The general test proce-
dure was as follows. The test sample was placed in the test chamber and the
test chamber was connected to the pressurization system. The sample was then
subjected to sequential pneumatic impacts by opening and closing the system
vent and impact valves. Each sample was impacted five times. Ignition of the
sample was indicated by an increase in temperature and/or pressure. The test
pressure was varied to determine the threshold for ignition by pneumatic im-
pact for each material.

Test Results and Discussion


The AIT data are shown in Table 2. These data indicate that oxygen concentra-
tion has an effect on the AIT and the overall trend is that AITs decrease with
increasing oxygen concentration. For some materials the oxygen concentration
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Test materials and oxygen compatibility data.

Limiting Heat
Tradename/ Oxygen Autoignition of

Downloaded/printed by
Chemical Index Temperature Combustion
Designation Chemical Type Batch Number 共%兲a 共°C兲 a 共J/g兲b
Buna-N/NBR poly共acrylonitrile-co-butadiene兲 elastomer RB 711 22 142 41,460
Teflon®/PTFE polytetrafluoroethylene ACCA241064 95 ⬎425 7,118
Silicone Rubber, polydimethyl siloxane elastomer R5789 21 262 17,370
Class 2,
Grade 70/MQ
Vespel®c aromatic polyimide B/N 050722519 53 321 31,810
SP-21/15 %
graphite-filled PI
Viton®cA/FKM poly共hexafluoropropylene-co-vinylidene 44893 32 155 16,714

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
fluoride兲
elastomer
Zytel® 42/PA aliphatic polyamide PC44G64 L/N 30 183 36,960
603930
a
Data shown are lowest values listed in ASTM Manual 36, 2nd Edition, Table 3-12.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


b
Data shown are highest values listed in ASTM Manual 36, 2nd Edition, Table 3-12.
c
Vespel® and Viton® are registered trademarks of E. I. Du Pont de Nemours and Co., Wilmington, Delaware.
SMITH, doi:10.1520/JAI102285 197
198 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—AIT test apparatus.

FIG. 2—Pneumatic impact test apparatus.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SMITH, doi:10.1520/JAI102285 199

FIG. 3—Pneumatic impact test system.

has a significant effect on the AIT, while for other materials the effect of oxygen
concentration is smaller. For instance, Teflon’s®2 AIT was not greatly affected by
oxygen concentration, with a change of less than 10° C from 21 to 100 % oxy-
gen. In comparison, Zytel®2 42 exhibited a change of nearly 70° C in its AIT
from 21 to 100 % oxygen. These results are consistent with other studies that
have been performed 关4,5兴.
The pneumatic impact data are shown in Table 3. The overall trend is that
materials ignite at lower pressures as oxygen concentration increases. The ef-
fect of oxygen concentration on the passing impact pressure is substantial for
all of the materials tested. Teflon exhibited the greatest differences in pneu-
matic impact ignition thresholds. In 21 % oxygen, Teflon did not ignite even at
9,500 psi; but as the oxygen concentration increased there was a dramatic
change in the ignition threshold, culminating in ignition of Teflon in 100 %
oxygen at 1,500 psi. The results for Vespel SP-21 were surprising in that it was
resistant to pneumatic impact ignition even in 100 % oxygen at 5,000 psi. This
was a much higher passing impact pressure than any other material. It is likely
that the graphite filler material increases the thermal conductivity of Vespel
SP-21, making it more difficult to raise any localized area to its ignition tem-
perature.

Conclusions and Limitations


As would be expected, materials are easier to ignite as oxygen concentration
increases. There can be dramatic effects on AITs and ignitability by pneumatic
impact, but the effects vary greatly from material to material. No obvious trend
was found in terms of which materials are more greatly affected by oxygen
concentration. The test results did indicate that all materials are vulnerable to
ignition by pneumatic impact as the pressure and oxygen concentration in-
crease, and even materials like Teflon 共which are typically considered among
the most oxygen compatible of polymers兲 will still ignite with a relatively low
test pressure in 100 % oxygen. This trend is logical given the extreme tempera-

2
Zytel® and Teflon® are registered trademarks of E. I. Du Pont de Nemours and Co.,
Wilmington, Delaware.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—AIT test results.
a
Average AIT, °C
Maximum Delta Between

Downloaded/printed by
Material 21 % O2 34 % O2 45 % O2 100 % O2 Average AIT Values 共°C兲
Buna-N 394 392 391 385 9
Teflon® 446 442 440 439 7
Silicone Rubber 306 301 301 302 5
Vespel® SP-21 420 376 368 342 78
Viton® A 312 305 299 293 19
Zytel® 42 272 255 247 203 69
a
Data shown are the average value from a total of five tests.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


200 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 3—Pneumatic impact test results.

Oxygen Passing Impact Failing Impact Number of Reactions/Number


Concentration Pressurea Pressure of Samples Impacted

Downloaded/printed by
Buna-N
21 % 2,500 3,000 1/20
34 % 2,000 2,500 1/13
45 % 1,000 1,500 1/17
100 % 1,000 1,500 1/1

Teflon®
21 % 9,500 N/Ab N/Ab
34 % 4,500 5,000 1/13
45 % 1,500 2,000 1/1

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
100 % 1,000 1,500 1/10

Silicone rubber
21 % 4,500 5,000 1/10
34 % 2,500 3,000 1/4

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


45 % 1,500 2,000 1/4
100 % 1,000 1,500 1/16

Vespel® SP-21
21 % 9,500 N/Ab N/Ab
34 % 9,500 N/Ab N/Ab
45 % ¯c ¯c ¯c
100 % 5,000 5,500 1/2

Viton® A
SMITH, doi:10.1520/JAI102285 201
TABLE 3— 共Continued.兲

Oxygen Passing Impact Failing Impact Number of Reactions/Number


Concentration Pressurea Pressure of Samples Impacted

Downloaded/printed by
21 % 3,000 3,500 1/1
34 % 2,000 2,500 1/10
45 % 1,500 2,000 1/1
100 % 1,000 1,500 1/6

Zytel® 42
21 % 2,000 2,500 1/1
34 % 1,000 1,500 1/1
45 % 500 1,000 1/2
100 % 500 1,000 1/8

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Note: N/A⫽not applicable.
a
Materials pass pneumatic impact testing when there are zero reactions in 20 tests.
b
The material passed testing at the highest possible impact pressure.
c
Pneumatic impact tests were not performed on Vespel SP-21 in 45 % oxygen.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


202 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
SMITH, doi:10.1520/JAI102285 203

ture rises that can occur as a result of a rapid pressurization event. The data
from this study will be helpful in assessing the ignition hazards in oxygen-
enriched systems, particularly those operating with less than 100 % oxygen but
greater than 21 % oxygen. Additionally, these data are helpful in understanding
the ignition hazards present in compressed air systems.

References

关1兴 ASTM G63, “Standard Guide for Evaluating Nonmetallic Materials for Oxygen
Service,” Annual Book of ASTM Standards, ASTM International, West Consho-
hocken, PA, 1999.
关2兴 ASTM G72, “Standard Test Method for Autogenous Ignition Temperature of Liq-
uids and Solids in a High-Pressure Oxygen-Enriched Environment,” Annual Book
of ASTM Standards, ASTM International, West Conshohocken, PA, 2001.
关3兴 ASTM G74, “Standard Test Method for Ignition Sensitivity of Materials to Gaseous
Fluid Impact,” Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA, 2001.
关4兴 Bryan, C. J., Hirsch, D. B., Haas, J., and Beeson, H. D., “Ignitability in Air, Gaseous
Oxygen, and Oxygen-Enriched Environments of Polymers Used in Breathing-Air
Devices, Final Report,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, Steinberg, T. A., Newton, B.
E., and Beeson, H. D., Eds., ASTM International, West Conshohocken, PA, 2000,
pp. 87–100.
关5兴 Hirsch, D., Hshieh, F., Beeson, H., and Bryan, C., “Ignitibility in Air, Gaseous
Oxygen and Oxygen-Enriched Environments of Polymers Used in Breathing-Air
Devices,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, Royals, W. T., Chou, T. C., and Steinberg,
T. A., Eds., ASTM International, West Conshohocken, PA, 1997, pp. 359–369.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102342
Available online at www.astm.org/JAI

Stephen F. Peralta,1 Keisa R. Rosales,2 and Joel M. Stoltzfus3

Liquid Oxygen Rotating Friction Ignition


Testing of Aluminum and Titanium
with Monel® and Inconel® for Rocket Engine
Propulsion System Contamination
Investigatio

ABSTRACT: Metallic contaminant was found in the liquid oxygen 共LOX兲 pre-
valve screen of the shuttle main engine propulsion system on two orbiter
vehicles. To investigate the potential for an ignition, NASA Johnson Space
Center White Sands Test Facility performed 共modified兲 rotating friction igni-
tion testing in LOX. This testing simulated a contaminant particle in the low-
pressure oxygen turbo pump 共LPOTP兲 and the high-pressure oxygen turbo
pump 共HPOTP兲 of the shuttle main propulsion system. Monel® K-500 and
Inconel® 718 samples represented the LPOTP and HPOTP materials. Alu-
minum foil tape and titanium foil represented the contaminant particles. In
both the Monel® and Inconel® material configurations, the aluminum foil tape
samples did not ignite after 30 s of rubbing. In contrast, all of the titanium foil
samples ignited regardless of the rubbing duration or material configuration.
However, the titanium foil ignitions did not propagate to the Monel and In-
conel materials. Based on these results, NASA determined that the risk of a
fire occurring in the turbo pumps was acceptably low.
KEYWORDS: rotating friction ignition testing, liquid oxygen
environment, aluminum, titanium, metallic contaminant particles

Manuscript received January 22, 2009; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
Mechanical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces,
NM 88004.
2
Chemical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces, NM
88004.
3
Project Manager, NASA White Sands Test Facility, P.O. Box 20, Las Cruces, NM 88004.
Cite as: Peralta, S. F., Rosales, K. R. and Stoltzfus, J. M., ⬙Liquid Oxygen Rotating
Friction Ignition Testing of Aluminum and Titanium with Monel® and Inconel® for
Rocket Engine Propulsion System Contamination Investigatio,⬙ J. ASTM Intl., Vol. 6, No.
10. doi:10.1520/JAI102342.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 204
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 205

Introduction

NASA Johnson Space Center White Sands Test Facility 共WSTF兲 performed
共modified兲 rotating friction ignition testing in liquid oxygen 共LOX兲. The testing
was performed because metallic contaminant particles were found lodged in
the LOX pre-valve screen of the shuttle main engine propulsion system on two
orbiter vehicles. The contaminant was considered, as worst case, to be either
aluminum or titanium. Particle size was estimated to be approximately 1.8
⫻ 1.8⫻ .08 mm3 共0.070⫻ 0.070⫻ 0.003 in.3兲. The concern was that a contami-
nant particle could cause an ignition if it were to become dislodged from the
screen and travel downstream into either the low-pressure oxygen turbo pump
共LPOTP兲 or the high-pressure oxygen turbo pump 共HPOTP兲. This paper sum-
marizes the testing and analysis performed to determine if the contaminant
was vulnerable to frictional ignition in the presence of LOX and if it could lead
to ignition of the turbo pump materials of construction.
Rotating friction ignition testing was one of four phases of testing per-
formed to support the investigation of contaminant in the pre-valve LOX
screen. Promoted combustion, modified mechanical impact, and frictional test-
ing were also performed in support of this effort.

Test Materials and Sample Preparation

Aluminum foil tape and titanium foil used in the processing of the shuttle main
engines were provided to WSTF from the shuttle program. Though the specific
alloy designations of the foils were unknown, the samples received were deter-
mined by scanning electron microscopy to be similar to 1145 aluminum and
unalloyed titanium grade 1, 2, 3, or 4. Simulated particle samples were con-
structed from aluminum foil tape and titanium foil, materials similar to those
used in and around the shuttle main engine. The sample length was determined
by a cut strip of the material with a measured mass between 0.3 and 0.4 mg for
the aluminum tape and between 0.5 and 0.6 mg for the titanium foil. The width
of the strip was 1.8± 0.08 mm 共0.070± 0.003 in.兲. Once the sample length was
determined, a blank piece with a tail was cut. The tail would be used to fasten
the particle sample to the stationary friction sample. The tail was no longer
than 6.4 mm 共0.25 in.兲 and no shorter than 2.54 mm 共0.1 in.兲. To simulate the
“crumpled” shape of the particle found on the space shuttle orbiter pre-valve
screens, the strip was folded into an “accordion” shape approximately 0.90 mm
共0.035 in.兲 long. Using tweezers, the tail end was twisted 90° 共Figs. 1 and 2兲.
Each simulated particle was then fastened to a stationary sample using a brass
flathead screw 共Fig. 3兲. The rotating and stationary samples were fabricated
from Monel®4 K-500 and Inconel®4 718 to represent the LPOTP and HPOTP
materials of construction. The stationary sample was made of Inconel; and the
rotating samples varied, being either Monel or Inconel. Figure 4 depicts the
rotating and stationary samples.

4
Monel® and Inconel® are registered trademarks of Inco Alloys International, Inc., Hun-
tington, West Virginia.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
206 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Aluminum foil tape sample 共top兲; titanium foil sample 共bottom兲.

FIG. 2—Aluminum 共top兲 and titanium 共bottom兲 foil samples, second view.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 207

FIG. 3—Aluminum foil tape simulated particle mounted to Inconel stationary sample.
共See Fig. 2 for a scale indicating the aluminum foil size.兲

Test Procedures and Conditions


The rotating frictional tests were conducted using the WSTF LOX frictional
heating test system to simulate what could occur inside the LPOTP and HPOTP
if contaminant were to reach those parts. The location and configuration of the
samples in the test system are depicted in Fig. 5 关1兴. The test chamber was
assembled and installed in the test system, and the data acquisition system was

FIG. 4—Rotating sample 共left兲; stationary sample 共with particle holder兲 共right兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
208 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—Location of sample holders in test system.

activated. After pressurizing the test system, the chamber was pressurized and
checked for leaks, then vented to ambient pressure. With flowing LOX, the test
chamber was chilled down and pressurized to the target pressure of approxi-
mately 400 psia. Once the LOX pressure stabilized, the drive motor was turned
on. The rotational speed of the rotating sample was approximately 30 000 r/min
to represent the linear velocities found in the turbo pumps. The data acquisi-
tion program operated a pneumatic cylinder that controlled the linear travel of
the rotating sample to and from the stationary sample. This allowed the foil
sample to be “crushed” to a target gap of 0.51 mm 共0.02 in.兲 for Monel and 0.76
mm 共0.03 in.兲 for Inconel. These crush heights were the gaps found in the turbo
pumps where the particle could become lodged and possibly ignite.
Each test with the aluminum foil tape was terminated after rubbing for a
maximum duration of 30 s. Tests with the titanium foil were terminated at
ignition, which was 1 s in most cases 共Table 1兲.
Because the Monel K-500 was not considered flammable at the test condi-
tions, only one titanium with Monel sample was tested. However, several tests
were run for titanium with Inconel samples because it was not certain whether
the titanium foil would propagate to the Inconel material.
At each test termination, the acquisition program reversed the motion of
the pneumatic cylinder, pushing the rotating sample away from the stationary
sample and foil test sample. After the test system was shut down and the test
chamber vented to ambient pressure, the sample specimens were removed,
visually inspected, and photographed. Observations were recorded, and the test
results were analyzed.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 209

TABLE 1—Summary of results for rotating friction testing.

Rub
Simulated Target Gap Pressure Duration Speed Sample
Sample Configuration Location 共mm 共in.兲兲 共psia兲 共s兲 共r/min兲 Ignition
Titanium/Monel ⫺ 1 LPOTP 0.51共0.02兲 512 30 30 100 Yes
Aluminum/Monel – 1 LPOTP 0.51共0.02兲 446 30 30 300 No
Aluminum/Monel – 2 LPOTP 0.51共0.02兲 400 30 30 500 No
Titanium/Inconel – 1 HPOTP 0.76共0.03兲 430 5a 30 600 Yes
Titanium/Inconel – 2 HPOTP 0.76共0.03兲 480 1 30 200 Yes
Titanium/Inconel – 3 HPOTP 0.76共0.03兲 440 1 30 000 Yes
Titanium/Inconel – 4 HPOTP 0.76共0.03兲 420 1 30 200 Yes
Titanium/Inconel – 5 HPOTP 0.76共0.03兲 356 1 30 000 Yes
Aluminum/Inconel – 1 HPOTP 0.76共0.03兲 445 30 30 300 No
Aluminum/Inconel – 2 HPOTP 0.76共0.03兲 443 30 30 000 No
Aluminum/Inconel – 3 HPOTP 0.76共0.03兲 446 30 29 800 No
Aluminum/Inconel – 4 HPOTP 0.76共0.03兲 448 30 30 300 No
Aluminum/Inconel – 5 HPOTP 0.76共0.03兲 435 30 30 200 No
Aluminum/Monel – 3 LPOTP 0.51共0.02兲 421 30 29 900 No
Aluminum/Monel – 4 LPOTP 0.51共0.02兲 492 30 30 000 No
Aluminum/Monel – 5 LPOTP 0.51共0.02兲 410 30 29 900 No
a
Bearing failure caused the rub duration to last longer than 1 s.

FIG. 6—Aluminum/Inconel—1 共no ignition兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
210 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Aluminum/Inconel—2 共no ignition兲.

FIG. 8—Aluminum/Monel—1 共no ignition兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 211

FIG. 9—Aluminum/Monel—2 共no ignition兲.

Test Results and Discussion


For the aluminum foil tape with the Monel and Inconel samples, five tests were
performed with each sample material in LOX at 400 psia. The aluminum did
not ignite after being rubbed for 30 s 共Figs. 6–9兲. One test was performed using
the titanium foil and a Monel K-500 sample. The titanium ignited after being
rubbed for 30 s 共Fig. 10兲. Five tests were performed using titanium foil and an
Inconel 718 sample. In all five of the tests, the titanium ignited 共Figs. 11–15兲;
however, the fire did not propagate to the adjacent Inconel 718 materials.
These results show a trend similar to those obtained by Benz et al. 关2兴 In
that work, the product of the loading pressure and the relative velocity between
the rotor and stator 共Pv product兲 required for ignition of aluminum 6061-T6 in
gaseous oxygen at 6.9 MPa 共1000 psi兲 was 0.063⫻ 108 N / m2 · m / s 共0.036
⫻ 106 lbf/ in.2 · in. / s兲 compared to 0.0039⫻ 108 N / m2 · m / s 共0.002
⫻ 10 lbf/ in. · in. / s兲 for Ti-6Al-4V. The Pv product required to ignite alumi-
6 2

num was more than an order of magnitude greater than the Pv product re-
quired to ignite a titanium alloy. In this work in which aluminum and titanium
foils were rubbed by Monel K-500 or Inconel 718 rotors in 400 psia LOX, the
result showed the same trend. The aluminum foil did not ignite when rubbed
for 30 s, while the titanium foil ignited when rubbed for as short a time as 1 s.
This result affirmed that titanium was significantly more vulnerable to ignition
by frictional heating than aluminum.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
212 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 10—Titanium/Monel—1 共ignition兲.

FIG. 11—Titanium/Inconel—1 共ignition兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 213

FIG. 12—Titanium/Inconel—2 共ignition兲.

Comparison of Properties
The thermal conductivity of a material affected the ability of the foil material
being rubbed to conduct the heat generated by friction throughout the test
sample. If the thermal conductivity was large 共Table 2兲, then hot spots are less
likely to form. If the thermal conductivity was small, then hot spots are more

FIG. 13—Titanium/Inconel—3 共ignition兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
214 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 14—Titanium/Inconel—4 共ignition兲.

likely to form. In this case, the thermal conductivity of aluminum was


2.2 W · cm−1 · K−1 共127 Btu· h−1 · ft−1 · ° F−1兲, which was more than an order of
magnitude greater 共12.2 times兲 than the thermal conductivity of titanium,

FIG. 15—Titanium/Inconel—5 共ignition兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 215

TABLE 2—Comparison of thermal conductivity and yield strength of titanium grade 1, 2, 3,


or 4 and 1145 aluminum.

Thermal Conductivity at 300 K Yield Strength

Foil Material W · cm−1 · K−1 Btu· h−1 · ft−1 · ° F−1 MPa ksi
Titanium grade 1, 2, 3, or 4 0.18 10.4 240–655 35–95
1145 aluminum 2.2 127 90–165 13–24

which was 0.18 W · cm−1 · K−1 共10.4 Btu· h−1 · ft−1 · ° F−1兲 关3兴. Therefore, one
would expect a significantly greater likelihood that the titanium sample would
ignite by friction.
Following a similar logic, it was surmised that yield strength affects the
resistance to deformation of the foil material when loaded by the spinning
rotor. The greater the yield strength, the more a test sample will be able to resist
deformation and incur a greater loading force, which will in turn generate
more heat. In this case, the titanium foil has a yield strength between 240 and
655 MPa 共35–95 ksi兲 关4兴, which was from 1.5 to 7.3 times greater than 90–165
MPa 共13–24 ksi兲 for 1145 aluminum 关5兴. Therefore, one would expect a signifi-
cantly greater likelihood that the titanium sample would ignite by friction.
Combining the two effects, one can predict that the titanium was 18
共12.2⫻ 1.5兲 to 88 共12.2⫻ 7.3兲 times more likely to ignite than the aluminum.
This prediction was consistent with the test results.

Propagation to Surrounding Hardware


A highly significant aspect of this analysis was to determine if the ignited alu-
minum or titanium foils would propagate a fire to the surrounding metal hard-
ware. Of course, if the metal foil parts did not ignite by friction, then the like-
lihood that they would propagate a fire to the surrounding hardware would be
zero. And, in the case of the aluminum foil, which did not ignite in the friction
tests, the likelihood that it would propagate a fire to the surrounding Monel
K-500 or Inconel 718 was affirmed to be zero. In the case of the titanium foil,
which did ignite in all five tests when rubbed with an Inconel 718 rotor, no
propagation from the burning titanium to the Inconel 718 stator or rotor oc-
curred. Based on these tests, it was concluded that even if a titanium particle in
the LOX pre-valve screen were to become dislodged and travel downstream
into the HPOTP and if that particle were to become trapped in the 0.76 mm
共0.03 in.兲 wide gap between the rotor and the stator, even though the particle
was likely to ignite, it was unlikely that it would propagate a fire to the Inconel
718 parts of the HPOTP. In addition, since the Monel K-500 was less flammable
than the Inconel 718 关6兴, it was also concluded that a similar ignition of a
titanium particle would not propagate a fire to the LPOTP parts, which were
fabricated from Monel K-500.
It was surmised that the small amount of heat released by the burning
titanium foil combined with the large heat losses to the 400 psia LOX in the test
chamber and to the cold Inconel 718 stator and rotor was the reason that the
burning titanium foil particle did not propagate a fire into the surrounding
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
216 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Inconel 718 parts. To evaluate this conclusion, the total heat released by the
burning particle in this test was compared to the total heat released in particle
impact tests reported in the literature. In the present work, the titanium foil
sample released approximately 8 J when burned completely 共0.0005 g
⫻ 16 000 J / g = 8 J兲. In the particle impact test work by Benz et al., Inconel 718
targets were not ignited in 14 tests by 1600 ␮m diameter aluminum particles,
even though the Inconel 718 targets were preheated to temperatures ranging
from 525 to 675 K 共485– 755° F兲 关1兴. The total heat release of the 1600 ␮m
diameter aluminum particles used in those tests was about 176 J. In addition,
Inconel 718 targets were not ignited in 24 tests by 2000 ␮m diameter alumi-
num particles, even though the Inconel 718 targets were preheated to tempera-
tures ranging from 340 to 450 K 共152– 350° F兲 关7兴. The total heat released by
the 2000 ␮m diameter aluminum particles used in those tests was about 342 J.
Comparing these values, it was apparent that the burning titanium foil
particle simply did not release enough heat to cause ignition of the cold Inconel
718 stator and rotor.

Determination of Risk of Ignition


The process of determining the risk of ignition by friction and burning of the
space shuttle oxygen turbo pumps as a result of a titanium or aluminum par-
ticle getting through the pre-valve screen and lodging in the gap between the
rotor and housing of the pump was a combination of quantitative and qualita-
tive evaluations. Some aspects of determining that the tests were performed at
conditions equal to or more severe than those in the HPOTP and LPOTP were
quantitative: The gap between the turbo pump rotor and housing was mea-
sured and duplicated in the test, the rubbing speeds were evaluated and simu-
lated in the test apparatus, and the fluid state, LOX versus gaseous oxygen
共GOX兲, was equivalent in the test and the actual flight hardware. On the other
hand, other aspects were qualitative: The size of the particles observed in the
pre-valve screen could not be measured, so the particle size for the test had to
be estimated. The test pressure was limited to 400 psia, while the pressure in
the HPOTP was higher, but it was determined that the lower pressure would
not significantly affect the heat transfer to the LOX from the heated and burn-
ing test samples. Because the particle material observed in the LOX pre-valve
could not be removed and analyzed, a worst-case scenario of titanium and
aluminum foil was tested to ensure that the test was conservative.
It was determined, using both quantitative and qualitative evaluations, that
the tests were conducted at conditions equal to or more severe than those in the
HPOTP and LPOTP.
Whether or not the particles ignited in the test apparatus was an objective
or quantitative aspect of the analysis. In this case, the aluminum foil particles
did not ignite and the titanium foil particles did ignite. If neither of the foil
particle materials had ignited, then because the conditions of the test were
equal to or more severe than the conditions in the flight hardware, it would be
appropriate to conclude that there was no risk of ignition in the flight hard-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
PERALTA ET AL., doi:10.1520/JAI102342 217

ware. However, since the titanium particle did ignite, it was necessary to deter-
mine if the burning titanium foil particle would promote a fire in the surround-
ing LPOTP and HTOPT materials.
In the five tests in which the titanium foil ignited, the burning did not
propagate into the Inconel 718 stator or rotor. This result was not a surprise
since the total heat released by burning the small titanium foil sample was in
the range of 8 J compared to the 176 and 342 J released in the burning of the
aluminum particles in the particle impact tests on Inconel 718 cited above. In
addition to the relatively small heat release of the burning titanium foil sample,
the particle impact tests were conducted in warm GOX, making it a much more
likely scenario to promote a fire in the test sample than in the titanium foil test,
which was conducted in LOX.
Therefore, because the ignited titanium foil samples did not propagate
burning to the Inconel 718 stator and rotor and because the test parameters
were more severe in the test than in the flight hardware, it was determined that
the likelihood of ignition and catastrophic burning in the flight hardware was
very low.
Finally, it was qualitatively or subjectively determined that the likelihood
that a particle would be dislodged from the screen, travel downstream, and
become trapped between a high rotational-speed turbo pump impeller tip and
the housing was extremely low. This low likelihood of occurrence combined
with the very low likelihood that a burning titanium particle would promote
burning in the HPOTP and LPOTP materials was judged to be very low and
enabled those responsible to determine that these risks were acceptable.

Conclusion
Metallic contaminant was observed on the LOX pre-valve screen of the space
shuttle main engine propulsion system. This contaminant was, in the worst
case, considered to be aluminum or titanium foil. The risk of frictional ignition
in the HPOTP and LPOTP was evaluated in the event that the contaminant
metal foil was to become dislodged and travel downstream. Small samples of
aluminum and titanium foil were tested in LOX using a modified rotating fric-
tional ignition test apparatus at NASA WSTF. The aluminum foil particle did
not ignite. The titanium foil particles did ignite in each of the five tests; how-
ever, the burning did not propagate from the burning titanium foil to the adja-
cent Inconel 718 stator and rotor. This lack of propagation was thought to be
due to the relatively small heat release from the burning titanium foil particle
and the very large heat sink provided by the LOX. When this test result, i.e., no
propagation of the fire to the Inconel 718 parts, was combined with the low
probability of the particle becoming dislodged, traveling downstream, and be-
coming trapped precisely between the rotor and housing of the turbo pumps,
NASA determined that the risk of catastrophic fire was acceptably low and
opted not to remove the contaminant.

References

关1兴 Benz, F. J., “Ignition of Metals and Alloys in Gaseous Oxygen by Frictional Heat-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
218 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ing,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:


Second Volume, ASTM STP 910, M. A. Benning, Ed., ASTM International, West
Conshohocken, PA, 1986.
关2兴 Benz, F. J., Williams, R. E., and Armstrong, D., “Ignition of Metals and Alloys by
High-Velocity Particles,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, Vol. 2, ASTM STP 910, M. A. Benning, Ed., ASTM Interna-
tional, West Conshohocken, PA, 1986.
关3兴 Monroe, R. B., Metal Combustion in High-Pressure Flowing Oxygen, Vol. 1, B. Wer-
ley, Ed., ASTM International, West Conshohocken, PA, 1983.
关4兴 The ASM Committee on Aluminum and Aluminum Alloys, Properties and Selection:
Stainless Steels, Tool Materials and Special-Purpose Metals, American Society for
Metals, Metals Park, OH, 1980.
关5兴 The ASM Committee on Aluminum and Aluminum Alloys, Properties and Selection:
Nonferrous Alloys and Pure Metals, American Society for Metals, Metals Park, OH,
1979.
关6兴 Beeson, H. D., Smith, S. R., and Stewart, W. F., Safe Use of Oxygen and Oxygen
Systems: Handbook for Design, Operation, and Maintenance, 2nd ed., ASTM Inter-
national, West Conshohocken, PA, 2007.
关7兴 Beeson, H. D., Safe Use of Oxygen and Oxygen Systems: Guidelines for Oxygen Sys-
tem Design, Materials Selection, Operations, Storage, and Transportation, 1st ed.,
ASTM International, West Conshohocken, PA, 2000.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102284
Available online at www.astm.org/JAI

Kyle Sparks,1 Timothy Gallus,2 and Sarah Smith3

Electrical Arc Ignition Testing for Common


Handheld Electrical Devices in
Oxygen-Enriched Atmosphere

ABSTRACT: Electrical arc ignition is a concern for materials used in low


pressure oxygen-enriched environments such as inside spacecraft. The Na-
tional Aeronautics and Space Administration 共NASA兲 Johnson Space Center
requested that the NASA White Sands Test Facility conduct arc ignition tests
to evaluate the hazard of electrical arc ignition of materials that could be in
close proximity to batteries for the Constellation Program. Wire-break elec-
trical arc tests were performed to determine the current threshold for ignition
of generic cotton with a fixed voltage of 3.7 V, a common voltage for hand-
held electrical devices. These tests were performed in 34 % oxygen at vary-
ing pressures.
KEYWORDS: oxygen, electrical arc ignition, flammability materials
testing, handheld electrical devices

Introduction

National Aeronautics and Space Administration 共NASA兲 Johnson Space Center


共JSC兲 Materials and Processes Branch requested that NASA JSC White Sands
Test Facility 共WSTF兲 perform testing for the Constellation Program to evaluate
the hazard of electrical arc ignition of materials that could be in close proxim-
ity to batteries. Specifically, WSTF was requested to perform wire-break elec-
trical arc tests to determine the current threshold for ignition of generic cotton

Manuscript received December 11, 2008; accepted for publication June 12, 2009; pub-
lished online August 2009.
1
Mechanical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces,
NM 88004.
2
Chemical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces, NM
88004.
3
Project Manager, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces, NM
88004.
Cite as: Sparks, K., Gallus, T. and Smith, S., ⬙Electrical Arc Ignition Testing for Common
Handheld Electrical Devices in Oxygen-Enriched Atmosphere,⬙ J. ASTM Intl., Vol. 6, No.
8. doi:10.1520/JAI102284.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 219
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
220 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Test system.

woven fabric samples with a fixed voltage of 3.7 V, a common voltage for hand-
held electrical devices. The wire-break test was developed during a previous test
program to evaluate the hazard of electrical arc ignition inside the Extravehicu-
lar Mobility Unit 关1,2兴.

Test System
All tests were performed in a test chamber in Cell 131 of the High Pressure Test
Area at WSTF. The humidity was kept to a minimum by the introduction of dry
gases into the test chamber. The test chamber was a 4 in. diameter stainless
steel cross with full diameter end ports at each leg of the cross. The chamber
ports accommodated a gas inlet and outlet, a view window for normal and
high-speed video recording, a test sample mounting block, a thermocouple for
temperature measurement, and power for the electrical arcing. The test system
is shown in Fig. 1.
The arcing and electrical test system included a power supply capable of
providing a maximum of 10 A at 28 Vdc. The power supply was designed to
emulate batteries while allowing flexibility for different voltages and currents
during testing. Two identical TENMA®4 Model No. TSX3510 power supplies
were used along with a transistorized current limiter. This power supply
scheme was very quick to respond to the rapid load changes produced by arc-
ing events.

4
TENMA® is a registered trademark of Premier Farnell Corporation, Independence,
Ohio.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102284 221

FIG. 2—Wire-break test setup.

Test Environment
Tests were originally requested to be performed in 34 % oxygen at 8.0 psi a.
However, the test system was not designed to allow testing at sub-atmospheric
pressures. Therefore, the request was changed to perform tests at 14.7 psi a
共+0.5, ⫺0.0兲 using a mixed gas supply containing 34 % oxygen, with the balance
being nitrogen. Increased pressure generally corresponds to greater ignition
hazards 关3兴, so it was assumed that the increased pressure would add conser-
vatism to the data generated.

Wire-Break Test Method


The wire-break test method was performed by clamping the ends of a single
strand of fine wire, and shaping the wire into a “U” so that the bottom of the U
was in contact with the test material 共Fig. 2兲. Applying current and voltage to
the wire caused the wire to heat up and eventually break. As the wire heated up,
the test material was preheated; and when the wire broke, an arc occurred. An
example data plot of a wire-break test is shown in Fig. 3.
It is logical that the wire needs to be in direct contact with the test sample
because of the importance of preheating. Each wire-break test can be divided
into preheating and arcing events. Calculations performed during a previous
study 关2兴 showed that the amount of available energy during preheating is
roughly three orders of magnitude greater than the available energy during the
arcing event. As the wire heats up while in contact with the test material, the
voltage begins to rise slightly because the electrical resistance of the wire in-
creases. With a fixed-current power supply, this higher resistance creates both a
higher voltage drop across the wire and more heat evolution at the wire. The
heat sinking ability of the wire holder assembly ensures that the hottest loca-
tion on the wire will occur somewhere near the bottom of the U. The hot wire
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
222 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—Example wire-break test plot.

preheats the test material, allowing the material to begin to vaporize before the
arcing event occurs. Once the material has been preheated, ignition occurs at a
much lower arc power since the vapors are easily ignited. The minimum igni-
tion energy for most hydrocarbon vapors is 0.28 to 1.15 mJ 关4兴, and this amount
of energy is easily provided by the arcing event 关2兴.
The batch of wires used in this test series were composed of silver-coated
copper, supplied by Phelps Dodge High Performance Conductors.5 All tests
were performed with 34 AWG 共American Wire Gauge兲 wire.
Each test consisted of one wire-break event, which was observed for visual
evidence of test material ignition. Visual evidence of ignition for fabrics such as
that used in this test series includes a fire that rapidly propagates, resulting in
consumption of the entire test sample. A material failed at the test current if it
ignited once in a maximum of 60 tests. A material passed at the test current if
there was no ignition in 60 tests. For each test, current and voltage measure-
ments were taken at a rate of 10 kHz. In addition, the ambient temperature in

5
Phelps Dodge High Performance Conductors is now IWG High Performance Conduc-
tors, a subsidiary of International Wire Group, Inman, SC.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SPARKS ET AL., doi:10.1520/JAI102284 223

the test chamber and the test pressure were recorded at a rate of 1 Hz. High-
speed video and normal speed video were recorded for each ignition.

Test Materials and Sample Preparation


Previous electrical arc ignition testing 关1,2兴 included testing of generic cotton
woven fabric test samples, which were cut from new unwashed white t-shirts.
Test data indicate that materials such as moleskin are easier to ignite by elec-
trical arc than generic cotton; however, generic cotton was selected as a mate-
rial that represented the worst-case of materials that could be in close proxim-
ity to batteries in the Constellation Program. The generic cotton woven fabric
samples used for this testing were from the same package of white t-shirts used
in the previous testing.
The size of the test samples was not standardized; however, it was ensured
that the wires were not placed in close proximity to the edge of the samples.
The test material was frayed using a wire brush, resulting in surfaces that were
fuzzier and easier to ignite 关1,2兴. The test samples could be used for multiple
tests as long as the wire did not contact the same area from test to test.
All of the samples were tested in the horizontal configuration, and each test
sample was mounted to the sample mounting plate using adhesive tape. It was
ensured that the tape was not in close proximity to the wire and arcing event. In
addition, when taping the materials to the mounting plate, the materials were
not stretched in such a way that they would shrink away from the wire once it
began to heat up.
As the wire heated up, it was possible for a small hole to be created in the
test material, which would allow a short between the wire and the sample
mounting plate. Therefore, a slide cover glass was placed underneath the test
sample when it was mounted to the sample mounting plate, as shown in Fig. 2.
During checkout tests, it was observed that the preheating of the wire prior
to the wire-break event was creating a hole in the test material, leaving no
material in close proximity to the arcing event. Previous testing indicated that
test materials must be in close proximity to the arcing event in order for igni-
tion to occur 关1,2兴. Therefore, the cotton was configured in three layers to en-
sure the test material was in close proximity to the arcing event. This method of
layering the test material was originally used in previous tests 关1,2兴. Layering
the test material results in more vaporization during the preheating of the wire
since there is more material available and in contact with the wire. Although
the precise amount of material being vaporized cannot be controlled from test
to test, it is presumed that layering the material allows the test to be as conser-
vative as possible since there is additional material available to be vaporized
共and it is the vapors that are directly ignited by the arcing event when the wire
breaks兲.

Procedures and Observations


Before testing commenced, the test voltage was set to 3.7 V and the current was
set to slightly lower than the desired test conditions. The test sample was
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
224 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

mounted to the sample mounting plate using tape, and the wire was clamped
into place.
Next, the test chamber was sealed, purged, and pressurized. Power was
then applied to the wire, and the test conductor manually increased the current
until the wire broke, creating an arcing event. If an ignition occurred, testing
was continued at a lower current level. If no ignition occurred, high-speed
video and visual inspection were used to verify that the wire broke in the de-
sired location 共i.e., in a location touching the test sample兲. The test sample
could then be retested on another untested location.
The current required to break each specific wire depended on several vari-
ables. The more heat that went into the sample, the more current was required
to break the wire. Therefore, the thermal properties of the material and the
quality of the contact between the wire and test sample affected the current
required to break the wire. Imperfections of the wire could cause the wire to
break at a lower current. Further, the longer the wire U, the less current was
required to break the wire.

Test Results and Discussion

Testing was performed with the largest wire size 共34 AWG兲 and the highest
current possible. There was no ignition of the test material in these conditions,
as summarized in Table 1. In order to determine the conditions that would lead
to ignition, some tests were performed with increased gas pressure inside the
test chamber. A total of three tests was performed with the test chamber pres-
surized to about 42 psi a, and two of these tests resulted in ignition of the test
sample. A total of 13 tests was performed with the test chamber pressurized to
about 32 psi a, and one of these tests resulted in ignition. Tests that resulted in
ignition at increased pressures are summarized in Table 2.
The test results clearly indicate that with 34 % oxygen at 14.7 psi a, com-
mon handheld battery-operated devices do not pose an electrical arc ignition
hazard for materials that would commonly be used in a spacecraft cabin envi-
ronment. Furthermore, the environment in the actual spacecraft will have a
reduced pressure of 8.0 psi a, which adds additional safety margin to the ap-
plication of this data. It would be expected that the generic cotton woven fabric
material is flammable in 34 % oxygen at 14.7 psi a; however, the test results
show that with a voltage of 3.7 V and the maximum current possible in the test
system, the amount of energy in a wire-break arc test is not sufficient to pro-
duce ignition. Increasing the pressure to 32.9 psi a did result in ignition of the
material. These results indicate that ignition can occur at increased pressures;
however it should be noted that the tests performed do not establish the lowest
pressure at which ignition could occur. That limit would be at a pressure be-
tween 14.7 and 32 psi a.
Additionally, test results indicate that the hazard of electrical arc ignition is
greatly reduced as pressure is decreased. This trend is consistent with data
from other electrical arc tests performed in oxygen-enriched environments
关1,2兴. This trend is also consistent with studies of other ignition mechanisms
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Results of testing in 34 共+0.5, −0.0兲 % oxygen at 14.7 psi a 共+0.5, −0.0兲.

Current Tested Voltage Tested


共A兲 共V兲

Downloaded/printed by
Test Material Test Results Average Maximum Minimum Average Maximum Minimum
Generic cotton No ignition in 60 tests 6.83 9.06 5.68 3.77 3.87 3.57

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


SPARKS ET AL., doi:10.1520/JAI102284 225
226 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 2—Tests resulting in ignition in 34 共+0.5, −0.0兲 % oxygen at increased test


pressures.

Test Pressure Current at Ignition Voltage at Ignition


Test ID 共psi a兲 共A兲 共V兲
Cev gc22 42.5 9.48 3.73
Cev gc23 42.4 7.36 3.85
Cev gc35 32.9 8.64 3.67

and ignition properties 关5–8兴, which showed that the ignitability of materials
decreases as oxygen pressure is decreased.

References

关1兴 Smith, S., Gallus, T., Harper, S., and Beeson, H., “Electrical Arc Ignition Testing of
Spacesuit Materials,” Proceedings of the 11th International Symposium on Flamma-
bility and Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 1479, Oc-
tober 2006, D. B. Hirsch, R. Zawierucha, T. A. Steinberg, and H. M. Barthelemy,
Eds., ASTM International, West Conshohocken PA.
关2兴 Smith, S. R., Gallus, T. D., Tapia, S., and Beeson, H. D., “Electrical Arc Ignition
Testing of Spacesuit Materials,” WSTF-TR-1035-001-01-05, NASA Johnson Space
Center White Sands Test Facility, Las Cruces, NM, 2007.
关3兴 Safe Use of Oxygen and Oxygen Systems, Handbook for Design, Operation, and
Maintenance, Manual 36, 2nd ed., H. D. Beeson, S. R. Smith, and W. F. Stewart,
Eds., ASTM International, West Conshohocken, PA, 2007.
关4兴 Ebadat, V., Electrostatic Hazards in the Chemical Process Industry, Chilworth Tech-
nology Limited, Beta House, Chilworth Science Park, Southampton, United King-
dom, 1996.
关5兴 Bryan, C. J., Hirsch, D. B., Haas, J., and Beeson, H. D., “Ignitability in Air, Gaseous
Oxygen, and Oxygen-Enriched Environments of Polymers Used in Breathing-Air
Devices, Final Report,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1395, Vol. 9, T. A. Steinberg, B. E. Newton, and
H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关6兴 Hirsch, D. B. and Bunker, R. L., “Effects of Diluents on Flammability of Nonmet-
als in High-Pressure Oxygen Mixtures,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres, ASTM STP 1197, Vol. 6, D. D. Janoff and J. M.
Stoltzfus, Eds., ASTM International, West Conshohocken, PA, 1993.
关7兴 Hirsch, D. B., Bunker, R. L., and Janoff, D., “Effects of Oxygen Concentration,
Diluents, and Pressure on Ignition and Flame-Spread Rates of Nonmetals,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP
1111, Vol. 5, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West Con-
shohocken, PA, 1991.
关8兴 Smith, S. R., “Effect of Oxygen Concentration on Autogenous Ignition Tempera-
ture and Pneumatic Impact Ignitability of Nonmetallic Materials,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1522, Vol.
12, H. Barthelemy, S. Smith, T. Steinberg, and C. Binder, Eds., ASTM Interna-
tional, West Conshohocken, PA, 2009.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102286
Available online at www.astm.org/JAI

Keisa R. Rosales1 and Joel M. Stoltzfus2

High Pressure Quick Disconnect Particle


Impact Tests

ABSTRACT: National Aeronautics and Space Administration 共NASA兲


Johnson Space Center White Sands Test Facility 共WSTF兲 performed particle
impact testing to determine whether there is a particle impact ignition hazard
in the quick disconnects 共QDs兲 in the environmental control and life support
system 共ECLSS兲 on the International Space Station 共ISS兲. Testing included
standard supersonic and subsonic particle impact tests on 15–5 PH stainless
steel, as well as tests performed on a QD simulator. This paper summarizes
the particle impact tests completed at WSTF. Although there was an ignition
in test series 4, it was determined that the ignition was caused by the pres-
ence of a machining imperfection. The sum of all the test results and analysis
enabled NASA to conclude that the risk of particle impact ignition hazard in
the ISS ECLSS QDs was acceptable.
KEYWORDS: quick disconnect, high pressure, particle impact testing,
stainless steel

Nomenclature

ECLSS ⫽ environmental control and life support system


ISS ⫽ International Space Station
QD ⫽ Quick disconnect
SS ⫽ stainless steel
WSTF ⫽ NASA White Sands Test Facility

Manuscript received December 11, 2008; accepted for publication June 12, 2009; pub-
lished online August 2009.
1
Chemical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces, NM
88004.
2
Project Manager, NASA White Sands Test Facility, P.O. Box 20, Las Cruces, NM 88004.
Cite as: Rosales, K. R. and Stoltzfus, J. M., ⬙High Pressure Quick Disconnect Particle
Impact Tests,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102286.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 227
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
228 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Introduction
Gas velocities inside the quick disconnects 共QDs兲 in the environmental control
and life support system 共ECLSS兲 on the International Space Station 共ISS兲 are
great enough to raise concerns about a possible particle impact ignition hazard.
These high velocities will occur any time the QDs are mated under pressure,
when the airlock tanks are used to repressurize the system, and when the re-
charge oxygen orifice bypass assembly is used to bypass a flow-limiting orifice
during pressurization of the airlock tanks. As a result of concerns that the high
gas velocities could cause ignition due to particle impact in the QDs, NASA
Johnson Space Center requested White Sands Test Facility 共WSTF兲 to perform
testing to determine whether there is a particle impact ignition hazard in the
QDs.
Testing included standard supersonic and subsonic particle impact tests on
15–5 PH stainless steel 共SS兲, as well as tests performed on a QD simulator. The
particulate mixture used was consistent with the system materials of construc-
tion. This paper summarizes the testing completed at WSTF.

Objective
The test objective was to determine if particle impact samples and a QD simu-
lator composed of 15–5 PH SS would ignite and burn when impacted by par-
ticulate and to determine the effect of the impact at the given temperatures,
pressures, and flow rates.

Test Methods
The selection of test materials, the test series, and the test procedures are de-
scribed.

Materials
The following paragraphs discuss the particulate mixtures, the test target and
QD simulator materials, and the sample preparation.

Particulate Mixtures—Two particulate mixtures were used, as described in


Table 1 and shown in Figs. 1 and 2, with quantity being the number of the
particular particles included in the mixture. The materials that made up the
particulate mixtures 共SSs, Inconel®3 alloys, and Monel®3 alloys兲 are native to
the ISS ECLSS. The ISS ECLSS was initially cleaned to level 200, which dic-
tates no particles greater than 200 ␮m 共0.008 in.兲 and a maximum allowed
distribution for particles smaller than 200 ␮m 共0.008 in.兲. Small amounts of
assembly and operationally generated particulates could be present in the

3
Inconel® and Monel® are registered trademarks of Inco Alloys International, Inc., Hun-
tington, West Virginia.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 229

TABLE 1—Particulate mixtures used for testing.

Particulate Mixture Material Size and Configuration Amount


A Annealed 304 SS 794 ␮m sphere QTY-1
Inconel 718®a ⫺100 meshb powder 3 mg
304 SS -20 meshb powder 3 mg
Inconel 718 250⫻ 250⫻ 2000 ␮m sliverc QTY-4
B Inconel 718 -100 meshb powder 3 mg
304 SS -20 meshb powder 3 mg
Monel®a 250⫻ 250⫻ 2000 ␮m sliverc QTY-4
a
Inconel® and Monel® are registered trademarks of Inco Alloys International, Inc., Hun-
tington, West Virginia.
b
Powders used are commercial products. 100 mesh corresponds to sieve openings of
149 ␮m; 20 mesh corresponds to sieve openings of 841 ␮m. The “⫺” before the sieve
mesh size indicates that the particles pass through the sieve. Typically, 90 % or more of
the particles will lie within the indicated range.
c
Slivers were manufactured in the WSTF machine shop.

system.
In previous particle impact tests on the space shuttle flow control valve 关1兴,
calculations were performed to determine the amounts of particulate allowed
in the propulsion system, external tank, and main engine, based on the surface
areas in those systems. The calculated amount of allowed particulate was then
multiplied by a safety factor, and the resulting amount used for testing was 10

FIG. 1—Particulate mixture A.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
230 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Particulate mixture B.

mg 共0.1 g兲. Similar calculations have not been performed for the ISS ECLSS
system; however, the surface area of the ECLSS system is much smaller. There-
fore, the materials and quantities of particulates used in mixtures A and B are
considered to be extremely conservative because the amount of particles is on
the same order as those used for the flow control valve tests. In addition, the
size range of the particles encompasses those that would be allowed in the
system by the cleanliness level, as well as larger particles.

Test Target and Quick-Disconnect Simulator Materials—The standard super-


sonic and subsonic configurations used 15–5 PH SS targets, corresponding to
the material used at the impingement points inside the ISS ECLSS QDs. For
the supersonic tests, these targets were configured in a cup shape with an out-
side diameter 关0.953 cm 共0.375 in.兲兴 that allowed flow over and around the
target surface. Each supersonic target sample was press-fitted onto a copper
sample holder, as shown in Fig. 3. The standard subsonic configuration targets
were flat discs 0.152 cm 共0.060 in.兲 thick with holes drilled radially to allow flow
over and through the target 共Fig. 4兲.
The QD simulator tests were performed using WSTF-machined parts com-
posed of 15–5 PH SS. The design of the QD simulator parts replicated the
relevant parts inside the flight QDs, based on information supplied by the
manufacturer. Figure 5 shows a cross section of the QD simulator used for
testing.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 231

FIG. 3—Standard supersonic target configuration.

Sample Preparation—All samples were prepared at WSTF and cleaned in a


mild detergent. The samples were then sealed in polypropylene bags until test-
ing. From this point, the targets and simulator parts were handled with latex
gloves to maintain cleanliness.

Test Series and Procedures


All tests were performed in the WSTF High Flow Test Facility. During particle
impact testing, the parameters of concern are temperature, pressure, particle
composition, target, and configuration. The maximum pressure in the ISS
ECLSS QDs is 20.68 MPa 共3000 psi兲, and the maximum temperature is 46° C
共115° F兲. These conditions represent the most severe in-use conditions. The test
temperature for all tests was 46° C 共115° F兲. Six test series were performed, as
described in Table 2.
For test series 1 and 2, the supersonic nozzle was installed on the outlet of
the subsonic injector assembly. The particulate was loaded into the injector,
and the injector cap was threaded onto the housing. The target sample and
copper posts were then positioned at the end of the nozzle 共Fig. 6兲. An untested
sample was used for each test.
For test series 3, the subsonic injector and nozzle were used. The particu-
late was loaded into the injector, and the injector cap was then threaded onto
the housing. The target sample and orifice 共for controlling the gas velocity兲
were then positioned on the end of the subsonic nozzle 共Fig. 7兲. An untested
sample was used for each test.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
232 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Standard subsonic flat target configuration.

For test series 4, 5, and 6, the QD simulator was installed on the outlet of
the flow straightener and subsonic particle injector assembly 共Fig. 8兲. The par-
ticulate was loaded into the injector, and the injector cap was then threaded
onto the housing. The same QD simulator was used for all tests.
For all test series, after system preparation was completed the test area was
cleared of personnel and placed in RED 共no access allowed兲 status. A video
camera was positioned to record any reaction visible at the end of the test
fixture. Heated gaseous aviator’s breathing oxygen 共ABO兲 at test pressure was
allowed to flow until the temperature of the target sample was achieved and the
gas flow stabilized. Upon command, the particulate was injected into the cham-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 233

FIG. 5—Cross section of QD simulator.

ber. After evidence of impact, the oxygen flow was terminated and the test
system was allowed to vent down to ambient pressure, ⬃4 s after particle
injection. After every test in test series 1, 2, and 3, each target sample was
visually inspected to verify that particle impacts had occurred. Each sample
was then individually bagged in its original bag and labeled with test informa-
tion. The test computer saved the test and system data. The test pressure, test
temperature, and average flow meter reading were recorded in the test area log
book. Videotapes of the test were recorded and stored. At the completion of
these tasks, the procedure was repeated.

Results and Discussion


The effect of impact on the test sample at the given temperatures, pressures,
and flow rate was determined visually and characterized in accordance with
NASA-STD-6001 关2兴, as follows:
• For testing materials
No Ignition—characterized by no evidence of combustion. A material that
does not ignite may show one or more particle indentations on the surface of
the material, which may include localized erosion.
Ignition—a portion or the entire sample is consumed; the target may not be
recoverable.
• For testing components
No Ignition—characterized by no evidence of combustion. A component
that does not ignite may show one or more particle indentations on the inside
surface of the component, which may include localized erosion.
Ignition—characterized by obvious consumption of the component, which
may include burn-out.
For a summary of all the test results, see Table 2.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—Particle impact test series and results summary.

Test Pressure Test Temperature Number


Particulate 共MPa兲 共°C兲 of

Downloaded/printed by
Test Series Test System Configuration Mixturea 关psig兴 关°F兴 Tests Number of Ignitions
1 Subsonic injector with supersonic nozzle A 7.24 关1050兴 46关115兴 100 0
2 Subsonic injector with supersonic nozzle A 11.72 关1700兴 46关115兴 85 0
3 Subsonic injector with subsonic nozzle A 20.68 关3000兴 46关115兴 60 0
4 QD simulator A 20.68 关3000兴 46关115兴 1 1
5 QD simulator B 17.24 关2500兴 46关115兴 60 0
6 QD simulator A 20.68 关3000兴 46关115兴 60 0
a
Particulate mixtures are described in Table 1 of this report.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


234 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 235

FIG. 6—Subsonic injector with supersonic nozzle used for test series 1 and 2.

Test Series 1
In test series 1, 100 tests were performed in ABO-grade oxygen with a target
pressure of 7.24 MPa 共1050 psi兲 and a target temperature of 46° C 共115° F兲. The
gas stream velocities were at least Mach 1 in the throat of the convergent/
divergent nozzle. There were no ignitions of the test samples in 100 tests. The
test pressure for test series 1 was chosen to simulate the environment in the low
pressure ECLSS QDs. It was not yet known that there would be particle impact
concerns in the high pressure ECLSS QDs, which have a maximum pressure of
20.68 MPa 共3000 psi兲 共see test series 3兲. Figure 9 shows a typical posttest test
sample.

FIG. 7—Subsonic injector and nozzle used for test series 3.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
236 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 8—QD simulator test system used for test series 4, 5, and 6.

Test Series 2
In test series 2, 85 tests were performed in ABO-grade oxygen with a target
pressure of 11.72 MPa 共1700 psi兲 and a target temperature of 46° C 共115° F兲.
The gas stream velocities were at least Mach 1 in the throat of the convergent/
divergent nozzle. There were no ignitions of the test samples in 85 tests. Al-
though higher test pressures were desired to simulate the maximum pressure in
the high pressure ECLSS QDs, 11.72 MPa 共1700 psi兲 was the maximum capa-
bility of WSTF’s supersonic particle impact test system due to the large pres-
sure drop through the supersonic nozzle. Figure 10 shows a typical posttest
sample.

Test Series 3
In test series 3, 60 tests were performed in ABO-grade oxygen with a system
pressure of 20.68 MPa 共3000 psi兲 and system gas temperature of 46.1° C
共115° F兲. Unlike the supersonic tests, these subsonic tests measured the pres-
sure and temperature upstream of the test sample. There were no ignitions of
the test samples in 60 tests. The orifice size was 0.874 cm 共0.344 in.兲, which led
to an average gas velocity of 130 m/s 共425 ft/s兲 共calculated from results obtained
by the system mass flow meter兲. This test series was performed because greater
test pressures could not be obtained in test series 2. Figure 11 shows a typical
posttest sample.

Test Series 4
In test series 4, one test was performed in ABO-grade oxygen with a pressure at
the inlet of the QD simulator of 20.68 MPa 共3000 psi兲 and system gas tempera-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 237

FIG. 9—Typical test series 1 supersonic posttest sample at 7.24 MPa 共1050 psi兲.

ture of 46.1° C 共115° F兲. There was an ignition of the QD simulator on the first
test. The flow direction in the QD simulator was from the female side to the
male side. Following the ignition, the QD simulator was sectioned to inspect
the internal parts and burned area. Figure 12 shows the burned QD simulator.
Upon inspection, a machining imperfection was found at the inlet of the throat
共Fig. 13兲. The velocity profile in the QD simulator with the machining imper-
fection was calculated using computational fluid dynamics 共CFD兲 and is shown
in Fig. 14. It is apparent that the machining imperfection created sonic veloci-
ties at the inlet to the female stop and sleeve. The velocity made the test unnec-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
238 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 10—Typical test series 2 supersonic posttest sample at 11.72 MPa 共1700 psi兲.

essarily severe and unrealistic since the inspected flight QDs would not have
similar imperfections. Therefore, a new QD simulator was machined for the
subsequent test series.

Test Series 5
A new QD simulator, verified to have no machining imperfections, was used in
test series 5. Sixty tests were performed in 100 % oxygen with a pressure of
17.24 MPa 共2500 psi兲 at the inlet of the QD simulator and a test gas temperature
of 46.1° C 共115° F兲. Test series 5 was the only series which used particulate
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 239

FIG. 11—Typical test series 3 subsonic posttest sample at 20.68 MPa 共3000 psi兲.

mixture B. Because the ignition in test series 4 raised many concerns about
upcoming ISS ECLSS operations, the test pressure was chosen to envelop the
worst-case pressure in near-term operations and the particulate mixture was
made less severe, as mixture A was considered to be overly conservative. There
were no ignitions of the QD simulator in 60 tests. Posttest inspections revealed
that the larger particulates were getting caught in various locations inside the
QD simulator, as shown in Fig. 15. Therefore, after each test was performed,
the QD simulator had to be disassembled, cleaned, and reassembled.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
240 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 12—Burned QD simulator.

FIG. 13—Location of machining imperfection.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 241

FIG. 14—CFD of velocity profile of QD simulator with machining defect.

Test Series 6
In test series 6, tests were performed to repeat the same test conditions where
the ignition occurred in test series 4, but without the machining imperfection.
Sixty tests were performed in 100 % oxygen with a pressure of 20.68 MPa 共3000
psi兲 at the inlet of the QD simulator and a test gas temperature of 46.1° C
共115° F兲. The velocity profile determined by CFD is shown in Fig. 16. There
were no ignitions of the QD simulator in 60 tests. Just as in test series 5, post-
test inspections revealed that the larger particulates were getting caught inside
the QD simulator in various locations. The results of test series 6 indicate that
the machining imperfection found in test series 4 was the cause of the ignition.

Statistical Analysis
The binomial distribution is used to describe results from tests of a “pass/fail”
or “go/no-go” nature, like those obtained from the particle impact test. From
the number of ignitions that occur in a certain number of independent particle
impact tests, upper and lower confidence limits for reliability 关3兴 can be calcu-
lated. The upper and lower limits for reliability can be used to predict the range
of probabilities that an ignition would occur if the ISS ECLSS QDs were to be
subjected to an ignition source as severe as those used in the tests. Obviously, if
zero ignitions occur in a certain number of particle impact tests, then the upper
confidence limit for reliability is 100 %, predicting that, in the best-case sce-
nario, zero ignitions will occur if the ISS hardware are exposed to an ignition
source as severe as the one used in the tests. On the other hand, the lower
confidence limit for reliability will be some value less than 100 %, predicting, in
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
242 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 15—Particulate caught in inlet of female stop.

the worst-case scenario, that no more than some maximum number of igni-
tions may occur if the ISS hardware are exposed to an ignition source as severe
as the one used in the tests. For example: If the lower confidence limit for
reliability is 97 %, then this indicates that no more than 3 % of the instances in
which the ISS ECLSS QDs are subjected to particle impact at the conditions
equal to those used in the test may yield ignitions. It is the lower confidence
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 243

FIG. 16—CFD of velocity profile of QD simulator without machining defect.

limit for reliability that is of particular interest to the user of the hardware and
it can be used, as will be demonstrated below, to evaluate the risk of fire occur-
ring in the ISS ECLSS QDs.
The lower confidence limit for reliability, or maximum probability of an
ignition occurring in ISS hardware, is a function of at least two parameters: the
confidence level chosen for the analysis and the number of tests performed. For
this analysis, the authors selected a 99 % confidence level because that value
enables one to perform a reasonable and cost effective number of tests and still
make an acceptably confident decision. The number of tests performed varied
from 60 to 100 in test series 1, 2, 3, 5, and 6. It is obvious that the greater the
number of tests performed without ignition, the lower the predicted maximum
probability will be that particle impact ignitions may occur in the ISS ECLSS
QDs.
Applying these principles to the particle impact tests described herein, it
was determined for test series 1 共supersonic nozzle, 1050 psig, particle mix A兲,
in which zero ignitions occurred in 100 tests, the predicted maximum probabil-
ity of an ignition occurring in the ISS ECLSS QDs at these conditions is 3 %
共using a 99 % confidence level兲. For test series 2 共supersonic nozzle, 1700 psig,
particle mix A兲, in which zero ignitions occurred in 85 tests, the predicted
maximum probability of an ignition occurring in the ISS ECLSS QDs is 4 %
共using a 99 % confidence level兲. For test series 3 共subsonic nozzle, 3000 psig,
particle mix A兲, test series 5 共QD simulator, 2500 psig, particle mix B兲, and test
series 6 共QD simulator, 3000 psig, particle mix A兲, in which zero ignitions oc-
curred in 60 tests, the predicted maximum probability of a fire event occurring
in the ISS ECLSS QDs is 5 % 共using a 99 % confidence level兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
244 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

It is possible to combine test results to obtain a lower predicted maximum


probability of ignition. For example, if one test series is conducted at a higher
pressure and another test series is conducted at a lower pressure, the results
can be combined to predict the maximum probability of a fire event occurring
at the lower pressure. Using this logic, the results from test series 1 and 2 can
be combined to determine the maximum probability of a fire event occurring in
the ISS ECLSS QDs at 1050 psig using the supersonic nozzle. In this case, there
were zero ignitions in 185 共100+ 85= 185兲 tests. This predicts that the maxi-
mum probability of ignition at 1050 psig is only 2 % 共using a confidence level of
99 %兲. Similarly, the results of test series 5 共QD simulator, 2500 psig, particle
mix B兲 and test series 6 共QD simulator, 3000 psig, particle mix A兲 can be com-
bined. In this case, particle mix B has fewer particles than particle mix A, so it
is considered the least severe case. With this combination, the maximum pre-
dicted probability of a fire event occurring in the ISS ECLSS QDs with particle
mix B, at 2500 psig is 3 % 共using a 99 % confidence level兲.

Determination of Acceptable Risk


To evaluate the overall risk of particle impact ignition occurring in the ISS
ECLSS QDs, it is necessary to consider three factors: 共1兲 The maximum prob-
ability of ignition in the ISS ECLSS QDs as predicted by the analysis of the test
results discussed in the previous paragraphs, 共2兲 the likelihood that the type
and quantity of contaminant used in the tests will accumulate upstream of the
ISS ECLSS QDs, and 共3兲 the likelihood that the collected contaminant will be
simultaneously released.
These factors were evaluated as follows:
共1兲 The predicted maximum probability of ignition was determined to be
less than 3 % 共using a 99 % confidence level兲 for test series 1 and 2 and
less than 5 % 共using a 99 % confidence level兲 for test series 5 and 6.
共2兲 The likelihood of the tested contaminant mixtures being accumulated
upstream of the ISS ECLSS QDs was determined to be very low due to
共a兲 the initial cleanliness of the ISS ECLSS system and 共b兲 the fact that
WSTF tested using quantities of contaminants that were significantly
greater than the maximum quantity of contaminant estimated to be
present in the entire ISS ECLSS.
共3兲 The likelihood of the collected contaminant being simultaneously re-
leased was determined to be very unlikely due to the low likelihood that
an ISS ECLSS filter will fail and release its contents into the down-
stream system.
Since the maximum predicted probability of an ignition event occurring
was low, and the likelihood of accumulation and simultaneous release were
both expected to be very low, NASA determined that the risk of particle impact
ignition in the ISS ECLSS QDs was acceptable.

Conclusions
Six series of tests were performed on 15–5 PH SS targets and a QD simulator at
subsonic and supersonic velocities to determine if a particle impact ignition
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102286 245

hazard existed in the QDs. In the 366 tests with a variety of particle mixes, no
ignition and sustained burning occurred except one in test series 4. The results
of test series 6 indicate that the ignition in test series 4 was caused by the
presence of the machining imperfection. The test results combined with the
low likelihood that particles will be accumulated and simultaneously released
enabled NASA to conclude that the risk of particle impact ignition in the ISS
ECLSS QDs was acceptable.

References

关1兴 Benz, F. J., Williams, R. E., and Armstrong, D., “Ignition of Metals and Alloys by
High-Velocity Particles,” Flammability of Materials in Oxygen-Enriched Atmo-
spheres, Vol. 2 STP 910, M. A. Benning, Ed., ASTM International, West Consho-
hocken, PA, 1986.
关2兴 NASA-STD-6001A Test 15, “Flammability, Odor, Offgassing, and Compatibility Re-
quirements and Test Procedures for Materials in Environments that Support Com-
bustion,” NASA, Marshall Space Flight Center, Alabama, April 2008.
关3兴 Coombs, C. F., Jr., Ireson, W. G., and Moss, R. Y., Handbook of Reliability Engineer-
ing and Management, 2nd ed., McGraw-Hill, New York, 1996.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102287
Available online at www.astm.org/JAI

Keisa R. Rosales1 and Joel M. Stoltzfus2

Advanced Crew Escape Suits „ACES… Particle


Impact Test

ABSTRACT: NASA Johnson Space Center requested NASA JSC White


Sands Test Facility to assist in determining the effects of impaired anodiza-
tion on aluminum parts in advanced crew escape suits 共ACES兲. Initial inves-
tigation indicated poor anodization could lead to an increased risk of particle
impact ignition, and a lack of data was prevalent for particle impact of bare
共unanodized兲 aluminum; therefore, particle impact tests were performed. A
total of 179 subsonic 共Test Series 1–3兲 and 60 supersonic 共Test Series 6兲
tests were performed with no ignition of the aluminum targets. Based on the
resulting test data and analysis, WSTF found no increased particle impact
hazard was present in the ACES equipment.
KEYWORDS: aluminum anodization, particle impact ignition testing,
advanced crew escape suit 共ACES兲

Introduction
NASA Johnson Space Center 共JSC兲 requested NASA JSC White Sands Test Fa-
cility 共WSTF兲 to assist in determining the effects of impaired anodization on
aluminum parts in the advanced crew escape suits 共ACES兲. The ACES equip-
ment has been used on the Orbiter at nominal pressures of 690 kPa 共100 psig兲
with a maximum pressure of 1,689 kPa 共245 psig兲. Previous particle impact test
results indicate that the hard 共Type III兲 anodized coating offers increased resis-
tance to particle impact ignition as compared to bare aluminum 关1,2兴. Due to
previous testing 关3,4兴, concerns were raised that poor anodization could lead to
an increased risk of particle impact ignition, and a lack of data was prevalent
for particle impact of bare 共unanodized兲 aluminum; therefore, particle impact
tests were performed. A total of 179 subsonic 共Test Series 1–3兲 and 60 super-
sonic 共Test Series 6兲 tests were performed at the most severe in-use condition

Manuscript received December 11, 2008; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
Chemical Engineer, NASA Test and Evaluation Contract, P.O. Box 20, Las Cruces, NM
88004.
2
Project Manager, NASA White Sands Test Facility, P.O. Box 20, Las Cruces, NM 88004.
Cite as: Rosales, K. R. and Stoltzfus, J. M., ⬙Advanced Crew Escape Suits 共ACES兲 Particle
Impact Test,⬙ J. ASTM Intl., Vol. 6, No. 10. doi:10.1520/JAI102287.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 246
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 247

with no ignition of the aluminum targets. This paper summarizes the test meth-
ods and results. It also highlights NASA’s approach to solving anomalies occur-
ring in active oxygen systems.

Objective

Testing was performed to evaluate whether aluminum test samples would be


subject to ignition and sustained burning in a given flow environment when
impacted by particulate. The effect of the impact at specified temperatures,
pressures, and flow rates was also evaluated.

Test Methods

Test variables were determined by the worst-case conditions that occur in the
components of concern. Because the components could be exposed to any de-
gree of particle impact velocity, subsonic and supersonic particle impact tests
were performed using a variety of particulate mixtures.

