Anda di halaman 1dari 167

Special Project

AIAA
SP-084-1999
Fire, Explosion, Compatibility, and Safety
Hazards of Hypergols - Hydrazine
AIAA standards are copyrighted by the American Institute of Aeronautics and
Astronautics (AIAA), 1801 Alexander Bell Drive, Reston, VA 20191-4344 USA. All rights
reserved.
AIAA grants you a license as follows: The right to download an electronic file of this AIAA
standard for temporary storage on one computer for purposes of viewing, and/or printing
one copy of the AIAA standard for individual use. Neither the electronic file nor the hard
copy print may be reproduced in any way. In addition, the electronic file may not be
distributed elsewhere over computer networks or otherwise. The hard copy print may only
be distributed to other employees for their internal use within your organization.
AIAA SP-084-1999
1
Correction Log
5 December 2001
1. Reference for ERP E515-80 on page 66 should be TR-226-001.
2. Expression for liquid density obtained from Yaws on page 132 should reflect proper conversion for
reduced temperature and proper exponent. The expression is replaced by the following equation:
2/7
80.0) 273.15K)/3 (T(K) 1 3
L
538 0.3171x0.2 ) (Mg/m

=
(

3. Table B.1 (that was actually for Monomethylhydrazine) is replaced with the correct table for hydrazine.
Please see page 2 of the correction log for the table.
AIAA-SP-084
2
Table B.1 Thermodynamic properties of hydrazine
T
K
P
KPa
V
L
m
3
/g
V
V
m
3
/g
H
L
(T,P)
Sat. liquid
J/g
H
V
(T,P)
Sat. vapor
J/g
S
L
(T,P)
Sat. liquid
J/g K
S
V
(T,P)
Sat. Vapor
J/g K
288 0.99 9.98E-07 7.52E-02 -29.91 1379.03 0.0258 4.79
298 1.86 1.00E-06 4.15E-02 0.00 1394.28 0.0000 4.68
308 3.34 1.01E-06 2.39E-02 28.90 1409.16 -0.0202 4.58
318 5.73 1.02E-06 1.44E-02 57.80 1424.59 -0.0311 4.49
328 9.48 1.03E-06 8.96E-03 86.47 1440.25 -0.0351 4.40
338 15.16 1.03E-06 5.77E-03 115.01 1456.14 -0.0332 4.33
348 23.51 1.04E-06 3.83E-03 143.51 1472.22 -0.0258 4.26
358 35.47 1.05E-06 2.61E-03 172.09 1488.47 -0.0135 4.20
368 52.16 1.06E-06 1.82E-03 200.84 1504.88 0.0030 4.15
378 74.97 1.07E-06 1.30E-03 229.87 1521.40 0.0235 4.10
388 105.51 1.08E-06 9.42E-04 259.30 1538.01 0.0475 4.06
398 145.65 1.13E-06 6.97E-04 289.75 1554.62 0.1422 4.02
408 197.54 1.10E-06 5.25E-04 319.79 1571.30 0.1048 3.98
418 263.59 1.11E-06 4.01E-04 351.05 1587.90 0.1376 3.95
428 346.47 1.13E-06 3.10E-04 383.14 1604.40 0.1729 3.92
438 449.14 1.14E-06 2.43E-04 416.15 1620.75 0.2104 3.89
448 574.77 1.16E-06 1.93E-04 450.18 1636.88 0.2500 3.86
458 726.81 1.17E-06 1.54E-04 485.30 1652.73 0.2915 3.84
468 908.90 1.19E-06 1.25E-04 521.62 1668.24 0.3350 3.82
478 1124.88 1.21E-06 1.02E-04 559.20 1683.33 0.3801 3.80
488 1378.76 1.23E-06 8.34E-05 598.11 1697.92 0.4270 3.78
498 1674.68 1.25E-06 6.90E-05 638.43 1711.92 0.4754 3.77
508 2016.89 1.27E-06 5.74E-05 680.22 1725.24 0.5254 3.75
518 2409.71 1.30E-06 4.80E-05 723.54 1737.79 0.5770 3.73
528 2857.46 1.33E-06 4.03E-05 768.43 1749.43 0.6301 3.72
538 3364.48 1.36E-06 3.40E-05 814.96 1760.04 0.6847 3.70
548 3935.03 1.40E-06 2.88E-05 863.17 1769.46 0.7409 3.69
558 4573.25 1.44E-06 2.44E-05 913.13 1777.52 0.7987 3.67
568 5283.16 1.49E-06 2.08E-05 964.90 1783.98 0.8583 3.66
578 6068.57 1.54E-06 1.77E-05 1018.57 1788.59 0.9198 3.64
588 6933.03 1.60E-06 1.51E-05 1074.28 1790.97 0.9835 3.62
598 7879.81 1.68E-06 1.28E-05 1132.21 1790.65 1.0496 3.59
608 8911.84 1.77E-06 1.09E-05 1192.69 1786.91 1.1188 3.57
618 10031.63 1.88E-06 9.14E-06 1256.32 1778.67 1.1919 3.54
628 11241.30 2.03E-06 7.61E-06 1324.26 1764.03 1.2705 3.50
638 12542.46 2.25E-06 6.21E-06 1399.38 1739.03 1.3584 3.44
648 13936.22 2.68E-06 4.81E-06 1493.07 1692.16 1.4695 3.36
Reference temperature = 298 K
T = temperature H
L
= enthalpy, liquid
P = pressure H
V
= enthalpy, vapor
V
L
= specific volume, liquid S
L
= entropy, liquid
V
V
= specific volume, vapor S
V
= entropy, vapor
AIAA
SP-084-1999
Special Project Report
Fire, Explosion, Compatibility, and Safety
Hazards of Hypergols - Hydrazine
Sponsored by
American Institute of Aeronautics and Astronautics
Approved
Abstract
This Special Project report presents information that designers, builders, and users of hydrazine systems
can use to avoid or resolve hydrazine hazards. Pertinent research is summarized, and the data are
presented in a quick-reference form. Further information can be found in the extensive bibliography.
AIAA SP-084-1999
ii
Published by
American Institute of Aeronautics and Astronautics
1801 Alexander Bell Drive, Reston, VA 22091
Copyright 1999 American Institute of Aeronautics and
Astronautics
All rights reserved
No part of this publication may be reproduced in any form, in an electronic
retrieval system or otherwise, without prior written permission of the publisher.
Printed in the United States of America
AIAA SP-084-1999
iii
Contents
Foreword ................................................................................................................................................ vi
Acronyms ..............................................................................................................................................viii
Glossary ................................................................................................................................................. ix
Trademarks ...........................................................................................................................................xiii
1 Introduction to hazard assessment .............................................................................................. 1
1.1 About this special report .............................................................................................................. 1
1.2 Approach to performing a hazard assessment ............................................................................. 1
1.2.1 Fire hazards................................................................................................................................ 3
1.2.2 Explosion hazards....................................................................................................................... 4
1.2.3 Material compatibility................................................................................................................... 8
1.2.4 Exposure hazards ....................................................................................................................... 9
1.3 Overall hazard........................................................................................................................... 10