Materials
A variety of particulate mixtures 共powder and particles兲 was used for testing.
Powders consisted of 共1兲 commercially pure −100+ 325 mesh titanium grade II
powder manufactured by Advanced Specialty Metals, Inc. 共Nashua, New
Hampshire兲; 共2兲 stainless steel powder with a maximum size of 150 ␮m; and
共3兲 aluminum powder containing a large spectrum of particle dimensions. The
smallest particles were 1 – 2 ␮m in diameter, and the largest spherical particles
were 1 mm in diameter. Slivers were 4–5 mm in length. One 1 , 587-␮m diam-
eter spherical aluminum particle and one 500-␮m diameter spherical stainless
steel particle were also injected in the subsonic tests. For the tests with the
subsonic injector and supersonic nozzle configuration, the aluminum powder
was sifted to have a maximum size of 200 ␮m.
The target sample material was unanodized aluminum. This material cor-
responded to the impingement points in the ACES where a possible particle
impact hazard existed within the oxygen manifold 共Fig. 1兲, quick-disconnect
共Fig. 2兲, and g-suit controller 共Fig. 3兲 components. The subsonic target samples
were used in two different configurations. The standard subsonic targets were
90°, 1 , 524 ␮m 共0.060 in.兲 thick flat disks with holes drilled near the outside
diameter to allow flow though the target 共Fig. 4兲. In addition, subsonic targets
were configured to simulate 118° drill points in passages of ACES equipment
共Fig. 5兲. The standard supersonic targets were configured in a cup shape with
an outside diameter that allowed flow around the 1 , 524 ␮m 共0.060 in.兲 thick
target surface. The sample was press-fitted onto a copper sample holder 共Fig.
6兲. Before testing, the samples were prepared at WSTF, cleaned, and then
sealed in polypropylene bags until testing. From this point, the targets were
handled with latex gloves to maintain the cleanliness level.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
248 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Oxygen manifold.

FIG. 2—Quick-disconnect.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 249

FIG. 3—G-suit controller.

Procedures
For the subsonic particle impact tests, the particulate was loaded into the in-
jector, and the injector cap was then threaded onto the housing. The target
sample and orifice were then positioned on the end of the subsonic chamber
共Fig. 7兲. The test conditions were 1 , 689 kPa+ 345/ 0 kPa 共245+ 50/ −0 psig兲
and 32+ 43/ −18° C 共90+ 110/ −0 ° F兲 with a velocity of at least 61 m/s 共200 ft/s兲.
For the supersonic particle impact tests, which used the subsonic injector
with the supersonic nozzle, the particulate was loaded into the injector, and the
injector cap was threaded onto the housing. The target sample and copper
posts were then positioned at the end of the supersonic nozzle 共Fig. 8兲. The test
conditions were 1 , 689 kPa+ 345/ 0 kPa 共245+ 50/ −0 psig兲 and 32+ 43/
−18° C 共90+ 110/ −0 ° F兲 with a velocity of at least Mach 1.
After system preparation was complete, the test area was cleared of person-
nel and placed in RED 共no access allowed兲 status. A video camera was posi-
tioned to record any reaction visible at the end of the test fixture. Heated gas-
eous aviator’s breathing oxygen 共ABO兲-grade oxygen at test pressure was
allowed to flow until the desired temperature of the target sample was achieved
and the gas flow stabilized. Upon command, the particulate was injected into
the chamber. After evidence of impact, which was indicated by a flash, the
oxygen flow was terminated. The test system was allowed to vent down to
ambient pressure ⬃4 s after particle injection. To verify that particle impact
occurred, the target sample was visually inspected after each test. Each sample
was then individually bagged, labeled with test information, and kept with its
original bag containing the remaining sample material. The test computer
saved the test data and system data. The test pressure, test temperature, and
average flow meter reading were recorded in the laboratory test log book. Video
tapes of the reaction were recorded and stored. At the completion of test data
storage and sample inspection, the procedure was repeated.

Results and Discussion


For unanodized aluminum target samples with 5 mg of aluminum and stainless
mixture or worse, 239 tests 共Test Series 1, 2, 3, and 6兲 were performed in 100 %
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
250 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Standard subsonic flat target configuration.

oxygen. The effects of impact on the test samples at the given temperatures,
pressures, and flow rates were determined visually and characterized by three
categories
• No burn: No particle indentations or erosion is apparent on the target
surface;
• Particle burn: One or more particle indentations are apparent on the
target surface, including damage from erosion; and
• Target burn: A portion or the entire sample is consumed; the target is
often not recoverable.
During particle impact ignition testing, the parameters of concern were
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 251

FIG. 5—Subsonic 118° drill point target configuration.

temperature, pressure, particle composition, target, and configuration. The


variables of particle size, particle material, mass flow rate, temperature, and
pressure were determined by the worst-case conditions that occur in the com-
ponents.
Three series of subsonic particle impact tests were performed on 6061-T6
aluminum targets, and another three series of tests were performed on 6061-T6
aluminum targets using the subsonic injector with the supersonic nozzle to
inject particulate and reach near-supersonic velocities. A total of six test series
were performed. The particulate mixtures are described in Table 1, and the test
results are summarized in Table 2.

FIG. 6—Standard supersonic target configuration.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
252 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—WSTF subsonic particle impact test system.

Test Series 1—Checkout Tests


The target configuration used was the flat standard sample configuration. The
particulate mixture was ⬃10 mg of titanium particulate 共with a maximum size
of 150 ␮m兲, plus one 1 , 587-␮m diameter spherical aluminum particle and
one 500-␮m diameter spherical stainless steel particle. The gas stream veloci-
ties ranged from 109 to 145 m/s 共357 to 475 ft/s兲. A total of 59 tests were
performed with no ignition of the aluminum targets 共Figs. 9–11兲.

Test Series 2
The target configuration used was the flat standard sample configuration. The
particulate mixture was ⬃10 mg of a mixture of aluminum 共mostly 100 to

FIG. 8—Subsonic injector with supersonic nozzle.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Particulate mixtures used for testing.

Particulate Mixture Material Size and Configuration Amount

Downloaded/printed by
A Titanium ⱕ150 ␮m ⬃10 mg
Aluminum 1 , 587-␮m-dia spherical particle QTY-1
Stainless steel 500-␮m-dia spherical particle QTY-1
B Mixture of aluminum and Aluminum, mostly 100 to 200 ␮m
stainless steel particulate stainless steel, ⱕ150 ␮m ⬃10 mg
Aluminum 1 , 587-␮m-dia spherical particle QTY-1
Stainless steel 500-␮m-dia spherical particle QTY-1
Mixture of aluminum and Aluminum, mostly 100 to 200 ␮m
C stainless steel particulate stainless steel, ⱕ150 ␮m ⬃10 mg
Mixture of aluminum and Aluminum—ⱕ200 ␮m stainless steel,
D stainless steel particulate ⱕ150 ␮m ⬃10 mg
Mixture of aluminum and

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
1.25 mg of aluminum, ⱕ200 ␮m 3.75 mg of
E stainless steel particulate stainless steel, ⱕ150 ␮m 5 mg
Note: QTY= Quantity

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 253
TABLE 2—Particle impact tests results summary.a

Gas Stream Number


Particulate Velocity Number of Target

Downloaded/printed by
Test Series Test System Configuration Target Configuration Mixtureb 关m/s 共ft/s兲兴 of Tests Ignitions
Subsonic injector with
1 Checkout subsonic nozzle Flat standard subsonic sample A 109–145 共357–475兲 59 0
Subsonic injector with
2 subsonic nozzle Flat standard subsonic sample B 114–141 共375–464兲 60 0
Subsonic injector with
3 subsonic nozzle 118° drill point subsonic sample A 114–121 共375–397兲 60 0
Subsonic injector with
4 Checkout supersonic nozzle Standard supersonic sample C Mach 1 5 0
Subsonic injector with
5 Checkout supersonic nozzle Standard supersonic sample D Mach 1 11 0

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Subsonic injector with
6 supersonic nozzle Standard supersonic sample E Mach 1 60 0
a
Test pressure 1689 kPa 共245 psig兲; test temperature 32° C 共90° F兲 for all tests.
b
Particulate mixtures are described in Table 1.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


254 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 255

FIG. 9—Test Series 1, Particulate Mixture A.

200 ␮m兲 and stainless steel particulate 共with a maximum size of 150 ␮m兲,
plus one 1 , 587-␮m diameter spherical aluminum particle and one 500-␮m
diameter spherical stainless steel particle. The gas stream velocities ranged
from 114 to 141 m/s 共375 to 464 ft/s兲. A total of 60 tests were performed with no
ignition of the aluminum targets 共Figs. 12 and 13兲.

FIG. 10—Typical subsonic flat target test results of a particle burn with titanium 共Test
Series 1, Particulate Mixture A兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
256 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 11—Typical subsonic flat target test results of a no burn with titanium 共Test Series
1, Particulate Mixture A兲.

Test Series 3—Checkout Tests


The target configuration simulated the 118° drill points in passages of ACES
equipment. The particulate mixture was ⬃10 mg of a mixture of titanium pow-
der 共with a maximum size of 150 ␮m兲, plus one 1 , 587-␮m diameter spherical
aluminum particle and one 500-␮m diameter spherical stainless steel particle.
The gas stream velocities ranged from 114 to 121 m/s 共375 to 397 ft/s兲. A total of
60 tests were performed with no ignition of the aluminum targets 共Figs. 14–16兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 257

FIG. 12—Test Series 2, Particulate Mixture B.

Test Series 4—Checkout Tests


The target configuration used was the standard supersonic configuration. The
particulate mixture was ⬃10 mg of a mixture of aluminum 共mostly
100–200 ␮m兲 and stainless steel particulate 共with a maximum size of
150 ␮m兲. A total of five tests were performed with no ignition of the aluminum
targets 共Fig. 17兲.

Test Series 5—Checkout Tests


The target configuration used was the standard supersonic configuration. The
particulate mixture was ⬃10 mg of a mixture of aluminum 共with a maximum
size of 200 ␮m兲 and stainless steel particulate 共with a maximum size of
150 ␮m兲. Four of the eleven tests used a 6,000-series aluminum target instead
of the 6061-T6 aluminum targets. A total of eleven tests were performed with
no ignition of the aluminum targets 共Fig. 18兲.

Test Series 6
The target configuration used was the standard supersonic configuration. The
particulate mixture was a 5-mg mixture consisting of 1.25 mg of aluminum
共with a maximum size of 200 ␮m兲 and 3.75 mg of stainless steel 共with a maxi-
mum size of 150 ␮m兲. A total of 60 tests were performed with no ignition of
the aluminum targets 共Figs. 19 and 20兲.

Statistical Analysis
The binomial distribution is used to describe results from tests of a “pass/fail”
or “go/no-go” nature, like those obtained from the particle impact test. From
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
258 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 13—Typical subsonic flat target test results of a no burn with aluminum 共Test
Series 2, Particulate Mixture B兲.

the number of ignitions that occur in a certain number of independent particle


impact tests, upper and lower confidence limits for reliability 关5兴 can be calcu-
lated. The upper and lower limits for reliability can be used to predict the range
of probabilities that an ignition would occur if the unanodized aluminum parts
in the ACES were to be subjected to an ignition source as severe as those used
in the tests. Obviously, if 0 ignitions occur in a certain number of particle
impact tests, then the upper confidence limit for reliability is 100 %, predicting
that, in the best-case scenario, 0 ignitions will occur if the unanodized alumi-
num parts are exposed to an ignition source as severe as the one used in the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 259

FIG. 14—Test Series 3, Particulate Mixture A.

tests. On the other hand, the lower confidence limit for reliability will be some
value less than 100 %, predicting, in the worst-case scenario, that no more than
some maximum number of ignitions may occur if the unanodized aluminum
parts are exposed to an ignition source as severe as the one used in the tests.
For example: If the lower confidence limit for reliability is 97 %, then this
indicates that no more than 3 % of the instances in which the unanodized
aluminum parts in the ACES are subjected to particle impact at the conditions
equal to those used in the test may yield ignitions. It is the lower confidence
limit for reliability that is of particular interest to the user of the hardware and
it can be used, as will be demonstrated next, to evaluate the risk of fire occur-
ring in the ACES.
The lower confidence limit for reliability, or maximum probability of an
ignition occurring in the ACES hardware, is a function of at least two param-
eters: The confidence level chosen for the analysis and the number of tests
performed. For this analysis, the writers selected a 99 % confidence level be-
cause that value enables one to perform a reasonable cost effective number of
tests and still make an acceptably confident decision. The number of tests per-
formed varied from 59 to 60 in Test Series 1 共checkout兲, 2, 3, and 6. It is obvious
that the greater the number of tests performed without ignition, the lower the
predicted maximum probability will be that particle impact ignitions may
occur in the unanodized aluminum parts in the ACES.
Applying these principles to the particle impact tests described herein, it
was determined for Test Series 1 共checkout兲, 2, 3, and 6, in which 0 ignitions
occurred in 59 or 60 tests, the predicted maximum probability of an ignition
occurring in the unanodized ACES parts, if they are impinged by particles at
these various conditions, is 5 % 共using a 99 % confidence level兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
260 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 15—Typical subsonic drill point target test results of a no burn with titanium 共Test
Series 3, Particulate Mixture A兲.

Determination of Acceptable Risk


To evaluate the overall risk of particle impact ignition occurring in the unan-
odized aluminum ACES parts, it is necessary to consider three factors: 共1兲 The
maximum probability of ignition in the parts as predicted by the analysis of the
test results discussed in the previous paragraphs; 共2兲 the likelihood that the type
and quantity of contaminant used in the tests will accumulate upstream of the
ACES; and 共3兲 the likelihood that the collected contaminant will be simulta-
neously released into the flowing oxygen and impact the ACES hardware.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 261

FIG. 16—Typical subsonic drill point target test results of a particle burn with titanium
共Test Series 3, Particulate Mixture A兲.

These factors were evaluated as follows:


共1兲 The predicted maximum probability of ignition was determined to be
less than 5 % 共using a 99 % confidence level兲 for Test Series 1 共check-
out兲, 2, 3, and 6;
共2兲 The likelihood of the tested contaminant mixtures being accumulated
upstream of the ACES was determined to be very low due to 共a兲 the
initial cleanliness of the space shuttle oxygen supply system and 共b兲 the
fact that WSTF tested using quantities of contaminants that were sig-
nificantly greater than the maximum quantity of contaminant esti-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
262 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 17—Test Series 4, Particulate Mixture C.

mated to be present in the entire space shuttle oxygen supply system;


and
共3兲 The likelihood of the collected contaminant being simultaneously re-
leased was determined to be very unlikely due the low likelihood that
an oxygen system filter or some other component will fail and release
its contents into the downstream system.

FIG. 18—Test Series 5, Particulate Mixture D.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 263

FIG. 19—Test Series 6, Particulate Mixture E.

Since the maximum predicted probability of an ignition event occurring


was low, and the likelihood of accumulation and simultaneous release were
both expected to be very low, NASA determined that the risk of particle impact
ignition in the ACES was acceptable.

Conclusions

Six series of tests were performed on unanodized aluminum 6061-T6 targets at


subsonic and supersonic velocities to determine if a particle impact ignition
hazard existed in the ACES components. In the 255 tests 共Test Series 1–6兲 with
a variety of target and test system configurations and particulate mixtures, no
ignition and sustained burning of the aluminum occurred. Although particle
burns were on the samples, it is surmised that the residence time of the par-
ticles in all tests was not long enough for the fire to propagate to the aluminum.
These test results, along with the very low likelihood that the type and quantity
of contaminant used in the tests will accumulate upstream and be simulta-
neously released into the ACES, enabled NASA to conclude that the risk of
particle impact ignition in the unanodized aluminum parts of the ACES was
acceptable.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
264 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 20—Typical supersonic target test results of a no burn with aluminum 共Test Series
6, Particulate Mixture E兲.

References

关1兴 Forsyth, E. T. and Stoltzfus, J. M., “Ignition Resistance of Hard 共Type III兲 Anodized
Aluminum to Particle Impact,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou,
and T. A. Steinberg, Eds., ASTM International, West Conshohocken, PA, 1997.
关2兴 Bahk, S., Chavez, D., Emery, B., and Wilson, B. “Protecting Aluminum Alloy From
Particle-Impact Ignition with an Al2O3 Film,” 1992 Hypervelocity Impact Sympo-
sium, Institute for Advanced Technology, 1992.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
ROSALES AND STOLTZFUS, doi:10.1520/JAI102287 265

关3兴 Benz, F. J., Williams, R. E., and Armstrong, D., “Ignition of Metals and Alloys by
High-Velocity Particles,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Second Volume, ASTM STP 910, M. A. Benning, Ed., ASTM
International, West Conshohocken, PA, 1986, pp. 16–37.
关4兴 Stoltzfus, J. M., Homa, J. M., Williams, R. E., and Benz, F. J., “ASTM Committee
G-4 Metals Flammability Test Program: Data and Discussion,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres: Third Volume, ASTM STP
986, D. W. Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp.
46–52.
关5兴 Coombs, C. F., Jr., Ireson, W. G., and Moss, R. Y., 1996, Handbook of Reliability
Engineering and Management, 2nd ed., McGraw-Hill, New York.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102299
Available online at www.astm.org/JAI

Elliot T. Forsyth,1 Barry E. Newton,2 Gwenael J. A. Chiffoleau,3


and Brendan Brophy4

Good Practices for Avoiding Fires in Steel


Mill Oxygen Systems

ABSTRACT: Over time, oxygen has been used with increasing frequency in
the steel industry. Applications include basic oxygen furnaces, electric arc
furnaces, scrap melting, lancing operations, and slab-cutting operations. Be-
cause of the inherently “dirty” environment of most steel mills, avoiding con-
taminants in the oxygen systems can be challenging and numerous fires
have occurred due to contaminant-promoted ignition sources. This paper
presents many good practices that can be practically implemented to reduce
the risk of oxygen fires in typical steel mill oxygen system installations.
KEYWORDS: oxygen, fires, steel mill, good practices, oxygen
cleaning, design, cutting torch, oxygen lance

Background
Oxygen is used extensively throughout the metals-refining and steel-making
industry to enhance melting and burning of metals in furnace applications or
to aid in cutting and lancing operations. In general, steel-making is an inher-
ently “dirty” process where environmental contaminants are ubiquitous. Many
of these contaminants are flammable in oxygen and can be easily ignited if
accidentally entrained in steel mill oxygen delivery systems. In fact, as the use
of oxygen in the steel-making industry has increased, several oxygen system

Manuscript received January 21, 2009; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
Technical Consultant, Oxygen Safety Consultants, Inc., Wendell Hull & Associates, Inc.,
Tulsa, OK 74104.
2
VP R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM
88007.
3
Senior Scientist, Test Facility Manager, Wendell Hull & Associates, Inc., 5605 Dona Ana
Rd., Las Cruces, NM 88007.
4
Process Engineer, North Star BlueScope Steel, 6767 County Rd. #9, Delta, OH 43515-
0128.
Cite as: Forsyth, E. T., Newton, B. E., Chiffoleau, G. J. A. and Brophy, B., ⬙Good
Practices for Avoiding Fires in Steel Mill Oxygen Systems,⬙ J. ASTM Intl., Vol. 6, No. 10.
doi:10.1520/JAI102299.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 266
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 267

fires have occurred. These fires substantiate the need for implementing good
practices in all oxygen system design, operations, and maintenance activities,
and for sound training in oxygen safety.

Typical Oxygen Use in Steel Mills


Due to the large quantity of oxygen used in steel mills, gaseous oxygen is cus-
tomarily delivered through a pipeline by a contracted gas supplier. Typically, an
underground pipeline will deliver oxygen to a central location at the mill; for
example, the pipeline could surface in the furnace building where it will be
fitted with a main isolation valve. The main supply pipeline is generally a large-
diameter carbon steel pipe, ND 300 共300 mm兲 12 NPS 共11 3/4 in.兲 nominal for
example, which operates at a maximum pressure between 15 bar 共220 psi兲 and
35 bar 共500 psi兲. From the header pipeline, there may be multiple smaller
branch lines 共also typically carbon steel兲 that then distribute oxygen to various
areas of use around the plant. The branch distribution lines are often located in
a pipe rack that runs along the facility ceiling and drop down to the specific
applications, as required.
For furnace applications, the oxygen distribution piping to the furnace gen-
erally includes an important series of components located on a “valve train,”
where the pressure is reduced and flow is controlled to the furnace burners.
The valve train typically has a series of isolation valves used for maintenance
purposes or emergency shut-down. It also contains a strainer, one or more
flow-control valves, flow transmitters, and other instrumentation used for con-
trolling oxygen flow to the furnace. Though the supply piping to the valve train
is often made of carbon steel, many of the valve train components are made of
more burn-resistant metals. Due to their operational severity, the metals used
are typically stainless steel, copper-nickel alloys like Monel, or brass. In this
more-severe application of the valve train, components typically use fluorinated
non-metal seats and seals such as PTFE, PCTFE, and FKM. However, some
hydrocarbon-based non-metals such as Buna-N or EPDM have been used in
more benign applications or highly protected configurations within compo-
nents.
For cutting or lance applications, oxygen is generally supplied to floor-level
cutting torches or lances through a service drop. The service drop typically
consists of a smaller diameter carbon steel pipe 共25 mm 关1 in.兴 or smaller兲 that
drops to the floor-level from the distribution header on the facility ceiling. The
service drop piping typically includes isolation valves, a regulator, and a pres-
sure gauge. A typical rubber welding hose then connects the piping directly to
a cutting torch or lance holder. Since the service drops often employ welding
regulators and other components common to the welding industry, they are
often made of brass or chrome-plated brass but occasionally are made from
stainless steel.

Oxygen Hazards in Steel Mill Oxygen Systems


As previously mentioned, oxygen systems in steel mills must operate success-
fully in an inherently contaminant-filled environment. Contaminants are per-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
268 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Photo of exterior contamination on steel mill oxygen piping.

haps the leading cause of fires in steel mill oxygen systems. Sources of both
metal and non-metal contaminants in steel mills can include iron or iron oxide
particulate from a carbon steel supply and distribution pipeline, lime and/or
coke dust from the raw materials used in blast furnaces to make molten iron,
oils and greases used in heavy machinery, and innumerable contaminants from
steel recycling in electric arc furnaces. Special measures such as those dis-
cussed in this paper must be taken in steel mills to minimize both metal and
non-metal contaminants in oxygen delivery systems. If these special measures
and safeguards are not implemented, external contamination will almost cer-
tainly infiltrate a steel mill’s oxygen pipeline. Figure 1 shows an example of the
high level of external contamination that can settle on steel mill oxygen piping
and components. This can be typical of installations proximate to blast or elec-
tric arc furnaces. Many of these contaminants are considered flammable and
easily ignited in oxygen. If they become entrained in the oxygen systems, the
risk of ignition and fire increases, especially where ignition mechanisms such
as compression heating, particle impact, and flow friction are active 关1兴. Igni-
tion of other fuels like the metals and non-metals of construction, can also
occur if a system is poorly designed or if good operations and maintenance
practices are not implemented.

Fire History in Steel Mill Oxygen Systems


Numerous fires have occurred in steel mill oxygen systems over the years.
Listed here are three examples of fires which underscore the importance of
implementing good practices in steel mill oxygen systems. The first incident
occurred at a stainless steel melt shop in Tornio, Finland, in 2003, as reported
by Lautkaski 关2兴. Two foremen were attempting to start a 35 bar 共500 psig兲
oxygen system by manually opening the ND 300 共NPS 12兲 main isolation but-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 269

terfly valve. When they had difficulty opening the valve manually, they used
pipe tongs to apply added torque to the hand wheel shaft. After several at-
tempts, the valve shaft turned approximately one revolution 共about 9° open兲
and a violent fire erupted, killing both foremen and a welder who had stopped
by to help. It was reported that the fire continued burning for several minutes
until the oxygen source was isolated. Lautkaski identified two possible causes
for the incident: Frictional heating between the butterfly disk and valve body, or
particle impact ignition. It is noteworthy that in this incident, a bypass valve
around the isolation valve was found to be in the closed position, indicating it
was not used to equalize pressure prior to opening. Had the pressure been
equalized around the valve prior to opening, which is a common procedure for
minimizing the severity of opening large-diameter isolation valves in oxygen,
this incident may have been prevented.
A second example occurred in a steel mill in Ohio in 2003 after a new
branch pipe was installed just upstream of a valve train. The piping was
cleaned for oxygen service and a nitrogen purge had been performed after
installation to mobilize particulates into a downstream “Y” strainer. When the
upstream isolation valve was opened under approximately 14 bar 共200 psig兲, a
fire erupted in the strainer. The fire burned out the stainless steel strainer body
and kindled ignition to surrounding stainless steel and carbon steel piping and
components. Figure 2 shows the Y strainer and associated piping both before
and after the incident. Nobody was hurt in the incident, but significant damage
was sustained. There was question as to whether or not the purge had adequate
flow and also whether the strainer should have been removed and cleaned after
the purge, prior to bringing oxygen on-line, This fire illustrates how Y strainers
can concentrate contaminants in a “sump” and present an ignition hazard in
oxygen if not properly used or maintained. It is noteworthy that if the strainer
had been serviced after the purge, or perhaps a different purge and start-up
protocol had been followed, this incident may not have occurred.
Perhaps the most common steel mill oxygen fires are associated with cut-
ting and lancing equipment. In rolling mills, for example, where flat rolled steel
is produced, oxygen lances are used to quickly cut and recycle large slabs of
steel that do not meet dimensional specifications. Oxygen lances, commonly
called burning bars, are sections of carbon steel pipe, sometimes stuffed with
steel wire and typically 3–4 m in length. They are ignited by an oxygen torch
and used to quickly cut thick cross-sections of metal. In 2005, a steel mill
employee was using a long lance to cut such a slab when a burn-out occurred
through the lance holder. The burn-out caused second and third-degree burns
to the workman’s hand and forearm which were supporting the bar down-
stream of the holder 共Fig. 3兲. Oxygen fires in lance holders can have multiple
causes, but a common contributing factor is contaminant in the form of small
metal fines or oils or greases that reside at the tip of the bar inside the holder
where it seals on an elastomer grommet. This fire, and many others like it, may
have been prevented had a thorough visual inspection for contaminants been
performed prior to assembly and had the manufacturer’s instructions for use,
including proper personal protective equipment, been closely followed.
As described earlier, these examples of oxygen fires in steel mills serve to
underscore the importance of understanding and implementing good practices.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
270 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Photo of valve train “Y” strainer pre-incident 共top兲 and post-incident 共bottom兲.

In each of these cases, the incident may have been prevented had the user been
more fully aware of good practices regarding design, operations, and mainte-
nance. Specific training on these good practices is important, too, to ensure
they are understood and implemented.

Applicable ASTM Standards and Training


The ASTM International Committee G-45 has developed and published several
standard guides and practices that are directly applicable to oxygen systems
used in the steel industry. For instance, ASTM G88 is a standard guide for

5
ASTM International Committee G-4 on the “Compatibility and Sensitivity of Materials
in Oxygen-Enriched Atmospheres,” ASTM, West Conshohocken, PA.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 271

FIG. 3—共a兲 Incident lance holder with shut-off valve; 共b兲 incident lance holder close-up;
共c兲 incident lance holder burn-through; and 共d兲 glove from incident.

designing oxygen systems and provides design factors that minimize the system
severity. Additionally, ASTM G63 and G94 provide evaluation criteria for the
selection of both non-metals and metals, respectively, for oxygen service. Fi-
nally, ASTM G93 provides a background on the need to minimize contaminants
in oxygen system and several methodologies for cleaning and inspecting oxygen
components and systems.
ASTM also provides technical and professional training in the areas of oxy-
gen system fire hazards and system operations and maintenance. Engineers at
Wendell Hull & Associates, Inc. and Oxygen Safety Consultants, Inc. help in-
struct these courses and have developed custom-tailored oxygen safety training
courses for the steel industry. One such training course was developed for
North Star BlueScope Steel and several of the good practices taught in the
training course are discussed next.

General Good Practices for Steel Mill Oxygen Systems


Given the fire hazards in oxygen systems in steel mills, several good practices
are discussed below. If implemented, these practices can minimize the severity
and risk of fire in these systems. The practices discussed include oxygen system
cleaning, design, and specific practices for cutting and lancing operations.

Good Practices in Oxygen System Cleaning


As mentioned earlier, due to the contaminant-rich environment of most steel
mills, good practices regarding system cleanliness are critically important.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
272 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ASTM G93 advocates an approach to oxygen system cleaning that is consistent


with Compressed Gas Association 共CGA兲 G-4.16 and European Industrial Gases
Association 共EIGA兲 Industrial Gases Council 共IGC兲 33-06,7 which has essentially
four main principles: Start clean, verify clean, assemble clean, and maintain
clean. Each of these four elements has its own good practices that can apply to
steel mill oxygen systems.

“Start Clean”
Starting clean, using an approved cleaning process, suggests that all oxygen
systems are cleaned to a known level before assembly. Starting with a clean
oxygen system in a steel mill is critical in order to avoid ignition of oil films and
greases by adiabatic compression or contaminant-promoted ignition, as well as
to avoid ignition of metallic particles by particle impact.
Table 1 provides a comparison between the maximum level of non-volatile
residue 共NVR兲 and particulate allowed by various cleaning standards. Steel mill
oxygen systems generally follow the cleaning levels prescribed in CGA G-4.1, or
IGC 33-06, which are oriented more specifically for industrial oxygen systems.
As a broad generalization, industrial gas applications use large pipe diameters,
have large volumes, and operate at lower pressures than other industries that
use oxygen. As such, they can generally tolerate more contaminant than
smaller, high-pressure oxygen systems such as aerospace or medical oxygen.,
This is reflected in the suggested cleaning levels in the standards for industrial
oxygen systems. Comparing NVR values, CGA G-4.1 recommends a maximum
allowable NVR of 500 mg/ m2 whereas IGC 33-06 differentiates based upon
pressure, allowing a 500 mg/ m2 maximum NVR for lower pressure systems
and, recognizing an increased risk of ignition at higher pressures, only allows a
maximum 200 mg/ m2 for pressures greater than 30 bar. ASTM G93 refrains
from giving specific recommendations, but shows NVR levels of between
11 mg/ m2 and 550 mg/ m2 as being typical of oxygen systems in industry.
One concern in oxygen systems that gives context to these NVR levels is oil
film migration. Data from oil film migration studies show that oil films in quan-
tities larger than approximately 215 mg/ m2 will migrate on smooth surfaces
关3兴. Thus, NVR acceptance levels greater than this would be expected to have
potential for oil migration in a given oxygen system, possibly leading to higher
concentrations of oil at a dead-end of cavity within a system. Thus, for systems
where oilfilm migration and concentration presents a risk of ignition, accep-
tance criteria for NVR is preferably set below 215 mg/ m2.
Several studies have been performed demonstrating the ignition of oil films
in oxygen service. Using a common ignition mechanism, adiabatic compres-
sion, studies have shown oil film ignition down to 65 mg/ m2 at pressures
down to 6.9 MPa 共1,000 psig兲 关4兴. However, in low-pressure oxygen systems,
data show ignition by adiabatic compression did not occur at 3.4 MPa 共500

6
“Cleaning Equipment for Oxygen Service,” G-4.1, Edition 5, Compressed Gas Associa-
tion, New York, 2004.
7
“Cleaning of Equipment for Oxygen Service,” Guideline IGC Doc 33/06/E, European
Industrial Gases Association, Avenue des Arts 305, B-1210 Brussels.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Acceptance criteria for particulate and NVR from various guides.

Standard Max NVR Max Particulate

Downloaded/printed by
CGA G-4.1 500 mg/ m2a Only if required—1,000 ␮ max,
ⱕ215 particles/ m2 b / n 500– 1,000 ␮
EIGA 33/06 For P ⬍ 30 bar, 500 mg/ m2b Single small chips or fibers tolerated
For P ⬎ 30 bar, 200 mg/ m2 Single very small chips or fibers tolerated
ASTM G93 No recommendations given, but No recommendations given, but specifications
specifications range from Level A to E, range from Level 175 to 500, corresponding to
corresponding to 11 mg/ m2 to 550 mg/ m2, no particles larger than 175– 500 ␮m, or level
or Level F being user specified as user specified
a
CGA G-4.1 states this value “could be more or less depending on specific application 共state of fluid temperature, and pressure兲.”
b
EIGA 33/06 also notes that no drops of water can be visible. Furthermore, it states this as a guide only and that “lower figures could
be requested depending on the specific application 共type of fluid, temperature, pressure, flow, velocity, product purity兲, or effects like
migration.”

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


FORSYTH ET AL., doi:10.1520/JAI102299 273
274 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

psig兲 even with contaminant levels as high as 1 , 080 mg/ m2 关5兴. These data do
not suggest that low-pressure oxygen systems, such as in steel mills, do not
require cleaning. They do indicate, however, that for the most common ignition
mechanisms for oil films in oxygen systems, namely compression heating and
flow friction, lower-pressure systems are more tolerant of oil film contamina-
tion. This is likely the reason that IGC 33-06 allows greater NVR levels at lower
pressures. Since few steel mill oxygen systems operate with a delivery pressure
greater than 30 bar, a 500 mg/ m2 NVR limit would be considered a common
upper bound for allowance of NVR for these systems.
Data for propagation of oil film ignition to metals further support this
guidance. Generally, high levels of hydrocarbon contaminants, active ignition
mechanisms, and elevated pressures are required to directly promote ignition
to bulk metals with oils. For example, a large quantity 共0.5 g兲 of Mobil 600W oil
was ignited by an electrical igniter, providing an energy release of approxi-
mately 10,000 calories to a 3.175-mm 共1/8 in.兲 rod. Carbon steel rods ignited
and burned at pressures down to 6.9 MPa 共1,000 psig兲 but stainless steel rods
did not ignite and burn even up to 34.5 MPa 共5,000 psig兲 关6兴.
These data suggest that in low-pressure oxygen systems, as found in most
steel mills, it is difficult to directly ignite bulk metals from hydrocarbon oil
sources. Under these conditions, ignition of bulk metals requires at minimum a
high ignition energy, or perhaps more commonly, another contaminant to form
a kindling chain. For example, oil films may be capable of igniting small metal
particulate which then form a kindling chain capable of igniting bulk metal
alloys. The gaseous combustion product of oil films may be sufficient to ignite
metal particulate, which then, due to their liquid slag combustion product, are
more effective at kindling ignition to bulk metal alloys. This is often the case
with fires involving bulk metals in steel mill oxygen systems. Thus, minimizing
particulate contaminant in steel mill oxygen systems is important to minimiz-
ing promoted ignition hazards. The oxygen cleaning standards shown in Table
1 all require minimizing particulate levels, as discussed next.

“Verify Clean”
The second aspect to oxygen cleaning is to verify that systems are clean prior to
assembly. This can be done both quantitatively and qualitatively. The burden of
quantitative clean verification is generally placed on the supplier of oxygen-
clean equipment, either by validating their cleaning process or by verifying the
actual clean level of every piece being cleaned. Verification that oxygen compo-
nents have been cleaned prior to assembly is important. For this reason, any
component that is not “bagged and tagged” as “oxygen clean” should not be
installed in an oxygen system unless special approval is given. By contrast,
qualitative clean verification should be performed by all end-users of oxygen
equipment as it involves visually inspecting for contamination prior to assem-
bly. It is important to remember that anyone performing assembly is the last
line of defense in assuring cleanliness as they are the last ones to see a compo-
nent before it is installed.
Visual inspection techniques using white and black 共ultraviolet, UV兲 light
can be effective for detecting large quantities of oil films and other NVR and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 275

materials. UV light causes many common hydrocarbon-based materials, or-


ganic oils and greases, and lint to fluoresce and thus be detectable. Approxi-
mate NVR detection limits for common visual inspection test methods used in
industrial gas applications were studied by Gilbertson and Lowrie 关7兴, and are
also published in IGC 33-06. That data show that UV light can be up to ten
times more effective at detecting some oil films than bright white light. Again,
it is important to note that visual inspection techniques provide only qualitative
results of contaminant levels, not quantitative results. Furthermore, the effec-
tiveness of UV light detection depends on the fluorescing characteristics of the
contaminant. Some contaminants, such as vegetable and some chemical oils,
do not fluoresce at all. For this reason, UV inspection alone cannot be relied on
as a test for cleanliness. Visual inspection techniques using white light are
typically effective for detecting particulate down to the visible limit, which is
approximately 40– 50 ␮m for personnel with 20/20 vision. Thus, for steel mill
applications, a combination of white and black light visual inspection used
liberally throughout a plant, but especially before any assembly is performed, is
an excellent practice for avoiding contaminant-promoted ignition. The philoso-
phy communicated by most steel mills to their personnel doing white and black
light visual inspection is simple, yet highly effective: If you can see contaminant
共oils, greases, particulates, etc. on the surfaces being inspected兲, it’s not clean. If
you can’t see contaminant, then you assume the approved cleaning process
worked.

“Assemble Clean”
Assemble clean means incorporating a host of good practices that will protect
the clean surfaces and open ports of piping and components during the assem-
bly process. This includes practices such as not allowing personnel to touch
oxygen-clean surfaces with bare hands or leather work gloves. Instead, if con-
tact with these surfaces is required, personnel can use clean, powder-free latex
or nitrile gloves combined with good assembly techniques to avoid oil film
contamination and particulate generation. Assemble clean also includes con-
cepts such as assembling systems in a “clean” area, as much as is practical. For
steel mills, being cognizant of environmental contaminants is critical during
assembly. Oxygen systems should avoid any assembly outdoors if the weather is
poor, unless a tent can be build to keep weather or other environmental con-
taminants out. Furthermore, it is generally preferred to assemble critical com-
ponent sections 共for a valve train, for example兲 in a cleaner environment first
共such as a warehouse, as opposed to the furnace building兲, then transport them
as an assembly for installation on the header. Finally, assemble clean includes
using good assembly practices to avoid generating contaminant during the as-
sembly process. For example, when assembling flange joints, ensure that pipe-
fitters minimize the friction or rubbing between the two flanges, which could
generate metal and non-metal particles.

“Maintain Clean”
And finally, maintain clean includes concepts like being aware of external con-
tamination on piping and components before opening a system for mainte-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
276 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Steel mill oxygen lines without proper protection of open ports.

nance, or before performing leak checks with a liquid leak detectant. It also
includes protecting open ports from environmental contaminants by installing
oxygen-compatible plugs or caps, or using plastic bagging. This is especially
important to steel mill oxygen systems. Figure 4 shows a photo of two oxygen
lines that feed an electric arc furnace in a steel mill. This section of the system
was temporarily down for maintenance and the oxygen flex-lines were left open
to environmental contaminant rather than maintained clean through the
proper use of clean blind flanges. The concept of maintaining clean also means
to purge with clean, dry, inert gas after any assembly or maintenance that
requires breaking into the system. The purge must be done prior to start-up to
mobilize and blow out particulate before introducing oxygen.

Good Practices for Oxygen System Design


In addition to system cleaning, another general good practice for steel mill
oxygen systems is to implement good design practices. As previously discussed,
ASTM standard G88 provides much design guidance applicable for steel mill
systems. Within G88 is an entire section 共Section 5.2兲 devoted to factors recog-
nized as causing fires in oxygen systems, and another section devoted to design
methodology to minimize these factors 共Section 7兲. It is a good design practice
to ensure that a steel mill oxygen system complies with the guidance in ASTM
G88.
Another design resource for steel mill oxygen systems is CGA G-4.4 共2003兲,
Oxygen Pipeline Systems.8 This is a harmonized design standard between the

8
CGA G-4.4–2003 共4th Edition兲/IGC doc 13/02, Oxygen Pipeline
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Systems, Compressed Gas
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 277

CGA and the EIGA for industrial oxygen systems. It provides specific guidance
on the pressures and flow velocities that restrict the use of various metal alloys
for piping and components to protect against particle impact ignition. This is
done through a set of pressure-velocity curves that metal alloys must comply
with if they are being used above their published exemption pressure. The
exemption pressure is the maximum pressure not subject to velocity limitations
in high purity oxygen 共nominal 99.7 %兲 where particle impingement may occur.
As such, it is considered good practice for the materials selected in steel mill
oxygen systems to be consistent with the guidance from CGA G-4.4.

Main Isolation Valves


As previously discussed, consider the large-diameter carbon steel supply piping
and primary isolation valve located at the battery limits of most steel mill oxy-
gen sources. Figure 5 shows an example of such an isolation valve, in this case
a large-diameter wafer-style butterfly valve. If the supply pressure was 10 bar,
for example, then per CGA G-4.4 carbon steel must conform the to pressure-
velocity curve for impingement or non-impingement 共depending on the appli-
cation兲 because the supply pressure exceeds the exemption pressure for carbon
steel. Without reprinting the CGA curves here, essentially this means that, at
this pressure, the use of carbon steel is restricted only to low gas velocity ap-
plications 共less than 30 m/s wherever particle impingements may occur兲 with-
out special justification.9 Thus, if the valve in Fig. 5 were carbon steel, to be
compliant with CGA G-4.4 it could not be opened under a pressure differential
since the gas velocity would likely exceed the 30 m/s limit during the opening
flow transient. During this transient, particle impact would be an ignition con-
cern. By comparison, at 10 bar the use of bulk stainless steel is not restricted to
the velocity curve because the exemption pressure of stainless steel is 14 bar for
thicknesses greater than 3.18 mm 共0.125 in.兲. Thus, under these conditions, per
CGA, stainless steel is allowed for transient high gas velocity applications like
opening an isolation valve under a pressure differential, or even for sustained
high gas velocity application like a control valve.
If carbon steel was highly preferred for the isolation valve in Fig. 5 共which
may be the case for economic reasons in large-diameter valves兲, then the par-
ticle impact ignition risk could be minimized by limiting gas velocity. One ap-
proach could be to install an oxygen-compatible bypass valve and equalize
pressure across the valve prior to opening. Castillo and Werley provided sound
guidance on the strategic use of bypass valves in oxygen service to comply with
CGA pressure-velocity criteria10 关8兴. This would allow the isolation valve to be
opened in a zero-flow environment, which is always less severe in terms of
ignition.
Figure 5 also provides an example of an oxygen flange joint in close prox-

Association, Inc., Chantilly, VA, European Industrial Gases Association, Brussels.


9
Such as an oxygen hazards analysis or compatibility assessment.
10
It is noteworthy that the CGA G-4.4 pressure-velocity curves changed with the 2003
publication and some of the data presented in this reference may be specific to the older
curves.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
278 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—Main oxygen isolation valve at supply battery limit.


imity to a flammable gas, in this case acetylene storage cylinders. It is consid-
ered good practice to avoid locating oxygen piping in the immediate proximity
of vulnerable areas and equipment such as flammable storage tanks or above-
ground flammable pipelines. Further, it is considered good practice to keep
mechanical joints, such as flanges, on oxygen piping away from similar joints
on above-ground flammable fluid piping to minimize the risk of fire in the event
of simultaneous leaks or failures.11

Service Drops
Most steel mills have an oxygen distribution system that includes a header pipe
along the ceiling with multiple service drops that supply oxygen, air, fuel gas,

11
CGA G-4.4-2003 共4th Edition兲/IGC doc 13/02, Oxygen Pipeline Systems, Compressed
Gas Association, Inc., Chantilly, VA, European Industrial Gases Association, Brussels.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 279

and other gases down to a cutting or lancing station. Good practices in the
design of these service drops are important to minimize fire hazards.
In general, it is preferred for service drop piping to branch off the top of the
header before dropping down to the service location rather than off the bottom.
This helps prevent particulate and other debris from directly falling into the
service drop and collecting on top of an isolation valve where it could be ignited
by particle impact upon opening. This is especially important where carbon
steel piping is used for the service drop. Filters can also be used upstream of
components that throttle flow 共i.e., regulators or valves opened to initiate oxy-
gen flow兲 to catch debris before it enters the high gas velocity component.
Another good practice is to avoid using ball valves for pressurization of
service drops. Much has been published regarding the oxygen hazards of ball
valves, specifically their fast-opening characteristics leading to adiabatic com-
pression ignition of non-metals 关9兴. Nonetheless, ball valves can be desirable in
applications where quick shut-off 共i.e., emergency shut-off兲 or high flow capac-
ity is required. Thus, if ball valves are used, personnel must first confirm that
they are specified and cleaned for oxygen service, as would be true for any
component. If carbon steel is used for the valve body or trim, personnel should
avoid opening the valve under pressure differentials to minimize particle im-
pact ignition. Personnel should also avoid placing them directly upstream of
any component that has exposed non-metals at a “dead-end” 共i.e., valve in
closed position, flexible hose with non-metal liner兲 or highly restricted flow
共i.e., regulator兲 where rapid pressurization could cause ignition 共ASTM G88,
Section 7.7兲. This practice is especially important in components where hydro-
carbon oils or greases may have migrated to the dead-end location.
In addition, service drops should use different styles of fittings for the dif-
ferent gases available in order to minimize the likelihood of connecting an air
tool to an oxygen line, or an oxygen component to a fuel line. This practice is
consistent with the outlet connections of welding regulators and mating weld-
ing hoses where the threads are always left-handed for fuel gas and right-
handed for oxygen. Fuel gas connections often have a “V” notch around the
inlet nut to further designate the connection.
Finally, service drops can be prime sources for the ingress of contaminants.
Open ports should always be protected from environmental contaminants
when equipment is not in use. Prior to use, equipment should always be in-
spected for contaminants.

Good Practices for Cutting Torches and Lances

The following good practices relate to the general use of cutting torches and
oxygen lances used for cutting large metal slabs or opening holes in pouring
tubes for example where molten metal has solidified. These good practices are
examples only and do not represent a complete list of good practices for these
devices. The equipment manufacturer should always be consulted for a com-
plete list of good practices for that equipment. Another source of information
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
280 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

on the safe use of oxygen for these applications can be found in NFPA 51:12
Standard for the design and installation of oxygen-fuel gas systems for welding,
cutting, and allied processes.
First, for cutting torches and lances, use only an industrial-grade welding
hose specified for the application. These are usually constructed of layers of
rubber or neoprene over a braided inner section and color coded green for
oxygen service and red for fuel gas service. Even though these hose materials
are flammable, they are designed to not support a flame once the heat source is
removed. Hoses should be properly cared for. For example, protect hoses from
any falling metal, slag or sparks during use, and from excessive contact with
oils, greases, and dirt, especially at open ends. Use of a spool is a good practice
to protect hoses from these hazards. Figure 6 shows a steel mill lance hose that
was left lying in a pool of grease as compared to one which had been properly
spooled when not in use. Before using any hose for welding or cutting, it is
important to visually inspect the hose for cuts, worn areas, or other damages. If
these characteristics are observed, the hose should be replaced.
The importance of “starting clean” to avoid the hazards of oil films and
particulates has been addressed in this paper. Thus, for cutting torches and
lance equipment, visually inspecting for contaminants prior to use is another
good practice. This includes inspecting external surfaces, especially proximate
to common ingress points like threaded connections or other fittings. If exter-
nal contaminants are observed on this equipment, such as on the cutting torch
handle in Fig. 7, oftentimes they can be easily removed with a solvent cleaner.
For oxygen cutting equipment used around grease-prone areas, such as in a
rolling mill, using an approved oxygen cleaner in an aerosol form can be a good
practice. The objective is to maintain all oxygen hardware in a visually clean
condition to minimize the likelihood of contaminating oxygen-wetted surfaces.
This guidance also applies to burning bars, which might be supplied in a visu-
ally clean condition from the vendor, but are often stored outside or in a
contaminant-prone area. Prior to assembly into the holder, it is considered
good practice to visually inspect the end of the bar that inserts into the holder
to ensure no oils, greases, or excessive particulate are present. If ignition of
contaminants occurs in the lance holder, a burnout may occur similar to that
shown in Fig. 3. The cleanliness of the opposite end of the burning bar is less
critical since it will be ignited and consumed in use. To maintain the visually
clean condition of burning bars as supplied by the vendor, they should be
stored in a clean area and protected from environmental contamination. If the
ends of the bars are wrapped, do not remove the wrap until ready for use.
Since oxygen leaks can cause localized oxygen-enrichment and lead to ig-
nition in certain configurations, it is considered good practice to leak check all
cutting torch and lance connections using a leak detectant approved for oxygen
service. Some steel mills have implemented a site-wide policy that leak checks
be performed before every oxygen cutting or lancing operation and even in-
clude a bottle of leak detectant at every service drop. Important locations to

12
National Fire Protection Association 共NFPA兲, 1 Batterymarch Park, Quincy, MA
02169–7471.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 281

FIG. 6—Steel mill lance hose lying in grease 共top兲 versus hung on a hose spool 共bottom兲.

check for leaks include the regulator-to-hose connection fitting and the hose-
to-torch or hose-to-lance holder connection. Also, since some lance fires have
been attributed to leaks past the grommet at the burning bar-lance holder in-
terface, leak checks at this location are also considered a good practice.
Another good practice for cutting torches and lance holders is to vent the
oxygen pressure back to an upstream isolation valve and fully reduce the regu-
lator 共if applicable兲 when the torch or lance is not in service. This practice not
only can increase the service life of component seats and seals, but also protects
personnel in the event that the equipment leaks or the cutting valve is inadvert-
ently opened to release oxygen. Thus, the sequence of operation each time the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
282 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Hydrocarbon grease contaminant on oxygen cutting torch handle.

equipment is used after being stored is to first open the isolation valve, then set
the regulator delivery pressure, and finally operate the torch or lance holder per
manufacture guidance.
One good practice specific to oxygen cutting torches is to implement a
periodic functional check on torch handle check valves, as recommended by
the manufacturer. Check valves are important components on cutting torches
as they provide a certain amount of protection against the backflow of oxygen
into the fuel gas hose and regulator, or vice versa, which can create a combus-
tion hazard. Check valves commonly fail in the event of a flashback, yet the
failure may not be detectable during normal operation. Thus, it is considered
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102299 283

good practice to test or replace check valves at regularly scheduled intervals. It


is also important to ensure all oxy-fuel cutting and welding applications use
flashback arrestors per manufacturer guidelines to help stop a flashback event
from propagating back into the upstream hose or regulator. Finally, another
good practice is to visually inspect the cutting tip of oxygen cutting torches
before each use. Cutting tips are subject to much abuse as molten metal can
clog or obstruct the small passages and restrict flow, which can lead to a flash-
back event. Most cutting tips can be cleaned with small round files per manu-
facture guidelines if clogged holes are observed.

Summary
Over time, the steel-making industry has used oxygen for both production and
cutting and lancing applications with increasing frequency. Because oils,
greases, particulates, and other contaminants are inherent to the industry,
avoiding contaminants in steel mill oxygen systems can be challenging and
numerous fires have occurred due to contaminant-promoted ignition sources.
Implementing several good practices can help avoid ignition in steel mill oxy-
gen systems. These include good practices in oxygen system cleaning such as
“Start clean, verify clean, assemble clean, and maintain clean,” each with its
own unique good practices to minimize the risk of contaminant-promoted ig-
nition. Implementing good design practices for steel mill oxygen systems is also
important, including proper design of main isolation valves and service drops.
ASTM G88 and CGA G-4.4 can provide helpful guidance for these systems.
Finally, good operating and maintenance practices for cutting torches and
lances are important, including the proper use and storage of hoses, visual
inspection of equipment for contaminants, and performing leak checks and
functional checks.

References

关1兴 Beeson, H. D., Stewart, W. F., and Woods, S. S., “Safe Use of Oxygen and Oxygen
Systems: Guidelines for Oxygen System Design, Materials Selection, Operations,
Storage, and Transportation,” ASTM Manual 36, ASTM International, West Con-
shohocken, PA, 2000.
关2兴 Lautkaski, Risto, “Investigation of a Large Industrial Oxygen Valve Fire,” J. Loss
Prev. Process Ind., Vol. 21共4兲, July 2008, pp. 466–471.
关3兴 Werley, B. L., “Oil Film Hazards in Oxygen Systems,” Flammability and Sensitivity
of Materials in Oxygen-Enriched Atmospheres, ASTM STP 812, B. L. Werley, Ed.,
ASTM International, West Conshohocken, PA, 1983, pp. 108–125.
关4兴 Pedley, M. D., Pao, J., Bamfod, L., Williams, R. E., and Plante, B., “Ignition of
Contaminants by Impact of High-Pressure Oxygen,” Flammability and Sensitivity
of Materials in Oxygen-Enriched Atmospheres, ASTM STP 986, D. W. Schroll, Ed.,
ASTM International, West Conshohocken, PA, 1987, pp. 305–317.
关5兴 Forsyth, E. T., Durkin, R. J., and Beeson, H. D., “Evaluation of Contaminant-
Promoted Ignition in Scuba Equipment and Breathing-Gas Delivery Systems,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
284 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

STP 1395, T. A. Steinberg, B. E. Newton, and H. D. Beeson, Eds., ASTM Interna-


tional, West Conshohocken, PA, 2000.
关6兴 McIlroy, K., Zawierucha, R., and Dmevich, R. F., “Promoted Ignition Behavior of
Engineering Alloys in High-Pressure Oxygen,” Flammability and Sensitivity of Ma-
terials in Oxygen-Enriched Atmospheres: Third Volume, ASTM STP 986, D. W.
Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp. 85–104.
关7兴 Gilbertson, J. A. and Lowrie, R., “Threshold Sensitivities of Teats to Detect Oil
Film Contamination in Oxygen Equipment,” Flammability and Sensitivity of Mate-
rials in Oxygen-Enriched Atmospheres: Second Volume, ASTM STP 910, M. A. Ben-
ning, Ed., ASTM International, West Conshohocken, PA, 1986, pp. 204–211.
关8兴 Castillo, D. G. and Werley, B. L., “Eliminating Bypass Valves in Selected Oxygen
Systems,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997.
关9兴 Koch, U. H., “Oxygen Piping Code-Where Knowledge Becomes Practice: Keynote
Address,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Eighth Volume, ASTM STP 1319, W. T. Royals, T. C. Chou, and T. A. Stein-
berg, Eds., ASTM International, West Conshohocken, PA, 1997.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102296
Available online at www.astm.org/JAI

Elliot T. Forsyth,1 Barry E. Newton,2 Gwenael J. A. Chiffoleau,3


and Brad Forsyth4

Oxygen Fire Hazards in Valve-Integrated


Pressure Regulators for Medical Oxygen

ABSTRACT: In recent years, medical oxygen regulators that incorporate a


cylinder isolation valve have increased in popularity. These devices, often
called valve-integrated pressure regulators 共VIPRs兲, essentially combine
several components into one highly compact manifold design, typically in-
cluding a fill valve, shut-off valve, residual pressure valve, regulator, and
relief valve. Combining these components into a single manifold block pre-
sents unique oxygen fire hazards that can differ greatly from those found in
the stand-alone versions of the components. In order to avoid fires, these
hazards must be understood and addressed in the design of the VIPR. This
paper presents the most common oxygen fire hazards found in VIPR devices
as well as the ASTM analysis and testing methods used to qualify a new
design, specifically considering oxygen compatibility of materials, toxicity of
combustion products, ignition mechanisms, and reaction effects of a fire.
KEYWORDS: oxygen hazards, fire risk, hazards analysis, valve-
integrated pressure regulators, VIPR, medical oxygen, ignition, toxicity

Introduction and Background


Portable medical oxygen is typically transported in high-pressure cylinders.
These cylinders either have a stand-alone valve or a valve-integrated with a

Manuscript received December 18, 2008; accepted for publication August 19, 2009; pub-
lished online September 2009.
1
Oxygen Safety and Forensic Engineer, Oxygen Safety Consultants, Inc., Tulsa, OK
74104.
2
VP R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM
88007.
3
Senior Scientist and Test Facility Manager, Wendell Hull & Associates, Inc., 5605 Dona
Ana Rd., Las Cruces, NM 88007.
4
Special Projects Engineer, Wendell Hull and Associates Inc., 5605 Dona Ana Rd., Las
Cruces, NM 88007.
Cite as: Forsyth, E. T., Newton, B. E., Chiffoleau, G. J. A. and Forsyth, B., ⬙Oxygen Fire
Hazards in Valve-Integrated Pressure Regulators for Medical Oxygen,⬙ J. ASTM Intl., Vol.
6, No. 10. doi:10.1520/JAI102296.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 285
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
286 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

pressure regulator 共VIPR兲. The stand-alone valve is used with a separate medi-
cal oxygen regulator to deliver near-atmospheric pressure oxygen at various
selectable flow rates. However, VIPRs integrate the valve and regulator into one
component to deliver the oxygen to the end user. Like stand-alone cylinder
valves, VIPRs are also the interface during cylinder filling. Due to the integra-
tion, VIPRs require the addition of a fill valve to the design, which is not
present in the cylinder valve design. Depending on regional requirements, extra
features of the cylinder valve and regulator are also integrated. These inte-
grated features include: Pressure gauge, burst disk, regulator pressure relief,
high-flow outlet, flow restrictor, and residual pressure valve 共RPV兲. The VIPR
design is relatively complex when compared to the individual components that
it integrates, and several variations of VIPR designs have surfaced on the mar-
ket.
Although the ratio of incidents when compared to the number of compo-
nents in service is extremely small, fires involving stand-alone components have
been well documented 关1兴. Ignition mechanisms such as compression heating,
particle impact, flow friction, and promoted ignition were associated with the
probable causes of these fires. These ignition mechanisms can also be active in
the VIPR design, but the integration of the stand-alone components into a
single manifold block presents unique oxygen fire hazards that can differ
greatly from those found in the stand-alone versions of the components. A
proper assessment of the hazards and fire risk is critical to avoiding fires and
ensuring the safety of personnel and equipment. This paper presents the most
common oxygen fire hazards found in VIPR devices, specifically considering
materials compatibility, ignition mechanisms, and reaction effects of a fire. The
toxicity risk of non-metallic materials is considered as well as the ASTM analy-
sis and testing methods used to qualify a new design.
This paper does not attempt to identify the superior design of an existing
VIPR. It merely evaluates the inherent hazards of various VIPR designs and
provides guidelines for designers and operators to mitigate these hazards. Fires
have occurred with VIPRs and although specific fire details are not available for
publication, Wendell Hull and Associates, Inc. 共WHA兲, an independent forensic
laboratory, was familiar with over 20 separate fires involving various VIPR de-
signs, at the time of this publication.

VIPR Design

The integration of the valve and regulator designed for use with an oxygen
cylinder allows for three main design variations as shown in Figs. 1–3. The
design variations involve the presence of the shut-off valve and whether it iso-
lates the cylinder 共Fig. 1兲, isolates the regulator 共Fig. 2兲 or is not included at all
共Fig. 3兲.
VIPR Design 1 共Fig. 1兲 uses the shut-off valve to isolate the high-pressure
oxygen in the cylinder. In this design, the valve is exposed to bi-directional flow,
first during filling, then during flow to the patient. VIPR Design 1 provides
manual isolation should a single point failure of either the fill valve or regulator
cause either of these components to leak and deplete the contents of the cylin-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 287

Regulator

Fill Valve
Shut-off
valve

Cylinder

FIG. 1—VIPR Design 1; cylinder isolated by shut-off valve.

der. It also allows external connection to the fill valve without the fill valve being
under oxygen pressure. This has made VIPR Design 1 a popular design option.
VIPR Design 2 共Fig. 2兲 uses the shut-off valve to isolate the regulator. The shut-
off valve in this design experiences oxygen flow in one direction only.5 Though
the regulator can be isolated, the fill valve is always under pressure once the
cylinder is filled. VIPR Design 3 共Fig. 3兲 does not use a shut-off valve in the
design regulator. It relies solely on the regulator to provide both gas-tight shut-
off and flow control to the patient, in essence designing the shut-off valve into
the regulator.

Oxygen Hazards
The following evaluates the oxygen hazards based on material flammability of
metals and non-metals, ignition risks, and reaction effect of ignitions. ASTM

5
Flow could occur in the opposite direction if the cylinder is vented and pressure is
trapped between the shut-off valve and the regulator. This would be momentary due to
the small volume of pressurized gas.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
288 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Regulator

Shut-off
valve

Fill Valve Cylinder

FIG. 2—VIPR Design 2; regulator isolated by shut-off valve.

standards such as G63, G94, and G88 provide the basis for the evaluation and
guidelines provided 关2–4兴. The evaluation assumes the VIPR parts are cleaned
for oxygen service including the high-pressure gauge.
The overall objective of an oxygen-compatible design for any component or
system is one that exhibits a low probability of ignition and a low consequence
of ignition. Both of these aspects are evaluated in a typical Oxygen Hazards &
Fire Risk Analysis 关5,6兴. The probability of ignition is evaluated by analyzing
the severity of the various characteristic elements of select ignition mecha-
nisms, combined with the material flammability. The consequence of ignition,
or fault tolerance of the design, largely considers the degree of fire propagation
within a component and the resulting damage.

Material Flammability
Materials become more flammable with increasing oxygen pressures, and the
energy required to ignite these materials is reduced. VIPRs are currently being
used with ⬎99.5 % oxygen and pressures up to 20 MPa 共2,900 psig兲 and being
considered for use at 30 MPa 共4,350 psig兲. Many metallic materials are consid-
ered flammable at these oxygen conditions depending on their thickness, and
all non-metallic materials are considered flammable. While the presentation of
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 289

Regulator

Fill Valve
Cylinder

FIG. 3—VIPR Design 3; no shut-off valve.

materials compatibility data next is not new, a review of this data as it pertains
specifically to applications in VIPRs is helpful to more fully understand the
oxygen hazards in VIPR designs.

Metals—Table 1 provides a summary of published oxygen compatibility


data for various metals 关7兴. The data listed was generated in approximately 100
% oxygen. The threshold pressure data, based upon ASTM G124, is the primary
data used to evaluate the flammability of metals. Generally, if a metal is used
above its threshold pressure, it is considered flammable in that environment. If
used in thin cross-sections 共i.e., ⬍3.2 mm thick兲, its threshold pressure would
be expected to be lower than the data shown, as indicated for thin cross-
sections of Monel and stainless steel 共Table 1兲. It is noteworthy, however, that
even if a material is flammable in its use configuration and operating environ-
ment, it may be selected if caution is used in avoiding ignition mechanisms. If
ignited, however, under conditions beyond its flammability threshold, self-
sustaining fire propagation is likely.
Brass, Monel, copper, nickel, and bronze all demonstrate superior ignition
and burn resistance. Their flammability threshold pressures are all above 55
MPa 共8,000 psia兲 in bulk 共3.2-mm thickness兲 configurations. Thus, unless thinly
divided 共e.g., particles or springs兲, they are considered non-flammable in this
component. Tin-bronze is often used as a sintered filter alloy and data shows
that even when finely divided, it is still fairly burn-resistant in oxygen 关8兴.
A market survey would show that the majority of VIPRs used today have a
brass body 共i.e., manifold兲 and brass parts where high-pressure oxygen exists.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 1—Oxygen compatibility data of various metallic materials.

Metal Thresholda Pressure 共MPa兲 ⌬Hc b 共J/g兲 Particle Impactc共°C兲 Frictional Heating 共W / m2兲

Downloaded/printed by
Brass 360 ⬎70 6,000 ⬎371 共NIe兲 0.70–1.19
Nickel 200 ⬎70 4,100 ¯ 2.29–3.39
Copper 102 ⬎55 2,500 ¯
316 Stainless 3.4 7,900 38 0.53–0.86
Tin-bronze ⬎70 ¯ ⬎371 共NI兲 2.15–2.29
Monel 400 ⬎70 3,642 ⬎371 共NI兲 1.44–1.56
Monel, stainless steel 共thin section兲 0.1d ¯ ¯ ¯
Carbon steel 0.2 7,500 ¯ 0.27–0.32
Aluminum 0.2 31,400 ⫺46 0.061
a
Threshold pressures of metals are based on 3.2-mm diameter rod data per ASTM G124.
b
Heat of combustion

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
c
Bulk sample temperature required for ignition by aluminum particle impact at supersonic velocities.
d
Thin sections of Monel and stainless steel have been shown to self-sustain combustion at ambient pressure 共e.g., wire mesh兲.
e
No ignition.

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


290 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
FORSYTH ET AL., doi:10.1520/JAI102296 291

These would be considered non-flammable for thicknesses greater than 3.2 mm


based upon the data discussed earlier. Despite the burn resistance of brass, thin
sections could be susceptible to melting and subsequent burn-out in the event
of a fire, especially if other flammable alloys within the VIPR burn and produce
liquid-phase combustion products.
Historically, some stand-alone regulators for medical oxygen have been
made of aluminum to take advantage of its high strength-to-weight ratio. Alu-
minum is a very desirable engineering alloy for this reason. However, due to the
flammability of aluminum in oxygen down to very low pressures 共Table 1兲, and
its ease of ignition if not adequately protected, many of these regulators exhib-
ited a high probability of ignition and a severe consequence of ignition, where
fires resulted in personal injury and even fatalities 关1兴. To the best knowledge of
the writers, no VIPR-style medical oxygen regulators use aluminum in the high-
pressure region of the component.
Downstream of the VIPR regulator’s main seat, where pressure is reduced
to three or four atmospheres, the internal metal parts are often made from
brass, aluminum or stainless steel depending on regional requirements. Brass
is generally considered nonflammable in this environment but any aluminum
parts would still be considered flammable. The flammability of stainless steel
parts at these lower pressures would be dependent on the thickness, configu-
ration, and exposure to a flowing oxygen environment. If flammable, these
parts of the regulator may be vulnerable to promoted ignition, or kindling, via
ignition of other energetic materials such as o-rings seals and non-metallic
valve seats. Fire in the high-pressure portion of the VIPR leading to ignition of
non-metals in the low-pressure regions could raise the pressure and tempera-
ture sufficiently to reach the flammability threshold of stainless steel parts in
the low-pressure region.
Sintered tin-bronze is typically used for filters and would be considered
non-flammable in this application. However, if the filter is damaged during
assembly or use, any dislodged sintered bronze particles that make up the filter
element could be considered flammable based on the particle size. Stainless
filters, either sintered or mesh type, are generally avoided in high-pressure oxy-
gen due to the alloy’s high flammability rating in these thin configurations.
Brass or bronze mesh filters are also considered flammable in this environment
and as such should be used with caution.
Various alloys are used for springs such as copper-beryllium, phosphor
bronze, stainless steel, and carbon steel. The flammability of the copper-based
alloy springs would be dependent on the spring wire thickness. Although po-
tentially flammable, these copper-based alloys would exhibit a significantly
higher burn resistance and lower heat of combustion than the ferric alloys
共Table 1兲. During the VIPR fire investigations by WHA, promoted ignition and
burning of stainless steel regulator springs 共⬃2-mm diameter兲 even in the low-
pressure region of a regulator were observed. When steel parts internal to the
regulator ignite, they can provide a substantial fuel source, which can melt-
through the brass body and lead to an energetic component burn-out.

Non-metals—Most non-metallic materials are considered flammable in


high-pressure oxygen systems. Because of this, all non-metals must be used
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
292 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 2—Oxygen compatibility data of non-metallic materials.

AIT ⌬Hc OI

Non-metal 共°C兲 共°F兲 共J/g兲 共%兲


EPDM 150–200 302–392 47,200 25.5
Nylon 6/6 共polyamide兲 178 352 32,200 36
Zytel 42 共polyamide兲 192 378 37,000
Buna-N 共Nitrile rubber兲 173 343 35,600 22
Hydrocarbon Greases 160–220 320–428 33,500–41,900 21
Polyurethane 181 358 21,800–27,200 22
PEEK 305 581 32,500 35
Viton 共FKM兲 300–325 572–617 12,600–20,900 56–100
Teflon 共PTFE兲 ⬎427 ⬎800 4,200–7,100 95–100
Krytox 共perfluorinated lubricant兲 ⬎427 ⬎427 3,800–4,200 DNI

with caution in oxygen systems and their successful use is application-specific.


Many fires in oxygen systems are the result of non-metallic materials igniting
first, then kindling ignition directly to surrounding flammable metals or via
contaminants such as metallic particles.
Table 2 provides a comparison of oxygen compatibility data for various
non-metallic materials used in VIPR designs 关7兴. Hydrocarbon contaminant is
also listed for comparison. Generally, preferred non-metals for oxygen service
have a high auto-ignition temperature 共AIT兲 共i.e., they resist ignition until ex-
posed to higher temperatures兲, low heat of combustion, ⌬Hc 共i.e., they have low
damage potential, as discussed earlier for metals兲, and high oxygen indices 共OI兲
共i.e., a high concentration of oxygen is required before they become flammable
at atmospheric pressure兲.
Materials like Viton 共FKM兲, polytetrafluorethylene 共PTFE兲, polychlorotrif-
luoroethylene 共PCTFE兲, and Krytox exhibit relatively high AITs and low ⌬Hcs,
resulting in reduced likelihood of ignition and lower consequence of ignition.
They are preferred materials for oxygen service based on oxygen compatibility
and fire hazards. Comparatively, EPDM, polyamide 共including Nylon 6/6 and
Zytel兲, and NBR 共Buna-N兲; all have low AITs and high ⌬Hcs, comparable to
hydrocarbon grease. These materials require relatively low amounts of energy
to ignite, and release relatively large amounts of energy when ignited.
Over the past decade, the selection of non-metals for oxygen breathing gas
systems has also been influenced by other design criteria, such as toxicity. The
toxicity of combustion products of non-metals in breathing gas applications is
important to consider because fires in oxygen breathing gas systems most-often
involve non-metals. Though halogenated materials containing fluorine and
chlorine, such as PTFE, PCTFE, and FKM, exhibit superior ignition and burn
resistance in oxygen, their combustion products are potentially toxic for hu-
man consumption if delivered in a sufficiently high dosage 共i.e., concentration兲
and rate 关9兴.
Though it is still to be determined whether ignition and burning of the
small halogenated non-metals internal to VIPRs would deliver sufficient dosage
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 293

at a rate to harm a patient, it is true that non-halogenated materials such as


EPDM and polyamide, which represent popular alternative material choices for
elastomers and plastics, respectively, release less toxic combustion products
共assuming the same dosage and delivery rate兲. Thus, the perception that halo-
genated non-metals present a greater toxicity risk than ignition risk to users of
oxygen breathing equipment has caused some VIPR designers to avoid the use
of halogenated non-metallic materials in these designs.
Unfortunately for a designer, the non-metals predicted to produce the least
toxic combustion products generally happen to be some of the most flammable
and easy to ignite in oxygen 共Table 2兲. This can present a conflict to a VIPR
designer: Use less oxygen-compatible non-metals to avoid toxic combustion
products but potentially induce a higher risk of ignition that may lead to burn-
out, or use more oxygen-compatible non-metals to minimize the risk of ignition
and fire propagation, yet risk potential ingestion of toxic combustion products
by the patient should ignition still occur. The historical approach to this con-
flict, at least from ASTM standards, has been to favor the use of oxygen-
compatible non-metals 共halogenated compounds兲, and ensure the risk of igni-
tion is sufficiently low through analysis and testing 关2,4,7,12兴. The rationale
used for this approach has been ignition prevention. It is considered important
that at the temperatures where halogenated non-metals begin to decompose
and off gas, non-halogenated replacement materials may have already ignited
and burned, releasing three or more times the amount of energy during com-
bustion than their halogenated counterparts.
Nonetheless, more recently there has been a trend for organizations and
regulating bodies to ban the use of halogenated non-metals with the argument
that there are documented patient fatalities resulting from the ingestion of
toxic products from halogenated polymers. The details of these incidents are
not widely available; but it cannot be questioned that the combustion products
of halogenated polymers present a toxicology risk if they are ingested in suffi-
cient dosage and delivery rate. However, the writers are not aware of any for-
mal study to compare and weigh the ignition risk to the toxicity risk of haloge-
nated non-metals in oxygen, except for the current study discussed next. This
type of industry-wide study is very much needed to make educated risk assess-
ments regarding the use of halogenated non-metals in oxygen.
The risk of toxic combustion products being ingested by a patient breath-
ing on a VIPR device ultimately depends on several factors, many of which are
currently unspecified. For example, the actual dosage and duration of exposure
that a patient may receive in the event of a non-metal oxygen fire is important
to understand in order to assess the risk. This exposure data then needs to be
compared to the toxicity concentration thresholds of the given combustion
products and a determination made by a qualified toxicologist as to the effect
on the human body. Then, the toxicology assessment should be incorporated
into a detailed risk analysis to weigh the toxicology risk against the ignition
risk, both of which are of great concern to a designer.
A cooperative, industry-sponsored, test program is now underway to fur-
ther characterize and quantify the toxicity risk of common halogenated and
non-halogenated non-metals used in oxygen breathing gas equipment, includ-
ing VIPRs. Test facilities are developing methods to ignite and burn a variety of
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
294 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

common non-metals, capture their respective combustion products, and deter-


mine the quantity of compounds present. A qualified industry toxicologist is
providing the required analysis as to how these amounts may affect human
health. Given that the overall mass of most non-metallic materials used in oxy-
gen breathing applications is relatively small, the configuration 共pressures, vol-
umes, flow rates兲 of these applications will significantly influence the dosage
and duration of exposure to an end user.
Unfortunately, in at least one case, a fire has occurred with a VIPR design
after undergoing design changes that included changing the non-metals from
halogenated materials to non-halogenated. Prior to these design changes, the
particular VIPR design exhibited a successful service history. Although it can-
not be shown that the material change directly caused this fire, the reaction
effect of the fire was no doubt made more severe as a result of the non-metals
ease of ignition and high energy release. In other cases, designers have removed
halogenated polymers to reduce the toxicity concerns, but have still specified
the use of halogenated lubricants 共albeit sparingly兲. Combustion of a haloge-
nated lubricant would be expected to produce similar combustion products as
halogenated elastomers and plastics. An alternative non-halogenated lubricant
is not advised because the remaining material options consist of highly flam-
mable hydrocarbon greases. Thus, further study of these issues are required so
that educated recommendations based upon data can be made.
Non-metals are an important part of any VIPR design and whether haloge-
nated or non-halogenated non-metals are used in the design, they are generally
considered the “weak link” in any oxygen component due to the ease of ignition
compared to metals. However, whether a designer chooses to use fluorinated
non-metals to minimize ignition, or protect the non-halogenated non-metals
from ignition through good design techniques 共i.e. fully encapsulating O-ring
seals兲, it should be understood that even with a VIPR constructed entirely of
burn-resistant metal alloys such as brass, ignition of a non-metallic material
can still present a hazard that leads to an external fire. Though the risk of
component burn-out is considered low if burn-resistant alloys are used, the
likelihood of non-metallic combustion products and fire exiting the VIPR is still
considered high either through the normal flow path or through the relief sys-
tem共s兲 if combustion temperatures sufficiently increase the pressure. With fire
or hot combustion products exiting these ports, the risk of kindling ignition to
exposed flammable plastic materials, such as a cannula on the VIPR’s barbed
outlet, a plastic guard covering the VIPR, or a plastic bag carrying the VIPR
and cylinder, may still be high. Although some of these external plastic materi-
als may not be flammable in ambient pressure air 共i.e., OI⬎ 21 % oxygen兲, the
continual supply of oxygen from the VIPR can increase the local oxygen con-
centration above the OI of these materials. Fires of this nature have been in-
vestigated. The risk of kindling these plastic materials and creating a large
external fire is directly dependent on the design of the VIPR and the heat of
combustion of the non-metallic materials. Table 2 shows the advantage of per-
fluorinated materials over the other material choices based solely on heat of
combustion.
While polyamide 共Nylon兲 has a long-standing history of use in oxygen cyl-
inder valve applications, fires have occurred in these valves 共oftentimes the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 295

cause of burn-out in welding and medical regulators兲 关1兴. The VIPR fires inves-
tigated by WHA exhibited similar evidence to the stand-alone cylinder valve
fires. The evidence was consistent with fire origin at the Nylon valve seat. If
Nylon is chosen as a sealing material, the risk and consequence of ignition can
be reduced by limiting the exposure of the material to ignition mechanisms and
reducing the mass of the material, thus minimizing the release of energy if
ignited. For example, a designer can reduce the exposed surface areas prone to
flow impingement, reduce frictional wear through ensuring a non-rotating
valve seat, and reduce the volume of sealing material that can combust. In
addition, ignition and promoted ignition testing can further support the safe
use of non-perfluorinated materials. However, it remains indisputable that flu-
orinated non-metals, like PTFE 共Teflon兲, PCTFE 共Neoflon兲, and FKM/FPM 共Vi-
ton兲 have superior resistance to ignition and burning in properly designed oxy-
gen systems as compared to hydrocarbon-based materials such as EPDM and
NBR.

Ignition Mechanisms and Reaction Effects in VIPR Sub-Components

Fill Valve—The fill valve is used to connect to a pressure source and fill the
oxygen cylinder. It functions as a one-way valve, allowing gas to fill the cylinder
but not escape from the inside out. The mating filling connector interfaces with
this valve with either a threaded connection or some other form of mechanical
locking mechanism. Both connection designs involve metal-to-metal contact,
friction, and hence, metallic particle generation upon assembly. Depending on
the location of these contact points, with respect to the external seals of the
connection assembly, the generated particle contamination could become en-
trained in the flow. O-ring mating seals do not completely prevent metal-to-
metal contact and are notorious, depending on their design, for experiencing
damage and generation of non-metallic particles. Often, the fill valve design
involves a poppet and spring that require physical actuation 共i.e., another
metal-to-metal contact兲 to open the valve for filling. This poppet and spring are
contained within the external seals. Depending on the design, the fill valve seat
can be exposed to high loads and configurations that cause damage to the
surfaces of the seat during valve actuation. Therefore, the potential for metallic
and non-metallic particle generation exists within the fill valve during filling
operations. These particles can then be entrained in the flow and transported to
critical locations within the body of the VIPR.
The use of a filter in the fill valve would appear to be a helpful design
feature to keep particulate from migrating further into the VIPR and help miti-
gate the risk of particle impact ignition and contaminant-promoted ignition in
the VIPR, assuming there are no other sources of particulate generation within
the device. However, filters are difficult to implement at this location in the
VIPR. They can become clogged and cause problems during the filling process.
Furthermore, clogged filters become susceptible to adiabatic compression igni-
tion if the filters are clogged with non-metallic contamination and are rapidly
pressurized. Another disadvantage of a filter in the fill valve is that it experi-
ences bi-directional flow. The filling process often begins with a cylinder vent
after the filling connector is attached to the fill valve. The cylinder is vented and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
296 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

sometimes vacuumed through the opened fill valve and a vent or vacuum line of
the filling system. Due to this reverse flow operation, any contaminants trapped
by the filter on the fill side can be ejected from the fill valve into the filling
system. This presents particle impact and contamination ignition risks in the
filling system or VIPR when filling of the cylinder commences after the venting
and vacuum stage.
Because of the risks associated with a filter in the fill valve, it is prudent for
the filling system to have a filter near the filling connector, instead of in the fill
valve itself. The location of this filling system filter is critical to ensure only
unidirectional flow. One method to prevent this might be to position the vent
line branch downstream 共i.e., cylinder-side兲 of this filter and add a check valve
to further mitigate reverse flow in the filling system filter. Purging the line and
filling connector before connection also helps mitigate the risk of particulate
entering the VIPR. Although the filling system filter is considered a good prac-
tice for minimizing particulate entry from the filling system, it does not address
the particles generated within the VIPR. The internal design of the VIPR should
allow these particles to easily flow into the cylinder by providing limited flow
restriction and limited opportunity for particle traps along the flow path to the
cylinder. By using an effective dip tube on the VIPR, particles can be safely
contained in the cylinder and gas is taken from the major diameter of the
cylinder.
The fill port valve is subject to adiabatic compression on the reverse-flow
side when the shut-off valve is opened in VIPR Design 1 共Fig. 1兲. Additionally,
adiabatic compression can occur if filling is attempted with high pressure al-
ready present in the component 共assuming internal pressure is not relieved
when the fill system is connected to the fill valve and the fill valve is a “pressure
to open” design兲. This produces a potentially severe condition for the seat and
seal of the fill valve. The industry standards that specify adiabatic compression
testing of VIPRs only subject the fill valve to adiabatic compression heating in
the filling direction 关10,11兴. These standards do not require testing in the re-
verse flow direction 共created by opening the shut-off valve兲 and, therefore, do
not evaluate this ignition risk 共an exception would be if the endurance cycling
test for the shut-off valve were performed in oxygen兲. Although no fires in
VIPRs investigated by WHA exhibited evidence that was consistent with adia-
batic compression ignition of the fill valve, fires did occur after opening the
shut-off valve during operation by an end user. The VIPR designs involved in
the fires had previously met the adiabatic compression testing requirements of
the current industry standard 关10,11兴. This suggests these tests, although criti-
cal and necessary, are not sufficient to address all active ignition mechanisms in
the VIPR design.
If the fill valve seat is ignited during filling, kindling of the spring is con-
sidered possible if the spring is not constructed of a burn-resistant alloy. The
added energy and flow direction could likely then kindle the RPV seat, regula-
tor seat, and shut-off valve seat. If the fill valve seat is ignited when the shut-off
valve is opened, fire would exit the fill valve and its intensity would be depen-
dent on whether the spring is kindled. Figure 5 shows potential kindling chains
should ignition occur in the fill valve. Kindling chains are broken if the mate-
rials in the chain are burn-resistant.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 297

Shut-Off Valve and Filters—Figure 4 shows a schematic of the typical com-


ponents included in a VIPR matching Design 1. Note that not all of these com-
ponents are included in every VIPR design, nor does the figure suggest that all
these components be installed. The shut-off valve functions to isolate gas flow
to the regulator from a pressure cylinder, and isolates the filling source from
the supply cylinder. During cylinder filling, if the pressure source were opened
rapidly with the VIPR shut-off valve closed, a potentially severe adiabatic com-
pression could occur at the seat. This ignition mechanism and configuration is
tested by industry standards 关10,11兴 but on new components “fresh out of the
box.” The risk of adiabatic compression is believed to increase with seat wear.
Performing adiabatic compression testing on a VIPR that has experienced
cycles of operation consistent with service life expectations may represent a
better assessment of the ignition risk as the internal parts wear.
The manner of high velocity flow across the exposed face of the seat, along
with probable wear from a rotating stem design, creates a risk of flow friction
ignition. A non-rotating stem produces less wear and less debris from friction
during actuation, which reduces the likelihood of flow friction ignition. Designs
with stem packing below the threads also prevent particle generation within the
threads from entering the flow paths. Metallic particles generated via filling, as
described earlier, or in the threads of the shut-off valve can further deteriorate
the surface of the shut-off valve seat and add to the quantity of flammable fuel.
Particles may also accumulate at the cylinder filter 共if present兲 during filling
共Fig. 4兲. Subsequent use of the regulator in patient supply mode, effectively
back-flows the filter, a poor practice in oxygen systems which should be
avoided. Under these conditions, particles entrained in the transient high-
velocity flow from the cylinder can impact at the sharp turn 共normally 90° turn兲
upstream of the shut-off valve before flowing through the shut-off valve. Igni-
tion of particles is possible during such impacts against the metallic flow path
walls, with subsequent deposition of the burning particulate on the shut-off
valve seat, kindling ignition to the non-metallic seat. If not ignited upstream of
the shut-off valve, the potential exists for particle impact ignition in the valve
itself, especially if the seat retainer/plunger is not made of burn-resistant alloys.
While filters are encouraged to minimize debris from being ingested into
high-velocity components, they are also a collection point for debris whereby
large accumulations of flammable contaminants may become trapped together
关4,7兴. If ignition occurs, the resulting fire can be more severe. Historically, oxy-
gen fires have occurred in industry as a result of bi-directional flow through
filters. Thus, filters must be used with caution and each application carefully
considered. In this design, if there was not a filter in the cylinder interface port,
these contaminants might be swept out of the device into the cylinder 共a more
benign location for particulate兲 during filling and safely contained by using an
effective dip tube.
The position of a burn-resistant filter between the shut-off valve and the
downstream 共i.e., regulator side兲 parts is important to minimize promoted ig-
nition to the regulator. Ignition mechanisms include adiabatic compression,
particle impact, and promoted contaminant ignition that could occur down-
stream of the shut-off valve upon opening, or by flow friction during oxygen gas
flow. The filter can help protect the downstream parts from particles and dissi-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
298 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Adjustable
Flow Head

Relief High Flow


Mechanism Outlet

Regulator

Residual
Pressure Valve

Regulator
Filter
Fill
Valve
Shut-off
Valve
Gauge Burst
Disc
Cylinder
Filter

Flow
Restrictor

Cylinder

FIG. 4—Schematic of typical integrated components of VIPR Design 1 共note: Not all of
these components are included in every design, nor does this figure suggest that each of
these components be installed兲.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 299

Shut-off Residual Adjustable


Fill Port Valve Pressure Valve Regulator Flow Head and
High Flow
Outlet

Contaminants

Seat Seat Seat Seat Seals

Spring
Diaphragm
Seals Springs,
flow orifices,
Seals
Relief Seat housing

Spring

Spring
Spring

Burn-Out
(Energetic Breach / Metal Fire)

Legend: High High and Low Low Nonmetallic Fire


Pressure Pressure Pressure Origin of Ignition Release Only

FIG. 5—Potential kindling chains in VIPR designs.

pate the energy created by compression heating, burning particles and/or burn-
ing shut-off valve seat. Since the energy release and delivery is difficult to model
and predict, testing the VIPR to evaluate its fault tolerance to this promoted
ignition event provides confidence that the filter design is capable of containing
the energy and breaking the kindling chain. Test methods such as ASTM G175
modified for VIPRs provides such information 关12兴. This is especially important
because VIPR fires exhibiting evidence consistent with these ignition mecha-
nisms have been investigated by WHA. The consequence of ignition, or a po-
tential kindling chain, within the VIPR Design 1 is shown in Fig. 5. As shown,
there are multiple paths that can lead to burn-out should ignition occur. A
burn-resistant filter placed upstream 共i.e., cylinder-side兲 of the regulator can act
to limit the kindling chain in many of these ignition scenarios.

Residual Pressure Valve—If included in the VIPR design, the RPV is gener-
ally upstream of the regulator seat and serves to prevent full depletion of gas
from the cylinder and thus reduces the risk of contamination in the cylinder. As
in other areas of the VIPR, the seat and seals are at risk of ignition from flow
friction, adiabatic compression, or promoted ignition from an upstream mate-
rial or component, particularly the shut-off valve seat. The risk of adiabatic
compression ignition is evaluated by industry standards 关10,11兴 but not the risk
of promoted ignition 共with the exception of ASTM G175 phase 2兲. The sealing
configuration of the RPV seat might produce wear on the o-ring over repeated
actuation and increase the potential for flow friction ignition or adiabatic com-
pression during opening of the shut-off valve or fill valve.
If the seat or seals are ignited, the spring behind the piston is at risk of
being kindled, depending on material flammability. Ignition of the seat could
kindle the regulator seat downstream. Ignition of the seals would expel non-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
300 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

metallic combustion products through the gaps between the RPV port and
parts. However, this scenario would be significantly less severe than if the
spring was kindled, causing the expulsion of metallic combustion products.

Pressure Regulator—The most common ignition mechanisms for regulators


in this type of application are adiabatic compression and promoted ignition of
contaminants in the upstream RPV and shut-off valve seats. For adiabatic com-
pression ignition to be credible, the shut-off valve or fill valve must deliver
high-pressure oxygen to the regulator seat sufficiently fast so that heat from gas
compression cannot be dissipated to the surrounding metals before ignition of
the regulator seat occurs. The plug style shut-off valves supplied with most
VIPRs have coarse stem threads and as such are considered fast-opening by
design. Therefore, compression heating will occur on the regulator seat each
time the shut-off valve is opened, assuming the regulator seat is depressurized
after use. Thus, ignition of the regulator seat is a risk if not properly designed to
mitigate this mechanism. Compression heating is evaluated by industry stan-
dards 关10,11兴 but the flow originates from the fill valve and not the cylinder.
Further, the test is performed on new components “fresh-out-of-the-box,”
which may not replicate the most severe service condition.
If the regulator seat fails in an ignition event, the kindling chain may con-
tinue into the low-pressure side of the regulator, potentially igniting the dia-
phragm, and possibly the regulator spring 共Fig. 5兲. Even though the brass body
is considered burn-resistant, the liquid combustion products of the metal
spring may allow vigorous melt-through under direct impingement of flame.
Likewise, combustion products can rapidly vent through the patient outlet
port, igniting or melting intermediate components, even burn-resistant alloys if
they are thin. To further evaluate the effect of promoted ignition, it is recom-
mended to perform testing per ASTM G175 Phase 2, modified so the ignition
pill is located at the cylinder inlet, and any cylinder filter or flow restrictor is
removed 关12兴. This allows the burning pill to directly kindle ignition to the
shut-off valve seat and carry combustion products up past the RPV and into the
regulator, thus evaluating the effectiveness of the component to dissipate heat
and protect parts downstream of the shut-off valve from ignition 共Fig. 4兲.

Adjustable Flow Head and High Flow Outlet—The low-pressure adjustable


flow head and high flow outlet connection are generally constructed of a variety
of materials. Due to the low pressure in these parts, the severity of ignition
mechanisms present is reduced. The dominant ignition risk is promoted igni-
tion from the combustion products of the regulator if ignited. Any flammable
parts such as springs, flow orifices, and seals would be prone to kindling, as
shown in Fig. 5. Depending on the flammability of the housing, the fire release
from these parts could be significantly energetic even if the pressure is low,
especially if the regulator seat ignites and oxygen flow increases.

Conclusions
Medical oxygen valves that integrate a regulator, or VIPRs, are now commonly
found in the marketplace. Their designs vary, but are generally quite complex
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
FORSYTH ET AL., doi:10.1520/JAI102296 301

as they integrate many components into a single design. The integration of the
components into a single manifold creates potential oxygen hazards that must
be mitigated to produce a safe, oxygen-compatible component, one that dem-
onstrates a low probability of ignition and a low consequence of ignition.
This paper presents the most common oxygen fire hazards found in VIPR
devices, specifically considering materials compatibility, ignition mechanisms,
and reaction effects of a fire. Hazards during filling and patient use are identi-
fied along with the risks associated with filters in the VIPR design and filling
systems. Material compatibility is evaluated along with a review of the toxicity
risk of non-metallic combustion products. The effectiveness of current industry
testing standards is discussed with respect to the fire hazards identified. Guide-
lines to mitigate ignition risks are provided, based on good practices and prin-
ciples from industry standards.

References

关1兴 Newton, B. E., Hull, W. C., and Stradling, J. S., “Failure Analysis of Aluminum-
Bodied Medical Regulators,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres: Ninth Volume, ASTM STP 1395, T. A. Steinberg, B. E. New-
ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关2兴 ASTM G63-99, 2007, “Standard Guide for Evaluating Nonmetallic Materials for
Oxygen Service,” Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA.
关3兴 ASTM G94-05, 2005, “Standard Guide for Evaluating Metals for Oxygen Service,”
Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA.
关4兴 ASTM G88-05, 2005, “Standard Guide for Designing Systems for Oxygen Service,”
Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA.
关5兴 Forsyth, E. T., Newton, B. E., Rantala, J., and Hirschfeld, T. “Using ASTM Stan-
dard Guide G 88 to Identify and Rank System-Level Hazards in Large-Scale Oxy-
gen Systems,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres: Tenth Volume, ASTM STP 1454, T. A. Steinberg, H. D. Beeson, and B. E.
Newton, Eds., ASTM International, West Conshohocken, PA, 2003.
关6兴 Shoffstall, M. S. and Stoltzfus, J. M., “Oxygen Hazards Analysis of Space Shuttle
External Tank Gaseous Oxygen Pressurization System,” Flammability and Sensitiv-
ity of Materials in Oxygen-Enriched Atmospheres: Tenth Volume, ASTM STP 1454, T.
A. Steinberg, H. D. Beeson, and B. E. Newton, Eds., ASTM International, West
Conshohocken, PA, 2003.
关7兴 Beeson, H. D., Smith, S. R., and Steward, W. F., Eds., Safe Use of Oxygen and
Oxygen Systems: Handbook for Design, Operation, and Maintenance, ASTM Inter-
national, West Conshohocken, PA, 2007.
关8兴 Stoltzfus, J. M., Lowrie, R., and Gunaji, M. V. “Burn Propagation Behavior of Wire
Mesh Made from Several Alloys,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres: Fifth Volume, ASTM STP 1111, J. M. Stoltzfus and
K. McIlroy, Eds., ASTM International, West Conshohocken, PA, 1991.
关9兴 ISO 15001-2003共E兲, 2003, “Anesthetic and Respiratory Equipment—Compatibility
with Oxygen.” International Standards Organization, Geneva, Switzerland.
关10兴 ISO 10524-3:2005, 2005, “Pressure Regulators for Use with Medical Gases—Part 3:
Pressure Regulators Integrated with Cylinder Valves,” International Standards Or-
ganization, Geneva, Switzerland.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
302 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

关11兴 CGA E-18, 2008, “Medical Gas Valve Integrated Pressure Regulators,” 1st ed.,
Compressed Gas Association, Chantilly, VA.
关12兴 ASTM G175-03, 2003, “Standard Test Method for Evaluating the Ignition Sensitiv-
ity and Fault Tolerance of Oxygen Regulators Used for Medical and Emergency
Applications,” Annual Book of ASTM Standards, ASTM International, West Con-
shohocken, PA.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102233
Available online at www.astm.org/JAI

William P. Schmidt,1 Michael Cawthra,2 Patrick A. Houghton,3


Ralph H. McDonald, Jr.,1 Robert J. Sherwood,1 and
Stephen J. Wieder1

Sealed Aluminum Cavity Reactions when


Submerged in Pure O 2 Reboiler Sump

ABSTRACT: Aluminum has a long history of safe service in cryogenic air


separation units; however, there have been some rare instances of
aluminum/O2 reactions. Aluminum ignition and propagation depends strongly
on the oxygen purity, oxygen pressure, aluminum geometry, and type and
energy of igniter. Over the years, the industry has experienced a particular
type of aluminum/O2 reaction: cavity incidents. These incidents are charac-
terized by the presence of a sealed cavity that is created when welding two
metal items together. Such a sealed cavity can lead to aluminum ignition
under certain specific conditions: 共1兲 The sealed cavity is submerged in a
bath of liquid oxygen. Over a long period of time, liquid cryogen can enter the
sealed cavity through slight imperfections in the weld. 共2兲 The sealed cavity
is warmed over a short period of time, typically a few hours. 共3兲 The liquid
oxygen vaporizes, and the pressure inside the cavity builds to very high
levels 共potentially over 50 bars兲. 共4兲 Ignition occurs in the high pressure, high
purity oxygen environment, and the resulting aluminum/oxygen reaction
burns through the relatively thick cavity walls. 共5兲 The oxygen and reaction
products exit the cavity through the hole, lowering the pressure and extin-
guishing the reaction. This paper discusses two such incidents, which have
occurred since 2001. Both incidents took place in brazed aluminum heat
exchanger 共BAHX兲 reboiler support beam systems. In both cases, the BAHX
reboiler was damaged, leading to a process leak, which required that the
plant be repaired. In one case, the damage occurred while warming a plant.

Manuscript received November 12, 2008; accepted for publication June 12, 2009; pub-
lished online July 2009.
1
Air Products and Chemicals, Inc., 7201 Hamilton Blvd., Allentown, PA 18195-1501.
2
Air Products and Chemicals, Ltd., Hersham Place Technology Park, 41-61 Molesey Rd.,
Walton-on-Thames KT12 4RZ, United Kingdom.
3
Ph.D., Air Products and Chemicals, Inc., 7201 Hamilton Blvd., Allentown, PA 18195-
1501.
Cite as: Schmidt, W. P., Cawthra, M., Houghton, P. A., McDonald, R. H.Sherwood, R. J.
and Wieder, S. J., ⬙Sealed Aluminum Cavity Reactions when Submerged in Pure
O2 Reboiler Sump,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102233.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 303
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
304 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

In the second case, the damage occurred during normal operation. This
second case does not appear to follow the sequence of events outlined
above; however, evidence is presented to support the scenario that the cav-
ity incident occurred during a previous warming of the plant. This initial dam-
age then impaired the normal operation of the reboiler, leading to a second
hydrocarbon related reaction during normal operation. The paper discusses
the potential ignition mechanisms, probable causes of the incidents, and
methods to prevent future re-occurrences.
KEYWORDS: aluminum combustion, air separation, ASU, BAHX,
brazed aluminum heat exchanger, cavity, cavity incident, reboiler,
sealed cavity

Introduction
Technology for separating air into its primary components 共oxygen, nitrogen,
and argon兲 by cryogenic distillation has been practiced for over 100 years. The
incidents discussed in this paper took place in the sump of the low pressure
column. The specific equipment item is the reboiler, which is a brazed alumi-
num heat exchanger 共BAHX兲. These are described in more detail in Refs 1 and
2.
A BAHX consists of alternate fluid passages 共see Fig. 1兲. The passages are
separated by parting sheets. Within each passage are corrugated fins, which
serve two purposes:
• They separate the parting sheets, while providing the mechanical integ-
rity to hold the heat exchanger together.
• They promote heat transfer between the fluids.
The BAHX is built up like a sandwich with alternating passages of differing
fluids 共see Fig. 2兲. The outermost parting sheets are thicker than the internal
parting sheets and are called “capsheets.” To introduce and remove the streams
from each passage, at the feed and product end of the exchanger, the fins are
angled toward a specific point. A half-pipe header is welded over the opening to
collect the fluid. For a reboiler, however, the oxygen end is open and does not

FIG. 1—BAHX passage construction.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 305

FIG. 2—BAHX construction.

have header tanks or distributors. While air separation unit 共ASU兲 reboilers
only have two fluids, BAHXs can be used to exchange heat between many
streams.
Figure 2 shows a reboiler schematic. The O2 passages are open at both
ends. The N2 passages are shown open on the sides in the left hand sketch. The
right hand sketch shows the header tanks attached. The actual passage heights
are 5–10 mm. A full size reboiler can be up to 1.4⫻ 1.4 m2 in cross section and
2–5 m long. Figure 3 is a schematic of the reboiler sump area. Here the BAHX
is submerged in a pool of liquid oxygen 共LOX兲. The oxygen passages are open at
the top and bottom of the exchanger. LOX enters the bottom of the reboiler and
boils as heat is added. The less dense vapor rises, bringing the remaining liquid
along. When the two-phase stream exits the reboiler, the vapor is removed, and
the liquid returns to the bath. The heat is provided from condensing nitrogen,
which is contained in passages alternating with the oxygen passages. Half-pipe
headers are welded over the openings to distribute and collect the N2 from its
passages.
There are some trace impurities in air. Those of concern in the ASU process
are CO2, N2O, and the low molecular weight hydrocarbons 共HCs兲. These com-
ponents boil at temperatures above oxygen so they will collect and concentrate
in the oxygen bath surrounding the reboiler. Trace components can accumulate
within unswept areas of the process equipment. To remove these, plants are
“defrosted” by warming them to ambient temperatures. This is done approxi-
mately every 3 years.

Typical Reboiler Incidents


The industry has experienced a small number of incidents involving reboilers.
These occur when HCs concentrate within the reboiler passages, and either
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
306 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—BAHX reboiler schematic.

reach the lower flammability limit or form a separated condensed phase 共i.e.,
solid or liquid兲. Either condition can then react with the LOX. The reaction is
typically strong enough that several passages are crushed due to the high pres-
sure, and a spherical or egg-shaped cavity is formed. These incidents typically
damage the BAHX to the point where leaks occur between passages, and the O2
purity is diluted with condensing N2. The damage is almost always caused by
overpressurizing the passages. There is typically little or no aluminum combus-
tion. However, in at least three cases, the HCs concentrated widely throughout
the reboiler, and a large fraction of the aluminum in the reboiler combusted.
The ASU industry has developed design and operating procedures to pre-
vent these types of incidents. These are discussed in detail in Refs 3–5. While
these incidents are infrequent and undesirable, they are known within the ASU
industry. However, the incidents discussed in this paper differ from the typical
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 307

reboiler incident because the initial overpressurization and damage are not
caused by HCs reacting with LOX during normal operation.

Cavity Incidents

In the late 1970s and early 1980s, Air Products had nine cavity incidents, with
the following similarities:
• A volume was completely enclosed, typically by a weld. This enclosed
volume is called the cavity.
• This cavity was submerged in LOX for an extended period of time 共typi-
cally several years兲 without being warmed up. It is believed that during
this time, LOX slowly diffused into the cavity through small cracks in
the weld, pinhole leaks in the weld, or perhaps through the weld itself, if
it had sufficient porosity.
• Of these incidents, four had some amount of aluminum combustion.
The other five incidents only showed evidence of overpressure leading to
bursting of the cavity.
• The LOX was always a minimum of 99.5 % pure.
In the early 1980s, recognizing this potential operating and safety hazard,
Air Products’ design standards were changed to eliminate sealed cavities in
areas that are submerged in LOX. Note that in many cases, the weld was placed
around the full perimeter of the cavity, even though this was not needed either
for strength or to intentionally create a sealed cavity. To eliminate the sealed
cavities, Air Products’ prohibits “welds that seal” or by positively venting sealed
cavities to a safe location. Additionally, existing plants were modified to elimi-
nate any cavities that could burn through the pressure containment envelope.
Air Products then had one more incident in a plant built before 1980. The
incident occurred in 1989, where a cavity was present in a reboiler and internal
supports were damaged.
In some cases where the aluminum combusted, the reaction burned
through nominal 13–25 mm thick aluminum plates. This was very surprising
because thick aluminum is relatively hard to ignite. In the early 1980s, an
experimental program was initiated to identify the ignition mechanism. Several
mechanisms were proposed but we were unable to reproduce ignition in labo-
ratory settings. Three mechanisms remain possible:
• Residual MIG welding dust might be present in the cavity. MIG dust is
relatively combustible, with a heat of combustion of 4700 cal/gm 共alu-
minum’s heat of combustion is 6700 to 7400 cal/gm, depending on the
data source兲. The fine MIG dust might provide the promoter 共kindling兲
that would ignite the thicker aluminum.
• Fine aluminum particles might be present in the cavity space and ignite
in the presence of the high pressure O2.
• As the pressure builds, small cracks develop in the weld or aluminum
surface. The surface of the crack may get locally very hot 共up to 800 to
1100 K兲, and the presence of a hot fresh aluminum surface might be
sufficient to begin the combustion. It should be noted that while this
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
308 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Reboiler, plan view.

mechanism has been postulated, it has never been produced in experi-


mental trials.

Incident Number 1

Background
The first incident occurred at a site hereafter referred to as site 1. Site 1 has
multiple plants and came online in the early 1980s. The plant was last defrosted
in March 1999 共26 months prior to this incident兲. There were no operating or
equipment problems in the weeks preceding the incident.
Each plant has four reboilers operating in parallel, designated A–D 共Fig. 4兲.
The reboilers are arranged in a square pattern. There is a single gaseous nitro-
gen 共GAN兲 pipe that comes up through the bottom head to feed the top of the
reboilers. Two of the four plant reboilers are shown in Figs. 5 and 6. There are
eight support plates on each reboiler, welded on by the reboiler vendor. Beams
are welded to these support plates, and the beams are then attached to the
column shell. The support plates are 300 mm tall by 12 mm thick and are
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 309

FIG. 5—Damage to reboilers, front view.

attached top and bottom to the reboiler by full welds. The support plates extend
across the entire reboiler width, and the vertical corner seams are seal welded
together. This creates a sealed “cavity space” under the support plates. Because
the vertical corner seams are welded to the other support plates 共not to the
core兲, there is probably a single cavity space at each elevation, extending
around all four sides of the reboiler. Note that this plant was built before Air
Products’ policy of eliminating sealed cavities was implemented.

Incident Description
One of the other plants at the facility was defrosted and returned to service.
Immediately following this, the plant in question was taken down for defrosting
at 07:45 on May 17, 2001. After completing the defrost and some routine main-
tenance, the plant was restarted at 22:00 on May 18. The initial observation was
that the plant ran very differently, with difficulty maintaining pressures. During
the next several hours, attempts were made to established liquid levels and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
310 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Reboiler damage, rear view.

purities. While some liquid levels were established, the plant never came close
to stabilizing and making purity. It was decided to enter the plant, inspect it,
and repair any damage.

Physical Evidence
On entering the column, significant damage was found. This is highlighted in
Figs. 5 and 6.

“A” Reboiler—
• On the corner closest to the D reboiler, a hole approximately 40 mm
diameter had burned through the support plate 共Fig. 5, item 3兲. The
location of the hole is significant, in that it is at the bottom of the sup-
port plate. A portion of the hole is actually through the weld which
attaches the support plate to the A reboiler capsheet. 共Fig. 7兲
• An open space approximately 150 mm⫻ 200 mm⫻ 700 mm was
found in the reboiler 共Fig. 5, item 1兲. It extends about 300 mm above
from the hole, and goes down to the bottom of the reboiler. In this area,
the fins and parting sheets were completely missing from seven O2 pas-
sages, either having been combusted or melted. The surface of the re-
maining metal was quite shiny. The third O2 passage was melted to the
bottom of the core. A small section of the reboiler was removed and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 311

FIG. 7—Closeup of cavity and reboiler damage.

brought back for analysis. This shows the extent of the damage. 共Fig. 5,
item 1; and Fig. 8兲.
• In the area under the reboiler and above the column shell, a large num-
ber of aluminum pieces were found. These were molten aluminum
which has melted, run out of the reboiler, and then resolidified. The

FIG. 8—Cavity in reboiler 共after being removed兲.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
312 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

largest piece was approximately 0.8 mm thick and 300–400 mm in di-


ameter.
• The top weld of the support plate on the side closest to the column shell
had completely failed, and the top edge of the support plate was bowed
out about 50 mm. 共Fig. 5, item 2兲 The bottom weld remained attached to
the reboiler. There was molten metal deposited on the “reboiler-side” of
the support plate, with a definite flow pattern. This pattern showed that
the molten metal sprayed in the direction from item 4 to item 5, as
shown in Fig. 5.

“D” Reboiler—
• The capsheet adjacent to the A core was severely scarfed at the elevation
of the A reboiler hole in the support plate 共Fig. 5, item 10兲. The scarfing
was most severe along the capsheet near to the 40 mm hole, extending
about 250 mm along the face of the core. The fire had scarfed through
the 6 mm capsheet and also had cut through a 5 mm tall passage. The
parting sheet side of the reboiler 共on the face of the D reboiler facing the
GAN pipe兲 had some scarfing, although this did not create any holes in
the endbars 共Fig. 6, item 13兲.

Reboiler Sump Area—


• A large amount of aluminum slag was found on the bottom head of the
low pressure column. Most pieces were 25–75 mm long and 1.5–6 mm
wide. These were found in the sump at the low points, presumably hav-
ing been washed down from their initial positions during the filling and
emptying of the low pressure 共LP兲 column during the attempted restart.
• Molten aluminum was sprayed in thin sheets or drops on the column
wall downstream of the scarfing of the D core 共Fig. 5, item 11兲.
• The support beams and column shell had molten aluminum sprayed on
them from the support plate failure 共Fig. 5, item 6兲.
• A small amount of molten aluminum spray was found on the bottom of
the bottom tray above the reboilers.
• On the side of the reboilers facing the GAN pipe from the high pressure
共HP兲 column, the support channel beams form a sealed box area. The
weld connecting the C and D reboilers had a burn hole through them,
apparently from fire that traveled along this box 共Fig. 6, item 12兲.

Repairs
A pressure test revealed that the first 11 O2 passages were leaking on the A
reboiler. To repair this core, 16 O2 passages were removed from service by
converting the adjacent N2 passages to O2 passages. The D reboiler first passage
obviously would not hold pressure because the first passage had been scarfed.
Three O2 passages were converted to N2 service. Two drain holes were drilled in
the bottom of the accessible support plates, which effectively vents the remain-
ing cavities on the bottom support plates. The top support plates were not
accessible so they were not drilled out.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 313

Note that when repairing reboilers, do not create 共a兲 sealed cavities or 共b兲
areas that can accumulate HCs by pool or dry boiling.

Analyses

Laboratory Analyses
A number of samples were obtained during the repair. The molten metal recov-
ered from the sump was analyzed and found to be resolidified aluminum, with
some samples also having Si and O. When transported, some gray dust came
loose from the samples. The elemental analysis of the dust shows Al, Si, and O,
with more O than in the slag material. Minor amounts of Al–N 共aluminum
nitride兲 and Mg were present, with the Mg believed to come from the braze
material used to fabricate the reboiler. The oxygen could have come from Al2O3
or SiO2 dust.
A sample was cut from the failed support plate 共item 2, Fig. 5兲. Resolidified
aluminum was found on the sample, with a definite flow direction. The molten
aluminum had been sufficiently hot to melt some of the parent metal of the
support plate. The failure of the top of the support occurred in the weld metal.
The small amount of weld that was attached to the sample showed some po-
rosity. No conclusions could be drawn on weld integrity due to the small
sample size.

Conclusions from Laboratory Analyses


Even with the lack of Al2O3 in the samples, aluminum combustion is still quite
possible. The Al/ O2 reaction is a vapor phase reaction, and the solid Al2O3
forms very small particles. These could then be carried out of the system during
the attempted restarts.
The presence of the gray Al dust is consistent with Al combustion since the
Al must be in the vapor phase to form such particles. The molten aluminum
slag with the gray dull surface is also consistent with Al combustion.
The presence of Al–N shows that very hot aluminum was exposed to N2.
The Al–N could either have come from Al combustion or it could have come
from the presence of very hot molten aluminum 共from HC or Al combustion兲 in
the presence of vapor N2.
In summary, the laboratory analyses were consistent with aluminum com-
bustion, and the strongest direct evidence of Al combustion was the presence of
the small dark gray Al dust particles.

Calculations
The pressure at which the support plate would fail on the A reboiler was cal-
culated to be approximately 13 800 kPa. This assumes that the support plate is
not further reinforced. Because the support plate did fail, this evidence shows
that high pressures did exist within the cavity.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
314 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

To estimate the amount of material that must have combusted, the follow-
ing line of reasoning was used:
• For each gram of Al that burns, enough heat is released to melt approxi-
mately 20 g of Al.
• About 750 g of aluminum were recovered from the column.
• It is estimated that one-half of the molten aluminum was recovered.
• It is assumed that 50 % of the heat of combustion went into melting
aluminum, and the remainder went to heating other components.
From these assumptions, it was calculated that to melt the observed alumi-
num mass, the following 共approximate兲 amounts of material must have com-
busted:
共1兲 150 g of aluminum 共needing 135 g of O2 to combust兲, or
共2兲 100 g of HC 共needing 300 g of O2 to combust兲.
The sealed cavity at 14 000 kPa 共the estimated minimum cavity pressure兲
would contain approximately 925 g of O2, which is seven times the stoichio-
metric amount of O2 needed to combust the aluminum and three times the
amount needed for HC combustion. Therefore, it is reasonable that the cavity
space could provide the needed O2 for the reaction.
A check was also done of the reboiler O2 inventory during the defrost. At
the conditions expected to exist at the time of the combustion 共i.e., sometime
while the plant was offline and being warmed兲, the reboiler could contain be-
tween 21 % and 99 % vapor O2 at ambient pressure. At these conditions, the
reboiler passages would contain only 8 % of the O2 needed to combust the
aluminum and 2.6 % of the O2 needed to combust HCs. Therefore, if gaseous
O2 were to support the combustion, a great deal of O2 would have to be circu-
lated through the burning area, which seems very unlikely.

Miscellaneous Evidence
There was no evidence that the plant operation prior to the defrost contributed
to the event:
• There was no evidence of CO2 accumulation during operation prior to
shutdown.
• The HC concentrations in the LOX surrounding the reboiler before the
shutdown were normal, 35–60 ppm as total HCs 共THCs兲.
• All acetylene tests prior to the shutdown were normal.
• The LOX purge rate was at the design flowrate.
• The reboiler submergence was at 100 %. After the incident, the trans-
mitter calibration was checked and found to be correct.

Scenario Description

Key Evidence
Summarizing the evidence that is most significant:
• There was much more molten aluminum than is seen in a typical re-
boiler incident.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 315

• Al/ O2 combustion is very unlikely with gaseous O2 or air at ambient


pressure.
• The location of the burnhole was very significant. The burnhole was low
in the core, and was on the bottom edge and weld of the support plate.
• The molten aluminum slag and fine dark aluminum powder are consis-
tent with aluminum combustion.
Based on the evidence, the most likely scenario is:
• During the time that the plant was running, LOX entered the area be-
hind the support plate, either through very small pinhole leaks, porous
welds, or through small cracks.
• The plant was shutdown, and the defrost began.
• As the defrost proceeded, the pressure in the sealed cavity increased
dramatically as the LOX warmed.
• At some point during the outage, aluminum combustion started behind
the plate. The timing and the ignition mechanism are not known. The
heat and combustion products raised the pressure, and the following
events happened essentially simultaneously:
o The fire burned through the capsheet, melting the fins. The hot burn-
ing gas traveled up and down the core, further melting the fins. The
molten aluminum ran down onto the intermediate head.
o The fire burned through the outside of the support plate. The pressure
was sufficient to cause a jet of hot burning gas to scarf the D reboiler.
The fire also entered the enclosed C channel behind the reboilers and
burned a hole in the corner of the C/D cores. Note that the enclosed C
channels are not a swept volume and were likely to contain low pres-
sure relatively pure oxygen.
o As pressure built up further behind the support plate, combustion
either started at the bottom corners of the support plate near the col-
umn shell, or traveled into the corners, and burned 12 mm holes. A
combination of high temperature and high pressure caused the support
plate to fail.
• When the support plate failed, this relieved the pressure and sprayed
aluminum up onto the column shell, reboiler core support beams, and
the LP column bottom tray.
• When the O2 pressure and/or purity were reduced, the combustion
stopped.
• The damage was immediately noticed when the plant was restarted.

Incident Number 2

Incident Description
This incident occurred at a different site 共hereafter referred to as site 2兲. The
site produces approximately 700 T/D of 99.6 % O2 in a reversing heat exchanger
ASU, starting up in 1981.
At approximately 18:30 on April 7, 2004, the oxygen purity suddenly
dropped to approximately 88 %. Prior to the incident, the plant had been run-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
316 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 9—Reboiler damage 共viewed from above兲.

ning normally. There were no process upsets, with no change in operation. The
LOX production was about 1.5 % of the airflow so the purge rate was very
satisfactory.
The plant HC levels had been normal, with no unusual excursions reported.
Typical readings are less than 30 ppm THC, with the maximum seen as 70 ppm.
The acetylene readings were also well within normal values, at or below the
LDL of approximately 0.1 ppm. The reboiler was running at 100 % submer-
gence.

Physical Evidence
The plant was shut down, and the LP column sump was entered. Of the three
reboilers, one reboiler was found to have significant damage near the lower
support 共see Fig. 9兲
There was a large bulge in the core just above the bottom reboiler support.
The capsheet had been moved out about 50 mm. Passages 1 and 2 had been
torn loose from the sidebars, and large gaps had opened up. O2 passage 1 had
a gap about 150 mm long, and O2 passage 2 had a gap about 375 mm long.
Near the support bar, N2 passage 1 and O2 passage 1 were pressed up against
the capsheet, even though the passage 1 sidebar had been pushed away from
the capsheet. This indicates that the O2 passage 1 had overpressurized first,
opening the gap. O2 passage 3 then overpressured, pushing the parting sheets
and passages against the capsheet, and opening the larger gap.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 317

FIG. 10—Support beam, viewed from underneath.

Passages 4 and higher were also crushed. The passages had been pushed in
about 50–75 mm, away from the capsheet. The first six O2 passages were dam-
aged, but the seventh O2 passage was in good condition.
During the outage, a portion of the capsheet was removed and further
studied. A portion was cut out, revealing passage 1. This showed an approxi-
mately 50 mm diameter area was missing fins. A dark gray material was
present, which was Al2O3. There was also a small amount of melted aluminum.
The melted aluminum shows that the temperatures were above the aluminum
melting point of 660° C. The Al2O3 shows that at least a portion of the heat was
supplied by combusting aluminum. Analysis showed that the fins suffered duc-
tile failure; that is, the area was overpressurized, and the fins simply tore apart.
There is a support beam welded to the reboiler capsheet. The beam ends
are attached to the column walls. The beam is approximately 300 mm from the
bottom of the reboiler. Inspection of the reboiler showed that the support beam
was welded to the capsheet on all sides, creating a sealed cavity between the
beam and reboiler capsheet.
The support beam was damaged. The weld along the top of the beam stayed
intact 共Fig. 9兲. However, the weld on the bottom of the support beam was split
open 共as seen from underneath, Fig. 10兲. The bottom of the beam was pushed
out approximately 40 mm from its initial position. The capsheet was pushed in
approximately 30 mm, and the passages underneath the capsheet must have
been crushed. The cavity between the capsheet and beam was visually in-
spected. It appeared clean. There was no dust or aluminum slag.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
318 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 11—共a兲 Removed portion, with passage 3 on top, passage 1 below. 共b兲 Sectioned
cutout showing passage 1.

Repair
The reboiler was repaired by removing 12 O2 passages from boiling service, and
converting 12 N2 passages to O2 service. As with the previous incident, the
repairs were made in such a manner as to prevent accumulating HCs or creat-
ing sealed cavities. The outer supports on all of the reboilers were modified to
be free draining to prevent LOX building pressure behind the supports in fu-
ture defrosts, which could lead to another potential incident.

Calculations
From the evidence gathered, calculations and inferences can be made.
The support beam was calculated to require 38 000 kPa to fail. The esti-
mated energy needed to rip the support beam weld and to deform the parting
and capsheets is 11.5 kJ. This can be compared to the energy stored in the
cavity. The cavity is assumed to have dimensions of 915 mm long
⫻ 175 mm wide⫻ 1.6 mm deep. The length and width are clearly known,
but the depth 共gap between the parting sheet and support beam兲 is an assump-
tion. Assuming that the cavity fails at 35 000 kPa, the mechanical energy re-
leased by adiabatic expansion is estimated at 15–30 kJ. This is reasonably good
agreement, given the approximate nature of the calculations and that not all of
the mechanical expansion energy would be consumed in tearing the weld.
There was clearly molten aluminum in the system. The molten aluminum
appears to have been generated in passage 1, and then sprayed elsewhere in the
column and heat exchanger. Figure 11共a兲 shows a portion of passages 1 through
three that were removed from the reboiler. Figure 11共b兲 shows the sample, with
passage 1 exposed. Passage 1 is clearly missing fins. Because the parting sheets
are present, the fins must have melted or burned.
As mentioned above, sidebars and parting sheets showed that during the
HC/ O2 reactions, there were two distinct energy releases, separated in time. O2
passage 1 would have overpressured first, followed by passage 2.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 319

The bright silver surface is fresh aluminum, which has been exposed by
scarfing away some of the other aluminum material. Close observation shows
what appear to be dark gray globules of Al2O3 deposited on the surface.
The timescale of the incident can also be determined. If the energy release
in the reboiler fins was relatively slow, it would be impossible for pressure to
build up in the fins. The only force holding the pressure in is the fluid above and
below the area where the reaction is occurring. One can estimate how long it
would take for the fluid to be released, and this can determine the time scale of
the event, which is approximately 30 ms for the liquid to be ejected. If the fluid
was a gas, instead, it would take 8.5 ms for the sonic wave to go up through the
core. One can estimate that the overpressurization must have taken place in
less than 30 ms. After this time, the pressure in the fin channels would fall, and
one could expect the Al/ O2 reaction to extinguish.

Most Likely Scenario

Several possible scenarios were considered. The most likely is:


• During a previous defrost, LOX trapped behind the reboiler support bar
vaporized. The support beam failed, crushing at least O2 passages 1 and
3. However, the N2 passages were not breached, or else the plant would
not have been able to make pure oxygen.
• The reboiler was returned to service.
• The crushed passages created a place for HCs to accumulate due to pool
boiling.
• Eventually, enough HCs had accumulated to react with the O2 in pas-
sage 1.
• The reaction created high pressure, and ignited the aluminum.
• LOX vaporized, creating high pressure. The core sidebars failed.
• This relieved the pressure, the O2 / Al reaction stopped.
Evidence for This Scenario:
• It is consistent with the physical evidence, including the support beam
failure and passage ruptures.
• There was clear evidence of a cavity incident: a sealed cavity 共between
the beam and capsheet兲 with a bulge, with the capsheet pushed inward,
crushing passages. For a cavity type incident, the only known method
for creating the very high pressure is to warm the plant to near ambient
temperature. There is no known mechanism for cavity incidents occur-
ring at cryogenic operating temperatures.
• This scenario explains how the crevice incident and passage leaks were
separated so far in time.
• This scenario provides a mechanism for HC accumulation, which could
then lead to the overpressure and aluminum combustion.
However, it is puzzling that there was a long time—several years—between
the last warmup of the reboiler sump and the incident. There is no known
cause as to what would have initiated the HCs accumulating at the time of the
incident, i.e., no change in air quality, operating procedures, etc. In other re-
boiler incidents caused by HC accumulations, it is typically a few days between
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
320 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

the creation of an accumulation site and sufficient HC accumulation to lead to


an energy release within the reboiler. Also, it is unusual for HCs to accumulate
so low in the reboiler, where there is little, if any, boiling of the LOX. 共At this
reboiler elevation, typically the LOX is de-subcooling, not boiling.兲

Summary: Prevention of Cavity Incidents


Air Products experienced approximately nine cavity incidents in the late 1970s
and early 1980s. Although the exact ignition mechanism was not identified,
cavity incidents have been prevented by changing our design and manufactur-
ing practices to eliminate cavities that are submerged in LOX. In manufactur-
ing new plants, a great deal of attention is paid to eliminating cavities by either
venting them or using stitch welds to allow fluid movement into and out of the
cavity.
In the 1980s, a significant program was undertaken to eliminate cavity
areas in existing plants. 28 plants were modified to eliminate various cavity
areas. The program paid special attention to eliminating cavities that could
breach the LP/HP column shells, and release cryogens from the columns into
the coldbox. Since that time, only three cavity incidents have been seen, all in
plants which were constructed before 1984. None of these have breached the
pressure envelope of the process. One of those 共discovered in 1989兲 had no
evidence of aluminum combustion. The two incidents discussed in this paper
clearly had aluminum combustion, although the combustion in incident num-
ber 2 is not believed to have occurred during the cavity incident.
Cavity incidents differ significantly from other reboiler incidents, which
accumulate HCs due to blockage of passages within the reboiler. In typical
reboiler incidents, there is little or no aluminum combustion, unless large
quantities of HCs accumulate in many places within the reboiler. Cavity inci-
dents can have aluminum combustion, but the extent of combustion is limited
by the small amount of high pressure O2 available.
The incidents described in this paper should remind all in the industry to
eliminate sealed cavities from areas which are submerged in liquid cryogens. It
is interesting to note that all of Air Products’ cavity incidents have occurred in
the reboiler sump area, which contains LOX. While liquid nitrogen 共LIN兲 would
certainly not combust aluminum, a sealed cavity submerged in LIN could po-
tentially suffer overpressure damage such as that seen in these incidents with
the support beams. The reason is not clear; perhaps it is because systems sub-
merged in LIN are less complicated, and hence do not have as many potential
cavities, or perhaps it is that cavity incidents have occurred, but have not
caused an operating problem 共such as loss of purity or leak into the coldbox兲,
which would bring the incident to the operators’ attention. In spite of the lack
of incidents in LIN, it is recommended that sealed cavities also be eliminated in
LIN service.

References

关1兴 Lunsford, K. M. “Advantages of Brazed Aluminum Heat Exchangers,” Hydrocar-


bon Processing, Vol. 75, 1996, pp. 55–64.0002-7820
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
SCHMIDT ET AL., doi:10.1520/JAI102233 321

关2兴 The Standards of the Brazed Aluminium Plate-Fin Heat Exchanger Manufacturers’
Association, 2nd ed., www.alpema.org 共Last accessed July 8, 2009兲.
关3兴 Schmidt, W. P., Kovak, K. W., Licht, W. R., and Feldman, S. L., “Managing Trace
Contaminants in Cryogenic Air Separation,” AIChE Meeting, March 5–9 2000, At-
lanta, GA, Paper T8001b.
关4兴 Schmidt, W. P. “ASU Reboiler Safety,” 15th EIGA Symposium, Production Safety,
February 2006, European Industrial Gases Association, Brussels, Belgium.
关5兴 IGC Document Doc 65/04, “Safe Operation of Reboilers/Condensers in Air Sepa-
ration Units,” European Industrial Gases Association, Brussels, www.eiga.org
共Last accessed July 8, 2009兲.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 6
doi:10.1520/JAI102252
Available online at www.astm.org/JAI

Alain Colson,1 Etienne Werlen,2 and Hervé Barthélémy3

Vacuum Superinsulated Liquid Oxygen Piping


and Vessels

ABSTRACT: Vacuum superinsulation is commonly used in industrial applica-


tions for liquid oxygen transfer piping and liquid oxygen vessels as an alter-
native to other solid insulation materials such as perlite or Foamglass®,
which have lower thermal performances. Tests have been conducted on in-
sulation materials according to ASTM D2512 共equivalent to ISO 11114–3, EN
1797兲 in LOX as well as according to ASTM G72 共equivalent to EN 1797兲 in
GOX. The superinsulation materials that have been tested are a number of
layers of aluminum foils of various thicknesses each being separated by
intermediate nonmetallic layers of various nature. Based on experience and
incidents in the industry, this paper is intended to discuss the recent test
results in pure oxygen and the possible ignition of superinsulation material in
the interspace and mechanism of propagation as well as the consequences
of a fire. Recommendations regarding material selection for vacuum super-
insulation materials in LOX applications are also provided.
KEYWORDS: insulation material, superinsulation, organic,
nonorganic, glass fiber, intermediate layers, contamination, aluminum
foil, polyester, interspace, compatibility, vacuum, adsorbent, getter,
impact, cryogenic, vessel, manufacturing, workshop, contamination,
bellow, thermal barrier, liquid oxygen, impact test

Introduction
The development of cryogenic equipment and transfer piping expanded over
the last decades in many countries. The vacuum superinsulation applied to

Manuscript received November 25, 2008; accepted for publication April 9, 2009; pub-
lished online June 2009.
1
Air Liquide International Senior Expert, Engineering and Large Industries Depart-
ment, Air Liquide, Champigny-sur-Marne 94503, France.
2
Air Liquide International Expert, Manager of the Engineering Technical Commission,
Air Liquide, Champigny-sur-Marne 94503, France.
3
Air Liquide International Fellow Senior Expert, Air Liquide, Champigny-sur-Marne
94503, France.
Cite as: Colson, A., Werlen, E. and Barthélémy, H., ⬙Vacuum Superinsulated Liquid
Oxygen Piping and Vessels,⬙ J. ASTM Intl., Vol. 6, No. 6. doi:10.1520/JAI102252.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 322
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 323

FIG. 1—Typical cross section of a vacuum superinsulated transfer line.

liquefied air gases as well as hydrogen and helium became common to many
companies and manufacturers. The purpose of vacuum superinsulation, which
was developed for liquefied gas application, is precisely to minimize heat trans-
fer by the elimination of each of the three heat transfer mechanisms, i.e., con-
vection, conduction, and radiation. This is achieved by using a high level of
static vacuum and appropriate materials for shielding against radiation, and
intermediate layers against conduction for the separation of shields. The ad-
sorption of residual gases and moisture in the volume under vacuum is
achieved using appropriate products. Therefore, the materials used for super-
insulation are selected according to their individual thermal performances and
the combination of these performances when submitted to vacuum and cryo-
genic conditions. However, some of the materials used for superinsulation must
be carefully selected for the reason that the submission of some of these mate-
rials to an oxygen enriched atmosphere can lead to violent reaction with pos-
sible ignition and catastrophic consequences. The liquid oxygen transfer piping
and vessels having vacuum superinsulation must be designed and constructed
so that such situation is avoided.

Background

Previous Work
Test results on materials used for superinsulation were provided by Barthélémy
et al., according to Ref 1. These tests were intended to test 7 ␮m and 15 ␮m
and associated Lydall® series 1303/243 for application on cryogenic vessels.
However, the development of various thicknesses of aluminum foils as well as
intermediate layers and recent incidents led to the development of a new test
program.

Typical Vacuum Superinsulated LOX Transfer Line


For high thermal performances, the design of superinsulated liquefied gases
addresses the three types of heat exchanges, i.e., convection, conduction, and
radiation. The appropriate performance is met using nonconductive materials
for intermediate layers and very thin aluminum foils having the role of screens
against radiation and vacuum for convection avoidance. A typical superinsu-
lated transfer line for LOX and other liquefied air gases is an assembly of
different components, as shown in Fig. 1:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
324 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Cut view.

共1兲 outer pipe,


共2兲 inner pipe,
共3兲 thermal barrier,
共4兲 expansion bellow,
共5兲 spacer,
共6兲 vacuum port,
共7兲 superinsulation,
As an example, the real construction of this type of superinsulated transfer
line taken from a workshop is shown in Fig. 2 using these components.

Fire with LOX Portable Tanks


Several accidents and incidents occurred with ignition and catastrophic conse-
quences with vacuum superinsulated LOX transfer piping and tank, although
this equipment is supposed to be used in cryogenic conditions for static instal-
lations as well as for transport. One accident occurred when a portable LOX
tank was filled with LOX. For some reason it toppled down, a few seconds after
being lifted up it exploded. This accident led to one fatality and injury. The
understanding of this accident is the following.
共1兲 The heater coil 共which is by construction within the vacuum insulation
space兲 had been weakened by some corrosion or stress problem 共poor
brazing or construction兲.
共2兲 The toppling down of the tank led to the rupture of this heater coil and
flow of oxygen gas within the insulation space. The oxygen immedi-
ately reacted with the palladium metal, which was present in the
vacuum space or more probably the palladium hydride, resulting in the
normal operation of the getter 共palladium oxide or similar兲. This
chemical reaction is violent and exothermic enough to initiate the com-
bustion of the aluminum foils of the superinsulation.
共3兲 The combustion of thin foils of aluminum in pure oxygen is violent and
similar to an explosion.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 325

FIG. 3—Exploded LOX transfer piping.

Fires with LOX Transfer Vacuum Jacketed Piping


Apart of LOX portable tanks, accidents occurred also on static LOX transfer
piping. In a previous incident, for some reasons presently not clarified, the
vacuum of a LOX transfer piping was very likely broken from the inside pipe of
this LOX transfer piping and the oxygen came in contact with the superinsula-
tion. Ignition and violent combustion occurred, causing the burst of large sec-
tions of the outer pipe. Figure 3 shows the LOX transfer piping just after the
incident.
However, in addition to the mechanical effect and consequent damages
that occurred to the LOX transfer piping and surroundings, as shown in Fig. 3,
it is noteworthy to consider that in one location where the outer pipe did not
burst, the inner and outer pipes burned through and that almost all the super-
insulation material were burned; Fig. 4 shows a piece of piping where the burn
through occurred. This situation has been seriously considered and used for
further analysis, test, and orientation for design and construction of LOX trans-
fer piping systems as it has been verified that the intermediate layer used for
superinsulation was flammable in air, that aluminum foil having a thickness of
8 ␮m was used.

Common Materials Used for Vacuum Superinsulation


Different materials are available and used for the design and manufacturing of
vacuum superinsulated LOX transfer piping and tank. These materials are sup-
plied by manufacturers for cryogenic application 共air gases, hydrogen, and he-
lium兲 using superinsulation and are known to have the capability to meet the
appropriate thermal performances when combined in a suitable number of
layers, i.e., typically from 15 to 25 layers for air gases and more depending on
the application. Although these materials are not specific for oxygen applica-
tion, some of them exhibit the capability to burn and propagate combustion in
air or in oxygen enriched atmosphere.
For example, the Lydall® series of materials are given to be recommended
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
326 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Burn through of a LOX transfer piping.

for use with liquid oxygen. Table 1 gives a nonlimited list of typical materials
available and used for superinsulation.
It is noteworthy that materials given in Tables 1 and 3 and the correspond-
ing information was provided directly by six different manufacturers of

TABLE 1—Typical materials used for vacuum superinsulation.

Identification Type of material


Paper “tee bag,” cotton, linen or other
Paper equivalent vegetal/organic fibers
Inorganic glass fiber nonwovens 共density of
0.19 g/cc, tensile strength of 0.31 kg/25 mm,
Lydall® 233 and basic weight of 16 g / m2兲
Intermediate layer Inorganic glass fiber nonwovens 共density of
0.18 g/cc, tensile strength of 0.27 kg/25 mm,
Lydall® 243 and basic weight of 12 g / m2兲
Inorganic glass fiber nonwovens 共density of
0.18 g/cc, tensile strength of 0.27 kg/25 mm,
Lydall® 1303 and basic weight of 14 g / m2兲

7 ␮m Aluminum foil, 7 ␮m thick


8 ␮m Aluminum foil, 8 ␮m thick
Shield 9 ␮m Aluminum foil, 9 ␮m thick
12 ␮m Aluminum foil, 12 ␮m thick
15 ␮m Aluminum foil, 15 ␮m thick

Other Mylar® Aluminized 900 Å polyester


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 327

TABLE 2—Elements found in an 8.2-␮m aluminum foil.

Elements found in an 8.2-␮m Percentage


aluminum foil 共%兲
Copper 共Cu兲 1.4
Iron 共Fe兲 0.6
Silicon 共Si兲 0.2
Zinc 共Zn兲 0.02
Manganese 共Mn兲 0.02
Chromium 共Cr兲 0.01
Titanium 共Ti兲 0.01
Silver 共Ag兲 ¯
Magnesium 共Mg兲 ¯
Aluminum 共Al兲 97.7

vacuum superinsulated LOX transfer piping. The suitability of these materials


is discussed below.
It should be noted that for vacuum insulated transfer piping of fluids hav-
ing a boiling point below the dew point of air 共e.g., liquid nitrogen below 2
barg兲, the compatibility of the materials used for the superinsulation should
also be considered. As a matter of fact, in case of loss of integrity of the outer
pipe, air would condensate and form an oxygen rich liquid mixture.

Purity of Materials Used


After the incident corresponding to Figs. 3 and 4 occurred, the analysis of the
superinsulation materials was performed. The results were that “paper” was
used as the intermediate layer and aluminum foils having a thickness of
8.2 ␮m with the elements given in Table 2.
Additionally, even if the materials used for intermediate layers are sup-
posed to be 100 % inorganic, it appears that a very low quantity of organic
binder can be present as it can be necessary for the mechanical stability of
structure and manufacturing of the material itself.
In other words, these investigations lead to the conclusion that some ma-
terials present in the superinsulation can be flammable when submitted to
oxygen enriched atmosphere and provided that energy is present for ignition.

Vacuum Pump and Evacuation


The evacuation is achieved using at least primary vacuum pumps and in most
cases secondary vacuum pumps for vacuum superinsulated LOX transfer pip-
ing depending on the required level of vacuum in given conditions. The vacuum
pumps used for such purpose utilize specific lubricants and the system for
evacuation is designed so that no contamination of the vacuum space is pos-
sible during the evacuation.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
328 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Sources of Ignition of a Vacuum Superinsulated LOX System

For LOX applications, the usual approach is to consider the mechanical impact
as the most common/important source of energy for ignition. However, it ap-
pears that at least three of the most common ignition mechanisms should draw
attention as they can be active in a LOX system when flammable materials are
present.

Mechanical Impact or Rupture


When a single or repeated impact loading occurs, the presence of porous ma-
terials and possibly a nonmetallic material increases the probability of ignition.
The rupture of a piping component or any other equipment wetted by oxygen
may provide the energy for ignition. This mechanism is one of the most com-
monly discussed with LOX systems.

Mechanical Friction
Two or more rubbing surfaces, rapid motion, and mechanical load can lead to
ignition. The probability is increased when high load is applied and/or with
aluminum alloys. Oscillation and vibration frequency also increase the prob-
ability of ignition.

Promoted Combustion and Kindling Chain


A source of energy is supposed to be present as well as easily flammable mate-
rials to kindle and propagate combustion. This usually applies to nonmetallic
materials, which can be capable to ignite a less flammable material.

Understanding the Ignition Mechanisms in Vacuum Superinsulated LOX Piping


The understanding of ignition mechanisms and propagation in vacuum super-
insulated LOX piping should consider the following elements for a scenario of
a fire, which depends on the materials used and on the mechanical design of
the LOX piping system. The unexpected possibility for LOX to come in contact
with superinsulation materials is the major hazard addressed.
• Rupture of the internal piping due to a weak point, i.e., bellows, instru-
ment, coil, weld, excessive mechanical stresses in static, and dynamic
conditions with consideration of the pressure thrust force.
• Dry vaporization or dead-end boiling of LOX leading to the accumula-
tion of trace hydrocarbons. The hydrocarbon issue with LOX is not ad-
dressed in this paper even if it can be stated that dead ends and low
point should be avoided for the design of a LOX transfer piping system.
• Mechanical impact from outside 共job site works, maintenance, etc.兲.
• Materials in the vacuum space, which are reactive with liquid or gas-
eous oxygen 共i.e., activated charcoal, palladium oxide, etc.兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 329

FIG. 5—Standard aluminum cups used for impact test in LOX with 50 J energy.

Testing Samples of Materials Used for Vacuum Superinsulation


Based on incidents reported and the above elements, a new test program has
been decided in order to evaluate the suitability of materials given in Table 1.

Testing Procedure
As said above, because the most common ignition mechanism is supposed to be
a mechanical impact, one of the selected test procedure was according to ISO
11114–3/EN1797 关4兴, which is equivalent to ASTM D2512, “Standard Test
Method for Compatibility of Materials with Liquid Oxygen 共Impact Sensitivity
Threshold and Pass-Fail Techniques兲” 关2兴 where a minimum of 20 drops is
required. Some unexpected results on the first samples obtained with the strict
application of the standard test methodology and using an energy of 100 J with
a standard striker pin led to perform a new series of tests in the exact same
conditions with the exception of the energy for the impacts that was reduced to
50 J. This is a deviation to the standard test procedure 关2,4兴, but violent reac-
tions have still been observed inside the cup. Figure 5 shows an example of
residuals inside the cup after being tested: no ignition at 50 J 共left兲 and ignition
at 50 J 共right兲.

Selection and Handling Samples to Be Tested


The test program used samples taken directly from workshops of manufactur-
ers of vacuum superinsulated LOX transfer piping. A number of 20 layers of
superinsulation were considered as a standard 共Fig. 6兲 and the samples were
prepared for tests using a cutter and white gloves for handling the samples and
avoiding contamination.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
330 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Typical superinsulation materials, 20 layers used for test samples.