2 Fire ........................................................................................................................................... 11
2.1 Hydrazine vapor ........................................................................................................................ 11
2.1.1 Flammability.............................................................................................................................. 11
2.1.2 Ignition...................................................................................................................................... 17
2.1.3 Flame velocity........................................................................................................................... 27
2.2 Liquid hydrazine........................................................................................................................ 32
2.2.1 Flash and fire points.................................................................................................................. 32
2.2.2 Burning rate and burning velocity .............................................................................................. 33
2.3 Hydrazine mists, droplets, and sprays ....................................................................................... 34
2.3.1 Flash and fire points.................................................................................................................. 34
2.3.2 Burning rates............................................................................................................................. 34
3 Explosion .................................................................................................................................. 37
3.1 Deflagration............................................................................................................................... 37
3.1.1 Hydrazine vapor ........................................................................................................................ 38
3.1.2 Liquid hydrazine........................................................................................................................ 40
3.2 Detonation ................................................................................................................................ 41
3.2.1 Detonation theory...................................................................................................................... 41
3.2.2 Hydrazine vapor ........................................................................................................................ 41
3.2.3 Liquid hydrazine........................................................................................................................ 47
3.3 Thermochemical reaction .......................................................................................................... 48
3.3.1 Thermodynamic instability......................................................................................................... 48
3.3.2 Thermal runaway....................................................................................................................... 49
AIAA SP-084-1999
iv
3.3.3 Rapid compression.................................................................................................................... 52
4 Hydrazine and material compatibility ......................................................................................... 58
4.1 Material degradation.................................................................................................................. 58
4.1.1 Test method.............................................................................................................................. 58
4.1.2 Test conditions.......................................................................................................................... 58
4.1.3 Material compatibility data......................................................................................................... 59
4.1.4 Alloy corrosion studies............................................................................................................... 60
4.2 Material effects on hydrazine..................................................................................................... 60
4.2.1 Test methods ............................................................................................................................ 60
4.2.2 Test conditions.......................................................................................................................... 71
4.2.3 Hydrazine decomposition data................................................................................................... 72
4.2.4 Chemical reactivity of hydrazine in air........................................................................................ 80
4.3 Assessment examples............................................................................................................... 80
4.3.1 Estimation of relative decomposition rates for material applications for which the
material response to hydrazine is well known ............................................................................ 81
4.3.2 Effects of materials on the heat generation rate......................................................................... 83
4.3.3 Hazard analyses........................................................................................................................ 85
5 Safety ....................................................................................................................................... 91
5.1 Hydrazine toxicity...................................................................................................................... 91
5.1.1 Results of hydrazine exposure................................................................................................... 92
5.1.2 Inhalation .................................................................................................................................. 93
5.1.3 Exposure to skin, eyes, and mucous membranes ...................................................................... 94
5.1.4 Ingestion ................................................................................................................................... 94
5.1.5 Carcinogenicity ......................................................................................................................... 94
5.2 Environmental fate of hydrazine ................................................................................................ 95