Additional Test Procedure


After the test results according to the above procedure were obtained, it ap-
peared that supplementary compatibility tests of nonmetallic materials used as
intermediate layers were necessary. The standard autoignition test methodol-
ogy was used according to ASTM G72, “Standard Test Method for Autogenous
Ignition Temperature of Liquids and Solids in a High-Pressure Oxygen-
Enriched Environment,” 关3兴. The reasons for using the AIT 共the autoignition
test兲 is that it gives a good indication of the flammability of nonmetallic mate-
rials in gaseous oxygen atmosphere and allows us to detect whether the inor-
ganic intermediate layers would contain organic and flammable material or
not. The AIT of aluminum foils has not been detected because this test is lim-
ited to a temperature up to 500° C and would not ignite aluminum foils.

Test Results and Interpretation


Tables 3 and 4 below give the AIT and impact test results. The test results given
in Table 3 are summarized below.
• For all combinations of aluminum foils and intermediate layers, reac-
tions have been observed when the standard impact energy of 100 J was
applied. According to the standard test procedure 关2,4兴, they would be
classified as noncompatible for use in LOX.
• A nonstandard energy of 50 J applied by impact in LOX led to much less
reaction. Some combinations did not ignite for a series of 20 impacts.
• The combinations of aluminum foils and intermediate layers have been
classified by order of decreasing compatibility, based on the impact test
results. It must be noted that due to the low number of tests, this clas-
sification does not have an absolute nature.
• The lowest number of reactions has been obtained with the thinner foils
共7 – 9 ␮m兲. It is believed that other parameters not related to aluminum
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 331

TABLE 3—Impact test results.

Reactions Reactions
with with
100 J 50 J Ref Origin
Aluminum Intermediate impacts impacts test of
thickness layer 共% nb兲 共% nb兲 CTEa samples
8 ␮m Paper 20 共4/20兲 0 共0/20兲 746 and 766 C
9 ␮m Lydall® 233 35 共6/20兲 0 共0/20兲 757 and 762 H
7 ␮m Lydall® 243 40 共8/20兲 0 共0/20兲 756 and 761 D
15 ␮m Glass fiber 50 共10/20兲 0 共0/20兲 768 and 776 IU
15 ␮m Lydall® 1303 40 共8/20兲 15 共3/20兲, 5 共1/20兲 763 to 765 F
N.A. Aluminized Mylar® 100 共20/20兲 ¯ 635 ¯
a
Centre de Technologie et d’Expertise of Air Liquide, Le Blanc Mesnil, France.

thickness explain this observation. Within the range tested, the alumi-
num thickness does not seem to be a significant parameter for the com-
patibility.
• The intermediate paper layer used for Test 766 was not ignited by im-
pact at 50 J. However, this material is flammable in air, has a low AIT,
and was used in the LOX pipes of the incident corresponding to Figs. 3
and 4. It is also easily wetted and impregnated by LOX. For all these
reasons, paper or similar organic intermediate layers must be avoided.
The result of Test 763 using 15 ␮m and Lydall with 100 J impact energy
was surprising because it is not consistent with the test results published in Ref
关1兴. A new test series was performed using a sample from the same origin but
only 50 J impact. Because the ignition still occurred despite a lower energy, a
second test series was performed with 50 J impact but the surface of the alu-
minum foil and intermediate layers were verified to be smooth and clean. One
ignition occurred under this last test series. It must be noted that the samples
used for these three series had been prepared at a workshop for assembly using
flammable adhesive tape. Contamination during fabrication is suspected to be
an aggravating factor, even with the combination having a priori the best com-
patibility. This is the demonstration that precautions must be taken and clean-
liness must be preserved while handling superinsulation materials to avoid con-
tamination.
Mylar exhibits 100 % of ignition in LOX at 100 J. Mylar is flammable also
in air. Aluminized polyester foils must be avoided.

TABLE 4—AIT test results.

AIT
Intermediate layer Origin of materials 共°C兲
Lydall® 243 ¯ None
Lydall® 1303 ¯ None
Paper C 237.8
Paper C 237.8
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
332 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Fragments and pleated superinsulation materials.

One element of the interpretation of test results as developed by Bar-


thélémy et al. according to Ref 1 was related to the concentration of energy
where glass fibers or particles are present under the striker pin due to micro-
scopic contact points sufficient to trigger a reaction between the fresh metal
surface and LOX. The particles can be created when dust, fragments, or pleats
of the aluminum or intermediate layers appear during the manufacturing pro-
cess and handling of the superinsulation, as shown in Fig. 7, or when the se-
lection of the intermediate layer leads to the presence of raw and undesirable
thick fibers, as shown in Fig. 8.

Adsorbent and Getter

Adsorbent
In addition to superinsulation materials, adsorbent and getter are used to
maintain the vacuum as high as possible over time after the liquid oxygen
piping system was delivered by the manufacturer, i.e., typically in the range of
10−3 – 10−4 mbar. The adsorbent has the function to capture residual moisture
and other impurities. For example, the use of Molecular Sieve Type 13X or
equivalent is suitable for liquid oxygen transfer lines as it contains only nonor-
ganic materials, such as silicon dioxide 共⬍60 %兲, aluminum oxide 共⬍40 %兲,
sodium oxide 共⬍20 %兲, and magnesium oxide 共⬍5 %兲.
On the other hand the use of activated charcoal as adsorbent for liquid
oxygen transfer lines presents major hazards if the vacuum of the liquid oxygen
transfer line is broken from the inner piping leading to contact with pure oxy-
gen. It is known that charcoal in the presence of liquid oxygen is an unstable
combination, which can result in a powerful explosion. Powerful explosions
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 333

FIG. 8—Aluminum foils and fibered layers.

can also occur with charcoal in the presence of some oxygen rich liquid mix-
tures 共e.g., liquid air兲 and therefore activated charcoal should not be used as
adsorbent for liquefied air gases.

Getter
The getter effectively removes hydrogen gas typically present in many hermeti-
cally sealed applications such as superinsulated LOX transfer lines and tanks
where the presence of such gas would affect the level of vacuum and perfor-
mances over time. The use of Pd–O 共palladium oxide兲 and other oxides are also
used alternatively as the getter. However, the use of Pd–O is commented from
the oxygen safety standpoint with regard to the physical properties of palla-
dium.
Pd – O + H2 → Pd + H2O
As an example for one standard superinsulated LOX transfer piping seg-
ment, based on the volume of vacuum the getter was comprised of 0.25 g of
palladium oxide 共Pd–O兲, which is a platinum group of metal. Figure 9 shows an
example of Pd–O getter wrapped with appropriate materials.
However, and as explained above, accidents and incidents occurred where
the getter using palladium oxide was reported as part of the root cause. The
Pd–O getter reduces back to nearly pure palladium metal as the hydrogen 共out-
gassing from aluminum foils and stainless steel walls兲 grabs the oxygen from
the getter and then is absorbed into the Molecular Sieve. The incidents, which
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
334 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 9—Getter installed and protected.

occurred with Pd–O, had a specific configuration where the Pd–O getter was
likely taped directly to or came in contact with the aluminum foils of the su-
perinsulation. In this configuration, if for any reason some liquid oxygen leaks
into the vacuum jacket, then the reduced Pd exothermically reoxidizes, gener-
ating enough local heat to cause some melting of the aluminum foil leading to
aluminum/oxygen ignition and combustion propagation. Experiments were
made with temperatures up to 400° C and red points appeared when oxygen
came in contact with palladium.

2Pd + O2 → 2PdO
For the reason explained above, a getter using PdO 共alternatively TiO兲
should be avoided where possible and especially on superinsulated LOX trans-
fer lines where the performance drop is less compared to a vacuum superinsu-
lated LOX tank and where the risk of rupture of vacuum is much higher com-
pared to a LOX tank with regard to the number of piping elements having
individual sealed vacuum. However, the use of the getter for a vacuum LOX
tank should be carefully addressed.
The installation of Pd–O was worked out using a wrapping getter in brass
or nonflammable metallic mesh as to prevent any heat to develop during the
reoxidation process if failure of the vacuum was to occur 共Fig. 9兲. Stainless steel
mesh should not be used for such purpose. In addition, the solution is also to
attach the getter to surfaces other than the aluminum foil to provide some
separation to the aluminum.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
COLSON ET AL., doi:10.1520/JAI102252 335

Discussion and Conclusion


In addition to the conclusion given by Ref 1, which stated that the likelihood of
ignition increases as the thickness of aluminum foil decreases, it appears that
different types of materials i.e., aluminum foils and type of intermediate layers
from various manufacturers and workshops, have the capability to ignite when
conditions are met regardless of the thickness of the aluminum foils and even
when lower energy is provided by impact. Once energy is provided by impact or
any other sources, such as those described above, and if oxygen enrichment is
present in the vacuum space, then ignition of thin aluminum foils may occur
with the capability to kindle close and thicker but flammable materials, leading
to extremely violent combustion of these materials with consequences such as
those described in the Introduction of this paper. Therefore, particular atten-
tion should be paid to the following issues for the design and construction of
vacuum superinsulated liquefied air gas transfer lines and vessels with consid-
eration of hazards related to possible presence of pure oxygen due to mechani-
cal problem leading to the rupture of the inner pipe. The possibility of oxygen
enriched atmosphere in the vacuum space due to rupture of the outer pipe and
condensation of air should also be considered but is less critical.

Basic Design Calculation and Selection of Components


As for any other vacuum superinsulated equipment, the basic design calcula-
tion should consider the mechanical stress calculation based on cryogenic ser-
vice temperature. The most critical parts where special attention should be
paid are the bellows installed on the inner pipe, spacers, and supports. In order
to minimize the risk of inner pipe loss of containment, the design calculation
should be done so that the maximum pressure thrust force is sustained, that
the acceptable loads on the bellows are not exceeded, and that inner pipe flex-
ibility is sufficient. The design and materials for bellows and spacers should be
carefully selected and verified. In addition, the connections between pipe sec-
tion 共e.g., bayonet connections, if any and instead of the most common welded
connections兲 and the bellows should be designed and mounted in such a way
that they do not create sealed or dead-end boiling areas.

Material Selection for Superinsulation


According to Refs 5 and 6 related to storage of cryogenic air gases, the compo-
nents used for insulating materials shall be such that the finished product is
suitable for oxygen service. Given the test results where the thickness of alumi-
num foils combined with intermediate layers, it appears that the situation and
cleanliness of these materials have a dominant role for the capability of an
aluminum foil to ignite. A low thickness such as 7 ␮m or 9 ␮m of aluminum
will be more problematic to handle and install for multilayer superinsulation,
where a greater thickness such as 15 ␮m has more mechanical resistance and
easier for manufacturing. However, the issue of heat losses by conduction is
opposed to this aspect. Based on the above and on the large industrial experi-
ence of safe operation of vacuum insulated LOX pipes and vessels using alumi-
num foils, having a thickness comprised between 7 ␮m and 15 ␮m, it is be-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
336 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

lieved that such aluminum foils can be safely used provided that the other
materials that are necessary for superinsulation are properly selected.
• The intermediate layer must be 100 % nonorganic and aluminized poly-
mers such as Mylar must not be used for LOX applications.
• The adsorbent must be inorganic for LOX applications and if oxygen
rich liquid mixture may be present in the vacuum space.
• The need and use of a getter as well as a type of getter should be care-
fully addressed for LOX transfer piping systems.

Manufacturing, Cleaning, Inspection, and Control

As for any oxygen application, manufacturing process including cleaning and


cleanliness preservation is a critical aspect for the vacuum space. Precautions
should be taken and inspection plan set in place in a workshop in order to
prevent contamination of any kind of the vacuum space equipment including
the insulation materials. This includes the verification that all nonsuitable resi-
dues from manufacturing steps 共i.e., working fluids and others兲 have been
eliminated.

References

关1兴 Barthélémy, H., Roy, D., and Malzoumian, N., “Ignition of Aluminum by Impact of
Lox—Influence of Contaminants,” Steinberg, T. A., Newton, B. E., and Beeson, H.
D., Eds., Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres:
Ninth Volume, ASTM STP 1395, ASTM International, West Conshohocken, PA,
2000.
关2兴 ASTM, Standard D2512, “Standard Test Method for Compatibility of Materials
with Liquid Oxygen 共Impact Sensitivity Threshold and Pass-Fail Techniques兲,” An-
nual Book of ASTM Standards, ASTM International, West Conshohocken, PA.
关3兴 ASTM, Standard G72, “Standard Test Method for Autogenous Ignition Tempera-
ture of Liquids and Solids in a High-Pressure Oxygen-Enriched Environment,”
Annual Book of ASTM Standards, ASTM International, West Conshohocken, PA.
关4兴 ISO 11114–3/EN 1797 Cryogenic Vessels—Gas/Material compatibility.
关5兴 EIGA IGC 115/04 Storage of Cryogenic Air Gases at User’s Premises.
关6兴 EN 13468 Cryogenic Vessels: “Static Vacuum Insulated Vessels” 共Part 1, “Funda-
mental Requirements;” Part 2, “Design, Fabrication, Inspection, and Testing;” Part
3, “Operational Requirements”兲.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102304
Available online at www.astm.org/JAI

Barry E. Newton1 and Ted Steinberg2

Adiabatic Compression Testing—Part I:


Historical Development and Evaluation
of Fluid Dynamic Processes Including
Shock-Wave Considerations

ABSTRACT: Adiabatic compression testing of components in gaseous oxy-


gen is a test method that is utilized worldwide and is commonly required to
qualify a component for ignition tolerance under its intended service. This
testing is required by many industry standards organizations and govern-
ment agencies. This paper traces the background of adiabatic compression
testing in the oxygen community and discusses the thermodynamic and fluid
dynamic processes that occur during rapid pressure surges. This paper is
the first of several papers by the authors on the subject of adiabatic com-
pression testing and is presented as a non-comprehensive background and
introduction.
KEYWORDS: adiabatic compression, pneumatic impact, gaseous fluid
impact, isentropic compression, oxygen, shock-wave heating

Introduction
The compressed gas industry and government agencies worldwide have utilized
one primary test methodology for qualifying high-pressure valves, regulators,
and other related flow control equipment for gaseous oxygen service. This test
methodology is known by various terms including adiabatic compression3 test-

Manuscript received December 17, 2008; accepted for publication June 12, 2009; pub-
lished online July 2009.
1
Vice President R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las
Cruces, NM 88007.
2
Professor and Director, Queensland Univ. of Technology, S-Block, LVL-10, GPO Box
2434, Brisbane, Qld, 4001.
Cite as: Newton, B. E. and Steinberg, T., ⬙Adiabatic Compression Testing—Part I:
Historical Development and Evaluation of Fluid Dynamic Processes Including Shock-
Wave Considerations,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102304.
3
While various terms are used for the type of testing discussed herein, adiabatic com-
pression testing is the term that will be used most frequently in this paper. This term is
chosen not because it is an accurate description, but because it is used most widely
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 337
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
338 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ing, gaseous fluid impact4 testing, pneumatic impact testing, and BAM5 testing
as the most common terms. Generally speaking, adiabatic compression is
widely considered as the most important ignition mechanism for directly kin-
dling of a non-metallic material in oxygen and has been implicated in many fire
investigations. The temperature rise by near-adiabatic compression has com-
monly been calculated by assuming ideal gas behavior through the polytropic
equation6

Tf ª Ti · 冉冊pf
pi
共k−1兲/k
共1兲

where:
Tf = final temperature 共abs兲,
Ti = initial temperature 共abs兲,
Pf = final pressure,
Pi = initial pressure,
k = ratio of specific heats for oxygen 共1.4兲,
considering isentropic behavior 共reversible and adiabatic兲.
The predominant test methodology that is normally utilized and a means of
evaluating the thermal profiles 共i.e., temperature versus time兲 for various test
systems has been discussed in another publication 关1兴 by these authors. This
paper outlines the historical development of the test method and discusses
some of the fluid dynamic processes that are being considered in an effort to
fully describe the test. This paper is not comprehensive but instead attempts to
broadly outline the historical development of the test method and especially
discuss the approaches that have been used by practitioners to estimate the
temperatures produced during a pressure surge cycle when the test is con-
ducted. This temperature profile and whether it differs from one test system to
another is of primary interest to the authors and will be the subject of both
measurement and modeling in subsequent papers on this subject.

within the industry. It is actually the methodologies irreversibility’s and non-adiabaticity


that this research program is evaluating.
4
Gaseous fluid impact is the officially balloted description in ASTM International Test
Method G74, Standard Test Method for Ignition Sensitivity of Materials to Gaseous
Fluid Impact 关5兴.
5
BAM stands for Bundesanstalt für Materialforschung und–prüfung and is the German
Federal Institute for Materials Research and Testing where the test methodology origi-
nated in the 1950s. The test method was also implemented by the National Aeronautics
and Space Administration 共NASA兲, in a somewhat different form, after the 1970s.
6
The temperature produced by adiabatic compression is usually calculated using isen-
tropic relationships assuming that the oxygen behaves like an ideal gas and that the
compression process is sufficiently rapid that heat transfer does not occur during the
short time of the pulse 共i.e., essentially adiabatic兲. This form of the equation is normally
used to calculate the final temperature.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 339

Historical Development and Background of Adiabatic Compression


Testing

The hazard associated with compression heating of oxygen in components and


systems has long been known in the industry. The 1983 keynote address by
Neary 关2兴 during ASTM G04’s first technical symposium celebrated the release
of ASTM Standard Guideline G63 关3兴, which was a guide for selecting materials
for oxygen service. Neary celebrated ASTM Guide G63 as “the industry’s first
guide” for evaluating materials for oxygen service. In this paper Neary reported
that the Compressed Gas Manufacturers Association 共CGMA, later shortened to
the Compressed Gas Association兲 formed an Oxygen Regulator Research Com-
mittee in 1921 due to fires caused by adiabatic compression of oxygen. Neary
reported that the first product of the newly formed CGMA industry committee
in 1923 was a report on oxygen regulator fires that recommended two principle
test methods, “the combustion 共autoignition兲 test” and “the heat of 共adiabatic兲
compression” test. Neary indicated that after the 1980 release of ASTM Guide
G63, the ASTM committee’s focus shifted to the release of ASTM Standard Test
Method G72, “Determination of Autogenous Ignition Temperature of Liquids
and Solids in a High-Pressure Oxygen-Enriched Environment” 关4兴 and ASTM
Standard Test Method G74, “Test to Determine Ignition Sensitivity of Materials
to Gaseous Fluid Impact” 关5兴. These standards essentially became the first
industry-wide implementation of the 1923 recommendations of the CGMA
Oxygen Regulator Research Committee. However, while the standards were a
positive step toward the implementation of test methods to improve oxygen
safety, the discussions in the standards do not go beyond the common isentro-
pic relationship 关Eq 1兴 for specifying the temperature of the compressed gas.
Werley 关6兴 provided an insightful review of adiabatic compression testing in
his 1993 paper “A Perspective on Gaseous Impact Tests: Oxygen Compatibility
Testing on a Budget.” In the background section of this paper Werley described
the substance of the ASTM G04 committee discussion pertaining to ASTM
Standard G74’s development. He indicated that in the early 1980s when the
committee was drafting G74, the members were aware of test apparatus uti-
lized by companies such as AIRCO, RegO, AGA, and Circle Seal as well as
government testing agencies such as BAM and National Aeronautics and Space
Administration 共NASA兲. He pointed out that some practitioners felt that adia-
batic compression testing “ignited everything” and other practitioners felt that
the test was insensitive and only ignited materials like polytetrafluorethylene at
elevated pressures. Werley indicated that in the early 1980s the only active
members of the ASTM G04 committee that conducted this test were NASA,
AGA, and Circle Seal and that among these institutions NASA had conducted
more extensive work. As a result, the ASTM G04 committee chose to depict the
NASA apparatus in the standard, but, the test parameters were selected to be
consistent with the other apparatus capabilities as well.
Adiabatic compression ignition was alleged in many fires in the industry
throughout the years and was an ignition mechanism utilized in much material
and component testing. In 1993 Koch 关7兴 reported in a paper on Oxygen System
Safety the results of five different fire investigations. He admits in this paper
that “the primary emphasis is on adiabatic compression, which has been iden-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
340 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

tified as a significant but often overlooked cause of oxygen fires.” In Koch’s


opinion, adiabatic compression should have been implicated in even more fires
than it had been. In this paper he provided the common methodology for cal-
culating the theoretical maximum temperature by use of the isentropic rela-
tionship. In 1997 Koch 关8兴 remembered the Neary reference to the 1923 CGMA
paper that implicated adiabatic compression as a “common cause” of fires.
Koch went on to identify adiabatic compression as the ignition source in sev-
eral other fires including U.S. Navy training facility dating to the 1970s and
opined that adiabatic compression as an ignition source must be “century-old
knowledge” since Linde, Hampton, and their peers, who developed air-
separation technology to produce oxygen, must have “understood the essentials
of what would cause an oxygen fire.”
The ASTM G74 test system was heavily utilized by NASA-White Sands Test
Facility 共WSTF兲 关9兴, who at that time was the only NASA center that conducted
adiabatic compression testing consistent with G74. In the early 1990s, at the
request of the Circle Seal Corporation, Wendell Hull and Associates, Inc.
共WHA兲, developed a similar test system patterned after the NASA system but
also consistent with the predominant industry standards 关10兴 in Europe. In
Europe, at that time, the test systems of prominence were operated by BAM
关11,12兴 and Air Liquide, Center for Technical Expertise 共CTE兲 关13–16兴. Binder et
al. 关12兴 at BAM provided a good description of their test system in their 1995
paper and included the statement that “this method has been well established
in evaluating oxygen equipment and is required in Germany by Deutches Insti-
tut für Normung, Comite Europeen de Normalisation 共CEN兲 standards, and
even by ISO standards.” Wegener et al. 关11兴 described the temperature rise in
the compressed gas and the influence on ignition as follows: “A compression of
oxygen at 20° C from 0.l to 2.5 MPa yields a temperature rise to 410° C 共this can
easily be calculated according to Poisson’s equation兲. This temperature is
higher than the ignition temperature of most organic substances, so that gas-
kets 共as seat gaskets, stuffing boxes, and piston rings兲, lubricants, hydraulic
fluids, and so forth are ignited and can burn in an explosive manner if exposed
to an adiabatic compression of oxygen. Such oxygen impacts may happen, for
example, in pipes if shut-off fittings under pressure are opened too rapidly or in
reciprocating compressors. In general, however, such compression processes
do not take place adiabatically so that temperature peaks are obtained that lie
between the initial temperatures and theoretically calculated maximum tem-
peratures.”
The temperatures indicated by Wegener et al. are easily obtained through
use of Eq 1, as provided earlier.
Air Liquide has made significant contributions to the way in which adia-
batic compression testing is currently being carried out and to the development
of criteria to increase the test severity and improve the reliability 关11–16兴. Bar-
thelemy et al. reported in 1988 while discussing flexible hose ignitions that
“another 共ignition兲 explanation proposed was an ‘adiabatic compression’ pro-
cess; when a gas is compressed rapidly, it increases in temperature. The theo-
retical final temperature when oxygen is compressed, assuming the process is
adiabatic 共for example, assuming no mixing with hose gases, no shock-waves,
and no heat transfer to the hose or containers walls兲, is calculated from the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 341

共familiar isentropic relationships兲.” The assumptions provided by Barthelemy


are commonly assumed and considered valid for very rapid compression pro-
cesses. However, the assumption of “no shock-waves” is important and will be
further discussed later. Indeed, Air Liquide performed a shock-wave analysis of
the compression process in 2000 关17兴 that will be discussed in a later section.
Air Liquide recognized in 1989 that results could differ between test labo-
ratories and therefore altered its internal test procedures to be more severe
than the predominant standards and achieve more conservative results 关13兴.
The work reported in this paper was foundational to several changes that were
eventually incorporated in the predominant International Standards 关18兴 and
European Norm Standards 关19兴 including a test pressure of 1.2 times the work-
ing pressure of the component and installation of the test article downstream of
an impact tube of specific dimensions.7 This research along with the advocacy
of Air Liquide led to the very wide subscription of the industry standards to
these provisions. Today, most industry standards that require adiabatic com-
pression testing 共see Table II in Ref 1兲 utilize the test parameters that originally
appeared in the 1983 version of ISO 2503 as modified by the recommendations
of Air Liquide after this work was published. The only other industry standard
that was not modified with these provisions was ASTM G74, which was main-
tained by NASA-WSTF. Further, no industry standard provided any guidance
pertaining to the calculation of the temperature or thermal energy in the com-
pressed gas other than the isentropic relationship 共Eq 1兲.
Adiabatic compression testing has been utilized heavily by NASA 关20–23兴.
NASA-WSTF used the ASTM G74 test methodology for individual non-metallic
materials and valves for both material selections, batch qualification of non-
metallic materials, and to evaluate components such as regulators and Teflon®-
lined flexible hoses 关9,21,22兴. Stradling et al. 关9兴 provided an early 共1983兲 dis-
cussion of the NASA uses for ASTM G74 and as the NASA designer/originator
of the test method provides his insights into the usefulness of pneumatic im-
pact testing.8 Hirsch et al. 关21,23兴 provide a history of how NASA used the test
method between the mid-1970s and 2003. While the statistical approach uti-
lized within NASA for testing non-metallic materials has been questioned 共dis-
cussed below兲, Hirsch et al. pointed out that as long as statistically rigorous
methods are utilized in the collection of ASTM G74 data, the results are not
only meaningful but even produce a strong correlation between the 50 % reac-
tion pressure9 and the autogenous ignition temperature 关4兴 of the material
being tested 关23,25兴.
As mentioned above, variability observed in the non-metallic material test

7
Test articles normally installed on a cylinder are tested at the end of a 5 mm inside
diameter tube that is 1 m long. Test articles normally installed on a manifold are tested
at the end of a 14 mm diameter tube that is 750 mm long.
8
Stradling was a charter member of the ASTM G04 committee and worked alongside
Neary and others to propel this ASTM committee and its standards into worldwide
prominence in oxygen.
9
The 50 % reaction pressure is the pressure at which 50 % of the test samples react as
determined by a statistically rigorous methodology known as the Bruceton up-down
method 关23,24兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
342 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

data produced by ASTM G74 in the late 1980s and early 1990s caused NASA-
WSTF to conduct several test programs to study the statistical aspects of quan-
tal 共go/no-go; ignition/no-ignition兲 type testing 关26–28兴. This testing changed the
way that NASA utilized ASTM G74 testing due to its clarification of the statis-
tically low confidence produced by the manner in which the tests were being
performed. Normally, the ASTM G74 testing was performed to rank a material
according to the pressure at which a non-metallic material achieves zero 共0兲
reactions in 20 successive pneumatic impacts.10 Hirsch summarizes the prob-
lem well, “An analysis of the cumulative binomial probabilities for the ASTM
G74 procedure indicated that for a probability of reaction of 0.05 共assumed兲 for
a single trial, the probability of obtaining zero reactions in the 20 trials pre-
scribed by the standard logic is about 36 % 关23兴. As a result, the lack of preci-
sion with the G74 test logic could be potentially misleading when results were
used to rank or qualify materials for oxygen service.”
For the purposes of this background, however, the statistical aspects of
ignition are not as interesting as the thermodynamic principles discussed in
this research. In 1988 Schmidt et al. 关26兴 attempted to evaluate the test meth-
odology by using an instrumented test chamber. In this instrumented test
chamber 共pressure surge volume兲, they included a fast response pressure trans-
ducer to record the pressure rise rate, a photocell to record the light emission
from an ignition, a special fast response thermocouple called an “eroding bead
thermocouple11” 关29兴 for measuring the temperature produced in the pressure
surge. Schmidt proposed that the following ignition mechanisms might be ac-
tive during the pressure surge to ignite a non-metallic material located at the
dead end of the pressure system:
共1兲 adiabatic compression of the oxygen in the test chamber before impact,
共2兲 adiabatic compression of a bubble of gas trapped within the test mate-
rial,

10
In reality, a “passing” pressure level was achieved by either zero 共0兲 reactions in 20
successive pneumatic impacts or one reaction in 60 pneumatic impacts.
11
This type of thermocouple is made by the NANMAC Corporation and is fabricated of
very fine films of two metals, such as chromel and alumel films for Type K, encased in an
aluminum-oxide and stainless sheath. The thermocouple sheaths are open at the end so
that the end can be polished thereby “smearing” the two metal films together to form a
junction. The film thickness once polished develops a junction with a time constant
proportional to the polished film thickness. In certain applications the time constant is
in the microseconds according to NANMAC. They are referred to as eroding bead ther-
mocouples since in an application measuring combustion temperatures, they will erode
or burn but will continuously re-make their junction. Based on WHA experience and
discussions with Nanigian, who holds the patent for these thermocouples, they do not
work well in the application envisioned since a film of cold gas forms over the junction
interfering with the sensation of heat in the compressed gas. The WHA tests with these
thermocouples included different shapes 共i.e., spherical ends and wedge shaped ends兲 in
an effort to resolve this problem. However, results similar to NASA-WSTF were achieved
where only a small temperature rise in the gas was measured. The principle of measure-
ment for these thermocouples is provided in Ref 34.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 343

共3兲 heating of the test material by mechanical compression or mechanical


shear,
共4兲 interaction of shock-waves with the test specimen,
共5兲 a combination of several of the above mechanisms.
For our purposes, the potential for shock-wave development during the
compression process is of interest since the gas velocity and temperature are
not the same behind a shock-wave as behind a compression wave having the
same pressure ratio 关30兴. Indeed, shock processes are fundamentally different
form isentropic compression and would lead to different features of a model
seeking to define the state conditions of the test gas. The NASA interest in
shock-wave development within the compressed gas was heightened during the
testing by Janoff et al. 关22兴, Pedley et al. 关31兴, and Forsyth et al. 关32兴, who had all
observed brief flashes of light within tubes undergoing a compression process
sometimes 100 or 200 ms before a combustion event developed.12 In 1987 Ped-
ley et al. discounted these flashes as ignition due to inadvertent contamination
in the tubes they were testing. Janoff et al. and Forsyth et al.13 however evalu-
ated the phenomenon further due to the unusual nature of the light emitted
when the installation of pre-cleaned empty tubes also produced light emission
on several occasions. Forsyth et al. theorized that the light emission could be
due to the emittance of sodium or potassium spectra, in visible wavelengths,
from the pre-cleaned stainless tubes. He indicated that “a related cause is a
phenomenon known as ‘double electron transfer,’ or the release of energy in the
form of photons resulting from electrons in the closely packed oxygen mol-
ecules changing states.” They went on to indicate that this phenomenon has
been theorized to occur in oxygen at pressures above 69 bars. They said that
“despite exhaustive efforts to characterize the emmittance, including installa-
tion of band pass filters of various wavelengths in front of the photocell, the
detection of the phenomenon was too inconsistent to characterize.”
Janoff et al. 关22兴 theorized that the light flashes resulted from shock ioniza-
tion of the oxygen and used band-pass filters corresponding to the ionization
wavelengths of 410, 440, and 480 nm, which corresponded to transitions in the
molecular structures of 01, 011, and O2+, respectively. They captured the flash
on high-speed film and provided a series of frames that demonstrate a flash
lasting about 24 ms in the visible-light spectrum. Their use of the three band-
pass filters indicated above along with a high-pass, ⬎700 nm, infrared filter
resulted in their concluding that the flash of light contained all three wave-
lengths expected from the shock ionization of oxygen. They further indicated
that the flash event contained little, if any, infrared emission and contained only
wavelengths of 700 nm and below. They ultimately opined that the flash could
be attributed to the shock ionization of oxygen during the compression process
where pressurization rates are fast. They theorized that the shock ionization of
the oxygen may play a role in the ignition process lowering the required acti-

12
Usually the lower the pressure, the longer the period between the flash of light and the
development of a visually observed combustion front.
13
Personal communication with Forsyth revealed that he had performed such rigorous
cleaning and cleaning verification of his test tubes that he was confident that the phe-
nomenon was not due to combustion of a contaminant within the tube.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
344 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

vation energy for ignition and making the oxygen more active. The pressuriza-
tion rates where these flashes were studied were of the order of 14 ms, the
fastest attainable with the WSTF system. By comparison, the pressurization
rate where the light flash was observed by Forsyth was 20 ms. Since Janoff’s
research involved the ignition of flexhoses by rapid compression and since ig-
nition of flexhoses by pneumatic impact was also observed at pressurization
rates of 200 ms, Janoff et al. concluded that adiabatic compression of the gas
probably provided the primary thermal energy for the ignition process and they
related the temperature rise to the isentropic relationships.
The NASA-WSTF G74 evaluations 关26–28兴 all ultimately concluded that the
thermal energy in the compression process was produced by a standard isen-
tropic compression of the gas rather than by shock-wave influences. Schmidt et
al. stated that “because the ignition occurs late in the pressurization cycle,
shock-waves, of which there is evidence only in the first 5 ms of pressurization,
are probably not responsible. Further evidence for this conclusion comes from
the actuation pressure study that suggests that relatively rapid pressurization
does not favor ignition.” Schmidt et al. had observed that “the pressure-time
curve measured by the dynamic pressure transducer was always steepest in the
first 3–5 ms, indicating possible incipient shock-wave formation.” Thus they
evaluated the influence on the valve opening time on the ignition frequency and
ultimately concluded that “the frequency of ignition in the instrumented cham-
ber was higher when the valve opening speed was slower.” However, the range
of opening speeds for the impact valves they used were from 1.6 to 6.85 ms,14
which are not considered substantially different when compared to the
⬃50 ms pressurization time normal to the ASTM G74 procedure and assumed
to have been used by Moffett based on the pressure rise graphs shown in his
paper.
Janoff et al. 关28兴 observed that increasing the volume of compressed gas
between the high-speed valve and the test sample significantly increased the
probability of ignition of the test sample by a pressure surge. They also showed
that the ignitions were achieved reliably at 180 ms pressurization rates even
though the reaction frequency decreased from the higher frequency at 18 ms
pressurization rates. These observations were related by Janoff et al. to the
theoretical temperatures produced by isentropic compression of the gas.

14
Later studies by Moffett et al. 关26兴 reported valve opening times from 7.8 to 16.4 ms.
The actual pressurization times were not reported. Janoff et al. 关27兴 reported pressuriza-
tion times of 18–180 ms. Janoff et al. also reported that “in the method currently used by
NASA, the pressurization time is between 50 and 60 ms.” He indicated that the 18 ms
pressurization time was the fastest that could be achieved in the system, although the
system Janoff et al. used was larger in volume than the one used by Moffett et al. or
Schmidt. The 180 ms pressurization time was accomplished by placing a metering valve
in the line between the high-speed valve and the test specimen. He showed the ignitions
were achieved reliably at 180 ms pressurization rates even though the reaction frequency
decreased from the higher frequency at 18 ms pressurization rates. The metering valve
would be expected to significantly degrade a coalescing shock-wave.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 345

Shock-Wave Heating or Isentropic Compression Heating

The role of shock-waves in a pressure surge consistent with the predominant


test systems utilized today is still unknown. It is understood that the NASA
project funding was limited and did not allow for research to be conducted
much beyond that stated above. However, the fact that light emission was ob-
served on at least three separate projects in pre-cleaned empty stainless tubes
and that band-pass filters detected the emission at wavelengths consistent with
shock ionization of oxygen indicates that at a minimum further evaluation of
shock processes would be appropriate.
The present research intends to further evaluate this question empirically
by appropriate tests in the future. However, since the influence on the thermal
profile applied to a test article could be substantial, depending on whether the
shock is weak or strong, the following brief background on shock-wave pro-
cesses pertaining to temperature rise in the gas was developed. Whether a fully
coalesced shock-wave can be produced in the process under consideration is
uncertain, but, based on NASA experiences, consideration of even weak shock
processes should be evaluated as part of this research. Indeed, the question of
shock processes has been raised by other oxygen practitioners, as indicated by
Ducrocq et al. 关17兴
In 2000 Air Liquide presented a fluid flow analysis of the gaseous impact
test conducted at CTE 关17兴, which considered the system as a shock tube. In
this research the investigators used both one- and two-dimensional numerical
computer codes to evaluate the reason for the pressure overshoot observed so
frequently in the pressure-time data for these tests 共see Fig. 1; discussed in Ref
1; note the pressure oscillation on each test system兲. The most common expla-
nation for this overshoot is an under-damped transducer responding to a step
input, as shown in Fig. 2 共reproduced from Ref 33兲.
The Air Liquide researchers sought to explain the behavior of these pres-
sure transducers instead through the use of flow processes considering the
superposition of running compression/shock-wave共s兲 and reflected expansion
or rarefaction wave共s兲, as in a typical shock tube analysis. They used a one-
dimensional numerical simulation code and successfully predicted the general
shape of the oscillating pressure pulse both in the overshoot amplitude and the
order of the oscillation frequency. Their simulation predicts the overall shape of
the pressure oscillation through superposition of multiple compression/
expansion wave interactions. Significantly, this approach considers the entire
system design including the driving gas accumulator, the tube connecting the
accumulator to the high-speed valve, and the impact tube to predict the tran-
sient pressure history in the system from a step change 共i.e., opening of the
high-speed valve兲. They pointed out that differences in the pressure oscillation
should be observed for differently designed systems. By reference to Fig. 1, this
is exactly what WHA has observed during its testing at the different laborato-
ries 关1兴. In fact, the pressure-time history appears to be unique for each system
tested thus far.
The Air Liquide approach also allowed for a temperature history to be
predicted utilizing one-dimensional shock tube theory; however, very steep
temperatures 共⬎2500 K兲 were predicted as a result of the propagation of the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
346 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

WHA 200 barg - Positions 1-4 (0.001 thermocouple)


500 250

225

400 200

175

Pressure (barg)
300 150

125

200 WHA Temp Position 1 100

WHA Temp Position 2 75


WHA Temp Position 3
100 50
WHA Temp Position 4
WHA Pressure 25

0 0
0.2 0.3 0.4 0.5 0.6 0.7

BAM 200 barg - Positions 1-4 (0.001 thermocouple)


500 200

180

400 160

140
Temperature (C)

Pressure (barg)
300 120

100

200 BAM Temp Position 1 80


BAM Temp Position 2 60
BAM Temp Position 3
100 40
BAM Temp Position 4
BAM Pressure 20

0 0
0.2 0.3 0.4 0.5 0.6 0.7

Time (sec)

FIG. 1—WHA and BAM pressure and temperature profiles.

normal shock. Temperatures of this magnitude are not expected; otherwise,


ignition of non-metallic test samples would occur during testing with much
higher frequency. Further, the Air Liquide researchers attempted to measure
the temperature rise with standard thermocouples and measured peaks of ap-
proximately 520 K 共247° C兲. Because of the lower temperature measured with
thermocouples, they also used a two-dimensional simulation program to study
the influence of mixing due to vortex generation during the reflection of a
shock-wave at the end of the impact tube. They theorized that this condition
would mix the hot “shocked” gas with the cooler gas along the boundary layer
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 347

150

Overshoot
Percent Output Change

100

Final Output
Value
63%

50
Initial
Output
Value

0
−2 −1 0 1 2 3 4 5 6 7 8 9 10

Time (units)

FIG. 2—Response of under-damped transducer to step change 关32兴.

of the tube. The result of this simulation, for their conditions 共200:1 pressure
ratio兲, predicted that the gas moving along the axis was cooled to approxi-
mately 600 K, while the gas in the hot plug moving along the wall was still
approximately 1357 K. These temperatures are greater than those measured
but have decreased as expected.
Certainly, a shock-wave analysis might be capable of explaining some fea-
tures of the transient process such as the pressure-time history of the system
and, if valid, would contribute to the thermal energy of the compressed gas.
WHA has observed additional support for this approach through its tempera-
ture measurements as shown in Fig. 3 共see Ref 1 for measurement details兲. This
figure depicts the response of a thermocouple array having bead diameters of
0.013 mm 共0.0005 in.兲, 0.025 mm 共0.001 in.兲, and 0.051 mm 共0.002 in.兲 placed at
the dead end of a volume being rapidly compressed. Each of these thermo-
couples, to varying degrees, seems to exhibit a tendency to respond thermally to
the pressure oscillation being recorded by the dynamic pressure transducer.
This is especially true for the 0.013 mm diameter thermocouple. Clearly, if the
oscillation on the pressure transducer was merely an under-damped response
of the transducer to the step pressure input, the thermocouples would not be
expected to record a corresponding temperature oscillation that rises and falls
somewhat in general agreement with the pressure oscillation. While this does
not indicate that a shock-wave produced the thermal variations, it does support
the conclusion that the pressure oscillation may be real and that an explanation
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
348 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

600 300

540

480 240

420
Temperature (C)

Pressure (barg)
360 180

300

240 120

180
1-0.013-mm TC
120 60
1-0.025-mm TC
60
1-0.051-mm TC
Pressure (barg)
0 0
0 0.056 0.111 0.167 0.222 0.278 0.333 0.389 0.444 0.5

Time (sec)

FIG. 3—WHA test: 180 barg. Test 13: Cycle 1.

for this profile should be part of the overall physical model that is developed
during this research.
A review of shock tube processes was undertaken to evaluate the nature of
the physical phenomenon that might develop during an adiabatic compression
test 关30,34–39兴. The adiabatic compression test system may be envisioned as a
simple shock tube with the high-speed valve acting as the diaphragm separat-
ing the high-pressure driving gases from the low pressure driven gas. In this
case, however, the diaphragm opening time is much longer and of the order of
10–15 ms as compared to diaphragm rupture times of 600 ␮s common to
shock tubes 关30兴.
In a simple shock tube the processes may be envisioned as shown in Fig. 4
关34,37兴. These processes may be imagined, to some extent, for the adiabatic
compression test. In a shock tube, when the diaphragm ruptures both shock
and expansion waves are generated. The shock-wave travels into the low pres-
sure gas 共driven gas section兲 and the expansion wave travels into the high-
pressure gas 共driving gas section兲. A contact surface is also formed across
which the pressure and velocity are constant, but the temperature and density
共hence the Mach number兲 are different. In Fig. 4, illustration “A” shows the
condition just prior to diaphragm rupture. Illustration “B” shows the condition
at time= t1, where the shock-wave and contact surface have traveled a distance
into the driven gas section and have influenced the gas properties according to
the generalized temperature and pressure graphs shown. In illustration B the
movement of the expansion waves is also shown as the pressure is disturbed in
the driving gas section to depress the total pressure somewhat. The expansion
waves move into the driving gas chamber at the velocity of sound for the un-
disturbed medium, region 4. The shock-wave moves into the driven gas, region
1, and depending on the initial pressure ratio across the diaphragm may accel-
erate to speeds greater than the speed of sound of the undisturbed driven gas.
When the shock-wave encounters the end of the tube section it will reflect
at more than twice the magnitude of the incident pressure step. The expansion
waves will also reflect when they encounter the end of the driving gas section
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 349

Contact Reflected
Reflected Expansion
Surface Shock
Waves
(2) C
(6) (3) (5)
(6) (5)

t Shock
t1
(3) (2)

(4) (1)
(2)
x
(3)
Expansion Contact
Waves Surface Shock

(4) (3) (2) (1) B


(4) (1)
T2
T4 T3 T1
P4
P3 P2 Reflected P5
P1 Shock
P2, P3
P6

(4) Driving Gas Section Driven Gas Section (1)


High Pressure Low Pressure P1
P4 A
Diaphragm
P1 C

FIG. 4—Simplified illustration of shock tube processes related to rapid compression.

and will travel at the velocity of sound of the medium plus the medium velocity.
The conditions for reflection and the resulting change in pressure are illus-
trated in “C” in Fig. 4.
It was the superposition of some of these dynamics that Air Liquide argued
caused the pressure oscillation observed. However, since the high-speed valves
do not open as rapidly as a diaphragm rupture, these processes cannot be
imagined to proceed completely as described. White 关30兴 indicated that in re-
ality even the rupturing of a diaphragm, fast though it is, would be expected to
produce a series of compression waves, which must coalesce into a shock-wave
at some distance from the diaphragm rupture. If that is true for a diaphragm
rupturing in 600 ␮s, then it is certainly of greater influence for a valve opening
in 10–15 ms. The process described by White is illustrated in Fig. 5.
If it is assumed that a shock-wave is formed by the coalescence of multiple
compression waves that have been formed by the rupturing of the diaphragm
or the opening of a valve, then White argued that the shock will form at a point
as shown in Fig. 5. White argued that as the diaphragm is rupturing, a series of
compression waves is sent out, each one heating the gas by compression as the
individual disturbances are traveling into the driven gas section. Since each
compression wave heats the gas slightly, the speed of sound for the next com-
pression wave is higher, and therefore that compression wave will have a
slightly higher velocity. Each compression wave produced as the diaphragm is
rupturing travels at a slightly faster velocity than the last. Eventually each of
these compression waves will coalesce with the first, and if the magnitude of
the initial pressure ratio across the diaphragm is great enough and as long as
the driven section is long enough, a shock-wave will form.
Figure 5 illustrates this process in a three-dimensional depiction. The
driver and driven sections are shown along with their respective initial pres-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
350 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Contact
Surfaces
4 Shock Wave

Time 3 E 2
C

1 Coalesced Shock
Pressure Driver
Section
Driven Section

Distance

FIG. 5—Illustration of shock formation after diaphragm rupture 关29兴.

sures. The time axis increases into the page showing the change that occurs
after the diaphragm ruptures. Each time step is illustrated along with the as-
sociated change in pressure and movement of individual compression waves
into the driven section. The development of a contact surface and the move-
ment of the expansion into the driver section are also illustrated starting at the
time the diaphragm is caused to rupture. The individual compression waves are
imagined to coalesce as shown in the diagram after several time steps have
occurred. Each compression wave is moving faster than the last due to the
increase in local gas temperature caused by the previous compression wave.
Over time, these compression waves catch the first and strengthen it until a
fully developed shock forms, if the driven section is long enough. Once the
shock coalesces, a new contact surface and expansion wave are formed as new
disturbances in the driven gas. At this point in the flow system, the usual prop-
erties as illustrated in Fig. 4 again apply where, with reference also to Fig. 5,
p2 = pe just as p3 = pc, but T2 ⫽ Te and Tc ⫽ T3. At this point in time, T1 = T4 but
T2 ⫽ Tc and now P2 ⫽ P3. After the shock-wave coalesces, the process is no
longer considered isentropic since part of the mechanical energy is converted
irreversibly to heat by the shock-wave.
White’s model allows for the calculation of the time required for the shock-
wave to build up through successive compression waves, one catching the
other; if the temperature of the compressed gas is calculated in small pressure
steps by Eq 1 and the sonic velocity is calculated by means of the usual rela-
tionship for local gas properties:

k·p
a2 = 共2兲

where:
a = local speed of sound,
k = ratio of specific heats 共Cp / Cv兲,
p = local gas pressure, and
␳ = local gas density.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 351

The model described by White is essentially that which Becker developed,


described in detail by Lewis and von Elbe15 关36兴 in the formation of a shock-
wave in a long tube. The relationships developed here will allow a comparison
of the temperature from a shock to be compared directly to isentropic compres-
sion. In the shock-wave there appear entirely different relationships of tem-
perature and pressure than those governing the usual adiabatic 共isentropic兲
compression.
If the unit of mass is compressed in an ordinary isentropic manner, which
may be envisioned as Becker did by enclosing it in a cylinder and moving an
imaginary piston against it sufficiently slowly so that the pressure, p, through-
out the gas is at each moment is equalized and smaller than the pressure on the
face of the piston by an infinitesimal amount, then the increase in the internal
energy, ⌬Eu, by the change in volume, dV, would be shown in Eq 3 关36兴 as
follows:

⌬Eu = 冕 v1

v2
− pdv. 共3兲

This is the usual relationship for energy change in a unit volume by “p-dv”
work assuming conditions are adiabatic. Therefore the energy change in the
gas 共ideal兲 and the temperature developed by compression are easily found by
these familiar terms.
However, for a shock-wave the different relationships of temperature and
pressure must be considered, and it is useful to evaluate the temperature dif-
ferences that might exist as compared to isentropic compression. Becker devel-
oped his equations by considering a unit mass of gas in front of the wave
having the volume v1 and pressure p1 before the shock-wave passes and v2 and
p2 after being compressed by the wave. In this case the work done is p2 共v1
− v2兲, since after the establishment of the wave the pressure on the piston is
always p2. This work would both increase the internal energy of the unit mass
and impart to it kinetic energy so that the change in internal energy is ex-
pressed as

w2
⌬Eu = p2 · 共v1 − v2兲 − 共4兲
2
where:
w = velocity change of the disturbed gas 共u1 − u2兲 to an observer moving
with the wave.
To this observer, the gas enters the wave with velocity u1 and leaves at a
smaller velocity u2.
Becker then developed his mass, momentum, and energy relationships
using fundamental steady state relationships as:

15
The discussion that follows pertaining to the Becker analysis is largely based on mate-
rial presented by Lewis and Von Elbe 关36兴, which discusses the model and analysis
originally presented by R. Becker in Z. Physik Journal in 1922 共Z. Physik 8, 321兲 and later
Z. Elektrochem Journal in 1936 共Z. Elektrochem. 42, 457兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
352 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

u1 u2
= 共5兲
v1 v2

u 12 u 22
+ p1 = + p2 共6兲
v1 v2

u 12 u 22
E1 + + p1 · v1 = E2 + + p2 · v2 共7兲
2 2
As can be seen, the change in energy from Eq 7 is very different from Eq 3 and
is not applicable to flow in which the pressure and volume changes are isentro-
pic 共reversible兲. Becker derived Eq 7 from the energy theorem for flow where
resistance occurs.
By substituting values of u21 and u22 from Eqs 5 and 6 into Eq 7, one can
obtain the famous Hugoniot equation for which this type of compression re-
places the integral in Eq 3 for isentropic compression

1
E2 − E1 = ⌬E = 2 共p1 + p2兲 · 共v1 − v2兲 共8兲
Lewis and Von Elbe 关36兴 pointed out that the physical interpretation of the
mechanism by which the gas entering the wave front is compressed according
to Eq 8 and not according to the ordinary adiabatic relationship follows as long
as it is remembered that during an isentropic compression process the com-
pression takes place so slowly that the pressure in the unit mass control volume
is at all times equal 共i.e., the external force on the piston imagined above is only
infinitesimally larger than the opposing force exerted by the gas兲. This will be
the case as long as the piston velocity is small compared to the average molecu-
lar velocity 共therefore the piston velocity can be quite high in actuality as long
as these conditions hold兲. However, when the piston velocity becomes on the
order of the magnitude of the molecular velocity, the degradation of the kinetic
energy of the piston into random molecular velocities 共i.e., thermal energy兲
contributes to the internal energy of the compressed gas. For very small volume
changes the Hugoniot equation reduces to the differential form of the isentro-
pic equation dE = −pdv.
Faeth and Meyer 关40兴, in an excellent discussion of isentropic compression,
related this condition to the wave relaxation time, L / a, where L is the length of
the driven gas section and a is the velocity of sound. For isentropic properties
to be valid, the rate of compression must be slow enough that the change in
pressure with distance 共␦p / ␦x兲 is negligible. Faeth and Meyer indicated that the
time of compression tc must be much longer than the quotient L / a 共tc Ⰷ L / a兲;
otherwise the pressure in the driven gas section cannot be assumed to be isen-
tropic and the pressure varies with position in the tube.
From Eqs 4–6 the velocity of shock propagation into the gas at rest, u1, and
the velocity w of the gas behind the wave, often referred to as the particle
velocity, is found from Eqs 9 and 10. Equation 11 is the ideal gas law, where n
is number of moles per unit mass and R is the molar gas constant, and Eq 12
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 353

TABLE 1—Shock waves in oxygen 共k = 1.4兲 for different pressure ratios.

w u1 T2_shock T2_isentropic
p2 / p1 v1 / v2 共m/s兲 共m/s兲 共K兲 共K兲
2 1.63 175 452 336 330
5 2.84 452 698 482 426
10 3.88 725 978 705 515
50 6.04 1795 2150 2260 794
100 7.06 2590 3020 3860 950
1000 14.3 8560 9210 19 100 1710
2000 18.8 12 210 12 900 29 000 2070
Note: Reference 35 points out that the values of Cv used by Becker in these calculations are
not accurate at very high temperatures, but the essential trend is the same.

relates the internal energy of an ideal gas to its change in temperature, with Cv
being the average specific heat at constant volume 共between T1 and T2兲, as
follows:

u1 − v1 · 冑 共 p 2 − p 1兲
共 v 1 − v 2兲
共9兲

w = 共 v 1 − v 2兲 · 冑 共 p 2 − p 1兲
共 v 1 − u 2兲
共10兲

p·v=n·R·T 共11兲

⌬E = Cv · 共T2 − T1兲 共12兲


From these relationships, Table 1 was presented by Lewis and Von Elbe and
provides the comparison between shock temperatures and isentropic tempera-
tures that we were seeking. As can be observed, the shock-wave produces a
temperature that is very different from ordinary isentropic compression for the
same pressure ratio. At low pressures the magnitude is similar but for higher
pressure ratios the difference is significant. Therefore, as a minimum, this re-
search must evaluate the presence and strength of any shock-wave that might
develop from the rapid opening of the high-speed valve. It is considered prob-
able that because the high-speed valve opens in a time much longer than a
diaphragm ruptures, shocks do not fully coalesce before the compression waves
reflect at the dead end of the driven gas section. However, certainly some com-
pression waves would be expected to catch and strengthen the leading com-
pression front and thereby create a pressure disturbance in the driven gas that
could be similar to a partially formed shock. Leslie pointed out that for a 1 m
long tube and for a sonic velocity of 350 m/s, the wave relaxation time ta
= 2.8 ms. So, for a 15–20 ms target pressurization time as required by the
present standards, tc ⬃ ta, and the condition for isentropic compression may not
be achieved. Based on this result also, the presence of partial shock conditions
共i.e., strong compression waves兲 should at least be evaluated in any model that
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
354 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

is developed to predict the thermal profile produced by the compression pro-


cess.

Real Gas Properties


One final adjustment to the temperatures estimated for the compressed gas in a
pressure surge has been suggested by several researchers. It has been recog-
nized that temperatures predicted by the polytropic equation 关Eq 1兴 and shown
in Table 1 共T2_isentropic兲 are based on ideal gas behavior and overestimate the
actual temperature if real gas properties were considered. Since an accurate
prediction of the thermal energy in the compressed gas tube 共driving gas
recompression+ driven gas compression兲 is desired in this research, then
evaluation of the state of the gas during the compression process utilizing real
gas relationships and equations of state would be useful. Recently, several re-
searchers 关41–43兴 have adjusted the polytropic exponent 共pvk = constant兲 by
empirical measurements or by considering the compressibility and change in
specific heats of oxygen 共Ref 41兲 to predict the temperature rise using the poly-
tropic relationship adjusted for some real gas properties. By adjusting the ex-
ponents, under very specific conditions, these researchers have shown that the
temperature developed by compression of a real gas may be calculated using
this simple relationship. However, the exponent derived by this approach is
only valid for the specific conditions under which it was developed and is not a
true equation of state for the real gas properties. Its use must be confined to the
circumstances in which it was developed. For instance, Leslie reported that for
the form of the polytropic equation shown in Eq 13, three values of the poly-
tropic exponent may be derived for Tinitial = 300 K and Pinitial = 100 kPa, as fol-
lows:

冉 冊冉 冊
Tfinal
Tinitial
=
Pfinal
Pinitial
n
共13兲

Leslie Eq 1: n = 0.2829: Ideal gas, ordinary isentropic value, k = 1.4.


Leslie Eq 2: n = 0.2599: Ideal gas with variable specific heats.
Leslie Eq 3: n = 0.2632: Real gas 共van der Walls兲, variable specific heats.
The resulting calculations, starting from the initial conditions given above,
are shown in Fig. 6, on a log-log chart to linearize the behavior. This figure
compares two different real gas approaches to calculating the compressed gas
temperature compared to the normal isentropic approach using Eq 1. As is
evident in this figure, the results of the calculation show that the polytropic
exponent for ideal gas with variable specific heats and a real gas model using
van der Walls’ relationship and variable specific heats result in very similar
temperatures, for the starting conditions chosen. At high pressures, the real gas
temperatures predicted diverge from the ideal gas isentropic predictions.
Barragan et al. 关44兴 also derived closed form equations of state for oxygen
from thermodynamic principles utilizing the Peng–Robinson equation of state
and the general entropy equation. Figure 7 compares the temperatures calcu-
lated by Barragan et al. 关44兴 and Leslie 关41兴 for isentropic compression of oxy-
gen considering real gas behavior and variable specific heat to the familiar
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 355

10000
Temperature (K)

1000

Ideal Gas - Constant Specific Heat


Ideal Gas - Variable Specific Heat
Real Gas - Varible Specific Heat
100
0.1 1 10 100

Pressure (MPa)

FIG. 6—Isentropic temperature–pressure relations 关40兴.

2500

2000
Temperature (K)

1500

1000

Tfinal = 552*P0.2474
R2 = 0.999
500
Ideal Gas - Constant Specific Heat
Real Gas - Barragan Approach [44]
Real Gas - Leslie Approach [41]
0
0 20 40 60 80 100

Pressure (MPa)

FIG. 7—Temperature for isentropic compression of oxygen using real gas 关44,41兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
356 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

isentropic relationship for ideal gas 共Eq 1兲. Thus, Fig. 7 directly compares the
van der Walls equation of state 共Ref 41兲 to the Peng–Robinson equation of state
共Ref 44兲.
The van der Walls equation of state is probably the best known and the
oldest such modification to the ideal gas law being originally presented in 1873
as a semi-theoretical improvement to the ideal gas relationship 关45兴. However,
the van der Walls equation is known to deviate from real gas behavior at tem-
peratures significantly over a gases critical temperature and pressure. The
Peng–Robinson equation of state was developed in 1976 to provide better ac-
curacy above the critical point, especially for calculations of the compressibility
factor. As can be seen in Fig. 7, the two approaches deviate from one another by
about 3–7 % between 20 and 60 MPa, respectively. However, the temperature
predicted by Eq 1 共ideal gas, constant specific heat兲 significantly over-predicts
the theoretical gas temperatures due to compression heating by 14–19 % be-
tween 20 and 60 MPa, respectively.
Barragan et al. argued that the calculation of Treal by this approach, con-
sidering both the heat capacity variation with temperature and the effect of
pressure, provides the best value that can be obtained by thermodynamic
analysis.

Summary and Conclusions from Background Research

This paper sought to outline the historical development of the gaseous fluid
impact 共or adiabatic compression兲 test method and discussed some of the fluid
dynamic processes involved, and to outline some of the considerations that will
be evaluated by the authors in further testing and research to estimate the
temperature and energy developed during a pressure surge. The temperature
profile in the compressed gas and whether it differs from one test system to
another are of primary interest to the authors and will be the subject of both
measurement and modeling in subsequent papers on this subject. This paper
sought to provide background for this research and is the first in a series of
papers planned on this subject.
The second paper in this series is presented herein as Ref 1 and presents a
measurement technique for the determination of the temperature profile in a
typical pressure surge. Initially, the research attempted to define differences in
the various test systems by the temperature profile; however, due to limitations
in the ability of thermocouples to respond quickly to transient thermal changes,
the presence or non-presence of shock-waves could not be verified by tempera-
ture alone, as will be discussed in Ref 1. Additional testing and modeling will be
utilized in later research to attempt a resolution of this important question.
Based on consideration of the background discussed above, the following
general conclusions have been drawn, pertaining to the estimation of tempera-
ture in the compressed gas:
共1兲 No research that we are aware of has empirically measured the thermal
energy in the driven gas section for the methodology required today by
the predominant standards. Several researchers have attempted mea-
surement including NASA-WSTF, Air Liquide, WHA, and Faeth and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 357

Meyer 关40兴 but temperatures do not compare favorably 共temperatures


are significantly lower兲 to the temperatures estimated by either isentro-
pic or shock methods. Faeth and Meyer used a unique approach, fur-
ther discussed in Ref 1, and produced measurements closer to those
expected than other researchers; however, their systems were larger
and pressurized much more slowly than the systems under consider-
ation here.
共2兲 An industry consensus has not been developed as to what thermody-
namic and/or gas flow processes are causing the increase in thermal
energy that leads to ignition of a non-metallic material by this test.
Heating by frictionless-adiabatic 共isentropic兲 compression and shock
tube methods are both alleged as the predominant processes involved.
However, neither the presence of shock-waves nor the irreversibilities
of the compression process have been defined so that the temperatures
actually produced have been determined.
共3兲 The fact that light emission was observed on at least three separate
NASA projects in pre-cleaned empty stainless tubes and that band-pass
filters detected the emission at wavelengths consistent with shock ion-
ization of oxygen indicates, at a minimum, that further evaluation of
shock processes would be appropriate. On the other hand, the rela-
tively long opening time for the valves commonly used in this test, as
compared to diaphragm rupture times for shock tubes, calls into ques-
tion whether a fully coalesced, or even strong, shock-wave could de-
velop in the distance between the valve and a test article 共usually 1 m兲.
Therefore, the role of shock-waves in a pressure surge consistent with
the predominant test systems utilized today is still unknown. Since the
existence and/or strength of a shock-wave in a typical pressure surge
cycle is unknown, additional testing and modeling are required to fur-
ther evaluate the presence and strength of shock processes in order to
determine their influence 共if any兲 on the temperature produced in the
compressed gas.
共4兲 The following temperature predictions for the compressed gas volume
that predominate in the literature can be summarized as follows:
共a兲 Isentropic compression calculations using ideal gas properties pro-
duce much higher temperatures than have been confirmed by mea-
surement.
共b兲 Isentropic compression calculations using real gas properties also
produce much higher temperatures than have been confirmed by
measurement.
共c兲 Adiabatic compression of oxygen using a real gas equation of state
derived from thermodynamic properties predicts much higher tem-
peratures than have been confirmed by measurement.
共d兲 Shock process calculations derived from first principles predict
temperatures well in excess of those measured and also exceed the
predictions from any form of isentropic compression.
共e兲 One-dimensional numerical methods used to predict the tempera-
tures produced in the test system by superposition of reflected com-
pression waves and expansion waves predict temperatures much
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
358 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

higher than have been confirmed by measurement.


共f兲Two-dimensional numerical methods used to predict the tempera-
tures produced by superposition of reflected compression waves
and expansion waves that include mixing predict temperatures
higher than measured but closer, within twice that measured.
共5兲 Empirical data supports that the pressure oscillation observed in the
pressure-time measurements is not due to instrumentation character-
istics such as an under-damped transducer. Instead, the oscillation can
be shown to produce a thermal response in the WHA thermocouples
共Fig. 3兲. Further, the Air Liquide shock model reproduced many of the
essential features of the pressure-time oscillation history. Therefore, a
proper understanding of the test system influence on the thermal en-
ergy should explain the pressure-time “fingerprint” developed by the
test system.

References

关1兴 Newton, B., Chiffoleau, G., Steinberg, T., and Binder, C., “Adiabatic Compression
Testing II–Background and Approach to Estimating Severity of Test Methodology,”
ASTM 12th International Symposium on Flammability and Sensitivity of Materials
in Oxygen-Enriched Atmospheres, October 2009, ASTM International, Berlin, Ger-
many.
关2兴 Neary, R. M., “ASTM G63: A Milestone in a 60-Year Safety Effort,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 812, B. L.
Werley, Ed., ASTM International, West Conshohocken, PA, Vol. 1, 1983, pp. 3–8.
关3兴 ASTM International Guide G63-99, 1980, “Standard Guide for Evaluating Nonme-
tallic Materials for Oxygen Service,” Annual Book of ASTM Standards, ASTM G04
Committee on Compatibility and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres, Vol. 14.04, ASTM International, West Conshohocken, PA.
关4兴 ASTM International Test Method G 72-96, 1982, “Standard Test Method for Autog-
enous Ignition Temperature of Liquids and Solids in a High-Pressure Oxygen-
Enriched Environment,” Annual Book of ASTM Standards, ASTM G04 Committee
on Compatibility and Sensitivity of Materials in Oxygen-Enriched Atmospheres, Vol.
14.04, ASTM International, West Conshohocken, PA.
关5兴 ASTM International Test Method G74, 2002, “Standard Test Method for Ignition
Sensitivity of Materials to Gaseous Fluid Impact,” Annual Book of ASTM Stan-
dards, ASTM Committee G4 on Compatibility and Sensitivity of Materials in Oxygen
Enriched Atmospheres Subcommittee G04.01 on Test Methods, Vol. 14.04, ASTM
International, West Conshohocken, PA., Current edition approved Sept. 10, 2001.
Originally published in 1982. Last previous edition G74-82 共1996兲.
关6兴 Werley, B. L., “A Perspective on Gaseous Impact Tests: Oxygen Compatibility Test-
ing on a Budget,” Flammability and Sensitivity of Materials in Oxygen-Enriched
Atmospheres, ASTM STP 1197, Vol. 6, D. D. Janoff and J. M. Stoltzfus, Eds., ASTM
International, West Conshohocken, PA, 1993.
关7兴 Koch, U. H., “Oxygen System Safety,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres, ASTM SIP 1197, Vol. 6, D. D. Janoff and J. M.
Stoltzfus, Eds., ASTM International, West Conshohocken, PA, 1993.
关8兴 Koch, U. H., “Oxygen Piping Code—Where Knowledge Becomes Practice: Keynote
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 359

Address,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-


spheres, ASTM STP 1319, Vol. 8, W. T. Royals, T. C. Chou, and T. A. Steinberg, Eds.,
ASTM International, West Conshohocken, PA, 1997.
关9兴 Stradling, J. S., Pippen, D. L., and Frye, G. W., “Techniques Employed by the NASA
White Sands Test Facility to Ensure Oxygen System Component Safety,” Flamma-
bility and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 812,
B. L. Werley, Ed., ASTM International, West Conshohocken, PA, 1983, pp. 97–107.
关10兴 Newton, B., Porter, A., Hull, W. C., Stradling, J., and Miller, R., “A 6000 psig
Gaseous Oxygen Impact Test System for Materials and Components Compatibility
Evaluations,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres, ASTM STP 1319, Vol. 8, W. T. Royals, T. C. Chou, and T. A. Steinberg, Eds.,
ASTM International, West Conshohocken, PA, 1997.
关11兴 Wegener, W., Binder, C., Hengstenberg, P., Herrmann, K.-P., and Weinert, D.,
“Tests to Evaluate the Suitability of Materials for Oxygen Service,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 986, Vol. 3,
D. W. Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp. 268–
278.
关12兴 Binder, C., Kieper, G., and Herrmann, P., “A 500 Bar Gaseous Oxygen Impact Test
Apparatus for Burn-Out Testing of Oxygen Equipment,” Flammability and Sensi-
tivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1267, Vol. 7, ASTM
International, West Conshohocken, PA, 1995, pp. 23–35.
关13兴 Barthelemy, H., Delode, G., and Vagnard, G., “Oxygen Compatibility of Pressure
Regulators for Gas Cylinders,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1040, Vol. 4, ASTM International, West Consho-
hocken, PA, 1989, pp. 267–285.
关14兴 Barthelemy, H. and Vagnard, G., “Ignition of PTFE-Lined Hoses in High-Pressure
Oxygen Systems: Test Results and Considerations for Safe Design and Use,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP
986, Vol. 3, ASTM International, West Conshohocken, PA, 1988, pp. 289–304.
关15兴 Vagnard, G., Barthelemy, H., and Delode, G., “Test Methods and Result Interpre-
tation for Selecting Non-Metallic Materials for Oxygen Service,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1111, Vol. 5,
ASTM International, West Conshohocken, PA, 1991, pp. 289–304.
关16兴 Barthélémy, H., Delode, G., and Vagnard, G., “Ignition of Materials in Oxygen
Atmospheres: Comparison of Different Testing Methods for Ranking Materials,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 1111, Vol. 5, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West
Conshohocken, PA, 1991.
关17兴 Ducrocq, J., Barthelemy, H., and Roy, D., “A Fluid Flow Analysis of the Gaseous
Impact Test,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres, ASTM STP 1395, Vol. 9, T. A. Steinberg, B. E. Newton, and H. D. Beeson,
Eds., ASTM International, West Conshohocken, PA, 2000.
关18兴 ISO 2503, 2009, “Gas Welding Equipment—Pressure Regulators for Gas Cylinders
Used in Welding, Cutting and Allied Processes up to 300 Bar,” International Orga-
nization for Standardization.
关19兴 EN 585, 1995, “Gas Welding Equipment—Pressure Regulators for Gas Cylinders
Used in Welding, Cutting and Allied Processes up to 200 Bar,” European Norm
Standard, standard sponsored by CEN.
关20兴 Bryan, C. J., “A Review of Test Methods Used in the Selection of Materials for
Oxygen Service: Keynote Address,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres, ASTM STP1395, Vol. 9, T. A. Steinberg, B. E. New-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
360 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ton, and H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关21兴 Hirsch, D., Peyton, G., Hshieh, F.-Y., and Beeson, H., “Review of Pneumatic Impact
Sensitivity Data Obtained on Selected Polymers for Aerospace Applications,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 1454, Vol. 10, T. A. Steinberg, H. D. Beeson, and B. E. Newton, Eds., ASTM
International, West Conshohocken, PA, 2003.
关22兴 Janoff, D., Bamford, L. J., Newton, B. E., and Bryan, C. J., “Ignition of PTFE-Lined
Flexible Hoses by Rapid Pressurization with Oxygen,” Symposium on Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM SPT 1040, Vol.
4, 1989, J. M. Stoltzfus, F. J. Benz, and J. S. Stradling, Eds., ASTM International,
West Conshohocken, PA.
关23兴 Hirsch, D., Skarsgard, E., Beeson, H., and Bryan, C., “Predictability of Gaseous
Impact Ignition Sensitivity from Autoignition Temperature Data,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1395, Vol.
9, T. A. Steinberg, B. E. Newton, and H. D. Beeson, Eds., ASTM International, West
Conshohocken, PA, 2000.
关24兴 Blackstone, W. R. and Ku, P. M., “An Assessment of Impact Test Techniques for
Determining the Fire or Explosion Hazards of Materials Exposed to Liquid Oxy-
gen,” Mater. Res. Stand., Vol. 11, No. 6, 1971, pp. 30–35.
关25兴 Blackstone, W. R., Baber, B. B., and Ku, P. M., “New Test Techniques for Evaluat-
ing the Compatibility of Materials with Liquid Oxygen Under Impact,” ASLE
Trans., Vol. 11, 1968, pp. 216–227.
关26兴 Schmidt, N., Moffett, G. E., Pedley, M. D., and Linley, L. J., “Ignition of Nonme-
tallic Materials by Impact of High-Pressure Oxygen II: Evaluation of Repeatability
of Pneumatic Impact Test,” Symposium on Flammability and Sensitivity of Materi-
als in Oxygen-Enriched Atmospheres, ASTM STP 1040, Vol. 4, 1989, J. M. Stoltzfus,
F. J. Benz, and J. S. Stradling, Eds., ASTM International, West Conshohocken, PA.
关27兴 Moffett, G. E., Schmidt, N. E., Pedley, M. D., and Linley, L. J., “An Evaluation of
The Liquid Oxygen Mechanical Impact Test,” Symposium on Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1040, Vol. 4,
1989, J. M. Stoltzfus, F. J. Benz, and J. S. Stradling, Eds., ASTM International,
West Conshohocken, PA.
关28兴 Janoff, D., Pedley, M. D., and Bamford, L. J., “Ignition of Nonmetallic Materials by
Impact of High-Pressure Oxygen III: New Method Development,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1111, Vol.
5, J. M. Stoltzfus and K. McIlroy, Eds., ASTM International, West Conshohocken,
PA, 1991.
关29兴 Nanigian, J. and Nanigian, D., “A Unique Thermocouple to Measure the Tempera-
tures of Squibs, Igniters, Propellants and Rocket Nozzles,” SPIE Paper 6222-3,
Nanmac Corporation, Framingham, MA.
关30兴 White, D. R., “Influence of Diaphragm Opening Time on Shock-Tube Flows” J.
Fluid Mech., Vol. 4, 1958, pp. 585–5999.
关31兴 Pedley, M. D., Pao, J., Bamford, L., Williams, R. E., and Plante, B., “Ignition of
Contaminants by Impact of High-Pressure Oxygen,” Flammability and Sensitivity
of Materials in Oxygen-Enriched Atmospheres, ASTM STP 986, Vol. 3, D. W. Schroll,
Ed., ASTM International, West Conshohocken, PA, 1987, pp. 305–317.
关32兴 Forsyth, E. T., Durkin, R. J., and Beeson, H. D., “Evaluation of Contaminant-
Promoted Ignition in Scuba Equipment and Breathing-Gas Delivery Systems,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 1395, Vol. 9, T. A. Steinberg, B. E. Newton, and H. D. Beeson, Eds., ASTM
International, West Conshohocken, PA, 2000.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON AND STEINBERG, doi:10.1520/JAI102304 361

关33兴 Norton, H. N., Sensor and Analyzer Handbook, Prentice Hall, Inc., Englewood
Cliffs, NJ, 1982.
关34兴 Schweppe, J. L., Shock Tube Methods, Department of Mechanical Engineering, The
University of Houston, 1963.
关35兴 Schafer, J. H. and Frohn, A., “Ionization Behind Shock-Waves in Nitrogen-Oxygen
Mixtures,” AIAA J., Vol. 10, No. 8, 1972, pp. 985–987.
关36兴 Lewis, B. and Von Elbe, G., “Combustion, Flames, and Explosion of Gases,” Deto-
nation Waves in Gases, 2nd ed., Academic Press, New York, 1961, Chap. 8.
关37兴 Hughes, W. F. and Brighton, J. A., “Fluid Dynamics,” One-Dimensional Compress-
ible Flow–The Shock Tube, McGraw-Hill Book Company, New York, 1967, Chap. 7.
关38兴 Griffin, D. K., 1996, “Performance Characteristics of Free Piston Shock Tunnels,”
Ph.D. Dissertation, Department of Mechanical Engineering, The University of
Queensland, Brisbane, Australia.
关39兴 Kendall, M, 1998, “A Study of High-Enthalpy, Entropy Raising Drivers for Impulse
Facilities,” Ph.D. Dissertation, Department of Mechanical Engineering, The Uni-
versity of Queensland, Brisbane, Australia.
关40兴 Faeth, G. M. and Meyer, W. E., “Spontaneous Ignition During Charging Processes
in Pneumatic and Hydraulic Systems,” U. S. Navy Bureau of Ships Contract No.
Nobs 78674; Index No. S-R001-0301, The Pennsylvania State University, Defense
Technical Information Center, Alexandria, VA, 1963.
关41兴 Leslie, I. H., “Thermodynamic and Fluid Mechanic Analyses of Rapid Pressuriza-
tion in a Dead-End Tube,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1111, Vol. 5, J. M. Stoltzfus and K. McIlroy,
Eds., ASTM International, West Conshohocken, PA, 1991.
关42兴 Thorncroft, G., Patton, J., and Gordon, R., “Modeling Compressible Air Flow in a
Charging or Discharging Vessel and Assessment of Polytropic Exponent,” AC 2007-
2695, American Society for Engineering Education, CA.
关43兴 Dutton, J. and Coverdill, R., “Experiments to Study the Gaseous Discharge and
Filling of Vessels,” Int. J. Eng., Vol. 13, No. 2, 1997, pp. 123–134.
关44兴 Barragan, M., Wilson, D. B., and Stoltzfus, J. M., “Adiabatic Compression of Oxy-
gen: Real Fluid Temperatures,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1395, Vol. 9, T. A. Steinberg, B. E. Newton, and
H. D. Beeson, Eds., ASTM International, West Conshohocken, PA, 2000.
关45兴 van Wylen, G. and Sonntag, R., Fundamentals of Classical Thermodynamics, Wiley
& Sons, New York, 1978, SI Version.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102297
Available online at www.astm.org/JAI

Barry E. Newton,1 Gwenael J. A. Chiffoleau,2 Ted Steinberg,3


and Christian Binder4

Adiabatic Compression Testing—Part II:


Background and Approach to Estimating
Severity of Test Methodology

ABSTRACT: Adiabatic compression testing of components in gaseous oxy-


gen is a test method that is utilized worldwide and is commonly required to
qualify a component for ignition tolerance under its intended service. This
testing is required by many industry standards organizations and govern-
ment agencies; however, a thorough evaluation of the test parameters and
test system influences on the thermal energy produced during the test has
not yet been performed. This paper presents a background for adiabatic
compression testing and discusses an approach to estimating potential dif-
ferences in the thermal profiles produced by different test laboratories. A
“thermal profile test fixture” 共TPTF兲 is described that is capable of measuring
and characterizing the thermal energy for a typical pressure shock by any
test system. The test systems at Wendell Hull and Associates, Inc., in the
United States and at the BAM Federal Institute for Materials Research and
Testing in Germany are compared in this manner and some of the data
obtained are presented. The paper also introduces a new way of comparing
the test method to idealized processes to perform system-by-system com-
parisons. Thus, the paper introduces an “idealized severity index” 共ISI兲 of the
thermal energy to characterize a rapid pressure surge. From the TPTF data

Manuscript received December 17, 2008; accepted for publication June 12, 2009; pub-
lished online July 2009.
1
VP R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM
88007.
2
Senior Scientist and Test Facility Manager, Wendell Hull and Associates, Inc., 5605
Dona Ana Rd., Las Cruces, NM 88007.
3
Professor and Director, Queensland Univ. of Technology, S-Block, LVL-10, GPO Box
2434, Brisbane Queensland 4001, Australia.
4
Head of Working Group “Safe Handling of Oxygen,” Bundesanstalt für Materialfors-
chung und–prüfung, Unter den Eichen 87 D-12205 BerlinGermany.
Cite as: Newton, B. E., Chiffoleau, G. J. A., Steinberg, T. and Binder, C., ⬙Adiabatic
Compression Testing—Part II: Background and Approach to Estimating Severity of Test
Methodology,⬙ J. ASTM Intl., Vol. 6, No. 8. doi:10.1520/JAI102297.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 362
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 363

a “test severity index” can also be calculated so that the thermal energies
developed by different test systems can be compared to each other and to
the ISI for the equivalent isentropic process. Finally, a “service severity
index” is introduced to characterize the thermal energy of actual service
conditions. This paper is the second in a series of publications planned on
the subject of adiabatic compression testing.
KEYWORDS: adiabatic compression, pneumatic impact, gaseous fluid
impact, isentropic compression, oxygen, shock-wave heating, thermal
profile, severity index

Introduction
The compressed gas industry and government agencies worldwide utilize “adia-
batic compression” test methodologies for qualifying high-pressure valves,
regulators, and other related flow control equipment for gaseous oxygen ser-
vice. This test methodology is known by various terms including adiabatic
compression5 testing, gaseous fluid impact6 testing, pneumatic impact testing,
and BAM7 testing as the most common terms. The test methodology will be
described in greater detail throughout this paper but in summary it consists of
pressurizing a test article 共valve, regulator, etc.兲 with gaseous oxygen within
15–20 ms. Because the driven gas8 and the driving gas9 are rapidly compressed
to the final test pressure at the inlet of the test article, they are rapidly heated by
the sudden increase in their internal energy to sufficient temperatures 共thermal
energies兲 to sometimes result in ignition of the nonmetallic materials 共seals and
seats兲 used within the test article. In general, the more rapid the compression
process, the more “adiabatic” the pressure surge has been presumed to be and
the more, like an isentropic10 process, the pressure surge has been argued to
simulate.

5
While various terms are used for the type of testing discussed herein, adiabatic com-
pression testing is the term that will be used most frequently in this paper. This term is
chosen not because it is an accurate description, but because it is used most widely
within the industry. It is actually the irreversibility and non-adiabaticity of the method-
ologies that this research program is evaluating.
6
Gaseous fluid impact is the official description given by ASTM Committee G04 共see Ref
11兲.
7
BAM stands for Bundesanstalt für Materialforschung und–prüfung and is the German
Federal Institute for Materials Research and Testing where the test methodology dates
back to the 1950s. The test method was also implemented by the National Aeronautics
and Space Administration 共NASA兲, in a somewhat different form, after the 1970s; and by
such companies as AIRCO, RegO, AGA, and Circle Seal as discussed in the first paper of
this series 共Ref 1兲.
8
Atmospheric pressure oxygen originally in the system piping or tubing upstream of the
test article.
9
Oxygen gas originally contained in a high-pressure accumulator and separated from the
driven gas by a fast operating valve.
10
It is noteworthy that while shock-wave processes are not discussed in this paper, the
faster the pressurization is, the more likely that shock processes may develop during a
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
364 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Generally speaking, adiabatic compression is widely considered the most


efficient ignition mechanism for directly kindling a nonmetallic material in
oxygen and has been implicated in many fire investigations. The temperature
rise by near-adiabatic compression has commonly been calculated by assuming
ideal gas behavior through the polytropic equation. The temperature produced
by adiabatic compression is usually calculated using isentropic relationships
assuming that the oxygen behaves like an ideal gas and that the compression
process is sufficiently rapid that heat transfer does not occur during the short
time of the pulse 共i.e., essentially adiabatic兲. This form of the equation is nor-
mally used to calculate the final temperature considering isentropic behavior
共reversible and adiabatic兲.

Tf ª Ti · 冉冊 Pf
Pi
共k−1兲/k
共1兲

where:
Tf = final temperature 共abs兲,
Ti = initial temperature 共abs兲,
Pf = final pressure,
Pi = initial pressure, and
k = ratio of specific heats for oxygen 共1.4兲.
However, the adiabatic compression process as required by the industry
standards has never been thermodynamically modeled and empirically verified,
although attempts have been made.
This research evaluates these questions:
共1兲 Can the compression process required by the industry standards be
thermodynamically and fluid dynamically modeled so that predictions
of the thermal profiles produced be made?
共2兲 Can the thermal profiles produced by the rapid compression process be
measured in order to validate the thermodynamic and fluid dynamic
models and estimate the severity of the test?
共3兲 Can a new industry standard be prepared to resolve inconsistencies
between various test laboratories conducting tests according to the
present standards?
This paper is the second of a series of publications that are planned to
evaluate these questions and will present the background and initial testing
that has been conducted in the current research. More complete system-to-
system comparisons, detailed analysis of the temperature-time histories of the
current test systems, modeling of shock-wave processes, and testing to evaluate
whether shock waves are present in the transient compression will be forth-
coming in later publications.
The first paper in this series, “Adiabatic Compression Testing I–Historical
Development and Evaluation of Fluid Dynamic Processes Including Shock-

pressure surge. Faeth and Meyer 关16兴 argued that if the time of the event 共pressure rise
time兲 is not much slower than the tube length divided by the local speed of sound 共i.e.,
trise Ⰷ length/ soundspeed兲, then shock processes are more likely and localized pressure
disturbances can be expected.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 365

FIG. 1—Simulated adiabatic compression against closed valve.

Wave Considerations” 关1兴, presents the authors’ understanding of the historical


development of the current test method and some of the fluid dynamic pro-
cesses that may influence the temperature of the compressed gas. The first
paper introduces the conclusion that shock-wave processes might be present
during a pressure surge but that neither their presence nor strength is currently
understood. The research anticipated in this series will attempt to resolve this
question and expects to use both measurement and computational fluid dy-
namic modeling. This paper, the second in the series, outlines the background
of the current test methods that are widely used and the importance of under-
standing the thermal profiles that are produced by the various test systems. It
also presents a measurement scheme that has shown promise in measuring the
thermal profiles that are produced by different test systems. The measurements
obtained by the time of publication had not resolved whether shock processes
are present in a pressure surge conducted according to the standards, but fur-
ther testing is currently planned and will be presented in a later publication in
this series. Historically, the oxygen safety community has focused its attention
on the heating that occurs in the driven gas 共i.e., gas being compressed by the
high-pressure slug兲 and has considered this process to be isentropic. This is the
perspective that will be taken in the material presented in this paper.

Testing Background
Historically adiabatic compression processes are often depicted by the illustra-
tion shown in Fig. 1. In Sequence 1, as illustrated, a volume of low pressure gas
at an initial pressure and temperature is isolated from a volume of high-
pressure gas by a valve 共or other isolating element兲. Another closed valve pro-
vides a dead end to the low pressure volume. If the upstream valve is opened
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
366 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1—Laboratories that carry out adiabatic compression testing.

High-Speed Pressurization
Test Laboratory Location Valve Design Rate Control
BAM Berlin, Germany Globe valve 3.5 mm orifice
CTE–Air Liquide Paris, France Sliding gate valve 4.3 mm orifice
Apragaz Brussels, Belgium Sliding gate valve Unknown orifice size
DNV Norway Unknown Unknown
WHA Las Cruces, NM, USA Ball valve Valve opening speed
NASA–WSTF Las Cruces, NM, USA Ball valve Valve opening speed
NASA–MSFC Huntsville, AL, USA Ball valve Valve opening speed

rapidly, as illustrated in Sequence 2, then the low pressure gas, hereafter de-
fined as the driven gas, suddenly undergoes a compression process by the high-
pressure gas, hereafter defined as the driving gas, which flows through the
newly opened valve. The “P-dV” work done by the driving gas causes a tem-
perature rise in the driven gas. This temperature rise is often considered to be
adiabatic as long as the pressure rise rate is sufficiently rapid, as compared to
the development time for conduction and convective heat transfer. During the
compression process, the driving gas also goes through state changes, both
expansion and recompression. Therefore an increase in temperature also devel-
ops in the driving gas, especially in the gas that flows into the impact tube in
the early stages of the compression process. The degree of mixing between the
driving and driven gases is an important element influencing the maximum
temperature achieved by the compression process.

Statement of Industry Problem


The test laboratories that commonly conduct this testing worldwide are indi-
cated in Table 1.11 While each test laboratory meets the requirements of the
predominant standards currently in use, subtle differences exist in the test
equipment operated at the different laboratories 共discussed further below兲,
which are believed to produce variations in the test results. Significantly, these
variations have been argued to result in some components passing the tests at
one laboratory while failing at another. This disparity in results is of great
concern to the industry since the adiabatic compression test is fundamentally a
test to ensure that safe and reliable components are placed into the public
marketplace.
Figure 2 shows a component that “passed” the current test method but was
withdrawn from the marketplace by a “safety recall” instituted by the U.S. Food
and Drug Administration’s Center for Devices and Radiological Health due to

11
It is noteworthy that Western Enterprises, Victor Equipment Company 共United States兲,
and the Cavagna Group 共Italy兲 also have the capability of conducting adiabatic compres-
sion testing on the components they manufacture. To our knowledge, however, their test
systems are not commercially available.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 367

FIG. 2—Regulator fire investigated by WHA.

ignitions in service. It is important to understand, however, that the ignitions


that occurred in the field were attributed more to design problems on this
device than to adiabatic compression testing problems. However, this example
does illustrate the importance of high fidelity in the testing methodologies.
One problem with properly defining the test methodology is the lack of a
thorough understanding of the state processes that the driving and driven gases
go through during actual service conditions or during the testing. To our knowl-
edge, while several attempts have been made, no thermodynamic or fluid dy-
namic model has been validated by testing that specifies the state conditions of
the gas and predicts the thermal profile 共e.g., temperature versus time profile兲
of the driving and driven gases during the compression process. As a conse-
quence, calculation of the thermal energy in the compressed gas has not been
utilized in the design of the test method to establish the safety margins pro-
vided by the test results. Further, no testing has been able to confirm the ther-
mal energies produced within the cylindrical tube sections upstream of a test
article due to the very rapid pressurization rates 共⬃15– 20 ms to full pressure兲
encountered in this testing, and then relate that thermal profile to the potential
for ignition or statistical variations between test laboratories.
An important outcome of this research will be to utilize this research in the
preparation of an ASTM International test standard that will specify the critical
control elements for test systems conducting adiabatic compression testing
worldwide.12

12
ASTM International Committee G04 formed a task group to develop a standard that
will specify the way adiabatic compression testing is conducted in the future. Most of the
test laboratories listed in Table 1 have agreed to participate in this evaluation and in the
development of a standard to specify the test system controls to be implemented. Several
industry working groups such as the ISO/TC 58/SC 2/WG 6 subcommittee responsible
for adiabatic compression testing of compressed gas cylinder valves and their counter-
parts from the Compressed Gas Association 共CGA兲 in the United States have requested
that the ASTM International standard development efforts be coordinated with these
ISO and CGA committees.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
368 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—WHA adiabatic compression test system and test cycle.

Test Method Background


The testing of interest here is conducted in different ways by different test
laboratories but the fundamental system requirements are few. For illustration
purposes the Wendell Hull and Associates, Inc. 共WHA兲 test system is shown in
Fig. 3 along with the pressure profile that is required by the predominant stan-
dards. The test is typically conducted by pressurizing a test article 共valve, regu-
lator, etc.兲 very rapidly by opening a high-speed valve 共impact valve兲, simulating
a sudden pressurization that might occur in service. When the impact valve is
opened, high-pressure oxygen stored in an accumulator at 1.2 times the test
article working pressure and pre-heated to 60° C pressurizes a test article posi-
tioned at the end of an impact tube within 15–20 ms. According to the stan-
dards, the impact tube 共volume of oxygen to be compressed upstream of the
test article兲 is either a 5 mm inside diameter 共i.d.兲 tube that is 1 m long or a 14
mm i.d. tube that is 0.75 m long depending on whether the test article is in-
tended for use on a cylinder or on a manifold. As shown in Fig. 3, after the
rapid pressurization, the test article is held at the test pressure for at least 10 s
to allow ignition and propagation to develop if the test article nonmetallic ma-
terials are vulnerable to this method. After this hold period, the test article is
vented to ambient pressure and allowed to cool for a minimum of 3 s before the
test cycle is repeated within 30 s. According to most standards, 20 test cycles
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 369

FIG. 4—Idealized temperature rise of small plastic sample due to heat content in com-
pressed slug originally in 5-mm 1-m long Impact tube 共n = 1.4 for oxygen兲.

are typically performed with the test article closed 共regulator reduced or valve
closed兲 and another 20 with the test article open 共regulator increased or valve
opened兲 and the discharge port plugged. Successful completion of the 40 cycles
completes the test series.
Figure 4 shows the relationship between the sensible heat developed by the
compression process 共Qcal in the driven gas兲 and the temperature rise devel-
oped by a small mass of a nonmetallic material 共considering an isentropic pro-
cess兲, assuming that all the sensible heat is used to uniformly raise the tem-
perature of the plastic. While near-adiabatic compression is known to readily
kindle most flammable nonmetallic materials in oxygen, the actual
temperature-rise rate and maximum temperature achieved in real systems have
never been measured in real time.
Recently some effort to correlate real-gas behavior to the compression pro-
cess has been made, but empirical measurements have not been successful in
large part due to the temperature rise occurring over such a small time incre-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
370 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ment 共⬍20 ms兲. Further, since empirical measurements have been largely un-
successful, no methodology has been developed to compare pressure surges
produced by two different test systems that utilize different components to
produce the pressure surge. So, the actual correlation of the temperature-rise
rate in any test system to the behavior shown in Fig. 4 is unknown.
Since adiabatic compression is such a common ignition mechanism in gas-
eous oxygen systems and has routinely been implicated as the primary reason
for component ignition failures, many industry groups including the Interna-
tional Standards Organization 共ISO兲, the U.S. CGA, Australian Standards Orga-
nization, and ASTM International require the performance of adiabatic com-
pression testing to qualify nonmetallic materials and pneumatic components
共primarily high-pressure valves, regulators, flexhoses, etc.兲 intended for use in
high-pressure oxygen systems, as illustrated in Table 2. Table 2 is a summary of
some of the international standards requiring adiabatic compression testing of
components. This table is not exhaustive and several more standards could be
included. As is evident, however, this test methodology has gained very wide
subscription throughout the world. It is rapidly becoming one of the most im-
portant test methodologies for high-pressure component validation in the oxy-
gen industry.
The historical development of the test method is traced in the companion
paper in this series 关1兴, but it is noteworthy that an early work was performed
by companies such as AIRCO, RegO, and Circle Seal. The German Federal
Testing Institute, BAM, developed a test capability in the 1950s and early 1960s
关1–3兴. The first German standard in which it was included was DIN 477: 1963-
11, which involved conducting 50 repetitive pressure surge 共pneumatic impact兲
cycles. Each pneumatic impact cycle was repeated every 10 s and exposed the
component to a pressure surge from ambient to its maximum working pres-
sure.
The test method was modified by the Air Liquide Corporation in the 1980s
关4–6兴 for component testing, which led to several changes in the way in which
adiabatic compression testing was performed. The most important of these
changes was the requirement to use a 5 mm internal diameter impact tube of 1
m in length. The Air Liquide contributions to the test methodology also led to
the incorporation of test criteria into many international standards described
in Table 2.
Presently, all prevalent test standards except ASTM G74 require 20 pressure
surge cycles to be performed. Two test configurations are generally required for
each component: closed and open/plugged, and generally the test 共i.e., required
number of cycles in each of two configurations兲 is repeated with three test
articles. The same two configurations are required on cylinder valves with the
exception that the pressure surge is applied to the outlet of the cylinder valve,
instead of its inlet, in order to evaluate the potential for ignition during cylinder
filling operations.
In the 1970s NASA-WSTF 共White Sands Test Facility兲 conducted adiabatic
compression testing of components in oxygen and was responsible for all quali-
fication of oxygen components for ground support and space shuttle opera-
tions. During the 1980s and up to the present, NASA required that all gaseous
oxygen handling components be qualified by passing adiabatic compression
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—International standards that include adiabatic compression test requirements 共not exhaustive兲.

Pressure Gas Cycle Impact Impact


Rate Temperature Interval Line Line ID

Downloaded/printed by
Standard Title of Standard Date Test Pressure 共ms兲 共°C兲 Failure/Cycles 共s兲 Length 共mm兲
ISO 2503 Pressure regulators for gas 1983 20 MPa 20 60± 3 0/20 30
cylinders used in welding, cutting, and
allied processes
EN 585 Gas welding equipment–pressure regulators 1994 24 MPa 20+ 0, ⫺5 60± 3 0/20 30 1m 5
for gas cylinders used in welding, cutting,
and allied processes up to 200 bar
CGA E-4 Standard for gas pressure regulators 1994 1.2 times 20+ 0, ⫺5 60± 3 0/20 30 1m 5 or 12
MWP
ISO 10524 Pressure regulators and pressure 1995 24 MPa 20+ 0, ⫺5 60± 3 0/20 30 1m 5
regulators with flow-metering devices for
medical gas systems
AS 4267 Pressure regulators for use with industrial 1995 MWP 20 60± 3 0/20 30 1m ⬎3
compressed gas cylinders
EN 849 Transportable gas cylinders–cylinders 1996 1.2 times 20+ 0, ⫺5 60± 3 0/50 30 1m 5
valves–specification and type testing MWP

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
EN 738-1 Pressure regulators for use with medical 1997 24 MPa 20+ 0, ⫺5 60± 3 0/20 30 1m 5
gases–part 1: pressure regulators and pressure
regulators with flow metering devices
BS-EN 849 Transportable gas cylinders–cylinders 1997 1.2 times 20+ 0, ⫺5 60± 3 0/50 30 1m 5
valves–specification and type testing MWP
ISO/DIS 2503 Gas welding equipment–pressure 1997 1.2 times 20+ 0, ⫺5 60± 3 0/20 30 1m 5
regulators for gas cylinders used in welding, MWP

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


cutting, and allied processes up to 300 bar
AS 3840.1 Pressure regulator for use with medical gases. 1998 1.2 times 20+ 0, ⫺5 60± 3 0/20 30 1m 5
part 1: pressure regulators and pressure MWP
regulators with flow-metering devices
EN 738-3 Pressure regulators for use with medical 1999 24 MPa 20+ 0, ⫺5 60± 3 0/20 30 1m 5
gases–part 3: pressure regulators integrated
with cylinder valves
ASTM G175 Standard test method for evaluating the 2003
ignition sensitivity and fault tolerance of
oxygen regulators used for medical and
emergency applications Specifies use of ISO10524 for adiabatic compression testing
CGA V-9 Compressed gas association standard for 2004 1.2⫻ MWP 20+ 0, ⫺5 60± 3 0/20 30 1m 5
compressed gas cylinder valves
NEWTON ET AL., doi:10.1520/JAI102297 371
TABLE 2— 共Continued.兲

Pressure Gas Cycle Impact Impact


Rate Temperature Interval Line Line ID

Downloaded/printed by
Standard Title of Standard Date Test Pressure 共ms兲 共°C兲 Failure/Cycles 共s兲 Length 共mm兲
ISO 10297 Gas cylinders–refillable gas cylinder 1999 1.2 times 20+ 0, ⫺5 60± 3 0/20 30 1m 5
valves—specification and MWP
type testing
ASTM G74 Standard test method for ignition 1982 To 69 MPa 50+/⫺3 20 0/20 or 1/60 12 238 7.9
sensitivity of materials to gaseous fluid mm
impact

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


372 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
NEWTON ET AL., doi:10.1520/JAI102297 373

testing 关7–10兴. The NASA-WSTF test system configuration was used as an ex-
ample of a suitable system in 1982 by the American Society for Testing and
Materials 共now known as ASTM International兲 in ASTM Standard G74, “Stan-
dard Test Method for Ignition Sensitivity of Materials to Gaseous Fluid Impact”
关11,12兴. However, ASTM G74 did not mandate design criteria for any specific
system and allowed some variation in the specific configuration.
In 1989, WHA, who conducts forensic investigations of fires and explo-
sions, including oxygen equipment fires, developed an adiabatic compression
test system similar to the NASA-WSTF system but was also consistent with the
predominant compressed gas industry adiabatic compression test methods that
were gaining wide subscription in the industry 关13兴. At that time, WHA was the
only commercially available test laboratory in the United States for this testing.

Variability Among Test Systems

A recent effort has begun within standards organizations to generate common-


ality between all the test methodologies within each of the various test stan-
dards. One of the test parameters being changed is the requirement for 50 test
cycles for cylinder valves, which is in the process of being reduced to 20 test
cycles to be consistent with the regulator requirements. However, the predomi-
nant test laboratories that conduct this testing 共see Table 1兲 report variations in
the pass/fail performance of identical test articles. In other words, test articles
that pass the testing at one laboratory sometimes fail the test at another labo-
ratory. Consequently, the statistical reliability and associated validity of the test
results has been questioned.
In an effort to better understand the test variances, WHA personnel visited
all of the test laboratories except DNV 共yet to be scheduled兲 to evaluate whether
the test methodology varied from laboratory to laboratory. While all of the
laboratories meet the limited standardized requirements for conducting this
test, such as pressurization rate and impact tube configuration 共length/
diameter兲, significant differences were observed in the hardware utilized and in
the system configurations. Some of the more important differences are listed in
Table 1.
One important difference observed was the design of the high-speed impact
valve utilized to produce the pressure surge 共see Table 1兲. This valve is very
rapidly opened at the start of a test cycle to suddenly pressurize a test article
共either a nonmetallic material or a component兲. Most importantly, the pressur-
ization profile could be very different due to the way in which the valve opens,
as shown in Fig. 5. This figure demonstrates the variability in the percent of
flow for different valve configurations and illustrates that since the different
laboratories do not use similar valve configurations, the pressurization profiles
downstream of the valve should exhibit different pressurization profiles. Fur-
ther, since the flow coefficients and flow turbulence of these valve types also
vary 共WHA Ball Valve: Cv ⬃ 27; BAM Globe Valve: Cv of 4.7–8.5兲, the pressuriza-
tion dynamics for these two systems would also be expected to be different and
produce different thermal events as a function of turbulence and mixing effects.
Indeed, as Fig. 6 demonstrates, the pressurization profiles recorded by the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
374 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

(a) (b) (c)

FIG. 5—Flow characteristics of valve configurations 共percent flow versus percent open兲
关14兴: 共a兲 ball valve configuration; 共b兲 poppet/globe valve configuration; and 共c兲 butterfly
valve configuration 共note that the valve cross-sections are for illustration only and are
not intended as accurate engineering drawings兲.

dynamic pressure transducers do exhibit differences, and the effect on the ther-
mal profiles produced upstream of the test article is the subject of this research.
Figure 6 depicts the results of testing performed by WHA to compare the WHA
and BAM thermal profiles and also shows the respective pressure profiles for a
200 barg pressure pulse obtained on a typical test cycle with a high-speed dy-
namic 共quartz crystal兲 pressure transducer. The valve used by WHA is a typical
ball valve. The valve used by BAM is a typical globe valve. Other test systems,
such as those of Centre for Technical Expertise 共CTE兲 and NASA-WSTF, have
also been characterized but the results of those evaluations are not included
herein and, if permission is received, will be addressed in a future paper in this
series. Figure 6 also demonstrates the temperatures that were obtained at four
different positions by the WHA thermal profile test fixture 共TPTF兲 to be de-
scribed later. The temperatures plotted in this figure were obtained by Type K
共chromel-alumel兲 thermocouples of 0.025 mm 共0.001 in.兲 diameter. It is note-
worthy that BAM utilizes an orifice to control the pressurization rate, whereas
WHA uses a variable speed valve to control pressurization rate. The influence of
these two approaches will be more thoroughly discussed in another paper that
will compare the thermal profiles of the various test systems.
As is also depicted in Fig. 6, the temperatures obtained during the pressure
pulse vary from system to system and do not obtain the temperatures calcu-
lated by the classical means using isentropic relationships 共1241° C兲
关3–6,13,15,16兴. It is, of course, recognized that the thermocouple response
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 375

WHA 200 barg - Positions 1-4 (0.001 thermocouple)


500 250

225

400 200

175

Pressure (barg)
300 150

125

200 WHA Temp Position 1 100

WHA Temp Position 2 75


WHA Temp Position 3
100 50
WHA Temp Position 4
WHA Pressure 25

0 0
0.2 0.3 0.4 0.5 0.6 0.7

BAM 200 barg - Positions 1-4 (0.001 thermocouple)


500 200

180

400 160

140
Temperature (C)

Pressure (barg)
300 120

100

200 BAM Temp Position 1 80


BAM Temp Position 2 60
BAM Temp Position 3
100 40
BAM Temp Position 4
BAM Pressure 20

0 0
0.2 0.3 0.4 0.5 0.6 0.7

Time (sec)

FIG. 6—WHA and BAM pressure and temperature profiles 共test methodology described
later兲.

times may not be sufficient to fully represent the transient temperatures


present in the pressure pulse 共another subject of this research兲.
As mentioned above, another parameter that varies among the test systems
is the method for controlling the pressurization rates, also recorded in Table 1.
WHA, NASA-WSTF, and NASA-MSFC 共Marshall Space Flight Center兲 use a
variable speed ball valve to control flow rate. BAM and CTE use an orifice 共of
different dimensions兲 to control the pressure rise. The thermodynamic states
undergone by the driving gas 共gas from the accumulator兲 and driven gas 共gas
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
376 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Si
Simplifie
mplifieddWHA
WHA Syste
Systemm Simplified
Simplified BAM
BAM Sche
Schema
matic
tic
Accumulator Ball Accumulator Gl
Globe
obe Ori
Orifice Dead
DeadEnd
Accumulator BallValve
Valve Dead
DeadEnd
End Accumulator fice End
Valve
Valve
33 11 22 33 44
11 22 55
00 00

0.5
0.5 m
m 66 m
m

Decreasing
Decreasing Decreasing
Decreasing
TT PP00 TT PP00
Pressure
Pressure Pressure
Pressure
55

PP33 PP33
33

33
44
0,1
0,1 22 0,1
0,1 22

WHA
WHA State
StateProcesse
Processess SS BAM
BAMSt
State
ate Processe
Processess SS

FIG. 7—Idealized T-S diagrams.

initially at 1 atm being compressed to test pressure兲 is expected to vary from


system to system because of these differences, as illustrated in Fig. 7. Figure 7
presents an idealized depiction of the state processes 共temperature-entropy兲
that the WHA and BAM driving and driven gases undergo during a pressure
pulse, if the flow differences between the valves are ignored and only state
processes are considered.
The WHA and NASA state processes are relatively straightforward. If it is
assumed that the perfect gas laws hold, no heat is transferred in the valves, and
no mass is stored, then the gas from the accumulator can be idealized to enter
the downstream pipe at its initial temperature and at the pressure of the down-
stream pipe 共P3, 1 atm兲. It is then recompressed to its original pressure 共P0兲 and
undergoes a corresponding isentropic temperature rise shown from state 2 to
state 3. The final idealized temperature is shown at position 3 on the
temperature-entropy diagram for the WHA state process in Fig. 7.
The BAM and CTE systems, by comparison, go through a similar state
change through the valve but recompress to a new intermediate pressure 共state
3兲 at the orifice before expanding again while flowing through the flow control
orifice 共states 3 and 4兲. The recompression process at state 3 could be expected
to increase the temperature of the driving gas before it expands through the
orifice to recompress again to its original pressure. The final state change is
reflected in Fig. 7 at position 5 for the BAM state processes. As shown in the
idealized diagrams, the temperature increase by the adiabatic compression
process in the BAM system could be expected by this analysis to be greater than
in the WHA system, even though the pressurization rate requirement is met by
both systems.
Another system difference between the WHA and BAM systems is the
length of the tubing between the accumulator and the high-speed impact valve.
The BAM system includes a length of tubing 6 m long between the accumula-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 377

SSim
implifie
plifiedd B
BAAM
M Sc
Sche
hem
matic
atic
AAccu
ccummuula
lato
torr GGlo
lobe
be Or
Orifific
icee D
Deeaadd EEnd
nd
VVaalv
lvee
11 22 33 44 55
00

66 m
m

Dec
Decrreas
easing
ing
TT PP00 Pr
Pres
essu
surree
55

PP33

55’’
33
44
0,1
0,1 22
22’’
11’’

BA
BAMM SSta
tate
te PPro
roce
cess
sses
es SS

FIG. 8—Idealized T-S diagram for pressure drop before impact valve.

tors and the impact valve. The WHA system is more closely coupled to the
impact valve and incorporates a length of tubing no more than 0.5 m long. If
the gas entering the impact valve decreases in pressure during the compression
process due to pressure expansion down the tubing run from the accumulator,
then the state changes for the driving gas entering the impact valve can be
idealized, as shown in Fig. 8. If the pressure drops at the inlet to the impact
valve, then the state processes are shown by the red 共dashed兲 lines in Fig. 8 and
lead to a final recompressed temperature of 5⬘, which could be substantially
lower than previously predicted.
At present, these uncertainties are being evaluated, but the potential
change in the outcome of an adiabatic compression test is readily evident by
the temperature predictions. More importantly, no test standard presently
available 共Table 2兲 specifies the test system configuration requirements that
would be expected to control these potential differences.
It is important to note here that the data we currently have available is
generally consistent with Fig. 6 and will be presented in detail in another paper
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
378 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

in this series. However, the thermal profiles for BAM and WHA do not vary
greatly from one another, as Fig. 6 illustrates, and the maximum temperatures
are generally within ⬃50° C for these two systems. The data cannot be fully
discussed in this paper, but it should be observed that the influence of the
orifice does not seem to predispose the BAM system to higher temperatures,
probably due to the close coupling of this orifice at the immediate outlet of the
high-speed valve, which minimizes the recompression influence. Mixing of the
driving and driven gases might occur downstream of the orifice due to turbu-
lence, but this influence is still under evaluation. Further, while the BAM tubing
length between their primary accumulators and their high-speed valve mea-
sures about 6 m, the i.d. of this line is large 共14 mm or greater兲 and therefore
does not restrict the re-supply of oxygen/pressure to the high-speed valve dur-
ing a pressure surge. This 6 m line is also heated so, for all practical purposes,
seems to function similar to a smaller accumulator. The idealized influences
that are pointed out in the temperature-entropy relationships in Figs. 7 and 8
are nevertheless valid, but this testing has shown these factors to be of less
importance on the BAM system. However, a system that did not closely couple
the orifice to the high-speed valve and/or utilized a supply line that choked the
flow would be expected to behave very differently, even though the two systems
would be schematically identical.

Methodology and Research Approach

This research anticipates that the thermal energy from a rapid pressure surge
in oxygen will generate sufficient sensible heat to ignite nonmetallic materials
either placed at the dead end of the impact tube or within the seat assemblies of
valves and regulators. Consequently, the rate at which this sensible heat is gen-
erated by the pressure surge and the maximum temperature that is developed
are measurable quantities sufficient to characterize the pressure surge itself
and the equipment used to create the pressure surge. Therefore the research
described here sought to both thermodynamically model the pressure surge
and measure empirically the temperature-rise rate as a function of time 共ther-
mal profile兲 during the period in which the pressure rise was occurring 共gener-
ally 20 ms according to the predominant test standards兲.
The challenges that this research encountered were twofold:
共1兲 Modeling: One challenge was to determine empirically the processes
undergone by the compressed gas 共isentropic or shock or a combina-
tion of both兲 and to then model those processes with sufficient fidelity
so that predictions of the thermal energy can be made and measured.
The background for the fluid dynamic processes considered and the
results of the modeling will be discussed in a separate paper.
共2兲 Measurement: Another challenge was to develop and validate a way to
measure the temperature rise as a function of time during the compres-
sion process and thereby to compare the performance of one test sys-
tem to another. If a method of measuring the thermal profile 共tempera-
ture versus time兲 of a pressure surge could be developed, then the
thermal energy contained within the pressure surge and the energy
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 379

Tmax 2

Tmax 1

Temperature
Temperature

Energy At
Energy At Ignition
Ignition

Tignition Time Tignition Time

FIG. 9—System specific thermal profiles allow comparison and calibration.

development rate could be directly compared between test systems, as


illustrated in Fig. 9. The second objective, measurement, will be dis-
cussed herein since much of this initial research has been focused on
developing a measurement approach.

Measurement
As shown in Fig. 9, characterization of the thermal energy in a pressure surge
can in principle be achieved by measuring the temperature versus time changes
and the maximum temperature attained. While the figure labels the area under
the temperature curve as “energy,” it is understood that the energy is really the
summation of the mass compressed 共m兲 times the specific heat of the gas 共Cp兲
times the temperature rise 共⌬T兲 for each increment in time. However, the major
contribution to the energy differences is expected to be the temperature rise or
thermal profile. The measurement of the thermal profile should allow system
comparisons to be made and integration of the total thermal energy required at
the time of ignition to be correlated. Regardless of the process that is producing
the temperature rise, ultimately, the measurement of the thermal profile should
allow for system comparisons to be made and process conditions to be evalu-
ated, for systems having sufficiently similar volumes undergoing compression.
The experimental approach that is suggested herein was first attempted by
Faeth and Meyer 关16兴 on systems of larger size and slower pressurization rates
than those presently utilized but was used with good success. Faeth and Meyer
关16兴 assumed that the temperature response of a thermocouple 共rise time兲 could
be considered to be infinitely fast if the thermocouple bead had essentially zero
mass.
Their approach was to take repeated measurements with two different
sized thermocouples and then extrapolate the temperatures measured in the
compression process to zero diameter. This approach was successful and com-
pared favorably to an isentropic model that included heat transfer influences,
but Faeth and Meyer 关16兴 utilized a linear extrapolation between two differently
sized thermocouple beads 关0.025 mm 共0.001 in.兲 and 0.076 mm 共0.003 in.兲兴,
each utilized on different test runs. No simultaneous measurements were made
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
380 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Now Using:
• 0.025-mm
• 0.051-mm
• 0.076-mm
NOTE: 0.013-mm is very delicate

0.076-mm

0.025-mm 0.013-mm

FIG. 10—Thermocouple array, end view.

and no validation of the extrapolation order 共i.e., linear first order兲 was at-
tempted.
A similar approach was utilized in this research except that a thermocouple
array has been designed to allow for simultaneous measurement of the tem-
perature in the driving and driven gas at three locations 共at the same plane兲
along at least four different planes in the impact tube. A typical thermocouple
array is shown in Figs. 10 and 11. Each array is comprised of three thermo-
couples of 0.025 mm 共0.001 in.兲, 0.051 mm 共0.002 in.兲, and 0.076 mm 共0.003 in.兲
bead diameters. Testing has also been conducted with thermocouple arrays
having 0.013 mm 共0.0005 in.兲, 0.025 mm 共0.001 in.兲, and 0.051 mm 共0.002 in.兲,
but due to the fragile nature of the 0.013 mm diameter thermocouples, the

FIG. 11—Thermocouple array, side view.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 381

FIG. 12—TPTF 共transparent view兲.

0.025, 0.051, and 0.076 mm arrays have been preferred for most of the testing.
The different sized beads provide different response times to the thermal pro-
cess at almost the same location, and, indeed, the 0.013 mm 共0.0005 in.兲 ther-
mocouple does provide faster response and thus measures higher tempera-
tures. Therefore, a means to improve the fragile nature of this thermocouple is
still being sought.
Several different types of thermocouple devices have also been evaluated,
including the NANMAC Corp. fast-response 共“eroding bead”兲 thermocouples
关17兴, Paul Beckman Co. microminiature thermocouples 关18兴, and fine-wire
beaded thermocouples. Several research articles describing the development
and use of fast-response thermocouples have also been referenced for back-
ground 关19–24兴 and aid in the construction of a measurement system. The
NANMAC and Beckman fast-response thermocouples were utilized and com-
pared to small diameter 0.013, 0.025, and 0.050 mm diameter exposed bead
thermocouples. For the testing conducted herein, the 0.0005 and 0.001 in. ex-
posed bead thermocouples provided faster rise times and reproduced the tem-
perature of a step input signal 共short pulse of hot air of known temperature兲
than either the NANMAC or Beckman thermocouples. Thus, the exposed bead
thermocouples were preferred.
The TPTF designed for these arrays is shown in Figs. 12 and 13. Each
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
382 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Position 4: 108-mm upstream


Position 2: 6.4-mm upstream

Position 1 – Dead end


Position 3: 31.8-mm upstream

FIG. 13—TPTF installed on WHA test system.

measurement position can collect temperature data on each thermocouple


array in real time. Two positions are provided to collect dynamic pressure data
during the short time of the pressure surge. The illustrations only show one
position for dynamic pressure, but two positions are actually designed into the
TPTF, one at the upstream end and one at the downstream end. A data acqui-
sition system has also been developed capable of taking temperature data every
70– 100 ␮s on each temperature and pressure channel 共total of 12 temperature
channels and two pressure channels兲. Pressure volume calculations confirm
that the driven gas will occupy approximately 2.3 cm at the end of the TPTF, if
mixing is ignored. Therefore, positions 1 and 2 should provide data pertaining
to the driven gas and positions 3 and 4 should provide data pertaining to the
driving gas. If mixing develops, which is expected, then position 3 will be ex-
pected to provide helpful data to evaluate the mixing influences.
The TPTF described here has been recently used to begin characterization
of the WHA and BAM test systems in an effort to evaluate the research ap-
proach. For these tests an identical TPTF was used on these four test systems
and the same thermocouple arrays were utilized on each. The position and
clocking of each array were maintained during this testing.
Figures 6 and 14 show actual data taken on the WHA and BAM test systems
with the TPTF and demonstrate that differences between the test systems are
measurable based on the methodology suggested here. Figures 6 and 14 pro-
vide actual data taken by the thermocouples in real time. While the tempera-
tures should not be considered as more than indications at this point in time,
the methodology does suggest that with sufficient refinement and validation of
the measurement, a means of characterizing the thermal profile of various sys-
tems and evaluating the effects of the different hardware can be achieved.
Figure 15 symbolically illustrates an extrapolation procedure that is being
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 383

WHA ~200 barg - Positions 1-4 (0.001 thermocouple)


500 250

225

400 200

175
Temperature (C)

Pressure (barg)
300 150

125

200 WHA Temp - Dead End 100

WHA Temp - 6 mm Upstream 75


WHA Temp - 32 mm Upstream
100 50
WHA Temp - 108 mm Upstream
WHA Pressure 25

0 0
0.25 0.3 0.35 0.4 0.45 0.5

Time (sec)

FIG. 14—Temperature profile in WHA system at four positions 共0.025 mm TC兲.

evaluated to produce an estimate of the temperature for an imaginary “zero-


diameter” 共infinitely fast response兲 thermocouple, similar to the method that
Faeth and Meyer 关16兴 utilized. The extrapolation uses the reading of each ther-
mocouple at each time increment to derive a temperature based on an instan-
taneous rise time from a pseudo-zero mass thermocouple. The extrapolation
should in reality be based on how the heat transfer develops an “electromotive
force” in the thermocouple bead, but the form and order of the heat transfer
are still being verified by a calibration process. However, the approach shown
in Fig. 15 illustrates a simple curvilinear fit to the data obtained from each
thermocouple at each time increment. Successful tests with the 0.013 mm di-

FIG. 15—Temperature extrapolation to zero diameter 共illustration兲 共theoretically instant


rise time兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
384 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

WHA ~ 200 barg - Position 1 (Cycle 10)


1000 250

900

800 200

700
Temperature (C)

Pressure (barg)
600 150

500

400 100
0.001-in TC
300
0.002-in TC
200 0.003-in TC 50
Extrap-2nd Order
100
Pressure - barg
0 0
0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45

Time (sec)

FIG. 16—WHA test at 200 barg.

ameter thermocouple indicated that the temperatures produced momentarily


by the adiabatic compression process do produce higher temperatures than
measured by the 0.025 mm diameter thermocouple. Further, when the extrapo-
lation scheme is applied to the 0.025, 0.051, and 0.076 mm data, and extrapo-
lations are carried out to predict the temperature of a 0.013 thermocouple, the
measured and extrapolated temperatures are within that do approximate the
temperatures calculated by the extrapolation method.
Figure 16 shows this procedure applied to the data taken at position 1 共the
dead end兲 for the WHA test results originally shown in Fig. 14. As shown, by
this procedure the maximum temperatures estimated are exaggerated by the
differences in the three thermocouple readings at each time increment due to
the response time of each individual thermocouple. This approach predicts a
maximum temperature for an infinitely fast near-zero mass thermocouple of
623° C, whereas the maximum temperature measured by the 0.025 mm ther-
mocouple was 391° C. Examination of the approach has indicated that the
greater the difference in the measurements, at each time increment, the greater
the predicted temperature when the extrapolation technique is used. Therefore,
additional work on the uncertainty in the temperature measurements must be
performed before the technique can be considered valid.
Figures 17 and 18 show the thermal profiles for WHA and BAM for 200
barg tests during one cycle of a 20 cycle test series. The actual temperatures
measured and the extrapolation predictions are shown for each of the four
positions of the TPTF. The clear differences in the pressurization profile dy-
namics 共i.e., pressure transducer response curves兲 are still under investigation,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 385

WHA ~ 200 barg - Position 1 (Cycle 10) WHA ~ 200 barg - Position 2 (Cycle 10)
1000 250 1000 250

900 900

800 200 800 200

700 700
Temperature (C)

Temperature (C)
Pressure (barg)

Pressure (barg)
600 150 600 150

500 500

400 100 400 100


0.001-in TC 0.001-in TC
300 300
0.002-in TC 0.002-in TC
200 0.003-in TC 50 200 0.003-in TC 50
Extrap-2nd Order Extrap-2nd Order
100 100
Pressure - barg Pressure - barg
0 0 0 0
0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45 0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45

Time (sec) Time (sec)

WHA ~ 200 barg - Position 3 (Cycle 10) WHA ~ 200 barg - Position 3 (Cycle 10)
1000 250 1000 250

900 900

800 200 800 200

700 700
Temperature (C)

Temperature (C)
Pressure (barg)

Pressure (barg)
600 150 600 150

500 500

400 100 400 100


0.001-in TC 0.001-in TC
300 300
0.002-in TC 0.002-in TC
200 0.003-in TC 50 200 0.003-in TC 50
Extrap-2nd Order Extrap-2nd Order
100 100
Pressure - barg Pressure - barg
0 0 0 0
0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45 0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45

Time (sec) Time (sec)

FIG. 17—TPTF data for all WHA thermocouples plus extrapolations 共⬃200 barg, Cycle
10兲.

but since the same transducer was used for each system, they are believed to be
real fluid dynamic differences between WHA and BAM.
The measurements do indicate that thermal energy in the compressed gas
can be measured at each of the four TPTF locations and that differences be-
tween the test systems can be characterized.

BAM Test - 200 barg (Cycle 3), Position 1 BAM Test - 200 barg (Cycle 3), Position 2
1000 250 1000 250

900 900

800 200 800 200

700 700
Temperature (C)
Temperature (C)

Pressure (barg)
Pressure (barg)

600 150 600 150

500 500

400 100 400 100

300 300
1-0.001 TC 2-0.001 TC
1-0.002 TC 2-0.002 TC
200 50 200 50
1-0.003 TC 2-0.003 TC
100 Extrap 1 100 Extrap 1
Pressure (barg) Pressure (barg)
0 0 0 0
0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45 0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45

Time (sec) Time (sec)

BAM Test - 200 barg (Cycle 3), Position 3 BAM Test - 200 barg (Cycle 3), Position 4
1000 250 1000 250

900 900

800 200 800 200

700 700
Temperature (C)
Temperature (C)

Pressure (barg)

Pressure (barg)

600 150 600 150

500 500

400 100 400 100

300 300
3-0.001 TC 4-0.001 TC
3-0.002 TC 4-0.002 TC
200 50 200 50
3-0.003 TC 4-0.003 TC
100 Extrap 1 100 Extrap 1
Pressure (barg) Pressure (barg)
0 0 0 0
0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45 0.25 0.27 0.29 0.32 0.34 0.36 0.38 0.41 0.43 0.45

Time (sec) Time (sec)

FIG. 18—TPTF data for all BAM thermocouples plus extrapolations 共⬃200 barg, Cycle
3兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
386 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Future Work

This research has shown that a method of characterizing the thermal profile of
an adiabatic compression test has been developed that is capable of determin-
ing whether differences in the thermal energy between different test systems
exist. WHA is currently undertaking to characterize and analyze the thermal
profiles of all of the predominant test systems that perform this testing and is
developing data analysis routines to deal with the massive amount of tempera-
ture data that is produced during these characterization tests. Further, fluid
dynamic modeling and shock-wave analysis of the compression process are
underway to aid in predicting the actual state processes undergone and the gas
temperatures that would be expected from a rapid pressurization test.
The system-to-system comparisons based on the thermal profiles are also
underway along with an uncertainty analysis on the thermocouple data. Since
the ignition propensity of a test article is ultimately being evaluated by the test,
differences in the thermal energy produced by the test systems must be mini-
mized. Therefore, the following approach to estimating the relative severity of
the test and the test systems is being proposed as a tool for making the com-
parisons.
Based on the temperature/energy data available from the WHA TPTF, a
calculation of three severity indices 共simple ratios兲 is proposed. The severity
index will be a variable used to relate the standardized test system “results” to
idealized compression events and to actual “in-service” circumstances. The fol-
lowing severity indices are proposed:
共1兲 Idealized Severity Index 共ISI兲. The ISI will be an index 共ratio兲 calculated
to compare purely adiabatic and reversible 共i.e., isentropic兲 compres-
sion of a mass of compressed gas to the thermodynamic and fluid dy-
namic predictions when real-gas properties are considered. This index
will establish an idealized limit for the potential thermal energy ex-
pected from an isentropic pressure surge on a test system.
共2兲 Test Severity Index 共TSI兲. From the TPTF data a TSI can be derived so
that the thermal energies developed by different test systems can be
compared to each other on the basis of the ISI as compared to the
equivalent idealized process. By this index, a particular test system can
be compared to the idealized behavior and then to other test systems
that have been evaluated in the same way. The TSI will provide a way to
directly compare one system to another.
共3兲 Service Severity Index 共SSI兲. A SSI can also be developed by utilizing
the TPTF to characterize the thermal energy of actual service condi-
tions, such as the opening of a cylinder valve with a regulator con-
nected. This is the most common service condition for which the adia-
batic compression testing is intended to qualify valves and regulators.
The SSI for this application, and potentially others, will be compared
to both the ISI and the TSI to help with the prediction of the statistical
reliability of the adiabatic compression test results. Once the SSI and
TSI are specified for a given service configuration and test system, then
a confidence interval for a “passing” result can be more readily derived.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
NEWTON ET AL., doi:10.1520/JAI102297 387

References

关1兴 Newton, B. E. and Steinberg, T. A., “Adiabatic Compression Testing I–Historical


Development and Evaluation of Fluid Dynamic Processes Including Shock-Wave
Considerations,” Proceedings of the ASTM 12th International Symposium on Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, October
2009, Berlin, Germany, paper ID JAI102304.
关2兴 Information obtained provided by BAM, Federal Institute for Materials Research
and Testing.
关3兴 Wegener, W., Binder, C., Hengstenberg, P., Herrmann, K.-P., and Weinert, D.,
“Tests to Evaluate the Suitability of Materials for Oxygen Service,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 986, Vol. 3,
D. W. Schroll, Ed., American Society for Testing and Materials, Philadelphia, PA,
1988, pp. 268–278.
关4兴 Barthelemy, H., Delode, G., and Vagnard, G., “Oxygen Compatibility of Pressure
Regulators for Gas Cylinders,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1040, Vol. 4, ASTM International, West Consho-
hocken, PA, 1989, pp. 267–285.
关5兴 Barthelemy, H. and Vagnard, G., “Ignition of PTFE-Lined Hoses in High-Pressure
Oxygen Systems: Test Results and Considerations for Safe Design and Use,” Flam-
mability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP
986, Vol. 3, ASTM International, West Conshohocken, PA, 1988, pp. 289–304.
关6兴 Vagnard, G., Barthelemy, H., and Delode, G., “Test Methods and Result Interpre-
tation for Selecting Non-Metallic Materials for Oxygen Service,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1111, Vol. 5,
ASTM International, West Conshohocken, PA, 1991, pp. 289–304.
关7兴 NASA Johnson Space Center, Nonmetallic Materials Design Guidelines and Test
Data Handbook, JSC-02681.
关8兴 NASA Handbook NHB 8060.1B, “Flammability, Odor, Offgassing, and Compatibil-
ity Requirements and Test Procedures for Materials in Environments that Support
Combustion,” Test Method 14 Pressurized Gaseous Oxygen Pneumatic Impact for
Nonmetals 共Test 14兲, NASA Office of Safety and Mission Quality, Washington, DC,
1991.
关9兴 Bond, A. C., Pohl, H. O., Chaffee, N. H., Guy, W. W., Alton, C. S., Johnston, R. L.,
Castner, W. J., and Stradling, J. S., “Design Guide for High Pressure Oxygen Sys-
tems,” NASA Reference Publication 1113, NASA Scientific and Technical Informa-
tion Branch, NASA Headquarters, Washington, DC, 1983.
关10兴 Stradling, J. S., Pippen, D. L., and Frye, G. W., “Techniques Employed by the NASA
White Sands Test Facility to Ensure Oxygen System Component Safety,” Flamma-
bility and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 812,
B. L. Werley, Ed., ASTM International, West Conshohocken, PA, 1983, pp. 97–107.
关11兴 ASTM Test Method G74, 2002, “Standard Test Method for Ignition Sensitivity of
Materials to Gaseous Fluid Impact,” Annual Book of ASTM Standards, ASTM Com-
mittee G4 on Compatibility and Sensitivity of Materials in Oxygen Enriched Atmo-
spheres Subcommittee G04.01 on Test Methods, Vol. 14.04, ASTM International,
West Conshohocken, PA, current edition approved Sept. 10, 2001. Originally pub-
lished in 1982. Last previous edition G74-82 共1996兲.
关12兴 Neary, R. M., “ASTM G 63: A Milestone in a 60-Year Safety Effort,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 812, B. L.
Werley, Ed., ASTM International, West Conshohocken, PA, 1983, pp. 3–8.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
388 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

关13兴 Newton, B., Porter, A., Hull, W. C., Stradling, J., and Miller, R., “A 6000 psig
Gaseous Oxygen Impact Test System for Materials and Components Compatibility
Evaluations,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres, ASTM STP 1319, Vol. 8, W. T. Royals, T. C. Chou, and T. A. Steinberg, Eds.,
ASTM International, West Conshohocken, PA, 1997.
关14兴 Aerospace Fluid Component Designers’ Handbook, Vol. 1, Revision C, G. W. Howell
and T. M. Weathers, Eds., TRW Systems Group, Redondo Beach, CA, 1968, Air
Force Contract 04共611兲-8385 and 04共611兲-11316.
关15兴 ASTM International Guide G63-99, 1980, “Standard Guide for Evaluating Nonme-
tallic Materials for Oxygen Service,” Annual Book of ASTM Standards, ASTM G04
Committee on Compatibility and Sensitivity of Materials in Oxygen-Enriched Atmo-
spheres, Vol. 14.04, ASTM International, West Conshohocken, PA.
关16兴 Faeth, G. M. and Meyer, W. E., “Spontaneous Ignition During Charging Processes
in Pneumatic and Hydraulic Systems,” The Pennsylvania State University, Depart-
ment of Mechanical Engineering, 1963.
关17兴 Nanigian, J. and Nanigian, D., “A Unique Thermocouple to Measure the Tempera-
tures of Squibs, Igniters, Propellants, and Rocket Nozzles,” SPIE Paper 6222-3,
Nanmac Corporation, Framingham, MA www.nanmac.com/handbook/e11.htm.
关18兴 Paul Beckman Company, 300 Series Microminiature Thermocouple Probe specifi-
cation sheet, 2727 Philmont Ave., Suite 101, Huntingdon Valley, PA 19006, 215-
663–0250, www.pbeckman.com.
关19兴 Moffat, R., “Notes Concerning Temperature Measurement for ISA–Thermocouple
Theory and Practice,” Proceedings of the 28th International Instrumentation Sym-
posium, Department of Mechanical Engineering, Stanford University, May 3–6,
1982, Las Vegas, Nevada.
关20兴 Moffat, R., “Gas Temperature Measurement,” Temperature–Its Measurement and
Control in Science and Industry, Vol. 3, Part 2, Reinhold Publishing, New York,
1962.
关21兴 Moffat, R., Gas Temperature Measurement: Direct Design of Radiation Shielding, 1,
6th edition, Measurement Engineering, 1970.
关22兴 Moffat, R., “Designing Thermocouples for Response Rate,” , Trans. ASME, Feb.
1958.
关23兴 Millisecond Response Thermocouples–Basic Theory, Paul Beckman Company, El-
kins Park, PA.
关24兴 Steinberg, T. A., “Temperature Measurement 共with Thermocouples兲,” ME698, New
Mexico State University, April 19, 1990.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 8
doi:10.1520/JAI102317
Available online at www.astm.org/JAI

Aditi Oza,1 Sudipto Ghosh,2 and Kanchan Chowdhury3

Tribocharging of Particle Contaminants


Evaluated as an Ignition Source in
Oxygen-Enriched Environments

ABSTRACT: Contamination particles make an oxygen system more vulner-


able to fire hazard as they provide a large surface area with reduced mini-
mum energy of ignition. Adiabatic compression, resonance, solid friction,
flow friction, and promoted ignition have acted as the source of ignition in
oxygen systems. In this paper, static electricity generated in solid particles is
proposed to be another possible source of ignition. It is known that the move-
ment of particles in a system results in the buildup of static charges in them.
This static electricity, if supported by suitable geometric configuration and
grounding of metal projections, may lead to propagating brush discharge
共PBD兲 and provide the energy needed for initiation of fire. Particles with low
auto-ignition temperature, particularly in dispersed state, may help in initiat-
ing the fire, while metal particles may aid the propagation because of their
high combustion energy. In this paper, we have determined the velocity pro-
file utilizing the FLUENT code and calculated the charge generated in par-
ticles, exploring the ignition probability of equipment in oxygen environment
through electrostatic discharge. We have also attempted to perform prelimi-
nary calculation to determine charge density and the electric field strength
required for brush discharge. Computational fluid dynamics analyses on
globe valves reveal the presence of probable areas of particulate accumula-
tion in the valve. The decrease in minimum ignition energy 共MIE兲 in the

Manuscript received December 31, 2008; accepted for publication June 13, 2009; pub-
lished online July 2009.
1
Research Scholar, Cryogenic Engineering Centre, Indian Institute of Technology,
Kharagpur 721 302, West Bengal, India, e-mail: oza.aditi@gmail.com
2
Assistant Professor, Dept. of Metallurgical and Materials Engineering, Indian Institute
of Technology, Kharagpur 721 302, West Bengal, India, email:
sudipto@metal.iitkgp.ernet.in
3
Professor, Cryogenic Engineering Centre, Indian Institute of Technology, Kharagpur
721 302, West Bengal, India, e-mail: chowdhury.kanchan@gmail.com
Cite as: Oza, A., Ghosh, S. and Chowdhury, K., ⬙Tribocharging of Particle Contaminants
Evaluated as an Ignition Source in Oxygen-Enriched Environments,⬙ J. ASTM Intl., Vol.
6, No. 8. doi:10.1520/JAI102317.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 389
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
390 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

environment of high pressure, high temperature, high purity of oxygen, low


humidity, small particles, high dielectric constant, low thickness of insulating
layer, and high resistance of the deposited dust layer facilitates occurrence
of PBD, which may cause ignition. Adiabatic heating may aid combustion by
lowering the MIE of particles. Experiments have been proposed for verifica-
tion of the hypothesis and analytical calculations presented by the authors.
The paper may give a new insight into the possible causes of oxygen fire.
KEYWORDS: oxygen, ignition, tribocharging, propagating brush
discharge, dust layer

Introduction
An alarming feature about oxygen fire is that it can occur in a system any time,
even after several decades of safe operation, with no apparent change in geo-
metric or operating parameters. The reports of investigations of several oxygen
fires, which based their analyses on forensic evidences and observations, could
not indicate the causes of fire in some instances 关1–7兴. However, the details of
the post-accident analyses presented in these reports have proved to be of im-
mense value to the researchers in the field of oxygen safety as they have thrown
up fresh issues to the fore and opened up new avenues of investigations in
different directions.
An accident, which sometimes may take place without any apparent or
deliberate cause, usually implies a negative probabilistic outcome, which re-
sults from a combination of many negative causes, such as design, manufac-
turing, material, and operating, occurring simultaneously. While an accident,
by definition, is not predictable, it is however preventable only if the root
causes and circumstances leading to the accident are recognized and acted
upon prior to its occurrences. It is quite clear from the accident reports that
some gap still exists in the comprehensive understanding of the mechanism of
ignition in oxygen environment. This paper is an attempt to contribute in this
direction discussing the role of contamination particles, which can become
charged with static electricity due to impact with pipes, bends, and valves; such
particles may eventually discharge their static electricity and thereby provide
the needed minimum ignition energy 共MIE兲 of the particles to start ignition and
thus complete the third arm of fire triangle.