5.2.1 Air ............................................................................................................................................. 95
5.2.2 Water ........................................................................................................................................ 95
5.2.3 Soil ........................................................................................................................................... 95
5.2.4 Bioaccumulation and biodegradation......................................................................................... 96
5.2.5 Remediation.............................................................................................................................. 96
5.3 Hydrazine exposure guidelines.................................................................................................. 96
5.3.1 Threshold limit values of the American Conference of Governmental
Industrial Hygienists .................................................................................................................. 97
5.3.2 Guidelines of the Occupational Safety and Health Administration .............................................. 97
5.3.3 The National Institute Of Occupational Safety And Health.......................................................... 98
5.3.4 Emergency planning requirements ............................................................................................ 98
5.3.5 Spacecraft maximum acceptable concentrations ....................................................................... 99
AIAA SP-084-1999
v
5.3.6 Environmental regulations ......................................................................................................... 99
5.4 Hydrazine exposure remediation and control ........................................................................... 100
5.4.1 First aid................................................................................................................................... 100
5.4.2 Personnel protection ............................................................................................................... 103
5.4.3 Smell....................................................................................................................................... 103
5.4.4 Medical surveillance................................................................................................................ 104
5.4.5 Evacuation procedures............................................................................................................ 104
5.4.6 Protective apparel ................................................................................................................... 104
5.4.7 Fire fighting............................................................................................................................. 105
5.4.8 Spill cleanup............................................................................................................................ 105
5.5 Hydrazine handling.................................................................................................................. 106
5.5.1 Engineering design.................................................................................................................. 106
5.5.2 Storage containers .................................................................................................................. 107
5.5.3 Storage areas.......................................................................................................................... 108
5.5.4 Transportation......................................................................................................................... 108
5.5.5 Monitoring equipment.............................................................................................................. 110
5.5.6 Waste disposal........................................................................................................................ 113
5.5.7 Current regulatory enforcement ............................................................................................... 114
5.5.8 Additional information.............................................................................................................. 115
5.6 Assessment example .............................................................................................................. 115
Annex A: Hazard assessment example............................................................................................... 121
Annex B: Chemical, physical, and thermodynamic properties of hydrazine.......................................... 130
Annex C: References.......................................................................................................................... 138
AIAA SP-084-1999
vi
Foreword
Hydrazine is a colorless, corrosive, strongly reducing liquid compound. Current aerospace applications
include its use in the Space Transportation System as a fuel for the auxiliary power units, in satellites as a
monopropellant for thrusters, and in jet aircraft as fuel for auxiliary power sources. Although hydrazine is
immensely useful in these applications, there are also drawbacks. For example, hydrazine vapor is
flammable and detonable; both liquid and vapor hydrazine are corrosive, react with many materials, and
are susceptible to catalytic decomposition; and hydrazine is highly toxic. The users and designers of
hydrazine systems must be aware of these hazards and safeguard against them.
This AIAA Special Report preserves the text of NASA document RD-WSTF-0002 Rev A, December 17,
1998, Fire, Explosion, Compatibility, and Safety Hazards of Hydrazine, developed by the NASA White
Sands Test Facility for the Propulsion and Power Division of the Lyndon B. Johnson Space Center and
the Air Force Space Division. In the interests of technology transfer, custody of the material was
assigned to AIAA through of Memorandum of Understanding dated February 1999. One of the purposes
of this Memorandum is to provide broader distribution of the valuable information developed and
published in the original manual.
The authors of the NASA Revision A manual, dated December 16, 1998 are: Stephen S. Woods, Donald
B. Wilson, Dennis D. Davis, Michelle Barragan, Walter Stewart, Radel L. Bunker, and David L. Baker.
Authors to the earlier 1990 edition are: Michael D. Pedley, David L. Baker, Harold D. Beeson, Richard C.