Background
Gallus and Stoltzfus 关1兴 observed that the occurrence of fire in whole oxygen
system component body could only be possible if the fire in polymer particles
ignited the metal contaminants, which in turn produced enough energy for
fire-engulfment of the metal body. They also pointed out that the presence of
finely divided materials, such as worn-out non-metal parts, hydrocarbon, and
sand contaminations, contributed to these fires. The three fire incidents dis-
cussed, namely, in a commercial airliner, Royal Australian Air Force, P-3 Orion
Spacecraft, and White Sands Test Facility test system brought out the following
important observations:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 391

• Three fires were caused under different operating conditions, different


materials, and geometries.
• All three systems were at high oxygen pressure.
• In all probabilities, the fire in the three systems originated in the poly-
mers.
• Systems that ran for decades without any problem suddenly caught fire
even in absence of any apparent or abrupt change in operating param-
eters or handling procedure.
Dicker and Wharton 关2兴 reviewed incidents occurring in the compressed
gas industry in Great Britain in the early years of 1980s and highlighted the
following:
• A large number of fires occurred in regulators, cylinder valves, manifold
check valves, and stop valves.
• Incidents are caused when contaminating materials are present and are
exposed to high pressure oxygen.
• These contaminating materials include hydrocarbon and silicone oils or
grease, fragments of polymers, deposits of rust, and sealing tape.
Grubb 关3兴 briefed that the root causes of the fires in RAAF have been iden-
tified as:
• The location of fire was a poppet valve in an aluminum manifold check
valve assembly.
• Sufficient quantity of contamination was present in the system, which
primarily consisted of iron oxide.
• Deterioration of silicone rubber seal believed to be the cause of failure.
McMahon and Nguyen 关4兴 discussed some fire accidents. Relevant for us is
the information that soft seat of cylinder valve was the location of fire and
adiabatic compression at the dead-end, which resulted in high temperature at
the seat, was thought to be the cause of fire. The study and analyses conducted
by Newton and Forsyth 关5兴 on two air-separation unit fire incidents concluded
that:
• The fire originated in valves 共ball and butterfly兲.
• The valves were only slightly opened when fire occurred.
• Particle and hydrocarbon contamination were present.
Gusky 关6兴 remarked that the combination of a contaminant accumulation
on seat and diaphragm and ignition mechanism such as adiabatic compression
makes the regulators the most vulnerable component for system failure. He
also observed that in the oxygen system failure, the components usually ignited
were valve seat, valve biasing spring, diaphragm, and pressure adjusting spring.
Newton et al. 关7兴 observed that the accumulation of contaminants occurs in
valves and regulators. They conducted experiments with contamination on
valves and regulators and observed that the ignition prone region is the prox-
imity of valve/regulator seat, non-metallic diaphragms, and the stem region.
What transpires from these reports are the following:
• Most of the oxygen systems, where fire developed, ran for several years
or decades. However, fire suddenly erupted one day, which may be due
to a dynamic event or change in operation.
• Valves are the prime suspect as the location for initiation of fire. Com-
ponents in reviewed papers, however, vary greatly in size and service,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
392 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

from a few millimeters diameter and high pressure 共⬎2000 psig兲 cylin-
der service to hundreds of millimeters in diameter but low pressure 共
⬍600 psig兲 pipeline service. Non-metal seat is one of the most vulner-
able parts to be ignited in a valve. Bends and attached instruments are
other possible locations. In regulators, the components ignited are valve
seat, valve biasing spring, diaphragm, and pressure adjusting spring.
However, particulate contamination could have been present in each of
these services.
• In some cases of fire, particles existed and contributed to ignition.
• Particles consisting of a mixture of polymers or non-metals are some-
times responsible for ignition in some of the fire incidents. High heat of
combustion arising out of large polymer mass in the component may
lead to ignition of the body without foreign contamination, as indicated
by laboratory-based adiabatic compression tests. However, a small mass
of polymer fire often cannot engulf valve body, and the presence of metal
particles may act as intermediary to provide the heat of ignition to the
valve body.
• Fires seem to start at the high pressure side of the valve.
It may be important to repeat the fact that while oxygen and fuel are ever
present in an oxygen system, it is the ignition energy that is missing to com-
plete the fire triangle for initiating fire. The laboratory experiments have mostly
provided heat energy to the promoter in promoted ignition test, while high
velocity large particles or mechanical impact of large bodies has provided en-
ergy in some cases. However, the scientific community has often not succeeded
in replicating firing incident in a laboratory in spite of setting all parameters
beyond the permissible values.
Considering the valve seat as the location, particles 共metal and non-metal兲
as the cause, and lubricating oil as an extra aid for initiation and propagation
of oxygen fire, there are three different hypotheses, which have so far been
forwarded in the literature as the possible causes of initiation and subsequent
propagation of fire:
• Particle impact and adiabatic compression: Newton and Forsyth 关5兴 re-
ported the causes of failure of valves as a combined effect of particle
impact and pneumatic impact in the seat-stem region. They also ob-
served that on the accumulation of oil layer over the seat contributed
towards ignition apart from the particle impact and adiabatic compres-
sion of gas providing the thermal energy to the particles.
• Mechanical Impact: Valves and regulators with spring loaded seals are
susceptible to ignition due to mechanical impact. The transient flow
conditions impose fluid pressure and force variations on the valve stem.
This might cause rapid throttling of the valve resulting in chatter or
repeated oscillatory impact on polymer seats, creating a mechanical im-
pact ignition hazard 关8兴.
• Flow Friction: The phenomenon of flow friction is also thought to be a
suitable candidate for ignition of non-metal valve seat, though not sup-
ported by a full configuration test 关1兴. Flow friction ignitions in service
were caused due to dynamic reasons like the leak past seal. However, it
has been proven by experiment that the presence of contaminants along
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 393

with any ignition source would accelerate the initiation of the ignition
process.
While the above three situations have been identified as the probable
causes of ignition in oxygen systems, in some cases, it has not yet been possible
to re-create it in laboratory environment whichever unsuitable material or op-
erating parameters one chooses to impose. This has sometimes left the scien-
tists in doubt about the sufficiency or adequacy of our understanding of the
phenomenon of initiation of oxygen fire. In this paper, the present authors
would like to present yet another probable cause of initiation of fire in an
oxygen system, particularly in the valves and regulators:
• Propagating brush discharge 共PBD兲: Flowing particles in an engineering
system get charged with static electricity due to their repeated impacts
inside pipes, bends, and valves. Valves, particularly the valve seats, con-
stitute potential locations for accumulation of these charged particles,
which create an electrostatic field around it. Operators are instructed to
crack open the valves at the start of operation of an oxygen system to
avoid sudden on-rush of oxygen gas and prevent consequent adiabatic
heating. However, this process of crack opening brings the stem very
close to 共yet not touching兲 the valve seat, which may be one of the rest-
ing places of the charged particles. In regulators, the non-metallic dia-
phragm may be the resting place. This configuration has the potential to
create a high voltage between the grounded stem and the charged par-
ticles accumulated on the seat backed by grounded conductor. The volt-
age may well exceed the breakdown voltage of the oxygen space and
result in PBD, which is proposed in this paper to be a possible source of
energy needed to start ignition.
Although there are various types of electrostatic discharges 共ESDs兲, corona
and PBD have relevance in our hypothesis. The description of these discharges
may help in understanding and identifying potential electrostatic hazards in an
oxygen system. The better known electric discharge, spark, is also described for
the sake of comparison 共though a spark is not expected in case of particle
tribocharging兲.

Spark or Capacitive Discharge


Spark or capacitive discharges occur between two conductors at different po-
tentials as soon as the electrical field in the gap reaches the breakdown value of
the surrounding medium. A prerequisite for spark discharge is a somewhat
homogeneous field such that the field strength is high enough across the gap
between the electrodes to effect breakdown.
The discharge channel in spark discharge is narrow and short and the
discharge is fast 共Fig. 1兲. The energy density in a spark discharge may be high
enough to ignite flammable mixtures. Therefore, presence of spark discharge is
usually enough to identify a hazard.

Corona Discharge
The electric field over the charged surface gets distorted due to the presence of
a grounded conductor in its vicinity. The field strength converges towards the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
394 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Generation of spark between two conductors.

surface of the conductor. The field strength is inversely proportional to the


radius of the conductor point. When the conductor point is sharp enough,
corona discharge begins and charges diffuse into the surrounding air 共Fig. 2兲.
Corona discharge is also referred to as one-electrode discharge 共in the field
of electrostatics兲 because this discharge frequently involves one electrode and a
charged insulating surface. However, corona discharge can also occur between
two electrodes when a sufficiently large potential difference exists across them
and the field is non-uniform. The energy density in corona discharge is low, and
the region of ionization is over a close space near the sharp electrode.

Propagating Brush Discharge


PBD occurs when a highly charged insulating surface is backed by a grounded
conductor. In such cases the surface of the insulator may be able to hold an

FIG. 2—Generation of corona discharge.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 395

FIG. 3—Schematic showing the outbreak of PBD.

extremely high charge density. At this highly charged state, when a grounded
conductor with blunt configuration is brought close to the charged surface, the
electric field due to the highly charged insulator surface gets distorted converg-
ing and getting concentrated towards the surface of the approaching conductor.
If this concentrated electric field exceeds the breakdown field strength, the re-
sulting discharge propagates over the surface of the insulative layer 共Fig. 3兲.
PBD, having higher energies than corona discharges, is capable of igniting
many flammable atmospheres. These discharges derive their energy from the
formation of a double layer charge on both sides of the insulating surface.
There are situations in a valve where a non-metallic valve seat is backed by
a grounded conductor. When electrostatically charged particles deposit on the
seat and the grounded valve stem 共which is blunt in shape兲 is situated at a small
distance away, it is an ideal condition for the onset of PBD. PBD may also take
place in situations where non-metallic particles deposit on a grounded metallic
valve seat. Charged particles deposited on non-metallic diaphragm, approach-
ing the grounded metallic push rod in a regulator, may also cause PBD.

Factors Resulting in PBD and Ignition


It has already been pointed out that a large number of negative factors have to
come together to cause an accident. Thus, we need to have a clear understand-
ing of all these factors and make sure that at least one of the factors is absent in
real system in order to avoid fire. We shall discuss the following factors in a
sequence of events that decide the occurrence of PBD and resulting fire:
• Oxygen system is not clean and metal and non-metal particulates 共either
or both兲 are present.
• Electrostatic charging of particles takes place due to impact while flow-
ing inside the system.
• Majority of charged particles accumulate on the valve seat. A small
quantity of particles may get dispersed in oxygen gas and stay in a dis-
persed state between the dust layers and approaching grounded metal.
The dispersed dust particles are the potential target for ignition.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
396 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

• Grounded stem situated very close to the particle layer and the non-
metallic seat backed by grounded conductor 共valve body兲 may result in
PBD, provided the quantity of charge resting on the seat is sufficient and
the dispersed particles have attained the required concentration. PBD
provides the required MIE for ignition of floating dust particles, which
may primarily consist of lighter non-metallic particles having a lower
auto-ignition temperature 共AIT兲.
• Heat of ignition of particles may be sufficient for ignition of metallic
valve body.
In the following sections, these aspects have been discussed and the avail-
able data in the literature have been presented and compared with our calcu-
lations, wherever possible. Sample calculations showing the possibility of igni-
tion have been presented. Proposed experiments shown at the end of the paper
may generate sufficient data to help in establishing the hypothesis as a poten-
tial cause of fire in oxygen systems.

Oxygen System Cleanliness

A system, through which oxygen is flowing over a long time, would contain dust
or particulates however small the amount may be. These dusts may be metallic,
polymeric, or ceramic in nature. They arise from the source of the fluid, from
the abrasion of moving parts such as valves, from unclean components, and
from generation of rust in steel pipes. These solid particles are carried by gas
with some velocity, which in turn generates electrostatic charges on impact or
rubbing against the inner wall of pipes, valves, and such other parts. Commit-
tees and organizations such as European Industrial Gas Association 共EIGA兲,
National Fire Protection Association, Bundesanstalt für Materialforschung und
Prüfung, Deutsches Institut für Normung, National Aeronautics and Space Ad-
ministration, and Compresses Gas Association have set the cleanliness levels
required in oxygen systems 关9–11兴. Table 1 shows the maximum allowable par-
ticles for various cleaning levels in oxygen environment. It has been shown that
the presence of bigger particles makes the system more unsafe than that of the
smaller particles, which is also clear from the requirement given in Table 1 关12兴.
The existing dust particles in oxygen systems attain velocity and momen-
tum from the flowing gaseous oxygen. Particles with diameter smaller than
100 ␮m follow the gas turbulences, whereas larger particles would influence
the gas turbulence due to their inertia 关13兴. The standards 关9,10兴 recommend
that at cleaning level 50, an unlimited number of particles less than 10 ␮m
diameters may be tolerated. Particles of sizes higher than 10 ␮m may be ig-
nited by ignition mechanisms such as particle impact, adiabatic compression,
etc., while particles smaller than 10 ␮m are more likely to be ignited by fric-
tion 共or other mechanisms兲 and then kindle the larger particles 关14兴.
Impact and rubbing of particles against the pipe or valve interior cause
generation 共strictly speaking, separation兲 of static charge in the particles, and
the process is called tribocharging of particles. Accumulation of tribocharged
particles on the valve seat and approaching of valve stem towards the seat
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 397

TABLE 1—Typical maximum allowable particles for different cleaning levels 关9兴.

Particle Size Range Number of Particles per


Cleaning Level 共␮m兲 0.1 m2
300 ⬍100 Unlimited
100–200 93
⬎250– 300 3
⬎300 0
100 ⬍25 Unlimited
25–50 68
⬎50– 100 11
⬎100 0
50 ⬍10 Unlimited
15–25 17
⬎25– 50 8
⬎50 0

result in PBD, which is proposed to be the cause of ignition of valve in oxygen


system.

Triboelectrification of Particles
Triboelectricity is the science of static electricity generated due to friction or
impact between two solids. The triboelectric effect is a type of contact electri-
fication in which certain materials become electrically charged after coming
into contact with another different material and are then separated. The polar-
ity and strength of the charges produced differ according to the materials,
surface roughness, temperature, strain, and other properties. Triboelectrifica-
tion of solids has both beneficial and hazardous effects. The beneficial aspect of
triboelectrification is widely used in many industrial applications, such as the
photocopying technique, powder coating, pre-charging of fabric filters, electro-
statically enhanced cyclone separators, oil mist filtration, aerosol particle col-
lection, and the electrostatic separation of materials. Triboelectrification of par-
ticles takes place in pipes, bends, valves, and sharp edges, while accumulation
may occur in valves, bends, and side branches. Charge accumulation beyond a
critical value with appropriate geometric configuration may lead to ESD.
Many researchers have investigated the charging of particles in fluid flow
systems emphasizing on the fact that the accumulation of charges and their
subsequent discharge might lead to fire and explosions 关15兴. In 1976, Masuda et
al. 关16兴 studied electrostatic characteristics in gas-solid flow in a metal pipe
theoretically and experimentally with particular attention to the collision be-
tween particles and the pipe wall. Their studies on the effects of gas velocity
and particle diameter showed that particles having diameter in the range from
10 to 100 ␮m generated higher current per unit powder flow rate
共1 – 20 ␮C / kg兲 than higher diameter particles for the same gas velocity. Mat-
susaka et al. 关17–23兴 measured the charge separation during the impact of a
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
398 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

sphere on a metal plate and analyzed the electrification by repeated impacts.


They also analyzed the effect of the impact deformation of the sphere on the
charge transfer, the method of simultaneous measurement of the mass flow
rate, and the charge-to-mass ratio of particles based on the characteristics of
the particle charging and the charge balance in the system.
The level of particle concentration in these experiments is usually much
higher than that encountered even in an unclean oxygen system. However, one
can utilize the concept of charge generation and the amount of charge gener-
ated in a single particle, which can be multiplied to the number of particles
present in an oxygen system considered.
The electrostatic measurement with non-contact method based on electro-
static induction by aid of ring sensors and ring probes was performed by Ga-
jewski 关24兴. Glor et al. 关25–31兴 gave an informative overview on theory and
experimentation on possible incendiary discharges from powders poured into a
heap. They compared the incendivity of ESD with the MIE of the particulates.
The MIE is the smallest electrical power stored in a condenser, which suffi-
ciently ignites an ignitable mixture of a combustible atmosphere by an electri-
cal discharge 共mJ兲. The MIE is determined under prescribed test conditions.
They showed that the equivalent energy for the ignition of propane/nitrogen/air
mixtures lies between 1 and 4 mJ. Their work has shown that ordinary brush
discharges cannot ignite flammable powders in the absence of solvent vapors in
air as the environment, even when they have a very low MIE. However, PBD
with the same MIE may cause ignition of flammable powders. MIE is likely to
decrease further in the presence of oxygen-enriched air or pure oxygen, thus
making PBD a potential candidate for ignition in particulates.
Masui and Murata 关32兴 studied the electrification of polymer powder trans-
ported through straight pipe and pipes with 45° and 90° elbows made out of
copper and stated that the charge/mass ratio of the powder passing through an
elbow pipe was larger than that through the straight pipe. They observed that
Nylon 12 particles with mean radius of 52.5 ␮m gained mean charge densities
of 462 ␮C / m2 and HDPE particles with mean radius of 70 ␮m gained mean
charge densities of 50 ␮C / m2. Nifuku and Katoh 关33兴 studied the relationships
between the discharge characteristics and the ignition of a dust in gas by ap-
plying electrical power. They also investigated the ignition characteristics of
dust and the gas explosions and stated that the ignition of dust is influenced by
the particle size and the MIE of the particles in the flowing medium. For con-
stant dust concentration, the MIE for particles with diameter of 10 ␮m was
found to be nearly 50 % less than that for particles with 150 ␮m diameter
based on energy per unit mass.
Eckhoff 关34兴 highlighted the numerous case histories of dust explosion and
indicated the relationships with different parameters, such as particle concen-
tration, humidity, oxygen purity, etc., that influence ignition. The occurrence
and incendivity of cone discharges in silos from 1 to 3 m diameter, containing
particles of MIE 2 mJ, were studied by Glor and Schwenzfeuer 关35兴. They
showed that the charge transfer of 300 nC in this situation was sufficient to
cause frequent discharges in the silos.
Mehrani et al. 关36兴 studied the effect of electrostatic charge generation and
ESD in fluidized bed. They observed that the finer particles contributed more
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 399

than the coarser particles in development of charge in the bed. Charged par-
ticles coated the vessel walls that required frequent cleaning. Triboelectrifica-
tion of pharmaceutical powders with a size range of 500– 600 ␮m by about 10
pC per particle was investigated by Watanabe et al. 关37兴.
To the best of the knowledge of the authors, the influence of electrostati-
cally charged particles on the possible ignition in oxygen-enriched environment
has not been taken up as a research subject by any research group so far. It is
clear from the existing literature that there is a distinct possibility of existence
of a correlation between ignition of particles in oxygen and electrostatic charg-
ing of particles in flowing gas.
As it has already been mentioned that the particle concentration in the
system reviewed is generally much higher than that found in oxygen systems.
However, it may be important to consider the fact that there is a possibility of
attainment of higher particle concentration in a valve or regulator compared to
the remaining system due to flow condition 共explained later in detail兲. More-
over, the literature cited may help in understanding hazards due to ESD in
systems with any level of particle concentration.

Saturation Charge
Static electrification of particles can generate large electric field in and sur-
rounding them. Tribocharging occurs due to the collision of particles with the
walls of the component. The compression and the decompression of particles
during collision give rise to charge separation. Therefore, the quantity of
charge developed on a particle depends on the surface areas, work function
共which decides the amount of charge separated in each impact兲, and number of
collisions. The number of times the particle collides with the wall is a function
of velocity, time, and geometry. However, a particle cannot acquire charge infi-
nitely since with each collision the charge separation is accompanied by charge
relaxation. The relaxation of charge increases with increasing charge accumu-
lation on a particle, and this tendency finally leads a particle to acquire what is
termed as “saturation charge” or “equilibrium charge.”
Cangiolosi et al. 关38兴 observed that with increased pipe length, the smaller
particles reached saturation more easily due to the higher number of impacts
they experienced. For a pipe 15 m in length, the charge accumulated on the
surface of a 100 ␮m particle is four times less than the equilibrium charge.
The relationship between the average specific charge of particles and the pipe
length is shown in Fig. 4. The rate of charge gain decreases with increasing
length, and after a certain length the particle attains its saturation charge. Fig-
ure 5 shows the observations of Gajewski 关39兴 that the value of saturation
charges depends on velocity of fluid and the particle diameter.

Electrostatic Discharge: A Possible Ignition Source


As the Slusser–Miller triangle explains the necessary condition for fire, an elec-
trostatic ignition leading to a fire or an explosion can be well explained by the
electrostatic fire quadrilateral. Figure 6 shows four arms required for the oc-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
400 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—Relationship between average specific charge and pipe length 关22兴.

currence of an electrostatic ignition of a fire or an explosion 关40兴:


共1兲 Ignitable mixture: The mixture of fine particles and oxygen serves as
one arm of the quadrilateral.
共2兲 Generation or separation of charges: The second arm is added when
tribocharging of these particles 共development of charge through con-
tact and friction兲 occurs. The total amount of charge is a function of
number of particles and the amount of charge that an individual par-
ticle can generate.
共3兲 Accumulation of charges: Accumulation of charged particles on insu-
lating base below prevents easy dissipation of the charges to the
ground; hence charge gets accumulated, forming the third arm of the

FIG. 5—Relationship between charge and velocity of particle for different particle diam-
eters 关39兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 401

FIG. 6—Electrostatic fire quadrilateral.

quadrilateral. Accumulation of low conductivity non-metallic particles


on a metallic base may act in the same manner.
共4兲 ESD in the ignitable mixture: The amount of static charge decides the
strength of electric field produced. The electric field associated with the
charge bound on the surface of the insulator becomes the cause for
charge separation on the conductor, situated at a small distance away.
The insulator might be the valve seat or the dust layer itself formed by
the non-metallic particles. When the accumulation of charges sur-
passes the accommodative capability of the medium, the accompany-
ing electric field causes ionization and subsequent breakdown of the
accumulation, completing the fourth and final arm of the quadrilateral
leading to electrostatic fire or explosion.
The accumulated electrostatic charge may get discharged by either charge
relaxation and leaking to body or ground or by specific mode of discharge
共spark, corona discharge, brush discharge, PBD, or bulking discharge兲, depend-
ing on the system confinement and the geometry.
If any one of the arms of the fire quadrilateral is successfully broken, the
fire cannot take place in an oxygen system. It may be accomplished by making
the system clean when the particles would not be present at all or be present at
a negligible level. This would break the first arm in Fig. 6 and make the system
fire-safe. In case there are particles present and are flowing in oxygen, it is
hardly possible to prevent them from being electrostatically charged. So it may
not be possible to break the second arm. However, in case the system can be
designed in such a way that accumulation of charged particles can be pre-
vented, one can break the third arm and make the system safe. Lastly, even if
the accumulation of charged particles cannot be avoided, one can probably try
to prevent ESD and thus break the fourth arm. Generally, grounding of the
system components is thought to be sufficient to prevent ESD. However, it may
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
402 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Electric field strength due to electrically charged insulator sheet: 共a兲 Isolated
insulator sheet; 共b兲 sheet adjacent to grounded metal plate; and 共c兲 sheet backed by
grounded conductor 关41,42兴.

not always be true. On the contrary, there are circumstances where grounding,
in fact, increases the likelihood of ESD.

Grounding May Not Mitigate Electrostatic Discharge in All Circumstances


The most common source of danger from static electricity is the retention of
charge on a conductor because virtually all the stored energy can be released in
a single spark to earth or to another conductor. The accepted method of avoid-
ing the hazard is to connect all conductors to each other and grounding them to
permit the relaxation of the charges. The electric field associated with the
charge bound on surface of the insulator 共or a layer formed by low conductivity
particles兲 becomes the cause for charge separation on the conductor brought to
its vicinity. The intensity of the electric field emerging from the surface of the
charged insulator decreases with increasing distance from the surface. The
charge relaxation in non-metallic particles takes comparatively more time than
in the metals. The charge relaxation time is more in insulator due to its high
resistivity than that in conductors.
Figure 7共a兲 shows an isolated charged insulator. The extent of its surface
charge density is limited by the breakdown field strength of gaseous oxygen. In
our case, this may be an insulator covered by layers of charged particles. The
presence of the grounded conductor 关Fig. 7共b兲兴 below the insulator 共albeit with
a gap兲 redistributes the field in such a way that a much higher charge density
can be accommodated before field breakdown. The grounded conductor may
be valve body in our case. This effect intensifies when the grounded backing
conductor comes in contact with the insulator. This effect, in reality, can be
observed when charged low conductivity particles accumulate on the non-
metallic seat, which is in direct contact with the valve body 关Fig. 7共c兲兴. The
situation is almost identical in cases where charged low conductivity particles
accumulate on a metallic seat of the valve.
Please note that there is one grounded conductor below the insulator serv-
ing as the base and another grounded conductor from above approaching the
insulator. In case of this stem touching the particles, all charges would be neu-
tralized through the grounded stem. A small gap is necessary for the onset of
ESD. As the conductor from above slowly approaches the insulator 共Fig. 8兲, the
field intensity around that conductor increases and it eventually may reach to a
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 403

FIG. 8—Highly distorted electric field in the presence of approaching grounded


conductor.

level when an electrical breakdown may occur. This electrical breakdown may
either lead to corona discharge, if the conductor has a sharp edge, or to PBD, if
the conductor is of blunt shape.
The presence of the grounded backing conductor below the insulator helps
in retaining higher charge on the insulator compared to what is possible in
absence of this backing. This leads to a much stronger discharge 共having higher
energy兲. In case the back-up metallic base is not grounded, the discharge would
take place at a much lower level of charge accumulation, thereby leading to
weaker discharge. This breakdown takes place when the field in front of the
metallic conductor above exceeds the “breakdown field,” a parameter that de-
pends on many factors such as type of gas 共gaseous oxygen in present case兲,
humidity level, impurity level in gas, etc.
When the electric field breaks down and discharge takes place, the intensity
of the discharge may be strong enough to run or “propagate” along the insulat-
ing surface neutralizing the high charge density. Discharges from the dust 共to
the approaching conductor from above兲 or to the dust 共from the conductor兲
共depending on the sign of the charge兲 are possible in grounded metal contain-
ers. Hence, this phenomenon is contrary to the very obvious solution to elimi-
nate static discharge: “ground the system and eliminate the danger of static
electricity.”
A similar situation can be realized in valves with charged particles getting
accumulated in intricate parts over the seal backed by valve body 共Fig. 9兲. The
concept of closing of valve or small opening 共crack兲 opening a valve to avoid the
sudden pressure rise in valves may cause the grounded stem to approach the
charged particles within the required distance to result in discharge. Also, when
the valve is just about to be closed, the electric field surrounding the particles
resting on the non-metallic seat or the layer formed by charged particles on the
metallic seat gets intensified. These might lead to a sudden outburst of the PBD.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
404 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 9—Cross-section of valve showing the possible area for outbreak of PBD 关43兴.

As already mentioned, the diameter of the approaching metal also influ-


ences the type of ESD. A small diameter would result in corona discharge even
at a very low voltage, while a larger diameter of the metal may lead to the
outbreak of PBD, as the breakdown electric field is surpassed due to intensifi-
cation of electric field lines around the approaching grounded metal.
It may be noted here that the present study does not question the general
practice of grounding to neutralize the charges on pipe and valve bodies and
the crack opening of valve in oxygen systems to avoid particle impact and
adiabatic heating. However, it is their potential to cause PBD that is needed to
be understood to prevent fire.

Dispersed Dust and Sufficiency of Energy from PBD for Dust Ignition
Sufficient energy can be stored before the outbreak of PBD, which is released as
a combination of heat, light, and sound energies, and the heat component of
PBD is sufficient enough to ignite vapors, gases, and dusts. The accumulated
dust with the smaller particles may be dispersed with slight turbulence in the
system. Though dispersed dusts are said to be less sensitive to ignition than
flammable gases/vapors, the difference of MIE is only two to three times that of
gas.
Like other ignition sources, ESD can act as ignition source only when the
minimum energy, which is sufficient for ignition, is attained by the discharge.
This minimum energy is termed as MIE. It is the energy below which an ex-
plosible mixture will not be ignited and above which there is a distinct possi-
bility for the ignition to take place. MIE depends on parameters such as particle
size, particle shape, particle concentration, turbulence, ignition delay time,
moisture content, electrode shape, electrode spacing, circuit inductance, circuit
resistance, supporting atmosphere, purity of oxygen, temperature, and pres-
sure. The MIEs in enriched oxygen atmosphere may be many times less than
those in air. As shown in Fig. 10共a兲, the minimum explosible dust concentration
in 50 % oxygen-enriched system is much less than that in air. It is expected to
be still lower as oxygen enrichment approaches 100 %. It may be observed in
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 405

FIG. 10—共a兲 Influence of oxygen content in air on minimum explosible dust concen-
tration and 共b兲 influence of oxygen content in air on MIE 关34兴.

Fig. 10共b兲 that MIE decreases with increasing oxygen content in the atmo-
sphere. In this case also, the value of MIE would decrease still further as oxygen
content in air approaches 100 %.
Figure 11 shows the relationship of MIE with the particle diameter. Smaller
particles are more vulnerable to ignition compared to the bigger particles as the
former may have the MIE, which are several orders of magnitude less than that
of the latter. This goes directly in contrast to the existing cleaning requirement
of oxygen systems, where smaller particles of unlimited quantity may some-
times be tolerated.
Figure 12 shows the decrease in MIE with increasing temperature and de-
creasing humidity. Unlike the case of corona discharge, where the energy re-
lease rate is small, the charge breakdown in PBD acts like a multipoint ignition
source 共though not as effective as spark兲. This discharge may contain up to
several milli-Joules of energy, so that nearly all gases and vapors and the ma-
jority of dusts under specific conditions could be ignited.
Flammable dusts of submicron sizes, when dispersed, are at greater risk of
ignition due to ESD. For a dust ignition to occur, the suspended solids concen-
tration must lie between the lower and upper limits, which vary from material
to material and are influenced by factors such as particle shape and size distri-
bution 共for example, flake shape may be more prone to ignition than spherical
one due to a higher surface area per volume兲. Figure 13 shows the influence of
dust concentration on the ignition sensitivity. It is clear that MIE has a lowest
value only at a particular dust concentration 共500 gm/ m3兲. However, optimum
concentration may vary with materials among other factors. Thus, both the
lowest dust concentration and the corresponding MIE are not sacrosanct but
vary with material, particle size and shape, oxygen purity, temperature, and
humidity. We have also mentioned that the charged dust particles have been
proven to have a lower MIE compared to that of neutral particles. The value of
MIE for dust layer is much higher compared to that of dispersed dust. This
makes a dust layer unlikely to get ignited at the low level of energy that a PBD
releases in an oxygen system. Hence, dispersed dusts have higher probability of
getting ignited by PBD.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
406 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 11—Influence of particle diameter on MIE 关34,44兴.

FIG. 12—Influence of: 共a兲 Temperature and 共b兲 humidity in MIE 关40兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 407

FIG. 13—Influence of dust concentration on MIE 关34,44兴.

Pyrolysis or devolatilization of particles, which occurs due to heat, pre-


cedes combustion. Therefore, the materials with higher vapor pressure with
combustible nature of its emanating gases are likely to have a lower MIE. With
the dispersed nature of dust, the energy from PBD is shared among less num-
ber of particles when compared with settled dust layer or with dispersed dust at
higher than optimum concentration, making ignition more probable. Addition-
ally, a larger number density makes heat transfer consume some amount of
energy making the value of MIE larger. A dust concentration lower than the
optimum may not be able to utilize the energy of PBD, as the discharge chan-
nels formed in PBD may not meet sufficient number of particles to transfer the
energy. Smaller non-metal particles with higher specific surface area are likely
to outgas more than bigger particles, making it more vulnerable to ignition.
FLUENT 关45兴 code has been used to generate the values of the velocity of
impact of particles and to use it for calculating separation of charges. Energy
from PBD arising out of accumulated charged particles has been calculated to
show the possibility of ignition in dispersed dust.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
408 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

To summarize, tribocharging of the dust particles in flowing gaseous oxy-


gen occurs due to their repeated collisions with the walls of the components.
The charged dust particles get accumulated at some locations, such as, valve
seats, bends, elbows, etc. At these locations electric field sets up around the
dust layer, which gets intensified with the presence of grounded metal backing
the insulator layer. The electric field increases with the decrease in the ap-
proach distance, i.e., the distance between the dust layer and the approaching
grounded metal such as the stem. When the electric field or the voltage exceeds
a critical value, i.e., the breakdown voltage, PBD may occur. The PBD supplies
the required energy to the dispersed dust resulting in ignition.
Dispersed dust may exist over the settled dust layer either due to flow of gas
over the settled dust at a very low velocity or under the circumstances when the
dispersed particles are getting settled. Settled and charged particles are neces-
sary because it would establish the electric field. Dispersed dust is necessary
because probability of its ignition is high due to its low MIE. If both of these
conditions are satisfied together, it makes the system vulnerable to ignition. A
computational fluid dynamics 共CFD兲 based mathematical model has been de-
veloped for the generation of charge in dust in air 共there is no difference in
charge generation between air and oxygen as the medium兲. Based on the cal-
culated average charge on the dust particles and the critical approach distance,
the corresponding energy release rate by the PBD was computed.

Numerical Simulation of Velocity and Charge Generation in Particles

Since the charge generation in the dust particles occurs due to their collision
with component the walls, the trajectories of the dust particles have been ob-
tained using discrete phase modeling 共DPM兲. DPM computes the trajectories of
the particles using a Lagrangian formulation that takes into account the dis-
crete phase inertia, hydrodynamic drag, and force of gravity. The hydrody-
namic drag is computed from the velocity field of the background fluid 共pri-
mary phase兲. The velocity field, in turn, is obtained by solving the Navier–
Stokes equation, which is an Eulerian formulation. Once the trajectory of the
dust particle is known, the charge generation in the dust particle is calculated
from the frequency of its collision with the walls, the difference in the work
function of particle and wall materials, the impact velocity, and other param-
eters related to the impact velocity.
The flow field within the air 共charge gained by particles in gaseous oxygen
would not be different兲 has been obtained by solving the steady-state Navier–
Stokes equations along with the continuity equation using control volume ap-
proach 关45兴 on FLUENT 6.3 platform. The flow is assumed incompressible
wherever the Mach number is below 0.3 关46兴. In case of charge calculation,
which has been done in pipes and bends, the Mach number was always below
0.3; hence, the flow is always incompressible. Since the Re number exceeds
10 000, the flow is assumed to be turbulent. In order to account for the turbu-
lence the k-␧ model has been chosen. The k-␧ model gives better results in the
free stream region and captures the features of the flow more accurately 关47兴.
Structured grids are used in all the simulations and different boundary condi-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 409

tions are used in the simulations as per the requirement. DPM essentially inte-
grates Newton’s second law equation with respect to time. The force on a par-
ticle having unit mass is given by 关45兴:

dup g共␳p − ␳兲
= F D共 u − u p兲 + + Fx 共1兲
dt ␳p
where:
up = particle velocity 共m/s兲;
u = fluid velocity 共m/s兲;
␳p = particle density 共kg/ m3兲;
␳ = gas density 共kg/ m3兲;
Fx = additional acceleration 共N/kg兲; and
FD共u − up兲 = drag force per unit particle mass 共N/kg兲.
This modeling has also been done on FLUENT 6.3 platform. The impact of
the particle on pipe results in an equal and opposite charge on them. The
charge acquired depends on the total potential difference between the contact
bodies, which is made up of surface work functions between the bodies, image
effect near the wall, and space charge effect. The current generated 共I in am-
peres兲 through the wall is given by 关15兴:

I = Wp 再冉 冊 冉 冊 冎
q
mp o

q
mp ⬁
n共⌬x兲
no
共2兲

where:
共q / mp兲o = average charge-to-mass ratio of particles at a reference point
共C kg−1兲;
共q / mp兲⬁ = charge-to-mass ratio of particles at x = ⬁ 共C kg−1兲;
n共⌬x兲 = number of collisions of a particle between x and x + ⌬x;
no = relaxation number of collisions 共which is a characteristic number兲; and
Wp = powder flow rate 共kg s−1兲.
共q / mp兲⬁ and no are given by following equations:

冉 冊q
mp ⬁
=
3␧OVc
共 1 + ␣ 兲 ␳ pD pz O
共3兲

␲ D ␳2
no ⬇ 共4兲
⌬t
2共1 + ␣兲S

where:
␧o = permittivity of air 共F/m兲;
Vc = contact potential difference based on work functions 共V兲;
Dp = particle diameter 共m兲;
zo = gap between contact bodies 共m兲;
S = contact area between particle and wall 共m2兲;
⌬t = duration of contact 共s兲;
␶ = time constant of electrification 共s兲; and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
410 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

␣ = ratio of space charge effect to image charge effect given by the following
equation:
3qzoWp␳Dū 2qzo 3Wp␳Dū 3m␳Dū
␣= / = = ⬎0 共5兲
2␲␧OQ␳pDp3␷¯ ␲␧ODp 2
4Q␳pDp␷¯ 4␳pDp␷¯
The simplifications of Eqs 1–4 are the following:
I
Wp
= a共1 + ␣兲
q
mp
冉 冊 o
+b 共6兲

2n共⌬x兲S⌬t
a= ⱖ0 共7兲
␲ D p2 ␶

6␧OVcn共⌬x兲S⌬t
b= 共8兲
␲ D p3␳ pz o␶

Result and Discussion on Charge Generation


Based on the above formulation, computation of charge separation has been
performed for following cases:
• Determination and comparison of charge separation as a function of
powder flow rate in a horizontal pipe for quartz and polyvinyl chloride
共PVC兲;
• Determination and calculation of charge separation as a function of
particle radius in a horizontal pipe;
• Determination and comparison of current generation in steel pipe with
bend in a vertical plane;
• Determination and comparison of fluid velocity profile in a vertical pipe;
and
• Determination and comparison of charge per unit length as a function
of length in a vertical pipe.
The results of numerical solution of Tanoue et al. 关46兴 are also compared
with FLUENT 6.3 simulation results. The geometry with meshing has been
modeled in GAMBIT. The continuous phase is solved with the k-␧ turbulent
modeling and DPM is used for particle tracking and particle velocity. The gov-
erning equations 共continuity, Navier–Stokes equations, k-equation and
␧-equation, and boundary conditions兲 are discretized over a control volume on
a staggered mesh and solved numerically by means of the SIMPLE method in
FLUENT.

Determination and Comparison of Charge Separation as a Function of Powder


Flow Rate in a Horizontal Pipe for Quartz and PVC
The geometry of the pneumatic conveying system and the type of particles used
in present simulations have been based on the experimental work of Masuda et
al. 关48兴 such that a meaningful comparison between the simulation and experi-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 411

TABLE 2—Material properties and system parameters 关48兴.

Experimental 共Paper兲 and CFD Parameters


Pipe internal diameter 6 mm
Pipe length 70 mm
Pipe material Nickel
Air velocity 30 ms−1
Particle 320 ␮m quartz
Shape of particles Spherical
Coefficient of restitution 0.1
Coefficient of friction 0.3
Computational cell size 0.12⫻ 0.07 mm2
Particle density 共solid density兲 2700 kg/ m3
Young’s modulus of elasticity E 共N / m2兲 300⫻ 1010
Poisson’s ratio of particle 0.21

mental outputs can be made. All parameters of our simulation work have been
kept the same as those used in experiments by Masuda et al. 关48兴. Table 2 shows
the input parameters for simulation. Grid independence test with structured
mesh size is performed where the cell number of 50 000 has been found to be
the size beyond which the parameters do not show any remarkable variations.
Further, the simulation is performed with PVC particles and the current
generated is obtained. Figure 14 shows the results of two-dimensional simula-
tion for current generation. It reveals that the current generation is directly
proportional to the powder flow rate of the particles. This may be understood
with the aid of Eq 6, where the current is directly proportional to the mass flow
rate. The consecutive collisions between the particles and wall intensify the

FIG. 14—Current generated on nickel pipe as a function of powder flow rate.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
412 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 3—Material properties and system parameters 关49兴.

Experimental 共Paper兲 and CFD Parameters


Pipe internal 53 mm
Horizontal pipe length 1400 mm
Vertical 2700 mm
Pipe material Steel
Air velocity 21.4 ms−1
Particle 51 ␮m quartz
Shape of particles Spherical
Coefficient of restitution 0.1
Coefficient of friction 0.3
Particle density 共solid density兲 2700 kg/ m3
Young’s modulus of elasticity E 共N / m2兲 300⫻ 1010
Poisson’s ratio of particle 0.21

charging of the particle and enhance surface charge redistribution. It can be


observed that the current generation for PVC is more than that for quartz.
The static electrification of particles brought about by collision leaves back
the surface of pipe wall charged. The model developed 关21兴 for calculation of
charge generation from particle impact velocity covers the phenomenon of re-
laxation time of particle material, the space charge effect 共arising from the
electric field of charged particles flowing in the pipe兲, and the image charge
effect 共near the wall兲, which have vital role in charge saturation and the charge
buildup in the particles. Relaxation time 共␶兲 of insulating materials is greater
than that of metals owing to their high resistivity. This results in less charge
redistribution. Hence the charge density and the electric field are more for PVC
than for quartz, making PBD more likely in the presence of PVC than in the
presence of quartz particles.

Determination and Comparison of Current Generation in Steel Pipe with Bend


in a Vertical Plane
The experimental results of Masuda et al. 关49兴 have been compared with the
results from simulation for current generated on steel pipes by quartz power of
51 ␮m diameter. Table 3 summarizes the parameters considered as input
simulation, which is the same as that of the experimental data.
Figure 15 shows the current generated on steel pipes as a function of axial
distance. Three-dimensional simulation of current generation in particles for
horizontal and vertical pipe has been done. The effect of inlet and bend gives
rise to higher current at the inlet and around the bend region. The current
generation in the pipe due to particle impact 关Eq 2兴 depends on the number of
particle impacts with the pipe wall 共n共⌬x兲兲.
Generation of higher current around the bend can be attributed to more
number of collisions of the particles with the wall of the pipe bend. Also, it can
be observed that the current generation is constant after certain length of pipe.
This can be explained by the concept of charge saturation, which is in turn
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 413

FIG. 15—Current generated on steel pipe as a function of axial distance.

dependent on the particle material and its mutual potential difference with the
pipe material 共work function兲, the field created by the charge generation in the
flowing gas and the image charge effect on the wall.
The occurrence of complex flow geometries like the valves, bends, joints,
etc. causes repetitive impacts of particles generating more charge.

Determination and Comparison of Fluid Velocity Profile in a Vertical Pipe


Tanoue et al. 关46兴 calculated triboelectrification of particles in gas-solid two-
phase flow numerically utilizing the fourth order Runge-Kutta methods of
Guass type method for solving the particle trajectories and the particle veloci-
ties. They performed the simulation under the following conditions, which
were again solved by the authors using FLUENT for comparison of generation
of static charge in the particles due to impact on the pipe wall.

Fluid Flow—
• The fluid flow is axisymmetric and two-dimensional.
• The incompressible fluid enters the pipe according to one-seventh
power law and the flow is turbulent.
• ␬ − 苸 turbulence model was used to model the turbulence of the flow.
Figure 16 shows the velocity profile for the gas in pipe as a function of

FIG. 16—Velocity profile for gas.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
414 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 4—Material properties and system parameters 关46兴.

Paper and CFD Parameters


Pipe internal diameter 6 mm
Pipe length 200 mm
Pipe material Aluminum
Air velocity 87 ms−1
Particle 50 ␮m glass beads
Turbulence model k-␧ model
Velocity inlet One-seventh power law
Shape of particles Spherical
Coefficient of restitution 0.1
Coefficient of friction 0.3
Particle density 共solid density兲 2480 kg/ m3
Young’s modulus of elasticity of particle
material 共N / m2兲 7.50⫻ 1010
Poisson’s ratio of particle 0.17
Young’s modulus of elasticity of pipe material
共N / m2兲 6.85⫻ 1010
Poisson’s ratio of particle 0.34
Computational mesh size 100⫻ 500

radius ratio r / R 共r: Radial distance from axis; R: Radius of pipe兲. The velocity
profile is parabolic with axial distance. Also, the turbulent kinetic energy and
eddy dissipation ratio are prominent near the walls.

Determination and Comparison of Charge per Unit Length as a Function of


Length in a Vertical Pipe
FLUENT simulations were performed under the following conditions.

Particle Motion—
• Particles are assumed to reflect on collision with the wall for simplifying
the simulation.
• The radial velocity just before impact is considered as the normal com-
ponent of impact velocity.
The continuous phase has the similar parameters and the particles were
tracked by DPM in FLUENT. Table 4 lists the input parameter.
The charge gained by the particles decreases with axial distance, as shown
in Fig. 17. This trend in the current with distance can be explained by the
charge saturation and charge holding capability of particle and wall. As the
particles move with the flow, they would gain and dissipate charge till they
attain the saturation value. The fluid flow affects the particle flow of all sizes.
However, the characteristic feature of Lagrange model is the presence of low
concentration to such a level so as not to influence the fluid flow.
The number of impacts of small particles is more than the large particles.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 415

FIG. 17—Profile for impact charge with length of pipe.

This is due to the fact that the turbulent energy fluctuations 共recirculation of
particles兲 are more effective for small particles. More number of impact at the
recirculation zones leads to more charge gain by the particles.

Determination of Charge Separation as a Function of Particle Radius in a Hori-


zontal Pipe
The variation in charge per unit mass of particle with particle radius has been
done for polytetrafluoroethylene 共PTFE兲, aluminum, brass, copper, and stain-
less steel. The particle velocity is obtained from FLUENT simulation for fluid
velocity 30 ms−1 and pipe diameter 6 mm and is inserted in Eq 3. The charge
per mass ratio increases with decreasing particle diameter as shown in Fig. 18.
It may be observed in Fig. 18 that the charge per unit mass of the particles
of diameters in the range of 25 ␮m is much more than that in higher diameter
range. Therefore, it may be observed that the probability of oxygen system
failure, due to ESD, increases with the presence of smaller diameter contami-
nation. However, the larger diameter particles may cause ignition due to im-
pact. Hence, both large and small diameter particles may get ignited in oxygen
system due to different ignition mechanisms.
It may be concluded that the combination of following factors would have
an enhanced probability of PBD in oxygen-enriched systems:
• A sharp rise in charge per unit mass may be observed for particles with
diameter less than 50 ␮m 共Fig. 18兲.
• High pressure, high velocity, and high temperature of gaseous oxygen or
of particles favor ignition.
• System component geometry and material influence ignition.
The probability of non-metallic particles ignition is high in due course of
PBD owing to their low AIT.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
416 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 18—Variation in charge per unit mass with particle diameter.

Accumulation of Dust
The flow of gases in the system is restricted by bends, joints, regulators, and
valves, which cause maximum velocity fluctuations around these regions. Lit-
erature on failure analyses of oxygen system shows that valves and regulators
are the vulnerable components. Figure 19 shows the flow pattern of gas ob-
tained by CFD simulations in ball valves performed by Chern and Wang 关47兴. It
shows that in the recirculation zones, the pressure reduces towards the centre.
Figure 20共a兲 shows the geometry of the two-dimensional model of globe valve
considered for CFD simulation performed by the authors 关43兴. The flow of
oxygen is considered to be compressible as the Mach number exceeds 0.3
around the seat-stem region. Figure 20共b兲 shows probable particulate accumu-
lation region at the seat and the location just behind the stem. It is also ob-
served that high velocity with sufficient support from the geometry of the com-
ponent can aid particle accumulation. Figure 21 shows that pressure decrease
in radial direction towards the centre of recirculation in the 共a兲 valve-seat re-
gion and 共b兲 the region just behind the stem. As the velocity also decreases in

FIG. 19—Flow pattern of gas flow in valve 关47兴.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 417

FIG. 20—共a兲 Axisymmetric view of grid for stem and seat region 共rotated 90°兲 and 共b兲
predicted particle accumulation in seat and back seat region in globe valve.

the recirculation zone, the entrapped particles, on not gaining sufficient mo-
mentum to get out of the vortices, may get accumulated within the recircula-
tion zone. This reduction in pressure gives rise to the possibility that more and
more dust particles would be entrapped in the vortices as the bulk flow of
oxygen with a very low concentration of dust flows past the recirculation zone.
This phenomenon may lead to a situation where the recirculation zone would
have a much higher concentration of particles than that present in the bulk
flow. Very small particles may be present within the valve/regulator in high
concentration in “dynamic condition” when the flow is on or just have been
stopped. The particles may be blown away after the reopening of the valve/
regulator. Therefore, the absence of small dust particles in the valve/regulator,
when opened for inspection, may not necessarily mean that high concentration
cannot exist in the dynamic condition. However, it is only a hypothesis and
needs to be substantiated by experiments.
The size of the recirculation zone and the extent of the reduction in pres-
sure and velocity would determine the level of accumulation of particles in the
vortices. As soon as the valve is closed, these particles 共already charged due to
impact兲 may deposit on the seat of the valve or the diaphragm of the regulator
favoring the occurrence of PBD. The accumulated charged particles form a
layer above the non-metallic surface backed by a grounded metal surface. In an
oxygen regulator, dusts may accumulate over the diaphragm at the given ori-
entation of installation shown in an EIGA leaflet 共Fig. 22兲.
Charged dust accumulation and dispersed dust formation above non-
metallic movable diaphragm make the system prone to PBD. The heating of

FIG. 21—Pressure profiles in the recirculation regions.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
418 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 22—共a兲 Schematic of pressure regulator. P1: High pressure; P2: Low pressure; Pa:
Atmospheric pressure; F: Power of the adjustable regulator spring; f: Power of the valve
spring; S: Area of the diaphragm; and s: Area of the valve seat. 共b兲 Regulator attached to
cylinder 关50兴.

metallic regulator body in the high pressure side due to adiabatic compression
may raise the temperature of charged fine dust particles moving around, thus
lowering their MIE further. Therefore, the ignition of these particles due to
outbreak of PBD becomes likely. The propagation of ignition to the accumu-
lated particles above the diaphragm is proposed to be a probable cause of
failure of regulators.

Calculation of Energy of PBD


Let us assume that the charged particles get accumulated on a surface to form
a dust layer having an area of A and thickness of ␦. Thus, the total charge 共q兲 of
the accumulated particles within the dust layer can be calculated from the
average charge per unit mass of particles 关Eqs 2–6兴 and mass of particles 共mp in
kg兲 as:

q = 共q/mp兲⬁mp 共9兲
Consider the fluid flow with velocity 30 ms−1
in the pipe of diameter 6 mm.
The fly ash particles have porosity of ␷ = 0.65, diameter of Dp = 3.4 ␮m, and a
solid density of 共␳p兲 = 2300 kg/ m3. The average current generated 共I / Wp兲 in
particles of 3.4 ␮m diameter= 0.4⫻ 10−3 C / kg 关48兴. Assuming the initial
charge to be negligible, the average charge per unit mass of particles 关Eq 9兴 is:

共q/mp兲⬁ ⬵ I/Wp = 0.4 ⫻ 10−3 C/kg


Volume of one particle 共Vp兲 = 共␲ / 6兲D3p = 共␲ / 6兲共3.4⫻ 10−6兲3 = 2.058
⫻ 10 −17 m . 3

Mass of one particle 共mp兲 = Vp␳p = 2.058⫻ 10−17 ⫻ 2300= 4.73⫻ 10−14 kg.
The average charge 共q 6 兲11:40:09
Copyright by ASTM Int'l (all rights reserved); Tue May p
on one particle= 共q / mp兲⬁mp = 1.89⫻ 10−17 C.
EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 419

FIG. 23—Model for effective capacitance.

Dusts are assumed to be deposited on the diaphragm of diameter 65 mm


and a surface area of 3318.3 mm2. Assuming that the thickness of powder
deposition is 0.2 mm 共weighing approximately 0.5 g兲, the number of particles
would be:
n = A共0.2 ⫻ 10−3兲 ⫻ 共1 − ␷兲/Vp = 3318.3 ⫻ 共0.2 ⫻ 10−9兲 ⫻ 共1 − 0.65兲/2.058 ⫻ 10−17
= 1.13 ⫻ 1010
The total charge 共q兲 on particles= nqp = 1.13⫻ 1010共1.89⫻ 10−17兲 = 2.14
⫻ 10−7 C = 214 nC.
The surface charge density 共␴兲 = q / A = 214⫻ 10−9 / 3318.3⫻ 10−6 = 64.28
⫻ 10−6 C / m2 = 64.28 ␮C / m2.
When the grounded metal approaches the charged dust layer, a situation
similar to a capacitor is formed as shown in Fig. 23 关51兴. The values of relative
permittivity of gaseous oxygen 共␧1兲, powder bed 共␧2兲, and insulator 共␧3兲 are
taken to be 1.0, 3.0, and 2.0, respectively. The lengths a1 = 2 mm, a2 = 0.2 mm,
and a3 = 2 mm, l = 20 mm, and B = 5 mm. The effective capacitance is given
by:
Ceff = C1 + C23 = 28.26 pF 共10兲
where:
C1 = capacitance of gaseous oxygen 关52兴 and is
2 ␲ ␧ 0l

再 冋冑 冉 冊 册冎
= 共2␲兲8.85 ⫻ 10−12 ⫻ 0.02/ln兵共0.022/共2 ⫻ 0.002
l2 4a1
ln 1+ −1
2a1共B兲 l
⫻ 0.005兲兲
⫻关冑 共1 + 共4 ⫻ 0.002/0.02兲兲 − 1兴其 = 0.73 pF 共11兲
C23 = effective capacitance of powder mixture and insulator sheet
= ␧0A␧2␧3 / 共␧2a3 + ␧3a2兲 = 27.53 pF.
Surface voltage with metal backing 共Vb兲 = q / Ceff = 214⫻ 10−9 / 28.26⫻ 10−12
= 7.57 kV.
Surface voltage without metal backing 共Vwb兲 = q / C1 = 214⫻ 10−9 / 0.73
⫻ 10−12 = 293.15 kV.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
420 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Electric field above the charged surface with metal backing 共Eb兲 = 7.57
⫻ 103 / 0.002= 3.8 MV/ m.
Electric field above the charged surface without metal backing 共Ewb兲
= 293.15⫻ 103 / 0.002= 146.58 MV/ m.
Castle 关51兴 reported that a surface potential of 4–8 kV is sufficient for out-
break of PBD with the thickness range of insulator from as small as 10– 20 ␮m
to 8 mm.
Surface potential generation of 7.57 kV is higher than the stated value,
hence laying the platform for PBD due to charge accumulation on insulator
backed by grounded metal.
To determine the true MIE of a material the duration of the igniting spark
has to be optimal. Eckhoff and Enstad 关53兴 showed experimentally that MIEs
for dispersed dust are strongly dependent on the discharge times of the electric
sparks used in the ignition tests. They observed that the long sparks with du-
ration of the order of milliseconds were less effective than shorter ones, which
are of the order of microseconds. Choi et al. 关54兴 observed that the minimum
value of spark energy that can ignite a polymer powder charged by corona
discharges was an order of magnitude less than the ignition energy of the neu-
tral powder. This observation is very interesting and deserves to be investigated
in case of PBD as well. They also noted that the longer the time for corona
discharge is 共within 700 ␮s兲, the less energy is needed to ignite the dust. In
case of PBD, the discharge time increases as the surface potential is increased
关55兴. Also, Nifuku and Katoh 关33兴 showed that the ignition energy decreases
with increasing spark duration time and that the contact time has a big influ-
ence on the ignitability of the particles. The time span 共t兲 of PBD has been
reported 关55兴 to be of the order of microseconds depending on the intensity of
the discharge.
Hence, considering the initial surface voltage= 7.57 kV, the final surface
voltage= 0 kV, and the effective capacitance to be 28.26 pF, the energy of the
discharge has been calculated 关Eq 12兴, and the rate of discharge energy of PBD
was calculated as 关54,56兴:
1
W = 0.5CeffV2 = 2 共28.26 ⫻ 10−12兲共7.57 ⫻ 103兲2 = 0.81 mJ 共12兲
This energy would raise the temperature of particles from 30 to 500° C for a
certain number of particles 共np兲 given by Eq 13:

np = 0.81 ⫻ 10−3/兵mpcp共500 − 30兲其 = 0.81 ⫻ 10−3/4.73 ⫻ 10−14共800兲470


⬵ 45 500共0.002 mg兲 共13兲
Similar calculations for particles of 320 ␮m diameter, having same thickness
of deposition, give following values: Charge on single particle= 3.36⫻ 10−13 C;
number of particles= 135.12⫻ 103; total charge= 45.4 nC; surface charge den-
sity of 13.68 ␮C / m2; surface voltage with backing of 共Vb兲 = 2.47 kV; and elec-
tric field with backing of 共Eb兲 = 1.24 MV/ m. It can be observed that the depo-
sition of larger particles with same thickness is safer than the smaller particles
as the electric field is lower than the breakdown field.
Yamaguma and Kodama 关55兴 in their experimentation of PBD on insulating
film with grounded metal observed that for the initial surface potential of 10 kV,
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 421

the discharge energy was nearly 1 mJ. The charge amount on the onset of PBD
with respect to the initial surface voltage was approximately 1 ␮C. It can be
observed that it is practically impossible to determine the energy of a PBD
precisely because the measurements of potential and charge distributions
throughout the area concerning the PBD, which are necessary to compute the
energy, are quite difficult from a technical viewpoint. Any equivalent circuit
representing the PBD may also be difficult to devise. However, for evaluating its
incendivity, it would suffice to estimate the energy level of a PBD with rather
simplified methods.

Possible Chain of Events Leading to Particle Ignition in a Cylinder


Regulator
• Metallic and non-metallic particles are generated inside an oxygen sys-
tem due to movement of valve stem and rubbing of seat. Each oxygen
cylinder has particles in them, however small in quantity. Because of the
presence of filter, only small particles, possibly less than 10 ␮m of size,
are allowed to go into the system. Unless it is cleaned properly, there is
a possibility that over the time span of years, a large number of particles
may accumulate inside the system. Smaller particles have more charge
density and less MIE than larger particles, making them more vulner-
able for ignition under ESD.
• As oxygen gas flows, these dusts acquire static electric charge due to
impact. As shown, valve seat/regulator diaphragm region forms recircu-
lation zone where pressure and velocity drop. Such conditions aid accu-
mulation of particles in these regions. The particle concentration in the
recirculation zone has the possibility to attain a much higher value than
that present in the bulk flow or in the rest of the system. However, as the
system is switched off and the flow of oxygen stops, the charged dust
particles inside the regulator start settling down on a metallic or a non-
metallic surface 共 non-metal dusts have very high resistivity preventing
rapid charge relaxation兲. As more particles accumulate on the seat or
diaphragm, the charge density and electrostatic field strength increase.
Dust concentration in the space between the charged layer and the con-
ductor may decrease to a level of optimum value at which MIE is lowest,
facilitating or increasing the chances of ignition.
• As the charged layer of dusts remain over the diaphragm and the regu-
lator is opened, the turning of the regulator screw brings the conductor
nearer to the charged layer and a very small amount of flow blowing
over the accumulated dust layer allows only the smaller particles to float
in the volume. Once the dispersed dust has an optimum density 共possi-
bly for a given flow兲, MIE may just match the energy from the ESD.
• Any adiabatic heating due to sudden opening of cylinder valve may heat
up the regulator body, which would in turn heat up the dispersed dust
decreasing the MIE further, thus aiding the process of ignition.
• The high purity of oxygen and the absence of humidity would decrease
the MIE of particles further.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
422 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 24—Experimental set up for tribocharging of particle in air. 1: Air compressor. 2:


Regulator. 3: Pressure gauge. 4: Rotameter assembly. 5: Hopper. 6: Humidity sensor. 7:
Faraday cage. 8: Temperature sensor. 9: Collecting chamber for charged particles. 10:
Electrometer. 11: Switch.

• The dispersed dust may contain non-metal particles, which have lower
mass density and thus able to float with a small flow of oxygen gas.
Non-metal particles have lower MIE than metal particles.
• Ignition of non-metal particles in the dispersed dust may set off a chain
of ignition, namely, to larger non-metal particles in the dust layer, then
to metal particles, and finally to the metal body.

Proposed Experiments
Figure 24 shows the experimental set up for the measurement of static charges
in particles flowing in air. The pipe material is copper with 8 mm inner diam-
eter, and a mixture of PVC and steel particles of diameter from 100 to 300 ␮m,
in various proportions, is considered for the experiment. The temperature and
humidity of air would be controlled and taken as parameters.
Particles would be injected in the pipe with compressed air flow. The
charged particles would induce equal and opposite charges in the pipe wall,
which would be measured by the assembly of Faraday cage and electrometers.
The effect of pipe length on the charge generation would be studied.
Figure 25 shows in situ tribocharging of particles where an experiment on
PBD experiment would be carried out. The experiment would be conducted in
two steps:
共1兲 The charging of particles and their accumulation on the insulator
sheet. A perforated copper tube would be inserted into a narrow Fara-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 423

FIG. 25—Tribocharging of particles and PBD experiment in oxygen environment.

day cage through which high pressure air would be passed. The inner
vessel of Faraday cage would contain particulates, which would be col-
liding with the inner wall of the vessel and between them, thus acquir-
ing tribocharge. A thin insulating sheet would be placed at the bottom
beforehand. During tribocharging, the inner vessel would be grounded
through a switch to emulate the reality. Just before turning off the air
flow, this switch would be disconnected and the perforated pipe taken
out with the help of a PTFE tong out of the inner vessel to allow the
Faraday cup to measure the amount of charge generated through tri-
bocharging.
共2兲 The perforated pipe and the particulates would acquire charges of
equal amount but of opposite sign. The pipe is gradually taken out of
the Faraday cage, grounded, and then inserted back gradually into the
Faraday cage controlling the movement by a fine thread and thereby
measuring the position of its tip accurately relative to the upper level of
the powder or the insulating sheet. This action may emulate the situa-
tion where the stem approaches the valve seat. The voltage and current
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
424 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

are measured by electrometers if and when PBD is produced due to the


charged particles.
The stepwise process in the experiment would replicate the effects of accu-
mulation of charged particles on the valve seat and the position of the stem very
near to it.

Conclusion

Tribocharging of particles in an oxygen system is suggested to be a potential


cause of ignition. Individual particles acquire necessary charge due to its im-
pact with the pipe and valves, which when accumulated on the valve seat may
generate an electric field sufficient to cause ignition. The knowledge in the
existing literature in the field of particle tribocharging and generation of elec-
tric field to cause ESD may be applied to an unclean oxygen system. Further,
the present study revealed the following:
• Contaminations in oxygen systems have sufficient potential to cause
ESD.
• The CFD results obtained by the authors matched well with the existing
literature of tribocharging of particles in pneumatic systems.
• The CFD analysis shows valve seats and some other spots to be possible
locations for accumulation of particles. Recirculation zones within a
valve or regulator have the potential to attract and retain particles, such
that the particle concentration within the recirculation zone may far
exceed the concentration in the bulk flow.
• Smaller particles could attain higher charge than larger diameter par-
ticles. Investigation in this area may have the potential to redefine the
existing international codes on the level of cleanliness required for a safe
oxygen system by limiting the number of fine particles as well.
• PBD has sufficient energy for igniting the dust particles under dispersed
condition in an environment of enriched oxygen, when several coexist-
ing factors, such as temperature rise due to adiabatic compression, low
humidity, higher oxygen purity, higher oxygen pressure, etc., cause fur-
ther lowering of MIE of particles.
• Dust layer without some dispersed dust may not be sufficient to cause
ignition as the MIE of dust layer is high. On the other hand, a system
with some dispersed dust but without a dust layer is safe; PBD can only
result from the accumulation of charge. Only a charge dust layer can
provide such a condition.
• The material of valve seat, which usually generates the non-metallic par-
ticles, has to be chosen based on its MIE and wear-resistance. Any pro-
jected metallic part even in the pipeline may have to be seen with sus-
picion, particularly if it is projecting towards the locations where
particulates may accumulate.
• An extensive study to understand the mechanism of tribocharging, con-
sequent ESD, and ignition of particles may be required. The generation
of sufficient data in this field could lead to a safer design of oxygen
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 425

valves and regulators through the selection of appropriate materials,


configuration, orientation, and operation of oxygen systems.

Acknowledgments
The enlightening discussion that one of the writers 共K.C.兲 had with Prof. Ted
Steinberg, during his short stay at Queensland University of Technology, with
Endeavors India Executive Award program by DEST 共June–September 2007兲, is
gratefully acknowledged. Contributions of N. Madhujith, Mechanical Engineer-
ing undergraduate student at National Institute of Technology at Warangal,
India, and Dr. P. Ghosh, Assistant Professor, Cryogenic Engineering Centre, IIT
Kharagpur, are appreciated. Funding given by Indian Space Research Organi-
zation 共ISRO兲 through a research project on Oxygen Safety is also gratefully
acknowledged.

References

关1兴 Gallus, T. D. and Stoltzfus, J., “Flow Friction Fire and Research,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 1479, Vol. 11,
ASTM International, West Conshohocken, PA, 2006, pp. 151–162.
关2兴 Dicker, D. W. G. and Wharton, R. K., “A Review of Incidents Involving the Use of
High-Pressure Oxygen from 1982 to 1985 in Great Britain,” Flammability and Sen-
sitivity of Materials in Oxygen-Enriched Atmospheres, ASTM STP 986, Vol. 3, D. W.
Schroll, Ed., ASTM International, West Conshohocken, PA, 1988, pp. 318–327.
关3兴 Grubb, J. W., “Case Study of Royal Australian Air Force P3B Orion Aircraft
Ground Oxygen Fire Incident,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 910, Vol. 2, M. A. Benning, Ed., ASTM Interna-
tional, West Conshohocken, PA, 1986, pp. 171–179.
关4兴 McMahon, J. B. and Nguyen, H., “Australia’s Contribution to Advancing the State
of Knowledge of Oxygen Systems–the Hard Way: A Review of 30 Years Incident
Experience Down Under, 1965-1995,” Flammability and Sensitivity of Materials in
Oxygen-Enriched Atmospheres, ASTM STP 1454, Vol. 10, ASTM International, West
Conshohocken, PA, 2003, pp. 258–267.
关5兴 Newton, B. E. and Forsyth, E., “Cause and Origin Analyses of Two Large Industrial
Gas Oxygen Valve Fires,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 1454, Vol. 10, ASTM International, West Con-
shohocken, PA, 2003, pp. 268–289.
关6兴 Gusky, F. J., “Oxygen Regulator Myths,” Flammability and Sensitivity of Materials
in Oxygen-Enriched Atmospheres, ASTM STP 986, Vol. 3, D. W. Schroll, Ed., ASTM
International, West Conshohocken, PA, 1988, pp. 359–367.
关7兴 Newton, B. E., Langford, R. K., and Meyer, G. R., “Promoted Ignition of Oxygen
Regulators,” Proceedings of the Symposium on Flammability and Sensitivity of Ma-
terials in Oxygen Enriched Atmospheres, ASTM STP 1040, Vol. 4, 1989, J. M. Stoltz-
fus, F. J. Benz, and J. S. Stradling, Eds., ASTM International, West Conshohocken,
PA, pp. 241–266.
关8兴 Beeson, H. D., Stewart, W. F., and Woods, S. S., “Safe Use of Oxygen and Oxygen
Systems: Guidelines for Oxygen System Design, Materials Selection, Operations,
Storage, and Transportation,” ASTM Manual 36, ASTM International, West Con-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
426 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

shohocken, PA, 2000.


关9兴 ASTM Standard G93-03, “Standard Practice for Cleaning Methods and Cleanliness
Levels for Material and Equipment Used in Oxygen-Enriched Environments,”
ASTM International, West Conshohocken, PA, 2004, pp. 13–14.
关10兴 ASTM Standard G127-95, “Standard Guide for the Selection of Cleaning Agents
for Oxygen Systems,” ASTM International, West Conshohocken, PA, 2000, pp. 1–4.
关11兴 Egoshi, N., Kawakami, H., and Fujita, I., “Oil Migration on the Structured Packing
by Evaporation and Recondensation During the Defrosting Operation in the Air
Separation Unit,” Flammability and Sensitivity of Materials in Oxygen-Enriched At-
mospheres, ASTM STP 1319, Vol. 8, ASTM International, West Conshohocken, PA,
1997, pp. 446–457.
关12兴 Benz, F. J., Williams, R. E., and Armstrong, D., “Ignition of Metals and Alloys by
High-Velocity Particles,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, ASTM STP 910, Vol. 2, M. A. Benning, Ed., ASTM Interna-
tional, West Conshohocken, PA, 1986, pp. 16–37.
关13兴 Gimie, M. M., “Turbulent Flow Precipitator for Combustion in Diesel or Gasoline
Engine Exhausts,” U.S. Patent No. 6029440 共2000兲.
关14兴 Fano, E., Faupin, A., and Barthelemy, H., “Selection of Metal for Oxygen Valves,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, ASTM
STP 1395, Vol. 9, T. A. Steinberg, B. E. Newton, and H. D. Beeson, Eds., ASTM
International, West Conshohocken, PA, 2000, pp. 38–56.
关15兴 NFPA 77, 2007, “Recommended Practice on Static Electricity,” National Fire Pro-
tection Association, Inc., St. Paul, MN, pp. 39–41.
关16兴 Masuda, H., Komatsu, T. and Iinoya, K., “The Static Electrification of Particles in
Gas-Solids Pipe Flow,” AIChE J., Vol. 22, No. 3, 1976, pp. 558–564.
关17兴 Matsusaka, S., Ghadiri, M., and Masuda, H., “Electrification of an Elastic Sphere
by Repeated Impacts on a Metal Plate,” J. Phys. D: Appl. Phys., Vol. 33, 2000, pp.
2311–2319.
关18兴 Matsusaka, S., Umemoto, H., Nishitani, M., and Masuda, H., “Electrostatic Charge
Distribution of Particles in Gas-Solids Pipe Flow,” J. Electrost., Vol. 55, 2002, pp.
81–96.
关19兴 Matsusaka, S. and Masuda, H., “Electrostatics of Particles,” Adv. Powder Technol.,
Vol. 14, No. 2, 2003, pp. 143–166.
关20兴 Matsusaka, S., Nishida, T., Yoshiaki, G., and Masuda, H., “Electrification of Fine
Particles by Impact on a Polymer Target,” Adv. Powder Technol., Vol. 14, No. 1,
2003, pp. 127–138.
关21兴 Matsusaka, S. and Masuda, H., “Simultaneous Measurement of Mass Flow Rate
and Charge-to-Mass Ratio of Particles in Gas–Solids Pipe Flow,” Chem. Eng. Sci.,
Vol. 61, 2006, pp. 2254–2261.
关22兴 Matsusaka, S., Oki, M., and Masuda, H., “Control of Electrostatic Charge on Par-
ticles by Impact Charging,” Adv. Powder Technol., Vol. 18, No. 2, 2007, pp. 229–
244.
关23兴 Matsusaka, S., Fukuda, H., Sakuraa, Y., Masuda, H., and Ghadiri, M., “Analysis of
Pulsating Electric Signals Generated in Gas-Solids Pipe Flow,” Chem. Eng. Sci.,
Vol. 63, 2008, pp. 1353–1360.
关24兴 Gajewski, J. B., “Non-Contact Electrostatic Flow Probes for Measuring the Flow-
Rate and Charge in the Two-Phase Gas-Solids Flow,” Chem. Eng. Sci., Vol. 61,
2006, pp. 2262–2270.
关25兴 Glor, M., “Ignition of Gas/Air Mixtures by Discharges Between Electrostatically
Charged Plastic Surfaces and Metallic Electrodes,” J. Electrost., Vol. 10, 1981, pp.
327–332.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
OZA ET AL., doi:10.1520/JAI102317 427

关26兴 Glor, M., “Hazards Due to Electrostatic Charging of Powders,” J. Electrost., Vol.
16, 1985, pp. 175–191.
关27兴 Glor, M. and Schwenzfeuer, K., “Tests to Determine the Ignition of Dust by Brush
Discharges,” J. Electrost., Vol. 30, 1993, pp. 115–122.
关28兴 Glor, M. and Schwenzfeuer, K., “A New Apparatus for Ignition Tests with Brush
Discharges,” J. Electrost., Vols. 40&41, 1997, pp. 383–388.
关29兴 Glor, M. and Schwenzfeuer, K., “Ignition Tests with Brush Discharges,” J. Elec-
trost., Vols. 51–52, 2001, pp. 402–408.
关30兴 Glor, M., and Schwenzfeuer, K., “Direct Ignition Tests with Brush Discharges,” J.
Electrost., Vol. 63, 2005, pp. 463–468.
关31兴 Glor, M., “Electrostatic Ignition Hazards in the Process Industry,” J. Electrost., Vol.
63, 2005, pp. 447–453.
关32兴 Masui, N. and Murata, Y., “Electrification of Polymer Powders After Passage
Through Metal Pipes,” Jpn. J. Appl. Phys., Part 1, Vol. 23, No. 1, 1984, pp. 110–117.
关33兴 Nifuku, M. and Katoh, H., “Incendiary Characteristics of Electrostatic Discharge
for Dust and Gas Explosion,” J. Loss Prev. Process Ind., Vol. 14, 2001, pp. 547–551.
关34兴 Eckhoff, R. K., Dust Explosions in the Process Industries, Gulf Professional Pub-
lishing, Elsevier Science, USA, 2003.
关35兴 Glor, M. and Schwenzfeuer, K., “Occurrence of Cone Discharges in Production
Silos,” J. Electrost., Vol. 40–41, 1997, pp. 511–516.
关36兴 Mehrani, P., Bi, H. T., and Grace, J. R., “Electrostatic Charge Generation in Gas-
Solid Fluidized Beds,” J. Electrost., Vol. 63, 2005, pp. 165–173.
关37兴 Watanabe, H., Ghadiri, M., Matsuyama, T., Ding, Y. L., Pitt, K. G., Maruyama, H.,
Matsusaka, S., and Masuda, H., “Triboelectrification of Pharmaceutical Powders
by Particle Impact,” Int. J. Pharm., Vol. 334, 2007, pp. 149–155.
关38兴 Cangialosi, F., Liberti, L., Notarnicola, M., and Stencel, J., “Monte Carlo Simula-
tion of Pneumatic Tribocharging in Two-Phase Flow for High-Inertia Particles,”
Powder Technol., Vol. 165, 2006, pp. 39–51.
关39兴 Gajewski, J. B., “Charge Measurement of Dust Particles in Motion-Part II,” J. Elec-
trost., Vol. 15, 1984, pp. 67–79.
关40兴 Pratt, T. H., Electrostatic Ignitions of Fires and Explosions, Center for Chemical
Process Safety of the American Institute of Chemical Engineer, New York, 2000.
关41兴 Mannan, S., Lee’s Loss Prevention in the Process Industries, Hazard Identification,
Assessment and Control, Butterworth Heinemann, Elsevier, Linacre House, Jordan
Hill, Oxford, 2004, pp. 16, 94.
关42兴 Ackroyd, G. P. and Newton, S. G., “An Investigation of the Electrostatic Ignition
Risks Associated with a Plastic Coated Metal,” J. Electrost., Vol. 59, 2003, pp.
143–151.
关43兴 Oza, A., Ghosh, S., and Chowdhury, K., “CFD Modeling of Globe Valves for Oxygen
Application,” 16th Australasian Fluid Mechanics Conference, organized by The Uni-
versity of Queensland, held at Gold Coast, Queensland, Australia, 2007.
关44兴 Bartknecht, W., Explosions, Course Prevention Protection, Springer-Verlag, Berlin,
1981.
关45兴 FLUENT 6.3, User’s Guide, 2006, FLUENT Incorporated, Centerra Resource Park,
10 Cavendish Court, Lebanon, NH 03766.
关46兴 Tanoue, K., Tanaka, H., Kitano, H., and Masuda, H., “Numerical Simulation of
Tribo-Electrification of Particles in a Gas-Solids Two-Phase Flow,” Powder Tech-
nol., Vol. 118, 2001, pp. 121–129.
关47兴 Chern, M. and Wang, C., “Effects of Control Devices on Flows in Ball Valve,”
Transactions of the Aeronautical and Astronautical Society of the Republic of China
, Vol. 36, 2004, pp. 291–300.0002-7820
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
428 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

关48兴 Masuda, H., Matsusaka, S., Akiba, S., and Shimomura, H., “Electrification of Fine
Particles in Gas-Solid Pipe Flow,” Kona, Vol. 16, 1998, pp. 216–222.
关49兴 Masuda, H., Komatsu, T., Mitsui, N., and Iinoya, K., “Electrification of Gas-Solid
Suspensions Flowing in Steel and Insulated Coated Pipes,” J. Electrost., Vol. 2,
1976–1977, pp. 341–350.
关50兴 SAG 85/06/E, 2006, “Safe Operations of Pressure Regulators,” European Industrial
Gas Association 共EIGA兲 Safety Newsletter, Brussels.
关51兴 Castle, G. S. P., “The Electrostatic Fields and Discharge Hazards of Insulating
Sheets Close to a Conductor: A Review,” IEEE Trans. Ind. Appl., Vol. 33, No. 1,
1997, pp. 274–278.
关52兴 Kaiser, K. L., Electrostatic Discharge, CRC Press, Taylor & Francis Group, Boca
Raton, FL, 2006.
关53兴 Eckhoff, R. K. and Enstad, G., “Why Are ‘Long’ Electric Sparks More Effective
Dust Explosion Initiators Than ‘Short’ Ones?” Combust. Flame, Vol. 27, 1976, pp.
129–131.
关54兴 Choi, K. S., Yamaguma, M., Kodama, T., Suzuki, T., Joung, J. H., Nifuku, M., and
Takeuchi, M., “Effects of Corona Charging of Coating Polymer Powders on Their
Minimum Ignition Energies,” J. Loss Prev. Process Ind., Vol. 17, 2004, pp. 59–63.
关55兴 Yamaguma, M. and Kodama, T., “Observation of Propagating Brush Discharge on
Insulating Film with Grounded Antistatic Materials,” IEEE Trans. Ind. Appl., Vol.
40, No. 2, 2004, pp. 451–456.
关56兴 Dahn, C. J. and Dastidar, A. G., “Propagating Brush Discharge Initiation of Dust
Layers–A New Test Method,” J. Phys. IV, Vol. 12, 2002, pp. 65–69.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 10
doi:10.1520/JAI102298
Available online at www.astm.org/JAI

Jared D. Hooser,1 Mingjun Wei,2 Barry E. Newton,3 and


Gwenael J. A. Chiffoleau4

An Approach to Understanding Flow Friction


Ignition: A Computational Fluid
Dynamics „CFD… Study on Temperature
Development of High-Pressure Oxygen Flow
Inside Micron-Scale Seal Cracks

ABSTRACT: Flow friction ignition of non-metallic materials in oxygen is a


poorly understood heat-generating mechanism thought to be caused by oxy-
gen flow past a non-metallic sealing surface. Micron-scale fatigue cracks or
channels were observed on non-metallic sealing surfaces of oxygen compo-
nents and could provide a leak path for the high-pressure oxygen to flow
across the seal. Literature in the field of micro-fluidics research has noted
that viscous dissipation, a heat-generating mechanism, may not be negli-
gible as the flow dimension of the channel is reduced to the micron-scale.
Results of a computational fluid dynamics study are presented and used to
determine if temperatures developed in high-pressure driven micro-channel
oxygen flows are capable of reaching the reported autogenous ignition tem-
perature of non-metallic materials in oxygen.

Manuscript received December 29, 2008; accepted for publication August 18, 2009; pub-
lished online September 2009.
1
MSME, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM 88007.
2
Assistant Professor, Dept. of Mechanical and Aerospace Engineering, New Mexico State
Univ., Las Cruces, NM 88003.
3
VP R&D, Wendell Hull and Associates, Inc., 5605 Dona Ana Rd., Las Cruces, NM
88007.
4
Senior Scientist, Test Facility Manager, Wendell Hull & Associates, Inc., 5605 Dona Ana
Rd., Las Cruces, NM 88007.
Cite as: Hooser, J. D., Wei, M., Newton, B. E. and Chiffoleau, G. J. A., ⬙An Approach to
Understanding Flow Friction Ignition: A Computational Fluid Dynamics 共CFD兲 Study on
Temperature Development of High-Pressure Oxygen Flow Inside Micron-Scale Seal
Cracks,⬙ J. ASTM Intl., Vol. 6, No. 10. doi:10.1520/JAI102298.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 429
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
430 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

KEYWORDS: flow friction ignition, viscous heating, oxygen fires, mi-


crochannel flow, cylinder valves, CGA 870 seals, computational fluid
dynamics, stagnation heating

Introduction
Flow friction ignition is a theorized heat-generating mechanism that has been
attributed to the cause of fires in oxygen systems. Flow friction ignition is
believed to occur when oxygen gas flows across or impinges upon non-metallic
materials. It may encompass a variety of gas dynamics such as stagnation heat-
ing and vortex flow 关1兴. Several other potential heat-generating phenomena
have been proposed to help understand flow friction ignition, but it remains a
poorly understood ignition mechanism. Torsion cold working of polymers, vor-
tex tube effects, and polymer friction were investigated by Gallus and Stoltzfus
with promising results, but flow friction ignition has yet to be recreated experi-
mentally 关2兴. In this paper, flow friction ignition is studied in high-pressure
micro-channels as a possible mechanism to ignite non-metallic seals in oxygen.
A proposed micro-channel gas flow scenario is demonstrated and used as the
basis to study frictional heating of high-pressure oxygen gas flow through
micron-sized cracks in non-metallic seals. Computational fluid dynamics 共CFD兲
is utilized to determine whether a critical crack configuration is capable of
developing autogenous ignition temperatures 共AITs兲 of select non-metallic ma-
terials in oxygen. This paper presents examples of fires attributed to flow fric-
tion ignition, micro-channel flow theory, a corresponding CFD model, and re-
sults of initial CFD simulations. The work is part of an ongoing research
program and this paper represents the first stage of the program.

Background
Flow friction ignition has been attributed to the cause of oxygen fires shown to
originate at non-metallic seals where no other known heat-generating mecha-
nisms were believed to be present. Common ignition mechanisms such as adia-
batic compression, particle impact, and promoted ignition are active only dur-
ing rapid transients. When an ignition of a non-metallic material occurs, well
after the end of rapid transients, flow friction ignition is often attributed to the
cause of the fire because of the lack of other credible ignition mechanisms.
The following examples are a selection of medical oxygen fires investigated
by Wendell Hull Associates, Inc. 共WHA兲 and other reported fires believed to be
caused by flow friction ignition. In these fires, non-metallic seals were the prob-
able points of origin and warrant the possibility of flow friction ignition.

WHA Medical Oxygen Flow Friction Ignition Fires

Oak Creek, Wisconsin Fire Department Fire—An oxygen fire occurred in


1999, when a firefighter opened a cylinder valve connected to a medical oxygen
regulator being used on an emergency medical service vehicle. The individual
indicated that the fire occurred immediately after opening the cylinder valve.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 431

FIG. 1—共a兲 Fire-damaged regulator; 共b兲 the cylinder valve seat retainer showing con-
sumption of the seat with melt flow erosion of the seat retainer.

There was a sudden eruption of flame from the regulator. An investigation into
the fire’s origin concluded that initial ignition occurred at the cylinder valve’s
non-metallic seat, which caused the fire to propagate from inside the cylinder
valve to the internal components of the regulator, resulting in fire damage to the
regulator’s aluminum body. The fire-damaged regulator and seat retainer are
shown in Fig. 1. The cylinder valve seat was identified as polyphenylene oxide
共PPO兲, a plastic material known to deteriorate, extrude, and exhibit material
loss during normal service. It was suspected that the PPO seat had degraded
over time, which caused surface irregularities in the oxygen flow stream. That
condition would have increased the exposure of the cylinder valve seat to the
gas flow and increased the likelihood of generating localized heating of the seat.

Alleghany Regional Hospital Incident—Another oxygen fire occurred in


1999, when a patient was receiving medical oxygen at a hospital. The medical
staff set a regulator to administer approximately 4 LPM of oxygen. The patient
was reported to take one to two breaths before coughing and removing the
nasal cannula. The investigation into the fire’s origin concluded that the fire
was focused inside the cylinder valve with the only fire damage resulting in
near total consumption of the cylinder valve’s non-metallic seat. The oxygen
regulator assembly, which safely contained the fire event, and cylinder valve
seat retainer are shown in Fig. 2. Burn residues on the cylinder valve seat
retainer revealed the presence of a plastic seat similar to polychlorotrifluoroet-
hylene 共PCTFE兲. PCTFE seats have been used successfully in oxygen systems,
but WHA and National Aeronautics and Space Administration 共NASA兲 testing
indicated some batches of the plastic were prone to substantial flow 共mechani-
cal instability兲 and tended to severely extrude within the seat retainer. Metal-
to-metal wearing of the seat retainer was also consistent with extreme extru-
sion of the cylinder valve seat. Seat wear and extrusion would have increased
leak paths and increased the opportunity for oxygen gas to generate heat as it
flowed across the seal.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
432 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—共a兲 Oxygen regulator assembly; 共b兲 the cylinder valve seat retainer showing
consumption of the cylinder valve seat.

Howard County Maryland Fire Department Fire—A third oxygen fire oc-
curred in 1999, at a fire station while conducting a leak test of a medical oxygen
regulator. A popping sound was heard shortly after opening the cylinder valve
and was followed by “fizzing” sounds. These audible sounds lasted for a few
seconds prior to the regulator fire. An investigation into the fire’s origin con-
cluded that the fire originated at the CGA 870 sealing gaskets at the inlet to the
regulator. The fire-damaged regulator and inlet seals are shown in Fig. 3. A
nylon crush gasket was found on top of an aluminum-bound rubber washer at
the regulator inlet. Normally, only one gasket is used at the CGA 870 to seal the
regulator and postvalve connection. Having two washers placed onto the regu-
lator inlet does not allow proper insertion of the regulator onto the outlet of the
cylinder valve. The improper installation of the sealing washers may have al-
lowed oxygen to flow past the seals 共i.e., external leak兲, thus increasing the
likelihood of a flow friction ignition. The likelihood of adiabatic compression
causing ignition of the seals was considered low due to the protection provided

FIG. 3—共a兲 Fire-damaged regulator; 共b兲 inlet sealing.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 433

FIG. 4—共a兲 Nylon seat removed from cylinder valve after the wear experiment; 共b兲 15⫻
microscope enlarged view of the nylon sealing surface.

by the regulator inlet, although the gas heating from compression probably
contributed to achieving the ignition temperature of the seal.

Additional Oxygen Fires Attributed to Flow Friction Ignition


Gallus and Stoltzfus reported on oxygen fires related to flow friction ignition
关2兴. One fire occurred in a dome loaded regulator at NASA White Sands Test
Facility in 1986. A Viton diaphragm inside the regulator ignited and kindled the
surrounding stainless steel body. Because the regulator was apparently static
共i.e., no gas flowing兲 at the time of the incident, the fire’s cause was attributed to
flow friction ignition occurring at the Viton diaphragm 共due to an external
leak兲, which was consistent with the fire patterns. Oxygen fires also took place
on commercial aircraft oxygen service carts. Investigators determined that
Vespel seats within manual oxygen valves ignited and kindled the stainless steel
valve stems, which caused fire to breech the brass valve housings. The cause of
the fire was attributed to flow friction ignition initiated by weeping flow past
the Vespel seats.

Observation of Micron-Scale Cracks on Nylon Oxygen Seals


Plastic materials are used to create gas seals over many cycles of use in oxygen
systems. Over time, the continual loading of these seals against metal surfaces
causes wear and surface irregularities to develop along the polymer surface.
The following experiment identifies micron-scale surface cracks that can de-
velop on the surface of a nylon valve seat through repeated use.
An experiment was conducted on a nylon cylinder valve seat to demon-
strate the development of surface micron-scale cracks, as shown in Fig. 4. A
new oxygen cylinder valve was opened and closed approximately 100 times in
order to expose the seat to a period of use and wear. Afterward, microscopic
observation of the seal revealed the presence of micron-scale cracks extending
radially across the surface of the seal 共Fig. 5兲. Measurements of the cracks at
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
434 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 5—60⫻ magnification of micron-scale cracking on the nylon cylinder valve seat
shown in Fig. 4共b兲.

various locations were obtained with a microscope to establish typical crack


widths on the nylon surface. The measurements are tabulated in Table 1.
The presence of micron-scale cracks on the surface of polymer seals could
allow high-pressure oxygen gas to flow across the sealing surface. This poses
the question: Could oxygen flow confined to a micron-scale channel generate
heat as it flows through these cracks? The presence of micron-scale cracks
justified a CFD study that investigated the temperature development of oxygen
gas flowing through micron-sized channels.

Viscous Dissipation in Micro-channel Flows


The flow of fluid through micron-scale devices has gained interest recently due
to advances in micro-machining technology, which has allowed for practical
development of micro-electromechanical systems and micro-fluidic devices.
Micro-heat-exchangers that cool electronics, micro-pumps used to transport

TABLE 1—Micron cracks measurements of nylon cylinder


valve seat.

Nylon Seal Crack Measurement Crack Width 共␮m兲


D1 20.4
D2 17.9
D3 16.0
D4 17.9
D5 30.6
Average width 20.6
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 435

medicine to patients, and micro-heat-engines that generate electrical power are


examples of these devices 关3–5兴. The need to predict the fluid behavior in micro-
fluidic devices is highly desirable to properly validate designs pre-fabrication
and has provided insight about the differences between micro-channel flows
and large scale flows. Most of the research work involving temperature effects
in micro-channel flows is related to heat transfer problems. Few research ef-
forts have been directed at investigating viscous dissipation effects and tem-
perature development generated by a fluid flowing through micron-sized geom-
etries.
Kundu and Cohen defined viscous dissipation as the rate of mechanical
energy loss and a gain of internal energy through deformation of a fluid ele-
ment 关6兴. The rate of viscous dissipation is proportional to the square of the
velocity gradients and is important where fluids undergo high shearing
stresses. The resulting heat generated by viscous dissipation can be intense;
meteorites entering the Earth’s atmosphere often burn up because of viscous
dissipation heating that takes place between the air and meteor surface.
Researchers of viscous dissipation effects in micro-channels have suggested
that the reduced flow dimension can increase the magnitude of viscous dissi-
pation and frictional effects otherwise considered negligible in macro-scale
flows. Koo and Kleinstreuer have analytically investigated the effect of diam-
eter on viscous dissipation 共i.e., temperature rise兲 in liquid micro-tubes and
found that tubes less than 100 ␮m in diameter produced noticeable changes in
fluid temperature, even for an isothermal wall condition 关7兴. The largest tem-
perature rise reported was 25 K for a 20 ␮m diameter adiabatic tube with a
Reynolds number of 2,000. Beskok and Karniadakis noted that the viscous
heating terms for micro-domains may not be negligible 关8兴. They developed an
analytical solution of a general micro-channel heat convection problem and
showed an increase in the tangential temperature gradient along the surface of
the pressure-driven channel as a function of increasing Mach number. The
temperature profiles of their micro-channels with insulated surfaces produced
the largest temperatures at the wall where shear stresses 共i.e., viscous dissipa-
tion兲 dominated. To investigate frictional heating at the channel wall Wu et al.
constructed a 20⫻ 2 ⫻ 4 , 400-micron channel with integrated temperature sen-
sors 关9兴. Nitrogen gas at 1.87 and 2.10 MPa was forced through the thermally
isolated channel. They reported that each temperature sensor reading was
lower than the ambient gas temperature 共297.87 K兲, but a sensor placed in
between the inlet and outlet of the channel read the highest temperature
共297.65 K兲. They attributed the highest temperature reading to frictional heat-
ing generated by the high-pressure gas flow.

Continuum Limit Considerations for Micro-channel Flow

Investigating pressure-driven oxygen flow through micron-scale cracks requires


an understanding of fluid behavior when the characteristic dimension of the
flow passage approaches the mean free path of the oxygen molecule. The mean
free path describes the average distance traveled by a molecule before impact-
ing another molecule. For oxygen, the mean free path is 2.98e−10 m which is
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
436 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 2—Knudsen number and corresponding flow regimes.

Knudsen Number Flow Regime


Kn⬍ 0.001 Continuum
0.001⬍ Kn⬍ 10 Slip/Transitional
Kn⬎ 10 Free Molecular

only one order of magnitude different when compared to the length scale of the
micron-scale cracks observed on the nylon seal in Fig. 5 共20 ␮m兲 关10兴. For this
reason, it was important to consider the limitations of continuum theory.
Continuum theory considers a continuous fluid and assumes all dependent
variables like temperature, velocity, and density can be described as average
quantities over a region that is large in comparison to the fluid particles but
small in relation to the scale of the physical problem being described 关11兴. The
Knudsen number 共Kn兲 is used to distinguish continuum flows from free mo-
lecular flows 共i.e., non-continuum flows兲. Table 2 outlines the difference be-
tween continuum and free molecular flows based on the Knudsen number.
The Knudsen number is defined as the ratio between the mean free path of
the fluid molecule and the characteristic flow dimension

Kn = 共1兲
L

kT
␭= 共2兲
冑2␲p␴2
where:
␭ = mean free path of the molecules,
L = characteristic length,
␴ = molecular diameter,
k = Boltzman constant,
T = temperature, and
p = pressure 关5兴.
Using Eq 2, a range of Knudsen numbers were calculated for different
micro-crack widths for the expected temperature and pressure conditions at
the inlet and outlet of the cracks5 given in Table 3.
By varying the widths of the cracks from 1 – 200 ␮m and using the pres-
sure and temperature values in Table 3, the Knudsen number was compared at
the inlet and outlet of the cracks in Fig. 6共a兲. At the inlet, the Knudsen number
was within the continuum limit for crack sizes larger than approximately
1 ␮m. At the outlet, the Knudsen number was greater than 0.001 for both
temperatures, implying a non-continuum condition.

5
Because oxygen tanks are typically filled with 13.8 MPa oxygen, this value was used as
the inlet crack pressure. The pressure at the outlet of the crack was taken as atmospheric
共0.1 MPa兲. The upper limit of temperature was chosen as 500 K because this is the
nominal AIT of nylon in oxygen.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 437

TABLE 3—Expected values of pressure and temperature at the


inlet and outlet of the micron-cracks.

Pressure, MPa Temperature, K


Inlet 13.8 298–500
Outlet 0.1 298–500

However, preliminary CFD simulations indicated that the gas near the
channel outlet maintained a higher pressure than the 0.1 MPa outlet boundary
pressure. The centerline CFD pressure distribution for a 20 ␮m channel is
shown in Fig. 6共b兲 highlighting the effective outlet pressure maintained near the
channel outlet, approximately 3.4 MPa. Using the effective outlet pressure, the
Knudsen number was within continuum limits at the outlet for crack sizes
larger than approximately 8 ␮m 共Fig. 7兲. Although gas expansion did not occur
instantaneously at the outlet boundary, a major reduction in pressure existed
after this point and the entire length of the micro-channel remained pressur-
ized well above ambient pressure. Therefore, the gas flow up to the outlet
boundary was expected to meet continuum conditions. Additional work is re-
quired to determine the crack size and channel pressure relationship with con-
tinuum flow.

Computational Fluid Dynamics Simulations with Fluent®


A steady two-dimensional turbulent compressible gas flow in a micro-channel
was simulated using the CFD software, Fluent®. Temperature development in
the micro-channels was investigated by varying the channel width, W, from
10– 300 ␮m and length, L, from 500– 10, 000 ␮m. A schematic of the channel
geometry is shown in Fig. 8.

FIG. 6—共a兲 Range of Knudsen numbers varying the channel width from 1 – 200 ␮m
with inlet conditions of 13.7 MPa, and outlet conditions of 101.325 kPa at two tem-
peratures; 298 and 500 K; 共b兲 preliminary CFD centerline pressure distribution for a
20 ␮m width channel and the effective outlet pressure, 3.4 MPa.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
438 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 7—Range of Knudsen numbers at the channel effective outlet pressure, 3.4 MPa.

Pressure-driven flow, consistent with an oxygen tank source feeding the


inlet of the channel and exiting to atmosphere, was used to define the inlet and
outlet boundary conditions of the model. Pressure boundaries were specified as
13.8 MPa and 0.1 MPa for the inlet and outlet, respectively, and gas tempera-
ture boundaries were maintained at 298 K. Boundary conditions at the channel
walls were non-slip and adiabatic, which did not allow for any heat transfer to
or from the surroundings. Heat transfer would normally occur in real micro-
channel flows, but the adiabatic wall condition was chosen to readily compare
heat development within each micro-channel flow.
A typical straight micro-channel configuration mesh is shown in Fig. 9. The
same geometry and meshing approach was used for an enlarged inlet and out-
let channel configuration by adding an enlarged region connected to the inlet
and outlet of the channel 共controlled by the channel height, h, illustrated in Fig.
10兲. To reduce computational time, only one-half the channel was simulated
utilizing a symmetric boundary at the channel centerline. This configuration
allowed for modeling of the oxygen gas flow entering and exiting the channel
and resembles a closer representation of the real micro-channel crack flow
situation.

Computational Fluid Dynamics Grid Independence Study


The simulation results depend on the size of the computational mesh chosen,
especially at boundaries where large gradients are present in the flow field. To
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 439

FIG. 8—Two-dimensional micro-channel crack model for the Fluent® CFD study.

ensure the results obtained for the micro-channel simulations were indepen-
dent of the computational mesh, it was necessary to perform several simula-
tions with different mesh sizes for a single micro-channel geometry.
A 40 ␮m height by 1 , 500 ␮m length channel was chosen for the mesh
independence study. A converged mesh was established when additional grid
points added to the computational mesh produced less than a 0.2 % change in
the simulation results. The computational mesh size was defined by grid points,
which created grid intervals along a micro-channel edge. For example, if N grid
intervals were desired along an edge, N + 1 grid points are required along that
edge. The computational mesh size was denoted by the number of grid inter-

FIG. 9—Typical straight micro-channel configuration mesh.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
440 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 10—Micro-channel configuration mesh with an enlarged inlet and outlet region
attached.

vals along the height and length directions 共e.g., 50⫻ 100兲. Mesh independence
was evaluated using velocity and temperature results obtained for each mesh
refinement size at the channel midplane 共a station located exactly between the
channel inlet and outlet兲. The successive grid refinements and results for cen-
terline velocity and temperature are given in Table 4. Refinement Level 7 was
selected as the converged mesh because additional refinements 共Levels 8 and 9兲
produced less than a 0.2 % change in the centerline velocity and wall tempera-
ture values.

Mesh Sizes for Different Channel Heights


To apply the converged mesh established for the 40 ␮m channel to channels of
different heights, separate meshing techniques were used. This was dependent
on whether the channel height was larger or smaller than the 40 ␮m height
channel. If the channel height was less than 40 ␮m, the same mesh size estab-
lished for the 40 ␮m channel mesh was used.6 This was justified by assuming

TABLE 4—Mesh refinements and corresponding velocity and temperature data.

Refinement Level Mesh Size Midplane Velocity 共m/s兲 Midplane Wall Temperature 共K兲
1 10⫻ 250 246.247 286.828
2 20⫻ 250 252.472 290.117
3 40⫻ 250 248.142 291.456
4 40⫻ 500 247.857 291.461
5 60⫻ 500 247.460 292.315
6 60⫻ 750 247.384 292.314
7 80⫻ 500 247.551 292.914
8 80⫻ 750 247.470 292.915
9 90⫻ 500 247.926 292.985

6
Only 50 mesh cells were used for the 10 ␮m height channels because additional cells
would result in a failed grid when imported into Fluent®.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 441

TABLE 5—Micro-channel simulation test matrix and mesh sizes used.

Channel Length 共␮m兲

Channel Height 共␮m兲 500 1,500 3,000 5,000 10,000


10 50⫻ 167 50⫻ 500 50⫻ 1 , 000 50⫻ 1 , 667
20 80⫻ 167 80⫻ 500 80⫻ 1 , 000 80⫻ 1 , 667 80⫻ 3 , 334
40 80⫻ 167 80⫻ 500 80⫻ 1 , 000 80⫻ 1 , 667
60 94⫻ 167 94⫻ 500 94⫻ 1 , 000 94⫻ 1 , 667
80 108⫻ 167 108⫻ 500 108⫻ 1 , 000 108⫻ 1 , 667
100 122⫻ 167 122⫻ 500 122⫻ 1 , 000 122⫻ 1 , 667
150 157⫻ 167 157⫻ 500 157⫻ 1 , 000 157⫻ 1 , 667
200 192⫻ 167 192⫻ 500 192⫻ 1 , 000 192⫻ 1 , 667
300 262⫻ 167 262⫻ 500 262⫻ 1 , 000 262⫻ 1 , 667

conservatively that the same amount of mesh cells within a smaller channel
height dimension would achieve the same or better resolution demonstrated
for the 40 ␮m converged mesh. If the channel height was larger than 40 ␮m,
the boundary grid resolution was kept the same as the 40 ␮m converged mesh
but additional grids were added to the center region of the channel to make up
the increase in height. This approach limited the computational time required

FIG. 11—Micro-channel height effect on the midplane velocity profile.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
442 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 12—Micro-channel height effect on the midplane temperature profile.

to perform mesh independence studies on all channel heights simulated. Table


5 shows the combination of micro-channel heights and lengths simulated with
mesh sizes used.

Fluent® Simulation Results


The Fluent® simulation results are presented for both the straight and enlarged
inlet and outlet attachment micro-channel configurations. Typical results for
the effect of channel height and channel length on the midplane velocity and
temperature profiles are given first followed by average wall and fluid tempera-
tures compared against the dimensionless micro-channel aspect ratios, L / h,
and Reynolds numbers, Re.
The effect of channel height and channel length on midplane velocity and
temperature profiles are shown in Figs. 11–14. Small channel heights tend to
reduce the magnitude of velocity and increase the magnitude of fluid tempera-
ture at the channel midplane. This result supports the increased influence of
viscous forces generated by the micro-channel wall for smaller height channels;
the closer the channel walls are to each other 共small height distance兲 the more
effect viscous forces will have across the small channel height. Conversely,
channels with large heights 共walls separated by a greater distance兲 were influ-
enced less by viscous forces, as seen by the greater magnitude in midplane
velocity and reduced midplane temperatures. The magnitude differences be-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 443

FIG. 13—Micro-channel length effect on the midplane velocity profile.

FIG. 14—Micro-channel length effect on the midplane temperature profile.


Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
444 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 15—Micro-channel aspect ratio and average fluid temperature.

tween midplane velocity and temperatures are noticeably different. Comparing


the 10 ␮m and 300 ␮m height channels, centerline velocity and temperature
values differed by 105 m/s and 26 K, respectively. A similar effect was seen
when the channel length was varied with a constant channel height 共Figs. 13
and 14兲. Long channel lengths tended to reduce the magnitude of the velocity
and increase the magnitude of the fluid temperature at the channel midplane.
This result corresponds to the additional friction force applied to the fluid flow-
ing through a channel of longer length. Conversely, channels of shorter length
have a reduced amount of frictional force applied along the wall and corre-
spondingly have higher velocity and lower temperature magnitudes at the
channel midplane. The differences in magnitude are again noticeably different.
Comparing the 0.5 mm and 10 mm length channels, centerline velocity and
temperature values taken from the midplane profiles differ by 126 m/s and 25
K, respectively.
The non-dimensional aspect ratio, L / h, and Reynolds number, Re, were
chosen to identify fluid temperature trends throughout the range of micro-
channel sizes simulated. The relationship between micro-channel aspect ratio
and average wall and fluid temperatures is shown in Figs. 15 and 16. Micro-
channels with short heights and long lengths 共i.e., large aspect ratios兲 maintain
higher average wall and fluid temperatures than channels with large heights
and short lengths 共i.e., small aspect ratios兲. This result corresponds to the pre-
viously presented channel midplane results in which channels with short
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 445

FIG. 16—Micro-channel aspect ratio and average wall temperature.

heights and long lengths had larger temperature magnitudes than channels
with large heights and short lengths. The relationship trend between aspect
ratio and fluid temperature tends to flatten for aspect ratios larger than 100.
This indicates an approach to an asymptotic limit, corresponding to approxi-
mately 290 K and 297 K for average fluid temperature and average wall tem-
peratures, respectively. This result suggests that the upper limit of fluid tem-
perature increase was approached with the micro-channel sizes simulated in
this study. Also, channels with the same aspect ratio but different height/length
combinations had comparable fluid temperatures. This indicated a strong cor-
relation between channel aspect ratio and the temperature within the micro-
channel flow field. The average fluid and wall temperatures were also compared
against the channel Reynolds numbers in Figs. 17 and 18. As the Reynolds
number decreases, average fluid and wall temperatures tend to increase, corre-
sponding to a reduction in channel height and an increase in channel length.
Because the Reynolds number is a ratio of the inertial and viscous forces in a
flowing fluid, this result suggested that the inertial forces diminished with de-
creasing channel height and increasing channel length, thus allowing viscous
effects to become more prominent 共i.e., higher average fluid and wall tempera-
tures due to friction at the channel wall兲.
An unusual trend was observed for the average wall temperature results
obtained for channels with the inlet and outlet regions attached as shown in
Figs. 16 and 18. Unlike the smooth trend observed for the average fluid tem-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
446 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 17—Micro-channel Reynolds number and average fluid temperature.

perature results, the average wall temperature trend was not smooth and did
not correlate well with different aspect ratios and Reynolds numbers. This un-
usual trend can be attributed to the possible influence of the inlet and outlet
regions on the wall temperatures, and the difficulty in obtaining a converged
solution for this channel configuration. The addition of the inlet and outlet
regions affected the stability of the solution process and many times would
diverge to an erroneous solution. Converging to a steady-state condition may
not have been achieved with this configuration and further work is needed to
fully understand the effect of the inlet and outlet attachments on the micro-
channel fluid temperature.
Finally, Mach number contours beyond the outlet of a 100 ␮m height
channel are shown in Fig. 19. The Mach number contours suggest that the
exiting flow expands supersonically to the ambient atmospheric air. This is an
interesting result because the high-speed flow leaving the channel outlet may be
of interest to future studies.

Conclusions
A new approach to understanding flow friction ignition that explores whether
gas flow could generate the heat required to ignite non-metallic materials in
oxygen was described. Oxygen fires attributed to flow friction ignition were
presented. A proposed micro-channel gas flow scenario was demonstrated for
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 447

FIG. 18—Micro-channel Reynolds number and average wall temperature.

the types of non-metallic seals that were identified as fire origin. Micro-channel
gas flow was used as the basis to study frictional heating of gas flow through
micron-sized cracks in oxygen seals. CFD simulations were shown to be appli-
cable to the study of crack wall temperatures developed by the high-pressure
oxygen gas flow and to investigate the heat development within micron-scale
channels 共i.e., cracks兲. For the range of micro-channels simulated 共h
= 10– 300 ␮m and L = 0.5– 10 mm兲, viscous effects were most dominant for
channels with large aspect ratios and small Reynolds numbers. Average fluid
temperatures were approximately 30 K higher for channels with large aspect
ratios and small Reynolds numbers than were the average fluid temperatures
for channels with small aspect ratios and large Reynolds numbers. Correspond-
ingly, the average wall temperatures were approximately 4 K higher for chan-
nels with large aspect ratios and small Reynolds numbers than they were for
channels with small aspect ratios and large Reynolds numbers. Although tem-
peratures developed were noticeably different for channels with large aspect
ratios, all average fluid and wall temperatures were well below the temperature
required to ignite nylon in oxygen 共approximately 500 K兲. Seal cracking prob-
ably occurs frequently in nylon seals that are used repeatedly in oxygen sys-
tems. Also, if temperatures developed in micro-channel cracks were high, oxy-
gen fires would be expected to occur more frequently. These results suggest that
the likelihood of an oxygen fire occurring due to an oxygen leak through a
micro-channel crack alone is unlikely and coincides with the rare occurrence of
oxygen fires observed in industry. The micro-channel cracks may be one factor
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
448 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 19—Mach number contours beyond the outlet of a 100 ␮m height channel.

that contributes to oxygen fires attributed to flow friction ignition. Other un-
known factors or a combination of different factors may join together to ex-
plain flow friction ignition in oxygen systems.

Future Work
The Fluent® simulation result of supersonic flow at the outlet of the micro-
channels may allow for shock wave development and the opportunity for stag-
nation heating downstream of the channel outlet. The random nature of crack
development in nylon seals could provide a configuration that emulates a
converging/diverging nozzle 共increasing the likelihood of supersonic flow兲 and
that allows nylon obstructions to stagnate the high-speed flow exiting the crack
outlet. The heating at these stagnation points may be significant and provokes
further investigation as an extension of the high-pressure micro-channel flow
friction ignition theory. The identification of micro-channel cracks and the ca-
pability for supersonic flow, as simulated by Fluent®, provides a significant
contribution and basis for future work.
Finally, interesting information related to localized heating of fatigue crack
tips in nylon polymers was discovered in the literature 关12兴. A hydraulic fatigue
test machine induced fatigue cracks in Nylon 6/6 and crack-tip temperatures
were recorded with a thermal imaging camera. For one test sample, the local-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
HOOSER ET AL., doi:10.1520/JAI102298 449

ized crack-tip temperature was reported as 55° C and was approximately 30° C
higher than the nylon temperature far from the crack-tip region. Future testing
may utilize the heat generated at the tip of a propagating nylon crack as an-
other means to explain flow friction ignition of non-metal seals in oxygen.

Acknowlegments

This flow friction ignition research effort was sponsored by Wendell Hull and
Associates, Inc. in part to fulfill the requirements of a master’s degree at New
Mexico State University 关13兴. Special thanks to Dr. Wendell Hull and the great
people of Wendell Hull & Associates, Inc. for their generous support of this
project. The writers would like to thank Dr. Chunpei Cai and Dr. Ian Leslie for
their contributions to this work.

References

关1兴 Stong, C. L., The Scientific American Book of Projects for the Amateur Scientist,
Simon and Schuster Publishing, New York, NY, 1960, pp. 514–520.
关2兴 Gallus, T. D. and Stoltzfus, J. M., “Flow Friction Fire History and Research,” J.
ASTM Int., Vol. 3, No. 9, 2006, Paper ID JAI13543.
关3兴 Ho, C. M. and Tai, T. C., “Micro-Electro-Mechanical Systems 共MEMS兲 and Fluid
Flows,” Annu. Rev. Fluid Mech., Vol. 30, 1998, pp. 579–612.
关4兴 Epstein, A. H. and Senturia, S. D., “Macro Power from Micro Machinery,” Science,
Vol. 276, 1997, p. 1211.
关5兴 Gad-el-Hak, M., “The Fluid Mechanics of Microdevices—The Freeman Scholar
Lecture,” ASME J. Fluids Eng., Vol. 121, 1999, pp. 5–33.
关6兴 Kundu, P. K. and Cohen, I. M., Fluid Mechanics, 3rd ed., Elsevier Academic Press,
New York, NY, 2004, pp. 105–106.
关7兴 Koo, J. and Kleinstreuer, C., “Liquid Flow in Microchannels: Experimental Obser-
vations and Computational Analyses of Microfluidics Effects,” J. Micromech. Mi-
croeng., Vol. 13, 2003, pp. 568–579.
关8兴 Beskok, A. and Karniadakis, G. E., “Simulation of Heat and Momentum Transfer
in Complex Microgeometries,” J. Thermophys. Heat Transfer, Vol. 8, 1994, pp. 647–
655.
关9兴 Wu, S., Mai, J., Zohar, Y., Tai, Y. C., and Ho, C. M. “A suspended microchannel
with integrated temperature sensors for high-pressure flow studies,” Proceedings of
the 11th IEEE Workshop on Micro Electro Mechanical Systems, 1998, pp. 87–92.
关10兴 Weast, R.C., Handbook of Chemistry and Physics, 54th ed., CRC Press, Cleveland,
OH, 1974, p. 195.
关11兴 Munson, B. R., Young, D. F., and Okiishi, T. H., 2002, Fundamentals of Fluid Me-
chanics, John Wiley and Sons, Inc., New York, NY, p. 4.
关12兴 Wyzgoski, M. G., Novak, G. E., and Simon, D. L., “Fatigue Fracture of Nylon
Polymers: Part 1 Effect of Frequency,” J. Mater. Sci., Vol. 25, 1990, pp. 4501–4510.
关13兴 Hooser, J. D., “A High-Pressure Driven Compressible Gas Flow Study Inside a
Two-Dimensional Uniform Microchannel,” MS Thesis, New Mexico State Univer-
sity, 2009.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 6
doi:10.1520/JAI102269
Available online at www.astm.org/JAI

Nicholas R. Ward1 and Theodore A. Steinberg2

The Rate-Limiting Mechanism for the


Heterogeneous Burning of Cylindrical
Iron Rods

ABSTRACT: This paper presents the findings of an investigation into the


rate-limiting mechanism for the heterogeneous burning in oxygen under nor-
mal gravity and microgravity of cylindrical iron rods. The original objective of
the work was to determine why the observed melting rate for the burning of
3.2-mm-diameter iron rods is significantly higher in microgravity than in nor-
mal gravity. This work, however, also provided fundamental insight into the
rate-limiting mechanism for heterogeneous burning. The paper includes a
summary of normal-gravity and microgravity experimental results, heat trans-
fer analysis, and post-test microanalysis of quenched samples. These re-
sults are then used to show that heat transfer across the solid/liquid interface
is the rate-limiting mechanism for melting and burning, limited by the interfa-
cial surface area between the molten drop and solid rod. In normal gravity,
the work improves the understanding of trends reported during standard
flammability testing for metallic materials, such as variations in melting rates
between test specimens with the same cross-sectional area but different
cross-sectional shape. The work also provides insight into the effects of con-
figuration and orientation, leading to an improved application of standard test
results in the design of oxygen system components. For microgravity appli-
cations, the work enables the development of improved methods for lower
cost metallic material flammability testing programs. In these ways, the work
provides fundamental insight into the heterogeneous burning process and
contributes to improved fire safety for oxygen systems in applications involv-
ing both normal-gravity and microgravity environments.

Manuscript received December 8, 2008; accepted for publication May 1, 2009; published
online June 2009.
1
School of Engineering Systems, Queensland Univ. of Technology 共QUT兲, Brisbane,
Queensland 4001, Australia.
2
School of Engineering Systems, Queensland Univ. of Technology 共QUT兲, Brisbane,
Queensland 4001, Australia.
Cite as: Ward, N. R. and Steinberg, T. A., ⬙The Rate-Limiting Mechanism for the
Heterogeneous Burning of Cylindrical Iron Rods,⬙ J. ASTM Intl., Vol. 6, No. 6.
doi:10.1520/JAI102269.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 450
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 451

KEYWORDS: iron, burning, melting, heat transfer, solid/liquid


interface, microgravity, rate-limiting mechanism

Introduction
This work presents the major findings of a research project relating to burning
metallic materials in microgravity and normal gravity environments. The ob-
jective of the work was to explain why burning metallic materials typically
exhibit a higher melting rate in microgravity compared to normal gravity
共under otherwise similar test conditions兲. However, through the investigation of
this problem, significant insight was also gained into the heterogeneous burn-
ing process in general, which is relevant to both microgravity and normal-
gravity environments. This paper provides a summary of heat transfer analysis,
microanalysis of quenched microgravity samples, and burn test results and
analysis. The results are shown to provide insight into the rate-limiting mecha-
nism for melting 共and burning兲 for metallic materials.

Background
Flammability is not an inherent material property but is instead dependent on
factors including specimen shape, size, composition, configuration and orien-
tation, oxygen pressure and concentration, and many other parameters 关1–4兴.
Flammability is therefore assessed in standardized tests, such as ASTM G124-
95. Relative flammability rankings of different materials are primarily based on
threshold pressure, which is the minimum pressure required to support self-
sustained burning. However, since testing is only performed at discrete sched-
uled pressure intervals, threshold pressure information is often insufficient to
distinguish between materials with similar flammability limits. Therefore, as a
secondary consideration, materials with the same threshold pressure are often
ranked according to the regression rate of the melting interface 共RRMI3兲, which
is a measure of melting rate and burning rate 共with assumptions about geom-
etry and reaction extent兲 关5兴. The RRMI is therefore an important parameter in
assessing the flammability of metallic materials and investigating the factors
that affect RRMI contributes to an improved fundamental understanding of the
processes involved.
Metal flammability experiments conducted in microgravity by Steinberg et
al. 关6–10兴 clearly demonstrated that gravity level affects both threshold pressure
and RRMI. For some metals, lower threshold pressure values were obtained in
microgravity than in normal gravity 关8–10兴. Also, RRMI was consistently ob-
served to be higher in microgravity than in normal gravity and, for iron, by a
factor of between 1.5 and 2 关6–10兴. These results showed that metals are typi-
cally more flammable in microgravity than in normal gravity 共when flammabil-
ity is indicated by threshold pressure and RRMI兲. This discovery raised doubts
about the validity and safety of applying the results of standardized flammabil-

3
RRMI is defined as the rate at which the external 共visible兲 boundary of the solid/liquid
interface 共SLI兲 proceeds along a test specimen.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
452 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—Upward and downward burning configurations for mild steel, from Ref. 12.

ity tests—which are typically performed in normal gravity—to material selec-


tions for spacecraft. The cause of higher RRMI values for metals burning in
microgravity has not been clarified although it is thought to relate to the reten-
tion of molten material 共since periodic droplet detachment does not occur in
microgravity兲 and the associated accumulation of thermal energy and/or al-
tered droplet shape. The work presented in this paper aims to clarify the cause
of the increase in RRMI in microgravity and, through this, to provide insight
into the heterogeneous burning process.

Importance of the Solid Liquid Interfacial Surface Area


An early indication that altered interfacial geometry between solid and liquid
phases may affect RRMI arose during analysis of the influence of sample ori-
entation. Benning et al. 关11兴 compared tests in which burning propagated up-
ward and downward along vertical aluminum test specimens. Although only
burn/no burn results were recorded 共and no RRMI data兲, clear differences in
flammability limits were observed between upward burning and downward
burning samples. It was also noted that the upward burning configuration “al-
lowed the molten slag and burning metal to fall away from the test specimen,
giving more reproducible results” 关11兴. This suggests that altering the way the
molten material interacts with the solid specimen can affect flammability lim-
its.
Sato et al. 关12兴 also investigated upward and downward burning configu-
rations. Figure 1共a兲 shows the position of the soild/liquid interface 共SLI兲
throughout tests at various pressures and in upward and downward burning
configurations. The gradients of these plots represent instantaneous RRMI. The
larger variability in instantaneous RRMI in the downward burning configura-
tion 共at pressure 0.5 MPa兲 is clearly visible in Fig. 1共a兲, consistent with the
observations reported by Benning et al. There is also a large difference in in-
stantaneous RRMI between upward and downward burning configurations at
pressure 0.5 MPa. Figure 1共b兲 presents average RRMI results at various pres-
sures in upward and downward burning configurations. Sato et al. noted that it
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 453

was difficult to track the SLI in the downward burning configuration since the
molten material ran down the side of the test specimen; however, despite this,
there is an obvious difference in RRMI between the two orientations. These
results are of interest because they demonstrate that simply igniting a uniform
sample at the top instead of the bottom caused a large difference in RRMI for
all other test parameters constant. This suggests that the geometric change may
have caused the large difference in RRMI. Indeed, Sato et al. later concluded
that the ratio of contact area 共between liquid and solid metal兲 to the sample
cross-sectional area “is an important parameter for discussing fire spread
mechanisms” 关13兴. In this paper, the importance of SLI surface area and its
effect on RRMI is analyzed and discussed in more detail.

Previous Work on the Rate-Limiting Mechanism


Further insight into the factors that influence flammability is obtained by de-
termining the rate-limiting mechanism. Of the many physical and chemical
processes that occur during burning, the rate-limiting mechanism is the most
restrictive. Benz et al. 关14兴 interpreted the existence of a region of melted but
unoxidized iron 共within the burning droplet兲 as an indication that fluid veloci-
ties within the molten droplet were low 共i.e., no mixing兲. Thus, the authors
concluded that the incorporation and transport of oxygen was the rate-limiting
mechanism. However, the existence of a separate unoxidized iron phase does
not necessarily imply that fluid velocities are low because Wilson et al. 关15兴
showed that since the surface tension of iron is higher than that of iron oxide,
there may be “multiple immiscible liquid phases” present within the molten
droplet. Thus, strong convective cells may exist within each phase, even though
mixing may not occur. Sato et al. 关13,16兴 also reported that the incorporation of
oxygen at the surface of the molten droplet was the rate-limiting mechanism, in
agreement with Benz et al. In contrast, Steinberg et al. 关7,17–20兴 clearly showed
that the molten droplet contains excess oxygen 共above the highest possible stoi-
chiometric requirement for burning兲, which suggested that oxygen incorpora-
tion was not rate limiting. Instead, the rate-limiting step was thought to be the
reaction between oxygen and iron at the iron/iron oxide phase boundary and,
hence, the energy supply rate to the solid. Dreizin 关21兴 noted that the differing
views offered by Sato et al. and Steinberg et al. were not necessarily inconsis-
tent, and could be related to the different experimental methods used. Clearly,
current understanding indicates that the rate-limiting mechanism for heteroge-
neous burning is related to either the energy availability or the process of heat
transfer between the molten droplet and the solid rod; however, the exact
mechanism has not been shown conclusively. Yet this information is vital to
improve the understanding of how various parameters affect the RRMI and
burning in general. This paper contributes to an improved understanding of
heterogeneous burning by identifying the rate-limiting mechanism for burning
iron.

Heat Transfer Analysis


A simple heat transfer analysis was performed by Ward et al. 关22–24兴 to model
the potential effect on RRMI of any change in SLI surface area in microgravity.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
454 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

The analysis was based on an energy balance along the SLI, which equated heat
input from the molten/burning droplet to the sensible and latent absorption of
heat in the solid rod. A brief summary is presented.
Melting of the solid metal occurs as a result of heat transfer from the high
temperature, liquid metal/oxide region. The rate of heat transfer across the SLI,
Q̇, is defined as the integral of incoming heat flux, q̇, over the surface area of the
SLI, ASLI

Q̇ = 冕 q̇dASLI 共1兲

Making the simplifying assumption that heat flux is constant over the SLI, Eq 1
can be written

Q̇ = q̇ASLI 共2兲
Energy input into the solid metal can be expressed in terms of the melting
rate. Heat transfer rate, Q̇, is equal to the product of the mass flow rate and the
sum of sensible and latent energy terms. It represents the energy required to
raise the temperature of the solid metal from ambient, T0, to the melting tem-
perature, TM, and complete the phase change from solid to liquid. As described
in previous work 关22兴, heat transfer rate to the solid can be expressed in terms
of thermophysical properties 共density, ␳, specific heat, CP, and latent heat, ␭兲,
sample cross-sectional area, S, and RRMI, vRRMI

Q̇ = 关␳共CP共TM − T0兲 + ␭兲SvRRMI兴 共3兲


It is reasonable to equate the energy transferred into the SLI from the
liquid phase 共Eq 2兲 to the energy entering the solid phase 共Eq 3兲, which gives
q̇ASLI = 关␳共CP共TM − T0兲 + ␭兲兴SvRRMI 共4兲
It is commonly assumed that the SLI is planar and perpendicular to the rod
centerline axis 关15,25兴. Under this assumption, the surface area of the SLI, ASLI,
is 共by definition兲 equal to the sample cross-sectional area, S, meaning these
terms are eliminated from either side of Eq 4, giving

vRRMI = 共5兲
␳ 共 C P共 T M − T 0兲 + ␭ 兲
This result wrongly implies that, for a given material and constant ambient
temperature, changes in RRMI can only be due to changes in heat flux, q̇. It has
been shown that, especially in microgravity, but also in normal gravity in the
case of altered sample cross-sectional shape 关26兴, that the SLI can attain a
convex, asymmetric, three-dimensional shape. By assuming a planar and per-
pendicular SLI, the effect on heat transfer rate of increased SLI surface area is
ignored 共as there is no ASLI term in Eq 5兲. In this way, the assumption of a
planar SLI can in some cases represent an unrealistic oversimplification of the
heat transfer process, causing insight and information to be lost. If this as-
sumption is not made, then Eq 5 can be rewritten, retaining the area terms
from Eq 4
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 455

ASLI ⫽ S

q̇ ASLI
vRRMI = 共6兲
␳ 共 C P共 T M − T 0兲 + ␭ 兲 S
When RRMI is expressed as shown in Eq 6,4 it is clear that changes in the
melting rate can be caused by changes in heat flux and/or changes in the SLI
surface area, for a given material and constant ambient temperature. This
means the melting rate is sensitive to variations in total SLI surface area, which
is dependent on the three-dimensional SLI shape. This result demonstrates the
importance of SLI geometry and its influence on the melting rate. This is con-
sistent with the earlier conclusion of Sato et al. 关13兴, who noted that the ratio of
the heat transfer surface area 共in this case, ASLI兲 to the sample cross-sectional
area is an important parameter for discussing fire spread mechanisms.
Importantly, this analysis predicts a proportional relationship between the
SLI surface area to the sample cross-sectional area ratio and RRMI. There are
no explicitly gravity-dependent terms in Eq 6 and it can be validly applied for a
first-order analysis in both normal gravity and microgravity. In this way, it is
not specified exactly how a change SLI surface area is obtained, merely that any
change in the ratio of SLI surface area to cross-sectional area will have a pro-
portional affect on RRMI. Following this result, experimental work was under-
taken to determine 共a兲 if the SLI shape was significantly different 共i.e., not
planar兲 in microgravity and 共b兲 whether this correlated in an approximately
proportional sense with an observed increase in RRMI.

Results and Analysis

Microanalysis of Quenched Samples


This section is an overview of microanalysis results that are presented in more
detail in other publications by Ward and Steinberg 关23,27,28兴. The purpose of
the work was to obtain quenched samples of iron rods burning under micro-
gravity conditions in order to confirm the existence of altered solid/liquid inter-
facial geometry, and to quantify the extent of the change in the SLI surface
area. Testing was performed in a ground based, 2-s drop tower facility operated
by the Queensland University of Technology, Brisbane, Australia. Onboard the
microgravity experimental apparatus, the 3.2-mm-diameter iron rods were
burned inside a high-pressure oxygen test system and extinguished by immer-
sion in a water bath. After each microgravity test, the quenched specimen was
recovered from the water bath, embedded in a resin block, polished to expose
the SLI cross section, etched with 2 % nitric acid and then viewed under mag-
nification. This produced images similar to that shown in Fig. 2, which revealed

4
It is possible to further expand Eq 6 by representing the heat flux term with the con-
duction heat transfer expression, based on the temperature difference between the reac-
tion zone and the melting temperature 共at the SLI兲, and an assumed effective heat trans-
fer coefficient. For simplicity, this work is not presented here.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
456 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Microscopy image of 3.2-mm-diameter iron rod that burned in oxygen at the
pressure of 2.21 MPa 共320 psia兲 under microgravity conditions and was quenched by
immersion in water. Note highly convex SLI shape.

four distinct zones: thermally unaffected metal, heat affected metal that never
melted 共heat affected zone兲, melted and resolidified 共but unoxidized兲 metal, and
a thin layer of oxide. The boundary between the heat affected zone and the
unoxidized metal that melted and resolidified is the SLI and, as shown in Fig. 2,
the interface is highly curved, which verifies the existence of a convex SLI in
microgravity.
To quantify the extent of the change in the SLI surface area, it was neces-
sary to obtain a three-dimensional view of the SLI. Using the iterative process
described in more detail in Ref. 27, photographs similar to that shown in Fig. 2
were obtained on multiple planes throughout each sample. Like assessing to-
pography through contours on a map, a three-dimensional representation of
the SLI was created. A numerical method was then used to estimate the area of
the three-dimensional SLI surface. This SLI surface area analysis process was
repeated multiple times per sample to estimate the error magnitude 共by stan-
dard deviation兲. The estimated SLI surface areas are presented in Table 1. In
the table, the area ratio represents the ratio of the estimated microgravity SLI
surface area to the cylindrical rod cross-sectional area. For the 3.2-mm-
diameter test specimens used, the cross-sectional area was 8.0 mm2. Since the
normal-gravity SLI has been shown to be approximately planar and perpen-
dicular to the test specimen centerline axis, the rod cross-sectional area is
therefore a good approximation of the normal-gravity SLI surface area
关14,15,25兴. Consequently, the area ratio values presented in Table 1 can be con-
sidered equivalent to the ratios of microgravity SLI surface area to normal-
gravity SLI surface area. In this way, the area ratio values represent the propor-
tional change in SLI surface area due to a change in gravity level. The estimated
areas and area ratios appear closely grouped, and a dependence on pressure is
not apparent. The average SLI surface area ratio is 1.65 with a standard devia-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 457

TABLE 1—Estimated SLI surface areas for 3.2-mm-diameter iron rods burning in
microgravity.

Pressure
Sample 共MPa/psia兲 Estimated SLI surface area 共mm2兲 Area ratio
1 5.87共850兲 12.5± 0.20 1.55± 0.05
2 6.18共895兲 12.3± 0.41 1.52± 0.08
3 3.31共480兲 13.0± 0.78 1.61± 0.13
4 3.59共520兲 14.6± 0.88 1.81± 0.15
5 3.97共575兲 13.3± 0.80 1.65± 0.13
6 2.28共330兲 14.1± 0.40 1.75± 0.08
7 2.21共320兲 14.2± 0.27 1.77± 0.07
8 2.25共365兲 ... ...
9 2.73共395兲 12.6± 0.76 1.56± 0.12
10 3.73共540兲 ... ...

tion of 0.11. Subsequent work then focused on determining whether this


change in SLI surface area in microgravity correlated with the observed in-
crease in RRMI, as predicted by Eq 6.

Regression Rate of the Melting Interface Test Data


The experimental techniques used to obtain RRMI data in microgravity and
some initial results are described extensively in a previous publication by Ward
et al. 关22兴. The experiment performed was referred to as a “transition test,”
since burning was observed during a transition in the gravity level from normal
gravity to microgravity. A 3.2-mm-diameter cylindrical iron rod was ignited in
normal gravity and allowed to burn 共in an upward burning configuration兲 for
approximately 2 s, then, while burning continued, it was introduced into mi-
crogravity for a further 2 s. This type of test was well suited to a drop tower
facility, which provides a step change in gravity level 共over a period of 0.05 s兲
from normal gravity to microgravity. These tests were significant because, in
addition to confirming that RRMI is higher in microgravity, they clearly
showed that the change in RRMI occurs very rapidly after the reduction in
gravity level. The increase in RRMI was apparent approximately 0.05–0.1 s
after the onset of microgravity and occurred over the same time scale as the
change in molten droplet shape from a “tear-drop” shape 共stretched by gravity兲
to a spherical shape. That is, a change in RRMI was observed when there was a
large change in the bulk shape of the molten material, consistent with a link
between droplet shape and RRMI. Further, it was shown that changes in other
system parameters, such as temperature or the rate of energy release could not
account for such a rapid increase in RRMI at the onset of microgravity.
Some tests were also conducted at constant gravity level, both in normal
gravity and microgravity. Tests were performed over a range of pressures from
1.4 MPa 共200 psia兲 to 6.9 MPa 共1,000 psia兲. The RRMI data obtained from all
tests are plotted in Fig. 3. Error bars represent an estimated 8 % error, which
was obtained by summing individual error terms 共scaling error and visual in-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
458 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 3—Summary of RRMI data obtained in normal gravity and microgravity. Hollow
markers included for comparison to previously published data for iron
关7,9,10,26,29,30兴.

terrogation errors兲 in quadrature. Previously published data by other research-


ers are also shown for comparison and there is general agreement at both
gravity levels, consistent with usual experimental variations.

Discussion

Comparison of Regression Rate of the Melting Interface Data and SLI Surface
Area Results
A general sense of the effect of microgravity on RRMI for burning iron is ob-
tained through further manipulation of the bold trend lines 共for normal gravity
and microgravity test data兲 in Fig. 3. Statistical regression is used to produce an
expression for each of the trend lines and since the proportional increase in
RRMI between normal gravity and microgravity is of interest, it is relevant to
take the ratio of these expressions 共i.e., divide the microgravity trend line ex-
pression by the normal gravity trend line expression兲. The elegant result is
presented graphically in Fig. 4. The resulting line provides an indication of the
general increase in RRMI under microgravity conditions and it clearly shows
that the ratio of microgravity RRMI to normal-gravity RRMI remains almost
constant at a value of 1.8 over the range of pressures considered, within the
experimental error bounds. The error bounds are based on the standard error
terms 共the differences between experimental data points and the trend line
values兲 for both normal gravity and microgravity trend lines. The error bounds
are wider at lower pressures because the standard error 共expressed in absolute
terms兲 is a larger proportion of the smaller RRMI values 共at low pressure兲.
An important question is whether the change in SLI surface area correlates
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 459

FIG. 4—Ratios of microgravity to normal gravity SLI surface areas and RRMI.

with the observed change in RRMI in microgravity, as predicted in the heat


transfer model 共Eq 6兲. To assess the correlation, the area ratio data points from
Table 1 are plotted alongside the RRMI ratio line in Fig. 4. As is clearly evident
in Fig. 4, the SLI surface area ratios lie within the RRMI ratio error bounds and
are similarly independent of pressure although some scatters are apparent. The
area ratios generally lie between the fitted RRMI ratio and the lower RRMI
error bound. Thus, there is excellent agreement between the microgravity to
normal-gravity RRMI ratios and SLI surface area ratios although the SLI sur-
face area ratios are generally slightly smaller. This suggests that the change in
SLI surface area accounts for most—but not all—of the variation in RRMI. This
is consistent with either 共a兲 a dependence of RRMI on SLI surface area to an
exponent greater than one, or 共b兲 the action and influence of one or many
secondary factors that affect RRMI in addition to a change in SLI surface area,
such as an increase in droplet temperature. However, it can be concluded that
the dominant factor that causes the change in RRMI at the onset of micrograv-
ity is the change in SLI surface area between normal gravity and microgravity.
Further, a proportional relationship between RRMI and the ratio of SLI surface
area to cross-sectional area is a good approximation.
In Fig. 4, it is highly noteworthy that the ratio of microgravity RRMI to
normal-gravity RRMI is apparently independent of pressure. The change in SLI
surface area is shown to be similarly independent of pressure. It is proposed
that the absence of a pressure dependency is related to the change in SLI shape
being driven by surface tension forces 共between the molten droplet and the
solid metal at the SLI兲. These surface tension forces are suddenly unopposed by
droplet weight at the onset of microgravity, with most other aspects of the
heterogeneous burning process remaining unchanged. By extension, if the
change in SLI shape is driven by surface tension forces then since the magni-
tude of the change is apparently independent of pressure, the magnitude of the
forces that caused the change must therefore also be independent of pressure.
Significantly, at the external boundary of the SLI 共where the outer edge of the
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
460 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

droplet attaches to the solid兲, where the surface tension forces act to change the
attachment geometry between liquid and solid phases, the temperature is and
must be, by definition, constant at the solid’s melting temperature 共which is
independent of pressure兲. Further, since the surface tension is primarily tem-
perature dependent, this value would be expected to remain constant at the
outer edge of the SLI, regardless of pressure. Therefore, even though the maxi-
mum droplet temperature may increase with increased oxygen pressure, this
variation cannot change either the temperature or the surface tension at the
SLI. Thus, it has been shown that the SLI shape and area changes in micro-
gravity, and further, Fig. 4 indicates that the magnitudes of these changes are
independent of pressure, which is consistent with a constant surface tension
“driving force” 共for shape change兲 at the SLI.

Rate-Limiting Mechanism
An important result from the transition tests was that a rapid change in RRMI
occurred at the onset of microgravity and this was accompanied by a rapid
change in the bulk shape of the molten droplet. The change in RRMI was
shown to be almost entirely due to the change in the SLI surface area at the
onset microgravity. It is useful to consider this in relation to the rate-limiting
mechanism. In an earlier work by Steinberg et al. 关7,8,10兴, it was shown that
surface tension induced convection 共which is not dependent on gravity level兲
provides sufficient mixing of fuel and oxidizer to enable burning to continue in
microgravity. This implies that transport phenomena and, by extension, the
basic physical and chemical processes of burning are generally similar in mi-
crogravity and normal gravity 共at least during the first 0.05–0.1 s following the
reduction in gravity level over which the change in RRMI was observed during
the transition tests兲. Therefore, since burning provides the energy supply that
causes melting and since the burning phenomena remain generally unchanged
in microgravity, the energy supply available within the burning droplet in nor-
mal gravity must be similar to that available in the early stage of microgravity
and sufficient to cause melting at the higher RRMI subsequently observed in
microgravity. That is, the energy supply rate available within the burning drop-
let and its ability or potential to cause melting of the solid rod was similar at
both gravity levels 共over the short period of time near the transition in gravity
level兲. This consideration provides great insight into the burning process be-
cause, despite there being almost no change in the burning phenomena after
the onset of microgravity, a rapid change in RRMI was observed, due to in-
creased contact surface area between liquid and solid phases at the SLI. In this
way, although a sufficient energy supply was available and the potential for a
higher RRMI existed in normal gravity, this potential was only realized when
the SLI surface area increased at the onset of microgravity. Clearly, heat trans-
fer across the SLI limited the RRMI in normal gravity and a rapid increase in
SLI surface area at the onset of microgravity 共due to the large change in SLI
shape兲 allowed for a proportional increase in RRMI 共through increased total
heat transfer rate兲. These results indicate that heat transfer across the SLI is the
rate-limiting mechanism for the melting, during burning, of cylindrical iron
rods in microgravity, and during a transition in gravity level from normal grav-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 461

FIG. 5—Video images of burning rods, with 共a兲 circular, 共b兲 rectangular, and 共c兲 and 共d兲
triangular cross sections. Major face visible in 共b兲; one face visible in 共c兲; two faces
visible in 共d兲. White lines added to show rod extent and corners/edges. Note differences
in SLI shape. Obtained from Ref. 26.

ity to microgravity. Since melting is a critical process to burning 共melting pro-


vides the liquid fuel source兲, heat transfer across the SLI is therefore the rate-
limiting mechanism for burning, limited specifically by the SLI surface area.

Applications—Normal Gravity
While the results clearly indicate that the SLI surface area available for heat
transfer limits the RRMI in microgravity, it can also be shown that this is the
case in normal gravity. Suvorovs et al. 关26,31兴 reported statistically significant
differences in RRMI between rods with different cross-sectional shapes but the
same cross-sectional area. Circular, rectangular, and triangular shaped mild
steel rods were tested and comparisons were made for five cross-sectional sizes.
Significantly, especially at larger cross-sectional sizes, the molten material was
consistently observed climbing the corners/edges of the samples during burn-
ing and thus clearly altering the shape of the SLI. In all cases, the magnitude of
the change in observed RRMI correlated 共in a qualitative sense兲 with the extent
of the change in SLI shape. The largest observed changes in SLI shape 共and
RRMI兲 occurred at the largest cross-sectional area tested, 28 mm2 共the cross-
sectional area of a 6-mm-diameter cylindrical rod兲. Images from tests at this
cross-sectional size are presented in Fig. 5. For the 28-mm2 cross-sectional
size, it was reported that RRMI for the rectangular samples was 12 % higher
than RRMI for the circular rods, and for the triangular samples it was 20 %
higher.
The samples tested by Suvorovs et al. 关26兴 were not quenched, making any
post-test analysis of the burned samples either impossible or invalid. However,
it is possible to infer an approximate SLI shape from the images presented in
Fig. 5 although assumptions are required because the rear side of the sample is
not visible and information is only available at the outside boundary 共and not
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
462 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—Test data 共from Ref. 26兲 and interpreted CAD model for a rectangular, mild steel
rod burning in oxygen at 6.9 MPa.

the interior兲 of the SLI. Similarity is assumed between the three faces of the
triangular shaped rods and between the two major faces of the rectangular
rods. No curvature of the SLI was observed when the rectangular rods were
viewed end on 关26兴. These assumptions enable the development of the three-
dimensional approximations of the SLI surfaces, using commercial computer
aided design 共CAD兲 software. The images in Fig. 5 were used to define control
points, which were connected using cubic splines. The CAD software was then
used to calculate the surface area of the inferred SLI surface. The analysis and
results are presented in Fig. 6 for the rectangular rods and Fig. 7 for the trian-
gular rods.
The SLI surface area estimates obtained from the CAD software for the
rectangular and triangular rods are presented in Table 2. The SLI in the circular
rods, as evident in Fig. 5共a兲, can be assumed to be planar and perpendicular to
the rod centerline. This means that for the circular rods, the SLI surface area is
equal to the rod cross-sectional area. Also included in Table 2 are the observed
RRMI values normalized to those observed for the circular rods.

FIG. 7—Test data 共from Ref. 26兲 and interpreted CAD model for a triangular mild steel
rod burning in oxygen at 6.9 MPa.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 463

TABLE 2—SLI surface area analysis.


a
ASLI ASLI − S / S vRRMI − vRRMI 兩CIRC / vRRMI 兩CIRC
Rod Shape 共mm2兲 共%兲 共%兲
Circular 28 0 0
Rectangular 31 11 12
Triangular 33 18 20
a
Obtained from Suvorovs et al. 关26兴.

It is important to note that the estimated SLI surface areas presented in


Table 2 for the rectangular and triangular rods are only first-order approxima-
tions. The process of interpreting video images and developing the CAD model
is likely to have introduced large errors and these have not been quantified.
However, this analysis may still provide a useful insight into the causes of
observed variations in RRMI between samples of different shapes. As shown in
Table 2, for the rectangular rods, an estimated 11 % change in SLI surface area
was associated with a 12 % change in RRMI, compared to the circular rods. For
the triangular rods, an estimated 18 % change in SLI surface area was associ-
ated with a 20 % change in observed RRMI. These results show good agreement
with Eq 6. When the SLI shape changed between the circular, rectangular, and
triangular rods, as shown in Fig. 5, the magnitude of the change in RRMI
matched the magnitude of the change in SLI surface area. Moreover, since test
conditions, sample cross-sectional area, and droplet volume did not change, it
is very unlikely that the change in rod cross-sectional shape caused changes in
the rate of oxygen incorporation, reaction extent, or maximum temperature
and hence the available energy supply rate or heat flux to the solid 关26兴. There-
fore, this analysis clearly demonstrates a case where a change in RRMI was
brought about not by a change in heat flux but by a change in attachment
geometry and available contact surface area at the SLI. The apparent correla-
tion provides strong evidence supporting Eq 6, even when the simple methods
used to estimate the areas are considered. This result suggests that, at this
pressure 共6.9 MPa兲, heat transfer across the SLI is the mechanism which limits
the melting rate during heterogeneous burning—in normal gravity.
This outcome and its analysis through the correlation of changes in SLI
surface area and RRMI are conceptually identical to the microgravity work. A
change in a single test parameter 共i.e., gravity level or cross-sectional shape兲
caused the shape of the SLI to be altered and a proportional effect on RRMI
was observed. The difference between the two 共seemingly unrelated兲 experi-
ments is that the extent of the change in SLI shape is much smaller in the work
presented by Suvorovs et al. because the gravity force is dominant compared to
the surface tension forces that alter the SLI shape.
It is also important to note that if the 共common兲 planar and perpendicular
SLI assumption had been made in the cases of the rectangular and triangular
rods, then the observed changes in RRMI could only have been explained by
changes in heat flux, which is not physically consistent with the system under
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
464 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

investigation. This demonstrates that the assumption of a planar SLI, perpen-


dicular to the sample centerline, must be made with great care as it is not
appropriate in all cases.
Finally, this example clearly illustrates that the work presented can be used
to understand the very important influence of geometric parameters 共like
sample shape, orientation, configuration, etc.兲 on metallic material flammabil-
ity. In addition to improving the general understanding of metallic material
flammability, it helps improve the way in which the results of standardized 共and
simplified兲 tests are applied in industrial situations. There are significant geo-
metric differences between the cylindrical rods typically used in flammability
tests and, for example, valve components. Further development of the work
presented may help improve the safety, reliability, and relevance of the use of
results from standardized testing when considering the flammability of the
more complex geometric shapes typically found in industrial situations.

Applications—Space Related
A major problem facing the designers of spacecraft oxygen systems is that
metallic materials are often more flammable 共as indicated by the threshold
pressure and RRMI兲 in microgravity than in normal gravity, meaning it is non-
conservative to apply normal-gravity flammability data to the microgravity en-
vironment 关6–10,32兴. This makes necessary the expensive, direct qualification of
metallic materials and components in a microgravity environment and often
forces spacecraft engineers to accept higher levels of risk. To minimize the need
for expensive material qualification testing in a microgravity environment,
there is clearly a need to improve the understanding of the factors that influ-
ence the way metallic materials burn under altered gravity conditions.
This work contributes to the general understanding of metallic materials
burning in microgravity, which is essential for the safe application of normal-
gravity flammability data in microgravity and the development of new test
methods and criteria. Figure 4 shows that the proportional increase in RRMI in
microgravity remains almost constant over the pressure domain considered.
This suggests that the extent of the increase in RRMI in microgravity may only
need to be assessed at a single pressure rather than at multiple pressures across
the entire pressure domain, which may reduce the need for expensive micro-
gravity experimentation. More work is required to determine 共a兲 whether this
trend holds at lower pressures closer to the threshold pressure and 共b兲 whether
a similar trend applies for other metallic materials that exhibit heterogeneous
burning. Validating the use of microgravity RRMI, obtained at only a single test
pressure to determine how to apply the extensive normal-gravity RRMI data-
base in microgravity 共for a given material兲, would be a major contribution to
flammability testing of spacecraft oxygen system materials.

Conclusion
It can be concluded that altered interfacial geometry at the SLI, which results
in a larger contact surface area for heat transfer between solid and liquid
phases, causes the higher RRMI values that are typically observed for metals
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 465

burning 共heterogeneously兲 in microgravity. The change in SLI shape is due to


the action of surface tension forces that become unopposed by gravity at the
onset of microgravity. Surface tension is temperature dependent and, since the
temperature at the outer boundary of the SLI is constant at the solid’s melting
temperature, the extent of the change in SLI shape—and, hence, the extent of
the change in RRMI—is constant over the pressure domain. There is an 共ap-
proximately兲 proportional relationship between RRMI and the ratio of SLI sur-
face area to cross-sectional area. Furthermore, it has been shown that the rate-
limiting mechanism for melting and burning 共during heterogeneous burning兲 is
heat transfer across the SLI, which is limited by the total SLI surface area. This
applies at least for iron burning at pressures of 1.4–6.9 MPa 共200–1,000 psia兲
and may apply for other metallic materials that exhibit heterogeneous burning.
This result was shown to be consistent with changes in RRMI due to geometric
changes 共e.g., sample cross-sectional shape兲 in normal gravity. In this way, the
work represents a contribution to understanding the link between geometric
parameters 共such as sample shape, size, configuration, orientation, etc.兲 and
flammability, which may improve the way in which results from standardized
flammability tests are applied in more complex real-world situations.

References

关1兴 Kirschfeld, L., “Rate of Combustion of Iron Wire in Oxygen Under High Pressure,”
Archiv fur das Eisenhuttenwesen, Vol. 32, 1961, pp. 57–62.
关2兴 Samant, A. V., Zawierucha, R., and Million, J. F., “Thickness Effects on the Pro-
moted Ignition-Combustion Behaviors of Engineering Alloys,” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 1454, Vol. 10, 2003,
ASTM International, West Conshohocken, PA, pp. 171–191.
关3兴 Sato, J., “Fire Spread Rates along Cylindrical Metal Rods in High Pressure Oxy-
gen,” Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres,
STP 1040, Vol. 4, 1989, ASTM International, West Conshohocken, PA, pp. 162–177.
关4兴 Endelman, L. L. et al., “Metal Combustion in High Pressure Oxygen Atmosphere:
Detailed Observation of Burning Region Behavior by using High Speed Photogra-
phy,” Proceedings of the 15th International Congress on High Speed Photography
and Photonics, San Deigo, CA, August 21–27, 1982, SPIE—International Society
for Optical Engineering, Bellingham, WA.
关5兴 Stoltzfus, J. M., Homa, J. M., Williams, R. E., and Benz, F. J., “ASTM Committee
G-4 Metals Flammability Test Program: Data and Discussion.” Flammability and
Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 986, Vol. 3, ASTM,
Philadelphia, PA, 1988, pp. 28–53.
关6兴 Steinberg, T. A. “The Combustion of Metals in Gaseous Oxygen,” Ph.D. thesis, New
Mexico State Univ., Las Cruces, NM, 1990.
关7兴 Steinberg, T. A. and Benz, F. J., “Iron Combustion in Microgravity,” Flammability
and Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP 1111, Vol. 5,
ASTM International, West Conshohocken, PA, 1991, pp. 298–312.
关8兴 Steinberg, T. A., Wilson, D. B., and Benz, F. J., “The Burning of Metals and Alloys
in Microgravity,” Combust. Flame, Vol. 88, 1992, pp. 309–320.
关9兴 Steinberg, T. A., Wilson, D. B., and Benz, F. J., “Microgravity and Normal Gravity
Combustion of Metal and Alloys in High Pressure Oxygen,” Flammability and Sen-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
466 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

sitivity of Materials in Oxygen-Enriched Atmospheres, STP 1197, Vol. 6, ASTM In-


ternational, West Conshohocken, PA, 1993, pp. 133–145.
关10兴 Steinberg, T. A. and Stoltzfus, J. M., “Combustion Testing of Metallic Materials
aboard NASA Johnson Space Center’s KC-135,” Flammability and Sensitivity of
Materials in Oxygen-Enriched Atmospheres, STP 1319, Vol. 8, ASTM International,
West Conshohocken, PA, 1997, pp. 170–188.
关11兴 Benning, M. A., Zabrenski, J. S., and Le, N. B., “The Flammability of Aluminum
Alloys and Aluminum Bronzes as Measured by the Pressurized Oxygen Index,”
Flammability and Sensitivity of Materials in Oxygen-Enriched Atmospheres, STP
986, Vol. 3, ASTM International, West Conshohocken, PA, 1988, pp. 54–71.
关12兴 Sato, K., Hirano, T., and Sato, J. “Behavior of Fires Spreading over Structural
Metal Pieces in High Pressure Oxygen.” American Society of Mechanical Engineers
and Japanese Society of Mechanical Engineers Thermal Engineering Joint Conf.
Proc., Vol. 4, ASME, New York, pp. 1983, 311–316.
关13兴 Sato, J., Sato, K., and Hirano, T., “Fire Spread Mechanisms Along Steel Cylinders
in High Pressure Oxygen,” Combust. Flame, Vol. 51, 1983, pp. 279–287.
关14兴 Benz, F. J., Steinberg, T. A., and Janoff, D., “Combustion of 316 Stainless Steel in
High-Pressure Oxygen.” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, STP 1040, Vol. 4, ASTM International, West Conshohocken,
PA, 1989, pp. 195–211.
关15兴 Wilson, D. B., Steinberg, T. A., and Stoltzfus, J. M., “Thermodynamics and Kinetics
of Burning Iron,” Flammability and Sensitivity of Materials in Oxygen-Enriched
Atmospheres, STP 1319, Vol. 8, ASTM International, West Conshohocken, PA, 1997,
pp. 240–257.
关16兴 Hirano, T., Sato, Y., Sato, K., and Sato, J., “The Rate Determining Process of Iron
Oxidation at Combustion in High Pressure Oxygen,” Oxidation Communications,
Vol. 6, 1984, pp. 113–124.
关17兴 Steinberg, T. A., Mulholland, G. P., Wilson, D. B., and Benz, F. J., “The Combustion
of Iron in High-Pressure Oxygen,” Combust. Flame, Vol. 89, 1992, pp. 221–228.
关18兴 Steinberg, T. A., Sircar, S., Wilson, D. B., and Stoltzfus, J. M., “Multiphase Oxida-
tion of Metals,” Metall. Mater. Trans. B, Vol. 28B, 1997, pp. 1–6.
关19兴 Steinberg, T. A., Kurtz, J., and Wilson, D. B., “The Solubility of Oxygen in Liquid
Iron Oxide During the Combustion of Iron Rods in High-Pressure Oxygen,” Com-
bust. Flame, Vol. 113, 1998, pp. 27–37.
关20兴 Wilson, D. B., Steinberg, T. A., and De Wit, J., “The Presence of Excess Oxygen in
Burning Metallic Materials,” Flammability and Sensitivity of Materials in Oxygen-
Enriched Atmospheres, STP 1395, Vol. 9, ASTM International, West Conshohocken,
PA, 2000, pp. 145–162.
关21兴 Dreizin, E. L., “Phase Changes in Metal Combustion,” Prog. Energy Combust. Sci.,
Vol. 26, No. 1, 2000, pp. 57–78.
关22兴 Ward, N. R., Suvorovs, T., and Steinberg, T. A., “An Investigation of Regression
Rate of the Melting Interface for Iron Burning in Normal Gravity and Reduced
Gravity,” J. ASTM Int., Vol. 3, 2006, Paper ID JAI13569.
关23兴 Ward, N. R. “The Rate-Limiting Mechanism for Heterogeneous Burning in Normal
Gravity and Reduced Gravity,” Ph.D. thesis, Queensland Univ. of Technology, Bris-
bane, Australia, 2008.
关24兴 Ward, N. R. and Steinberg, T. A. “Thermal Analysis of Iron Rods Burning in Nor-
mal Gravity and Reduced Gravity,” Annals of the Assembly for International Heat
Transfer Conference 13, Begell House, Inc, Redding, CT, 2006.
关25兴 Steinberg, T. A. and Wilson, D. B., “Modeling the NASA/ASTM Flammability Test
for Metallic Materials Burning in Reduced Gravity.” Flammability and Sensitivity
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
WARD AND STEINBERG, doi:10.1520/JAI102269 467

of Materials in Oxygen-Enriched Atmospheres, STP 1395, Vol. 9, ASTM Interna-


tional, West Conshohocken, PA, 2000, pp. 266–291.
关26兴 Suvorovs, T., Ward, N. R., Steinberg, T. A., and Wilson, R., “Effect of Geometry on
the Melting Rates of Iron Rods Burning in High Pressure Oxygen,” J. ASTM Int.,
Vol. 4, 2007, Paper ID JAI101004.
关27兴 Ward, N.R. and Steinberg, T. A., “Geometry of a Fast Moving Melting Interface in
Cylindrical Metal Rods Under Microgravity Conditions.” French Academy of Sci-
ence, Special Publication: Comptes Rendus Mecanique, C. R. Mec, M. El Ganaoui,
and R. Prud’homme, eds., Vol. 5–6, 2007, pp. 342–350.
关28兴 Ward, N. R. and Steinberg, T. A., “Iron Burning in Pressurized Oxygen Under
Microgravity Conditions,” Microgravity Sci. Technol., Vol. 21, No. 1–2, 2008, pp.
41–46.
关29兴 Edwards, A. P. R. “Modelling of the Burning of Iron Rods in Normal Gravity and
Reduced Gravity,” Ph.D. thesis, The Univ. of Queensland, St. Lucia, Australia,
2004.
关30兴 Hart, M., “Analysis of the Burning Temperatures of 1/8-in. Diameter Iron Rods in
Enriched Oxygen Atmospheres,” Internal Special Test Data Report No. WSTF 99-
34536, NASA JSC White Sands Test Facility, Las Cruces, New Mexico, 1999.
关31兴 Suvorovs, T., Ward, N. R., Wilson, R., and Steinberg, T. A., “Effect of Sample
Geometry on Regression Rate of the Melting Interface for Carbon Steel Burned in
Oxygen,” J. ASTM Int., Vol. 3, 2006, Paper ID JAI13565.
关32兴 Chiffoleau, G. “Ultrasonic Investigation of Burning Metals in Normal and Reduced
Gravity,” Ph.D. thesis, The Univ. of Queensland, St. Lucia, Australia, 2002.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Reprinted from JAI, Vol. 6, No. 9
doi:10.1520/JAI102215
Available online at www.astm.org/JAI

James VanOmmeren1

Chlorine Trifluoride Exposure Testing and


2
Oxidizer Reactivity Results

ABSTRACT: Chlorine trifluoride 共ClF3兲 was first synthesized in the 1930s and
is one of the most oxidizing and reactive halogen fluorides known. In fact,
with the exception of elemental fluorine, ClF3 may represent one of the most
oxidizing and reactive materials in existence. Active research and commer-
cial use of ClF3 began in the late 1940s, and ClF3 has been utilized in such
diverse applications as military weapons usage, rocket fuel oxidant, nuclear
fuels processing, oil well rod cutting, mineral analysis, and electronics manu-
facturing equipment cleaning. Such fluoride compounds present significant
toxic, oxidizing, highly reactive, and environmental challenges in their manu-
facture, delivery, customer use, handling, disposal treatment, and emer-
gency response. Air Products has been the primary manufacturer of ClF3 in
North America for about 40 years, and ClF3 represents one of the strongest
oxidizing and highest reactivity materials that Air Products handles world-
wide. However, little information has been available in public literature or the
industry regarding its oxidizing potential and reactivity strength with specific
materials it may routinely or accidentally come in contact with. Therefore, Air
Products conducted extensive release testing to expose various materials of
construction 共metals, plastics, and elastomers兲, building materials, personal
protective equipment, and other common materials to both vapor and liquid
phase releases of ClF3. This paper discusses the synthesis, hazards, varied
uses, and major incidents involving ClF3. For example, one major ClF3 liquid
spill accident from the 1960s is discussed where an eyewitness stated, “The
concrete was on fire!” The paper details controlled release exposure testing
methods Air Products employed, results, and lessons learned regarding im-
proved handling and usage recommendations for ClF3 suppliers and custom-
ers. For example, a significant potential hazard was discovered regarding

Manuscript received October 30, 2008; accepted for publication July 31, 2009; published
online September 2009.
1
Process Safety Engineering Associate, Air Products and Chemicals, Inc., Allentown, PA
18195.
2
Chlorine Trifluoride Release Testing and Reactivity Results, Managing Chemical Reactiv-
ity Hazards and High Energy Release Events, American Institute of Chemical Engineers.
Cite as: VanOmmeren, J., ⬙Chlorine Trifluoride Exposure Testing and Oxidizer Reactivity
Results,⬙ J. ASTM Intl., Vol. 6, No. 9. doi:10.1520/JAI102215.
Copyright © 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Conshohocken, PA 19428-2959.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by 468
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 469

use of a safety shower if a person’s clothing was exposed to ClF3 liquid,


which caused Air Products to change their Material Safety Data Sheet and
Product Safety hazards training.
KEYWORDS: chlorine trifluoride, release, testing, applications,
properties, hazards, handling, usage, MSDS, PPE

Introduction

“Chlorine Tri-What?”
Chlorine trifluoride 共ClF3兲 is one of those molecules that your high school
chemistry teacher would claim cannot exist. Whenever I mention the name of
this compound to someone for the first time, their response usually is, “Chlo-
rine tri-what?” The valences for the constituent atoms to bond together and
produce this molecule do not seem to make sense. How can you oxidize el-
emental chlorine, which is already a strong oxidizer? It is possible because
fluorine is a more powerful oxidizer than chlorine and it can oxidize almost any
element or material even including some of the noble gas family. One result of
this unlikely union of chlorine and fluorine is chlorine trifluoride. ClF3 is a
toxic, corrosive, and very reactive liquefied compressed gas packaged in cylin-
ders as a liquid under its own vapor pressure of 1.55 kg/ cm2 at 21° C. ClF3 is
a very useful chemical in operations requiring a high-energy fluorinating agent
or incendiary material, especially since it can be handled at room temperature.
However, those same factors that make it quite useful also contribute to several
high hazard potentials for the product.
Air Products has been the primary manufacturer and distributor of ClF3 in
North America for the past 40 years, and the compound represents one of the
highest reactivity materials that Air Products currently manufactures or
handles worldwide.

Initial Discovery
Fluorine is recognized as the most powerful oxidizing agent of all known ele-
ments. However, fluorine 共F2兲 has the inherent disadvantage that the liquid
must be stored and handled in a cryogenic state, with a normal boiling point of
−188° C. Therefore, it is impractical or exceedingly difficult to use liquid-phase
F2 to provide maximum density exposure, thus enhancing its fluorinating reac-
tivity performance. Gaseous usage of F2 utilizes a much more dilute density
共vapor-phase兲 contacting the reactants and thus yields lower reactivity perfor-
mance than liquid F2 contact. Therefore, various fluorinating substitutes were
evaluated in the late 1920s that possessed the highly reactive properties of F2
but posed fewer handling difficulties. ClF3 was first successfully isolated by
Ruff and Krug in 1930 after experimental tests with chlorine monofluoride
suggested the presence of a higher fluoride species 关1兴. Liquid ClF3 is consid-
ered more reactive than vapor-phase F2 reactions since more moles of fluori-
nating agent are present per unit area of reactant surfaces. Also, liquid ClF3
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
470 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

may demonstrate even higher reactivity in certain circumstances than liquid


fluorine since the F2 liquid temperature is cryogenic, thus reducing its activity
potential.

First Bulk Application—Military Weapons


German interest in ClF3 during World War II prompted the first industrial bulk
production capability for the material. The Germans produced ClF3 in tonnage
quantities for military use during World War II for use in flame-throwers due to
the liquid’s extreme “hypergolic” nature with fuels and as a general incendiary
material. A hypergolic material is self-igniting on contact with a fuel; no igni-
tion source is required. The concept is essentially the “inverse” of a pyrophoric
material, which is self-igniting on contact with an oxidizer requiring no igni-
tion source. ClF3 was originally produced in large quantities using a two-step
synthesis method developed by the Germans. This process caused fluorine and
chlorine to react under heat to form chlorine monofluoride. The second step in
the reaction brought chlorine monofluoride together with excess fluorine to
produce ClF3 on a scale of approximately 135 kg/h with 96–98 % yield results.
Following the war, interest in the use of ClF3 for organic synthesis work in-
creased, although the material was eventually considered to be too reactive for
practical use and mostly abandoned for these applications 关2兴. Synthesis reac-
tions proved difficult to control and usually led to a wide variety of reaction
by-products that were hazardous.

Early History—Rocket Science and the Atomic Age


Many fluorinating compounds were evaluated as potent oxidizers for liquid-
fueled rockets in the late 1940s and early 1950s to overcome the storage and
handling disadvantages of liquid F2. Some of the combinations of compounds
considered were chlorine-fluorine, nitrogen-fluorine, and fluorine-oxygen. Dur-
ing this time, various U.S. agencies and companies did considerable testing
using ClF3 as the oxidizer for liquid-fueled rocket research. ClF3 was first tested
in the United States in 1948 on a liquid propellant 45 kg thrust rocket motor
using hydrazine as the fuel. Later the performance of hydrazine-nitrogen
tetroxide and hydrazine-chlorine trifluoride was compared in a larger 135 kg
thrust rocket motor. In 1951, ClF3 was tested with both ammonia and hydra-
zine as liquid rocket fuels. Further related work with ClF3 included its use as an
ignition source in various non-hypergolic rocket fueled systems 关2兴. It was de-
termined to be so rapidly hypergolic when exposed to various fuels that no
ignition delay was ever successfully measured during experiments. Rocket per-
formance results with the compound were very positive, yielding smooth en-
gine operation, immediate reaction 共which allowed a small reaction chamber
size兲, and near theoretical energy output. ClF3 was concluded to approach
solid-propellant rocket convenience since it could be stored stably in liquid-
phase at room temperature and relatively low pressure. Also, it was deemed as
one of the most promising storable oxidizers available at the time since it con-
tains 62 % by weight elemental fluorine. However, all rocket materials of con-
struction 共including metals and seals兲 that could contact ClF3 had to be scru-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 471

pulously selected, cleaned, and passivated to prevent the components from


burning during reaction 关3兴. ClF3 was also recognized as an extremely hazard-
ous propellant due to its reactivity, toxicity, and toxic by-products of fluorina-
tion.
Beyond ClF3’s use as an oxidizer for liquid-fueled rockets, other interest in
chlorine trifluoride increased in the late 1940s due to various potential appli-
cations in the developing nuclear industry. For example, ClF3 was investigated
for use in uranium enriching applications during the early atomic Manhattan
Project and later Oak Ridge nuclear effort.

Famous Industrial Incident—“The Concrete Was on Fire!”


There is one major incident involving ClF3 that occurred during the liquid
rocket propellant era that is relatively famous in the industry. A domestic
chemical company was a major supplier of ClF3 for U.S. governmental appli-
cations. They had an incident at one of their chemical facilities when personnel
for the first time loaded a 1 ton steel container with liquid ClF3 for bulk ship-
ment. The container had been cooled with dry ice to perform the liquid transfer
and help make the product safer to handle since ClF3 vapor pressure is only
0.007 kg/ cm2 at dry ice temperature. However, the dry ice bath embrittled the
steel container wall, and while maneuvering the full container onto a dolly, it
split open and instantaneously released 907 kg of cold ClF3 liquid onto the
building floor. The material “consumed through” 共reacted兲 the 30 cm thick con-
crete floor and through about another 90 cm of gravel underneath the spill. The
fumes that were generated 共chlorine trifluoride, hydrogen fluoride, chlorine,
hydrogen chloride, etc.兲 severely corroded everything that was exposed 关3兴. One
eyewitness described the incident by stating, “The concrete was on fire!”

Recent Past Uses—The Oil Crisis


During the decades of the 1960s and 1970s, ClF3 was utilized in several diverse
industries, including use for performing mineral analysis by actually dissolving
rock or mineral samples, thus facilitating analytical testing.
One interesting application during the increased oil exploration activity of
the 1970s and 1980s involved use of ClF3 as a chemical cutting agent primarily
for oil drilling rigs. The device that was used to perform the cutting had three
chambers stacked in a vertical column configuration. The top chamber con-
tained a small explosive charge, the middle chamber contained liquid ClF3, and
the bottom chamber housed a catalyst. The device would be lowered into an oil
well to any desired depth and the explosive charge was detonated. The explo-
sion pushed the liquid ClF3 through the catalyst chamber where its tempera-
ture was rapidly increased and the explosive energy then forced the heated ClF3
out of small nozzles located at the bottom of the device, which were aimed
radially outward. The high pressure heated stream of ClF3 contacted the well
tubing wall and/or drill rod where an extremely rapid and vigorous reaction
cleanly cut the metal in less than 1 s. The tube or rod could then be withdrawn
from the well and recovered 关4兴.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
472 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Modern Usage—Integrated Circuit Chip, Thin-Film Transistor 共TFT兲 Liquid


Crystal Display 共LCD兲, and Photovoltaic Cell Manufacturing
The only surviving major industrial application for ClF3 in the present is its use
in the nuclear industry for the reclamation of uranium from irradiated fuels.
This is accomplished by reacting the nuclear fuel with ClF3 and forming the
corresponding fluoride compounds. The gaseous uranium hexafluoride is then
recovered from the other compounds present via distillation 关4兴.
However, in the 1990s the semiconductor industry developed a high-purity
application that uses ClF3 in the cleaning process for certain chemical vapor
deposition 共CVD兲 tool chambers. CVD involves decomposing source gases 共such
as silane, dichlorosilane, silicon tetrafluoride, and tungsten hexafluoride兲 using
heat or a plasma as the energy source. These film forming reactions in CVD
tools are driven to the wafer surface and are precisely deposited to impart
specific electrical characteristics to the integrated circuit 共IC兲 device. However,
during CVD, material is deposited not only on the silicon substrate but also on
the walls of the process chamber. Periodically, the tool chamber requires clean-
ing to preempt particle shedding and maintain an ultra-high purity environ-
ment for the wafers being processed. In situ cleaning of the tool is desirable
because the solid residues on the chamber interior are removed from the walls
without dismantling the tool or risking personnel exposure to the hazardous
residues or cleaning agents. In situ cleaning also yields quicker turn-around
time for the tool to resume wafer processing. Fluorine atoms readily react with
these solid residues to form volatile reaction products that can be purged and
evacuated from the system 关5兴. The primary advantage ClF3 has over other
gaseous fluorine cleaning sources 共such as nitrogen trifluoride, hexafluoroet-
hane, or carbon tetrafluoride兲 is that the high reactivity of ClF3 allows the
operation to be accomplished at relatively low temperatures without requiring
plasma or high temperature heating to disassociate it for use during the clean-
ing process. The use of vapor-phase ClF3 for CVD chamber cleaning has dem-
onstrated the ability to prolong chamber component life through the lack of
high temperature or plasma use and tool dismantling requirements.
Two analogous usages of ClF3 involve TFT LCD manufacture and photovol-
taic 共PV兲 solar cell manufacture. Each application utilizes similar fabrication
tools, process gases, and process chemicals as does semiconductor chip manu-
facture. Therefore, ClF3 has potential in situ tool cleaning applications in the
TFT LCD and PV manufacturing industries.
Because the use of ClF3 in semiconductor applications has started to ex-
pand due to its effective tool cleaning performance, Air Products personnel and
customers involved in serving these applications are required to handle increas-
ing amounts of ClF3. The various safety reviews that were used to design and
operate Air Products’ ClF3 manufacturing facility and develop the safety re-
quirements for our operators do not translate well to a consumer’s site. Also,
most of the publicly available information regarding ClF3 was dated and of
questionable value since the tested product had impurities that are no longer
present in Air Products’ electronics grade material 共99.9 % minimum purity兲.
Finally, because past applications using ClF3 were mostly government-
controlled, many of the handling procedures that were utilized for governmen-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 473

tal services were classified. Therefore, Air Products decided to empirically test
the reactivity of ClF3 with currently used materials of construction, personal
protective equipment 共PPE兲, contaminants commonly found in systems and
equipment, and other materials that may come into contact with the product.
The following paper discusses the test results of the exposure of various mate-
rials to both vapor-phase and liquid-phase contacts with ClF3. The results of
this analysis were then used to re-evaluate the materials of construction uti-
lized in ClF3 service, improve Air Products’ knowledge of PPE performance,
and establish improved recommendations to those manufacturing, handling,
and performing emergency response with the product.

General Properties of Chlorine Trifluoride

Physical and Chemical Properties

Molecular weight 92.447


Boiling point 共1 atm兲 11.75° C
Melting point −76.32° C
Gas density 共21.1° C兲 3.913 kg/ m3
Specific volume 共21.1° C兲 0.2556 m3 / kg
Specific gravity 共air= 1兲 3.260
Vapor pressure 共21.1° C兲 1.55 kg/ cm2
Appearance Gas Colorless
Liquid Pale green
Solid White
Odor Varies with hydrolysis by-products; low
concentrations are described as bleach-like,
while higher concentrations are described as
acidic or suffocating

Toxicological Properties

299 ppm for 1 h rat 共death due to respiratory


LC共50兲 failure兲
OSHA PEL 0.1 ppm ceiling
ACGIH TLV 0.1 ppm ceiling
NIOSH IDLH 20 ppm
AIHA ERPGs ERPG-1 = 0.1 ppm
ERPG-2 = 1 ppm
ERPG-3 = 10 ppm
EPA AEGLs 共60 min兲 AEGL-1 = 0.12 ppm
AEGL-2 = 2.0 ppm
AEGL-3 = 21 ppm

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
474 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

TABLE 1—Gibb’s free energies.

⌬G0
共KJ/mol兲

Compounds State 298 K 500 K 1000 K 1500 K Source


F2 g 0.00 0.00 0.00 0.00 JANAF-85
HF g ⫺274.65 ⫺275.98 ⫺278.44 ⫺280.03 JANAF-85
NF3 g ⫺90.58 ⫺62.22 7.42 75.50 JANAF-85
BF3 g ⫺1119.03 ⫺1107.36 ⫺1076.36 ⫺1044.03 JANAF-85
CF4 g ⫺888.51 ⫺857.84 ⫺781.34 ⫺705.48 JANAF-85
CHF3 g ⫺662.59 ⫺638.32 ⫺575.10 ⫺511.24 JANAF-85
C 2F 6 g ⫺1258.13 ⫺1200.07 ⫺1058.61 ⫺920.17 JANAF-85
SF6 g ⫺1116.45 ⫺1044.53 ⫺855.77 ⫺647.29 JANAF-85
SF4 g ⫺722.02 ⫺693.07 ⫺610.32 ⫺505.85 JANAF-85
PF5 g ⫺1520.70 ⫺1469.67 ⫺1342.48 ⫺1199.78 JANAF-85
SiF4 g ⫺1572.71 ⫺1543.75 ⫺1471.48 ⫺1399.58 JANAF-85
WF6 g ⫺1632.30 ⫺1572.09 ⫺1426.69 ⫺1284.96 JANAF-85
ClF3 g ⫺118.88 ⫺91.93 ⫺26.75 36.78 JANAF-85
BrF3 g ⫺229.38 ⫺204.00 ⫺139.11 ⫺75.95 JANAF-85
AgF s ⫺186.60 ⫺175.15 ⫺150.75 ⫺128.76 HSC
AgF2 s ⫺302.18 ⫺272.70 ⫺204.52 ⫺148.21 HSC
AuF3 s ⫺292.83 ⫺246.39 ⫺136.18 ⫺29.41 HSC
AlF3 s ⫺1431.12 ⫺1377.91 ⫺1250.07 ⫺1121.27 JANAF-85
CuF s ⫺259.55 ⫺246.19 ⫺217.75 ⫺192.04 JANAF-85
CuF2 s ⫺491.64 ⫺460.35 ⫺388.60 ⫺321.66 JANAF-85
NiF2 s ⫺604.10 ⫺583.88 ⫺512.70 ⫺441.20 HSC
FeF2 s ⫺663.17 ⫺634.94 ⫺568.05 ⫺501.46 JANAF-85
FeF3 s ⫺972.29 ⫺926.55 ⫺817.57 ⫺709.32 JANAF-85
ZnF2 s ⫺713.30 ⫺805.70 ⫺876.06 ⫺979.54 HSC
CrF2 s ⫺804.96 ⫺709.05 ⫺644.53 ⫺583.76 HSC
CrF3 s ⫺1088.00 ⫺1225.74 ⫺1313.74 ⫺1428.36 HSC
CrF4 s ⫺1285.25 ⫺1318.06 ⫺1433.98 ⫺1587.68 HSC

Oxidizing Potential
Evaluation of various oxidizing gas’ Gibbs free energies of formation with dif-
ferent metals yields a relative indication of the reactivity or oxidizing potential
of the gases. ClF3 is similar to that of pure fluorine and exceeds the predicted
reactivity of F2 at elevated temperatures 共greater than about 700° C兲 关6兴. See
Table 1 for Gibbs free energies of various fluoride gases and metal fluoride
compounds.

Safe Handling and Use


Chlorine trifluoride is a strongly oxidizing, toxic, and corrosive liquefied gas at
typical storage temperatures. It is normally packaged in specially cleaned and
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 475

prepared carbon steel or stainless steel cylinders. ClF3 is most safely used as a
vapor to limit its potential reactivity with system components and other mate-
rials. Unless special precautions are taken, ClF3 should only be removed from
the cylinder as a vapor and care must be taken to prevent its condensation in
piping or other equipment. Employees working with ClF3 should be specially
trained to assure they understand the system requirements and the physical
and exposure hazards of ClF3 and its reaction products.

Main Hazards of Chlorine Trifluoride


The main hazards of ClF3 are the following.
• It is an extremely vigorous fluorinating agent and unstable when ex-
posed to easily fluorinated materials.
• It reacts with water and many oxide compounds to produce chlorine
oxides, which are also unstable and toxic.
• It is a toxic product and its by-products of fluorination or reaction are
often also toxic.
• Exposure to moisture generates acids, which can be very corrosive.

System Preparation
Systems used for chlorine trifluoride must be very carefully cleaned to remove
readily oxidized impurities and must be meticulously maintained to prevent
contamination. Cleaning agents must be thoroughly removed prior to admit-
ting ClF3 into systems as they can also become fuels in the presence of ClF3.
This is normally conducted by extensive purging with high purity nitrogen or
other inert gases. Heating of system components during purging should be
considered to ensure removal of low-volatility cleaning agents and moisture
adhered onto the walls of metallic or elastomer components.
To minimize potential problems, ClF3 systems should avoid excessive use
of mechanical connections 共to limit potential leakage sites兲 and elastomers 共to
limit potential reactivity and contamination兲. Similarly, system valves incorpo-
rating metal seats should be used if possible to reduce the chance of ignition of
elastomeric seats. Only fluorinated elastomers should be used in chlorine trif-
luoride systems and only if a metal cannot be substituted. Cleanliness of these
materials is critical, as any contamination can quickly compromise the mate-
rial. It is recommended that elastomers not be used in service where they will
be exposed to the liquid phase since even these elastomers may deteriorate
rapidly and possibly ignite at elevated temperature 共130° C兲. Elastomers are
more difficult to clean for oxidizer service because they can absorb the solvents
or detergents used to clean them; therefore, all elastomers should be degassed
after cleaning in a vacuum oven. Most non-fluorinated elastomers show little or
no resistance to ClF3. Others like Neoprene™, rubber, polyethylene, and PVC
have shown resistance during short exposures but are very susceptible to any
contamination—the reaction of ClF3 with the contaminant rapidly propagates
to the elastomer.
Chlorine trifluoride systems should be passivated before use with increas-
ing vapor concentrations of either ClF , fluorine, or a fluorine/inert gas mix-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014 3
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
476 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

ture. Passivation allows the formation of a thin metal fluoride surface that is
resistant to further reaction with ClF3 and, most importantly, allows a con-
trolled reaction with any remaining contaminants left behind by imperfect
cleaning. The system should be observed during passivation for unexpected
overheating of any components that would indicate an excessive reaction with
the components.

System Operations
PPE selection is important to identify before handling ClF3 systems to prevent
exposure to vapors or liquid. Because PPE may be compromised by the pres-
ence of dirt or water, PPE used in normal operations must be clean and free of
contamination. In emergency situations where the product is leaking and re-
sponse time is critical, the PPE must be as clean as possible to prevent reaction
with the product. Therefore, only new gloves and acid-splash suits or totally
encapsulating suits should be used in emergency situations to minimize the
chance of reaction of the product contaminants on PPE. Because the possibility
of a reaction exists and the external PPE may melt if exposed to heat, natural
fiber clothing should be worn under the PPE to minimize any melted material
binding to the skin. For normal operations polycarbonate faceshield over safety
glasses, PVC splash suit, inner gloves smooth leather, outer gloves 17 mil nitrile
is recommended. All PPE must be clean and dry. For emergency response self-
contained breathing apparatus, totally encapsulating chemical protective suit,
natural fiber clothing only to be worn under the PPE, smooth leather inner
gloves, and 17 mil nitrile outer gloves are recommended. Gloves and splash suit
must be new. Earplugs should also be considered due to the possibility of a
loud vigorous reaction.
Important reminders about PPE selection are the following.
• PPE that comes into contact with liquid may ignite.
• Always assume that at least the outer gloves have been exposed to HF;
exposure can cause burns.
• Knowing the limitations of PPE in emergency operations, responders
should think about how they may come in contact with the material and
minimize their exposure.
Chlorine trifluoride systems must be kept dry to minimize corrosion and
contamination from acids that will form on contact of ClF3 with moisture.
Equipment should be de-pressurized when not in use and purged or kept sealed
under positive pressure with dry inert gas for extended out-of-service periods.
As with any hazardous gas cylinder, operators should confirm the cylinder
valve is firmly closed before loosening the valve outlet seal to connect a cylinder
to the system. The outlet seal should be loosened slowly to limit the release rate
of any ClF3 that may have leaked into the valve outlet space. The same proce-
dure should be used to disconnect the cylinder from the system, and the pigtail
must be vented and purged thoroughly before loosening the connection.
When connecting new ClF3 cylinders, care is needed to prevent contami-
nation of the valve outlet connection, especially if a gasketed connection is used
共e.g., Compressed Gas Association 670 or Compressed Gas Association 728兲.
New gaskets must be thoroughly degreased and dried prior to installation un-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 477

less specially cleaned and packaged gaskets are used directly from the manu-
facturer. New cleanroom gloves or equally clean alternatives should be used to
install replacement gaskets. Note that oils from skin contact on gaskets or other
wetted components may cause reaction with ClF3. Similar precautions must be
taken when changing other system components to avoid introduction of easily
ignited contaminants.
When ClF3 supply cylinders are initially opened, the operator should al-
ways be prepared to quickly re-close the valve in case any evidence of reaction,
overheating, or leaks develop. If any uncertainty exists about the cleanliness of
system components, the cylinder valve should only be opened to admit a mini-
mal amount of ClF3 vapor and then immediately closed. This will limit the
amount of ClF3 available to sustain ignition if a problem develops.
It is very important that chlorine trifluoride systems be kept well below the
ClF3 condensation pressure of the coolest component in the process to prevent
liquefying ClF3 vapors. This is primarily to avoid the heightened reactivity of
liquid-phase ClF3 in the system and to allow proper flow control. Use of an
absolute pressure regulator is recommended whenever possible to control sys-
tem pressure below the ClF3 condensation pressure. With its relatively high
boiling point, ClF3’s vapor pressure is low at typical use temperatures, which
can result in unacceptably low system delivery pressure and flow rate, espe-
cially from cylinders with little remaining inventory. If the supply cylinder is
heated to permit higher flows, it is even more critical to protect against con-
densation in cooler downstream components. System heating can also be con-
sidered to avoid condensation, but it can be difficult to maintain uniform heat-
ing throughout the entire system.
Prior to any maintenance, including supply cylinder changes, special care
must be taken to assure no hazardous quantity of ClF3 remains in the equip-
ment. Thorough purging with inert gas should always be done after the ClF3
cylinder valve is firmly closed. If there is any suspicion of condensed ClF3 in the
system, the process should be heated and ideally evacuated to confirm no ClF3
remains before the system is opened.

National and Regional Code Requirements


ClF3 is a challenging material to properly manage from a code compliance and
usage standpoint. This is because ClF3 possesses both physical hazards 共very
strong oxidizer and very reactive兲 and health hazards 共toxic and corrosive兲. The
physical hazards of ClF3 often cause local authorities to require the source
containers be positioned remote from the consumer’s main facility or occupan-
cies. However, the low vapor pressure of the product at room temperature
provides challenges to distribute the vapor any significant distance.
ClF3 is classified as a “hazardous production material” 共HPM兲 per regional
Code definition based on its NFPA 704 ratings 共Fig. 1兲.
An HPM is a solid, liquid, or gas associated with semiconductor manufac-
turing that has a degree of hazard rating in health, flammability, or reactivity of
Class 3 or 4 as ranked by NFPA criteria. Also, an HPM requires that the mate-
rial is used directly in research, laboratory, or production processes that have as
their end product materials, which are not hazardous 共e.g., integrated circuits兲.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
478 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 1—NFPA hazard diamond for chlorine trifluoride.


The physical hazards or ClF3 often cause low threshold or exempt quanti-
ties for the material based on local code requirements. However, proper facili-
tation and usage of ClF3 does provide a safe supply and distribution system if
the hazards and code requirements are adequately addressed. One method to
effectively manage ClF3 hazards is through the use of properly designed and
facilitated gas cabinets that are specially engineered to house ClF3 cylinders
and distribution control piping and components. One example of the unique
design hazards for ClF3 equipment is that Air Products, in the early semicon-
ductor experimentation phase with the product, did not recommend installing
a water sprinkler inside of the gas cabinet for fire protection. Chlorine trifluo-
ride is so water reactive that alternate means were deemed preferred for pre-
vention and mitigation of fires, both internal and external to a ClF3 gas cabinet.
This position was supported in 1997 by a major industrial insurance carrier’s
specific recommendation to prohibit sprinklers in ClF3 gas cabinets used for
the semiconductor industry 关7兴.

Gas Cabinet Supply Safety Considerations


When chlorine trifluoride is used in semiconductor or research facilities, con-
sideration should be given to the following gas cabinet system design elements:
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 479

• Special materials of construction should be used for metal components


and non-metallic soft-goods to protect against oxidizing potential and
corrosivity.
• Automatic sprinkler protection for ClF3 gas cabinets should be evalu-
ated due to the potential for violent reaction with leaking ClF3.
• Pneumatically operated cylinder valves should be considered to allow
automatic and immediate supply shutdown should there be a down-
stream loss of containment.
• A gas detector should be located in the cabinet to monitor and cause an
automatic shutdown alarm on either hydrogen fluoride 共HF兲 or chlorine
dioxide 共ClO2兲 indication.
• A heat or smoke detector should be located in the gas cabinet to monitor
and cause an automatic shutdown alarm on an internal or external fire
condition.
• The delivery pipe or tubing should be co-axial 共double contained兲 with a
monitoring alarm for loss of primary containment.

Test Conditions for Exposure and Reactivity Study

Vapor-Phase Exposure Testing


The testing was conducted at Air Products’ manufacturing facility in a large
walk-in vent booth that was continuously exhausted to a caustic wet scrubber.
The ambient temperature averaged 26.7° C during the testing with relative hu-
midity approximately 40 % for the duration. A 2.8 L DOT cylinder containing
1.6 kg of chlorine trifluoride was used as the source. The initial vapor pressure
of the chlorine trifluoride was 1.87 kg/ cm2 at the test ambient temperature.
The vapor-phase exposure testing was conducted using a 51 cm long piece
of 0.312 cm diameter stainless steel tubing connected directly to the cylinder
with the use of a CGA 670 nut and nipple. The cylinder was mounted upright in
the vertical position inside the vent booth on a metal stand. The outlet of the
vapor discharge tubing was directed at the target piece, which was held by a
laboratory clamp and stand 共Fig. 2兲. For added safety, a local vent snorkel was
positioned behind the target piece to assist in vapor removal. The flow of the
vapor was controlled by manual operation of the cylinder valve.

Liquid-Phase Exposure Testing


The liquid exposure testing was conducted by inverting the cylinder to access
the liquid phase and mounting it above the metal stand. Two metal-to-metal
seat pneumatic nickel valves were connected to the cylinder. The first valve was
connected directly to the valve outlet. The second valve was coupled as close as
possible to the first valve with about 2.5 cm length of 0.635 cm diameter stain-
less steel tubing. The second valve discharged to a 20 cm length of 0.635 cm
diameter copper tubing, which directed the liquid-phase chlorine trifluoride to
the target piece by gravity. The target pieces were positioned on the metal stand
immediately below the tubing outlet. The testing was done by remote operation
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
480 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 2—Chlorine trifluoride vapor-phase exposure testing apparatus.

of the two pneumatic valves. The cylinder valve was opened admitting liquid
chlorine trifluoride up to the first valve seat. The first valve was cycled opened
and closed. This trapped approximately 2 mL of liquid between the pneumatic
valves. When the second valve was opened the trapped liquid was dropped onto
the target piece through the copper tube 共Fig. 3兲. The tests were conducted in
this manner to limit the amount of liquid ClF3 available to the sample being
exposed.
It is important to note that all equipment items used in this testing were
specially selected, assembled, and cleaned to meet Air Products’ standards for
fluorine and its oxidizing compounds. Also, the testing was performed and
attended by personnel very experienced in the hazards and behavior of ClF3
reactions.

Exposure Test Results


Chlorine trifluoride exposure testing was conducted over a 2-day period, and
the results were recorded using digital video and still cameras. The results are
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 481

FIG. 3—Chlorine trifluoride liquid-phase exposure testing apparatus.

summarized in five major categories of target materials, as follows:


• PPE 共Table 2兲
• Metals 共Table 3兲
• Plastics/elastomers 共Table 4兲
• Building materials 共Table 5兲
• Miscellaneous 共Table 6兲
In some cases where no reaction occurred with the target sample, a drop of
water was added to the liquid to observe if a reaction could be initiated. The
heat of reaction between ClF3 and water is often sufficient to ignite a combus-
tible sample that was exposed to liquid-phase chlorine trifluoride.
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
TABLE 2—PPE exposure test results.

Personal Protective Equipment

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Faceshield–polycarbonate No reaction No reaction; liquid left white
deposit on plastic
Glove, latex–new No reaction No reaction until liquid contacted
a contaminant and then material
ignited
Glove, nitrile–new No reaction Ignited immediately after liquid
exposure and burned vigorously
after first drop charred it
Glove, nitrile–oil Film Ignition as soon as exposed to Not tested
contamination vapor

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Glove, smooth leather–new Burning at point of impingement Instant flash on liquid contact
and shriveling of leather and leather shriveling
Glove, smooth leather–used Not tested Flashes and flames were more
intense and glove shriveled
significantly
Glove, rough heavy leather Orange flame that went out when Liquid immediately flashed on

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


–new flow stopped; exposure burned the surface and the second drop
hole through leather charred leather
Glove, rough heavy leather– Intense flame at oil spots and Not tested
oil film contamination charring of general area
Suit, acid resistant First test produced intense flame Explosion upon contact that left
and burn through of material; charred material where liquid
second test had no reaction impacted
482 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Suit, acid resistant–oil film No reaction Not tested


contamination
TABLE 2— 共Continued.兲

Personal Protective Equipment

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Suit, CPF4® material No reaction No initial reaction with small
strips, which then ignited when
liquid flowed underneath;
using a full suit, liquid ClF3
puddled and evaporated; no
reaction with slight discoloration
of material
Suit, Kappler Responder® No reaction No initial reaction with small
strips, which then ignited when
liquid flowed underneath
Suit, Kappler light material No reaction No initial reaction with small

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
strips, which then ignited when
liquid flowed underneath
Suit, neoprene material No reaction No immediate reaction but liquid
appeared to contact a
contaminant

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


that ignited material
Suit, Nomex Smooth and intense burning at Instant and intense flash with
the point of contact; burned flame extinguishing after liquid
completely through suit in 2 s; totally reacted
same result for three tests
VANOMMEREN, doi:10.1520/JAI102215 483
TABLE 2— 共Continued.兲

Personal Protective Equipment

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Suit, Tychem® No reaction Liquid had no reaction and
collected as a puddle;
droplets of water created small
flashes
and reaction was sufficient
to ignite material, which
continued to burn
Suit, Tyvek® Vapor traces remaining in Liquid exposure on material had
discharge tube immediately no reaction; liquid soaked into
burned the Tyvek® fibrous material;
Tyvek® as it was moved a drop of water introduced onto

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
around the tube outlet the saturated material triggered
a major detonation

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


484 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 3—Metals exposure test results.

Metals

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Aluminum chips Not tested No reaction; test chips were displaced by
liquid contact
Aluminum plate Not tested No visible reaction or discoloration; liquid
ClF3 just bubbled and evaporated
Aluminum plate–water on surface Not tested Slight brown fuming and white deposit left on
surface
Brass chips Not tested No reaction; one chip did ignite but could
have been due to oil contamination on test
stand
Brass fitting Not tested No reaction

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Carbon steel plate–epoxy painted Slight browning of the paint surface No significant reaction; liquid just puddled
and evaporated, and paint appeared to bubble
Carbon steel wool Vapor traces remaining in discharge Immediate intense flash totally reacting
tubing were enough to burn material as material with very little residue remaining
it was moved around tube outlet
Copper chips Not tested No reaction; one chip did ignite but could

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


have been due to oil contamination on test
stand
Copper tube Not tested No reaction; copper tube was used for all the
liquid drip tests and the tube after 40+ tests
had no evidence of reaction
Stainless steel chips Not tested Immediate flash with the second drop causing
a more intense flash
Stainless steel plate No reaction No reaction
VANOMMEREN, doi:10.1520/JAI102215 485
TABLE 4—Plastics/elastomers exposure test results.

Plastics/Elastomers

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
FRP, Derakane 470 No reaction Immediate flash on surface—liquid appeared
to be reacting with plasticizer
Gasket, red rubber Not tested Muffled pop and orange flame

Gasket, Teflon-filled–new No reaction Not tested


Gasket, Teflon-filled–old Not tested No reaction
Gasket Teflon-filled–oil film No reaction No reaction
contamination
Polyethylene bag–emergency No reaction during direct No reaction after 6 s of liquid contact until
response item impingement or filling of liquid flowed underneath Teflon block on test
bag with vapor stand; then bag ignited and burned

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Polyethylene netting–cylinder Not tested Material did not react until liquid contacted a
wall shipping protection contaminant, which initiated reaction and fire
Polyethylene tubing–new Not tested No reaction as the liquid dripped through the
tube
Polyethylene tubing–used Not tested Immediate bright orange flame entire tubing

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


length, which continued to burn with intense
yellow flame; same result for two tests
PVC Intense burning at the point of contact, Immediate surface flash and charring
similar to wood; burned circular
hole in test piece
Teflon block Not tested No reaction
486 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS
TABLE 5—Building materials exposure test results.

Building Materials

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Asphalt Intense burning at point of contact Small pieces totally exploded in both tests;
larger piece flared intensely with yellow flames
Cinder block Not tested Surface sparking with no visible damage
Concrete No reaction No reaction
Concrete–wet No visible reaction other than lightening of Not tested
area where sample was wetted
Duct tape No reaction. Delayed reaction when liquid flowed
underneath and contacted adhesive that
ignited sample
Wood, large timber Intense burning with sparks ejected at point Immediate surface flashing reactions. Then

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
of contact, similar to welding torch. Burned volatile gaseous compounds exploded above
circular hole in test piece the wood sample

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


VANOMMEREN, doi:10.1520/JAI102215 487
TABLE 6—Miscellaneous materials exposure test results.

Miscellaneous Materials

Downloaded/printed by
Material Vapor Exposure Liquid Exposure
Chicken, raw piece Immediately flared intensely during entire Liquid drops hitting skin immediately
vapor exposure; generated three large flares exploded with any unreacted splattered drops,
radially outward from point of contact resulting in
vigorous sparks and pops; fat appeared to
vaporize and volatile compounds exploded
above skin
Cotton rag Not tested Cotton burned immediately and vigorously;
became more intense when second drop
impacted charred material
Manila tag Burned with yellow flame at point of Not tested
contact and ignited remainder of tag

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Oil film on epoxy painted metal Not tested Instant explosion when liquid contacted
sample; circular char pattern created, which
traced the
liquid droplet splashes
Paper towel Paper burned immediately on contact and Paper immediately burst into flames but was

Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.


continued burning after vapor flow stopped; not a vigorous reaction
not a
vigorous reaction
Soda ash No reaction; solids were displaced by Not tested
vapor contact
Water, ice Rapid melting of ice with no fuming Rapid melting of ice with no fuming or noise
or noise evident evident
488 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Water, liquid Introduced above water surface; loud Dripped into water with immediate aggressive
popping and bubbling but no sparking flames popping and heating; water was ejected from
container with some sparking
Water, spray No reaction Not tested
VANOMMEREN, doi:10.1520/JAI102215 489

Liquid Exposure
Not tested
Not tested
Miscellaneous Materials

Vapor Exposure
No reaction
No reaction
Water film on aluminum plate
TABLE 6— 共Continued.兲

Water film on concrete


Material

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
490 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 4—共a兲 Vapor ClF3 exposure to nitrile gloves—clean glove. 共b兲 Vapor ClF3 exposure
to nitrile gloves—oil contaminated glove.

Exposure Testing Conclusions


As expected the exposure test results confirmed much of the industry reactivity
knowledge for ClF3; however, they did refute certain expected behaviors. Some
of the general conclusions were as follows:
• The ClF3 vapors were colorless and there was no fuming as reported in
some literature. This result occurred even when the vapor or liquid was
contacted onto water, onto ice, and sprayed with water.
• Liquid ClF3 does not evaporate readily even at an ambient temperature
of 26.7° C.
• Certain PPE had to be tested with large pieces since small pieces ignited
when liquid ClF3 contacted the uncoated underside of the sample.

Importance of Maintaining Cleanliness


One of the most significant findings was how reactive chlorine trifluoride can
be with common contaminants, such as water or hydrocarbon 共e.g., oil兲. In
some cases where there was no reaction between liquid ClF3 and the sample, if
liquid contacted water or oil contamination near the test sample, the reaction
heat was sufficient to ignite the sample material. This phenomenon was very
evident in liquid exposure testing on contaminated nitrile gloves 共Fig. 4兲, used
polyethylene tubing 共Fig. 5兲, and oil contaminated surfaces. Therefore, main-
taining cleanliness with any materials that may come into contact with ClF3,
and especially PPE, is of paramount importance.

Most Violent Test Result—“Bunny Suit” Explosion


The most severe reaction occurred with the Tyvek® suit material. Tyvek® is
routinely employed as the material for cleanroom garments at semiconductor
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 491

FIG. 5—共a兲 Vapor ClF3 exposure to polyethylene tubing—clean tubing. 共b兲 Vapor ClF3
exposure to polyethylene tubing—used tubing.

facilities, which are sometimes referred to as a bunny suit. The liquid ClF3 did
not initially react to Tyvek®; however, it started to soak into the material form-
ing a fuel-oxidizer mixture with intimate contact. A single drop of water caused
the saturated material to explode and damage the liquid testing rig. This explo-
sion result was of great concern to Air Products since it is feasible that a worker
handling ClF3 could accidentally spill some liquid onto their PPE, which would
permeate into the material. This would cause the worker to feel an immediate
irritation/burning in the exposed area, and workers are typically trained to
immediately use a safety shower.
However, to protect workers exposed to liquid ClF3 spills from serious re-
actions and chemical burns, they must be trained to first remove exposed cloth-
ing before stepping into a safety shower.

Exposure Results with Water


As expected, the reaction of ClF3 with water was vigorous, producing consid-
erable heat. In one test where liquid ClF3 was dripped onto a small pool of
water, there was no sound, just the immediate disappearance of the water
puddle. Vapor ClF3 that was passed above the liquid water surface actually
reacted with a prominent sound.

Importance of System Purging


Even trace amounts of chlorine trifluoride remaining in a tube that was not
purged properly were sufficient to cause immediate reactions. Samples of steel
wool, Tyvek®, and paper were placed against the end of a tube containing re-
sidual traces of ClF3 vapor, and the gas diffusing from the tubing caused an
immediate reaction 共Fig. 6兲. This was also evident in the raw chicken tests
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
492 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

FIG. 6—共a兲 Residual vapor exposure on materials—steel wool. 共b兲 Residual vapor ex-
posure on materials—Tyvek® suit.

where brief flashes were noticed in the tubing outlet probably due to vaporized
chicken fat severely reacting with the residual ClF3 vapor.

Vapor versus Liquid Reactivity


As expected, liquid ClF3 exposure was much more reactive than gas exposure.
Chlorine trifluoride is a low pressure liquefied gas, which can easily condense
in a system. If a system leak should then develop, the resulting liquid drip onto
clothing or another surface can cause an immediate and violent reaction.

Exposure Test Results on Fuels


As a powerful oxidizer, liquid ClF3 reacts immediately with any fuel, such as oil
or fat. The most visually dramatic test was the liquid chlorine trifluoride expo-
sure onto a piece of raw chicken. This test was conducted to observe the ap-
proximate reaction of ClF3 contact on exposed skin. In slow motion of the
video you can see the liquid contacts the skin and immediately reacts, dispers-
ing smaller ClF3 droplets around the point of contact area in a donut pattern.
Even the gas release onto a piece of raw chicken created a very loud and intense
flame 共Fig. 7兲.
The reaction with asphalt was also as expected. Two of the three liquid tests
produced detonation of the asphalt sample. Operators who have worked with
ClF3 have reported that leaking cylinders can sometimes sound like firecrackers
going off every few minutes due to the liquid drops contacting the ground. In
the design of a ClF3 system, one must ensure that there are no combustible
materials on the floor surface in the vicinity of the system.

Exposure Tests with No Reaction


The testing team was surprised to find no significant reaction with the concrete
or cinder block samples even when they were wetted with water. These results
are in conflict with what is reported in certain literature. However, just because
a reaction did not occur with certain materials 共especially some organics兲, the
same materials may ignite and react vigorously given enough energy or quan-
tity of exposed liquid. For example, the clean polyethylene bag did not imme-
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
VANOMMEREN, doi:10.1520/JAI102215 493

FIG. 7—共a兲 Vapor and liquid release on raw chicken—vapor exposure. 共b兲 Vapor and
liquid release on raw chicken—liquid drop first contact. 共c兲 Vapor and liquid release on
raw chicken—liquid splatter and sparking.

diately react with liquid exposure until some of the puddled liquid contacted a
contaminant on the Teflon block securing the bag. Once ignited, the bag burned
intensely and completely. Also, if the drop height for liquid contact was much
higher, more energy would be imparted to the surface and an immediate reac-
tion might have initiated with certain organic samples.

Final Exposure Testing Conclusions

In conclusion, the Air Products’ testing has demonstrated and confirmed that
chlorine trifluoride is a very reactive material in both vapor and liquid phases.
Manufacturers, consumers, and emergency responders working with ClF3
should make absolutely sure that the systems and PPE they use are meticu-
lously clean and that all personnel are very familiar with the hazards of ClF3.
As a direct result of this testing, Air Products has made significant changes
to our chlorine trifluoride Material Safety Data Sheet. Also, Air Products has
developed and provides a special training session for all of our employees and
customers handling ClF3. 6 11:40:09 EDT 2014
Copyright by ASTM Int'l (all rights reserved); Tue May
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
494 JAI • STP 1522 ON FLAMMABILITY AND SENSITIVITY OF MATERIALS

Acknowledgments

The writer would like to thank D. Croll, S. Coombes, D. Fisher, R. Martrich, A.


Santay, G. Stianchie, and C. Whitehead, all of Air Products and Chemicals, Inc.,
and E. Ngai formerly of Air Products, for their efforts on the original testing
and documentation for this study.

References

关1兴 Stacey, M., Sharpe, A. G., and Tatlow J. C., Advances in Fluorine Chemistry, Vol. 1,
Academic Press, New York, 1960.
关2兴 Cloyd, D. R. and Murphy, W. J., “Handling Hazardous Materials,” Technology Sur-
vey No. SP-5032, Technology Utilization Division, National Aeronautics and Space
Administration, Washington, D.C., September 1965.
关3兴 Clark, J. D., Ignition! An Informal History of Liquid Rocket Propellants, Rutgers
University Press, New Brunswick, NJ, 1972.
关4兴 Gugliemini, C. J., Chlorine Trifluoride Technical Data, Safety and Handling, Air
Products and Chemicals, Inc., Allentown, PA, 1997.
关5兴 Gugliemini, C. J. and Johnson, A. D., “Properties and Reactivity of Chlorine Trif-
luoride,” Semiconductor International Magazine, June 1, 1999.
关6兴 Do, V. Q., “Oxidizing Potential of F Gases—Thermochemical Analysis,” Internal
Memorandum, Air Products and Chemicals, Inc., Allentown, PA, January 29, 1997.
关7兴 FM Global, Loss Prevention Data 7–7R, Semiconductor Fabrication Facilities 17–
12R, Factory Mutual Insurance Co., Johnston, RI, January 2003.

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
STP1522-EB/Jan. 2009
495

Author Index
A Hirsch, D., 175-94
Hooser, J. D., 429-49
Arlt, K., 133-43 Houghton, P. A., 303-21

B L

Barthélémy, H., 144-74, 322-36 Lehné, S., 124-32


Binder, C., 124-32, 133-43, 362-88 Lombard, S., 144-74
Brock, T., 124-32, 133-43 Longuet, O., 49-67, 144-74
Brophy, B., 266-84 Lynn, D., 21-37, 80-95
Bruat, J., 144-74
M
C
McDonald, R. H., Jr., 303-21
Carré, M., 144-74 Million, J. F., 10-20
Cawthra, M., 303-21
Chiffoleau, G. J. A., 96-108, 266-84, N
285-302, 362-88, 429-49
Chowdhury, K., 389-428 Newton, B. E., 96-108, 266-84,
Colson, A., 322-36 285-302, 337-61, 362-88, 429-49
Crayssac, F., 49-67
O
D
Odom, G. A., 96-108
Davis, S. E., 68-79, 109-23 Oza, A., 389-428

F P

Fielding, J. R., 96-108 Peralta, S. F., 204-18


Forsyth, B., 285-302
Forsyth, E. T., 266-84, 285-302 R
Frisby, P. M., 109-23
Robbins, K. E., 68-79
G Rosales, K. R., 204-18, 227-45,
246-65
Gallus, T., 219-26
Ghosh, S., 389-428 S

H Samant, A. V., 10-20


Schaaff, J. P., 144-74
Harper, S., 175-94 Schmidt, W. P., 303-21
Hartwig, P., 133-43 Sherwood, R. J., 303-21
Herald, S. D., 68-79, 109-23 Slusser, J. W., 1-9
Hesse, O., 124-32, 133-43
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Smith, S., 175-94, 195-203, 219-26
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Copyright © 2009 by ASTM International www.astm.org
496

Sparks, K., 21-37, 80-95, 219-26 W


Steinberg, T., 21-37, 38-48, 80-95,
337-61, 362-88, 450-67 Ward, N. R., 38-48, 450-67
Stoltzfus, J. M., 21-37, 80-95, 204-18, Wei, M., 429-49
227-45, 246-65 Werlen, E., 49-67, 322-36
Wieder, S. J., 303-21
T Willot, F., 49-67
Tillack, T., 124-32, 133-43

V Z

VanOmmeren, J., 468-94 Zawierucha, R., 10-20

Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
STP1522-EB/Jan. 2009
497

Subject Index
A components, 133-43
computational fluid dynamics, 429-
adiabatic compression, 337-61, 49
362-88 contamination, 322-36, 322-36
adsorbent, 322-36 copper, 49-67
advanced crew escape suit 共ACES兲, crew escape suit, 175-94
246-65
cryogenic, 49-67, 322-36
air separation, 303-21
cutting torch, 266-84
air separation unit, 49-67
CW6M, 10-20
aluminum, 49-67, 204-18
cylinder valves, 429-49
aluminum anodization, 246-65
aluminum combustion, 303-21
D
aluminum foil, 322-36
applications, 468-94
Dalton, 68-79
argon, 49-67
design, 266-84
ASU, 303-21
diffusion, 175-94, 175-94
autogenous ignition temperature,
dust layer, 389-428
195-203
autoignition temperature, 124-32
E
B
electrical arc ignition, 219-26
BAHX, 303-21 Elgiloy, 10-20
batch testing, 133-43 exchanger, 49-67
bellow, 322-36 Extravehicular Mobility Unit, 175-94
brazed aluminum heat exchanger,
303-21 F
burn criteria, 21-37
burn length, 21-37 filling process, 124-32
burning, 80-95, 450-67 fire hazard, 124-32
fire risk, 285-302
C fires, 266-84
flammability, 1-9, 10-20, 21-37,
cavity, 303-21 49-67, 80-95, 124-32, 133-43, 175-94
cavity incident, 303-21 flammability materials testing,
CGA 870 seals, 429-49 219-26
chlorine trifluoride, 468-94 flow friction ignition, 429-49
combustion, 144-74 fluorinated polymers, 109-23
compatibility, 322-36
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
foil, 49-67
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
Copyright © 2009 by ASTM International www.astm.org
498

G isentropic compression, 337-61,


362-88
gas analysis, 144-74
gaseous fluid impact, 337-61, 362-88
gaseous oxygen enrichment, L
175-94
getter, 322-36 layer, 49-67
glass fiber, 322-36 liquid, 49-67
good practices, 266-84 liquid oxygen, 322-36
liquid oxygen environment, 204-18
H

handheld electrical devices, 219-26 M


handling, 468-94
maintenance, 133-43
Hastelloy B-3, 10-20
manufacturing, 322-36
Hastelloy W, 10-20
MAPTIS, 109-23
hazards, 468-94
maximum oxygen concentration
hazards analysis, 285-302
共MOC兲, 175-94
heat affected, 80-95
medical oxygen, 285-302
heat affected zone, 21-37
melting, 38-48, 450-67
heat conduction, 21-37
metal combustion, 21-37
heat transfer, 38-48, 450-67
metallic contaminant particles, 204-
heterogeneous burning, 38-48
18
hexadecane, 49-67 metals, 10-20
high pressure, 227-45 microchannel flow, 429-49
high purity oxygen atmosphere, 68-
microgravity, 450-67
79
model, 38-48
hydrocarbon, 49-67
MP35N, 10-20
MSDS, 468-94
I

igniter effects, 21-37 N


ignition, 49-67, 80-95, 285-302,
389-428 nitrogen, 49-67
impact, 49-67, 322-36 NITROX, 124-32
impact test, 322-36 non-metallic materials, 144-74
insulation material, 322-36 nonmetallic materials,
intermediate layers, 322-36 124-32, 133-43, 195-203
interspace, 322-36 nonmetals, 175-94
iron, 450-67
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
nonorganic, 322-36
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
499

O propagating brush discharge, 389-


428
organic, 322-36 properties, 468-94
oxidizer, 1-9 PTFE, 144-74
oxygen, 1-9, 49-67 purge cycles, 68-79
Oxygen, 80-95 purge equation, 68-79
oxygen, 144-74, 219-26, 266-84, purity, 49-67
337-61, 362-88, 389-428
oxygen cleaning, 266-84 Q
oxygen compatibility, 1-9, 133-43
oxygen concentration, 68-79, quality assurance, 133-43
195-203 quick disconnect, 227-45
oxygen enrichment, 124-32
oxygen fires, 429-49 R
oxygen hazards, 285-302
oxygen lance, 266-84 RA 330, 10-20
oxygen mixtures, 124-32 rate-limiting mechanism, 38-48,
450-67
oxygen percentage, 68-79
reboiler, 49-67, 303-21
oxygen systems, 109-23
release, 468-94
rods, 80-95
P rotating friction ignition testing, 204-
18
partial pressures, 68-79
particle impact ignition testing, 246- S
65
particle impact testing, 227-45 SCUBA diving, 124-32
PCTFE, 144-74 sealed cavity, 303-21
permeation, 175-94 severity index, 362-88
PICT behavior, 10-20 shock-wave heating, 337-61, 362-88
pneumatic impact, 124-32, 133-43, solid/liquid interface, 450-67
195-203, 337-61, 362-88 stagnation heating, 429-49
polyester, 322-36 stainless steel, 227-45
polytetrafluoroethylene, 109-23 steel mill, 266-84
PPE, 468-94 Stellite 6, 10-20
promoted combustion, 21-37, 80-95 superinsulation, 322-36
promoted combustion chamber,
68-79 T
promoted ignition, 21-37
promoted ignition-combustion, 10- temperature, 80-95
20 testing, 468-94
promoter, 80-95
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
textiles, 175-94
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.
www.astm.org
ISBN: 978-0-8031-7508-2
Stock #: STP1522
Copyright by ASTM Int'l (all rights reserved); Tue May 6 11:40:09 EDT 2014
Downloaded/printed by
Rochester Institute Of Technology pursuant to License Agreement. No further reproductions authorized.

Potrebbero piacerti anche