Wedlich, Frank J. Benz, Radel L. Bunker, and Nathalie B. Martin.
AIAA Special Reports are a part of the AIAA Standards Program and frequently serve as precursors to
formal consensus documents. This publication is under the purview of the Liquid Propulsion Committee
on Standards, the group responsible for determining the future of the publication and for maintaining it in
a technically current state.
The AIAA Standards Procedures provide that all approved standards, recommended practices, and
guides are advisory only. Their use by anyone engaged in industry or trade is entirely voluntary. There is
no agreement to adhere to any AIAA standards publication and no commitment to conform to or be
guided by any standards report. In formulating, revising, and approving standards publications, the Liquid
Propulsion Committee on Standards will not consider patents that may apply to the subject matter.
Prospective users of the publications are responsible for protecting themselves against liability for
infringement of patents, or copyrights, or both.
At the time of publication, the members of the AIAA Liquid Propulsion Committee on Standards were:
Robert Ash Old Dominion University
Kyaw Aung Georgia Institute of Technology
C. T. Avedisian Cornell University
Curt Botts Air Force, 45
th
Space Wing, Patrick AFB
Patrick Carrick Phillips Laboratory
Fred Cutlick California Institute of Technology
Tom Draus NASA Kennedy Space Center
Irvin Glassman Princeton University
Howard Julien AlliedSignal Technical Services
Chad Keller Hughes Aerospace
Charles Leveritt Army Research Laboratory
Dennis Meinhardt Primex Aerospace Company
AIAA SP-084-1999
vii
Mark Mueller Aerospace Corporation
Gregory Nunz Los Alamos National Laboratory
Bryan Palaszewski NASA Lewis Research Center
Steven Schneider NASA Lewis Research Center
Joseph Shepherd California Institute of Technology
Timothy Smith NASA Glenn Research Center at Lewis Field
Bill St. Cyr NASA Stennis Space Center
Ray Traggianese Olin
Mark Underdown NASA Goddard
Stephen Woods AlliedSignal Technical Services

The Standards Executive Council accepted the document for publication on April 30, 1999.
AIAA SP-084-1999
viii
Acronyms
A, E
a
Arrhenius parameters
ACGIH American Conference of Governmental and Industrial Hygienists
AHJ Authority Having Jurisdiction
APU Auxiliary power unit
ARC Accelerating rate calorimeter
ASME American Society of Mechanical Engineers
CERCLA Comprehensive Environmental response, Compensation, and
Liability Act
C-J Chapman-Jouguet
CPIA Chemical Propulsion Information Agency
DDT Deflagration to detonation transition
DOT Department of Transportation
DOT Department of Transportation
EIS Electrochemical Impedance Spectroscopy
EPA Environmental Protection Agency
FDCA Food, Drug, and Cosmetics Act
GI Gastrointestinal
HAZMAT Hazardous Materials
HZ Hydrazine
IDLH Immediately Dangerous To Life or Health
IRFNA Inhibited Red Fuming Nitric Acid
JANAF Joint-Army-Navy-Air Force
JANNAF Joint-Army-Navy-NASA-Air Force
LEPC Local Emergency Planning Committee
LFL and UFL Lower and Upper Flammability Limit
MIE Minimum Ignition Energy
MMH Monomethylhydrazine
MSDS Material Safety Data Sheet
NIOSH National Institute of Occupational Safety and Health
NPDES National Pollutant Discharge Elimination System
OSHA Occupational Safety and Health Administration
PEL Permissible Exposure Limits
RCRA Resource Conservation and Recovery Act
REL Recommended Exposure Limit
RQ Reportable Quantity
SARA Superfund Amendments and Reauthorization Act
SERC State Emergency Response Commission
SI International System units
SMAC Spacecraft Maximum Acceptable Concentration
STEL Short Term Exposure Limit
STS Space Transportation System
TLD-1 Toxic Level Detector 1
TLV Threshold Limit Value
TLV-TWA Threshold Limit Value - Time Weighted Average
TOMES Toxicology, Occupational Medicine, and Environmental Series
TPQ Threshold Planning Quantity
TRI Toxic Release Inventory
TSCA Toxic Substances Control Act
AIAA SP-084-1999
ix
Glossary
Activation Energy (or Apparent Activation Energy): In absolute-rate theory, the energy associated
with the formation of an activated complex intermediate between the reactant(s) and product(s) of an
elementary reaction. An apparent activation energy is used as the parameter E
a
in an Arrhenius
function when the exact kinetic mechanism is unknown.
Adiabatic: A process in which the system changes state without thermal energy exchange between the
system and the surroundings.
Adiabatic Compression: Mechanical work transferred to a system under conditions where there is an
increase in the internal energy of the material for a static system or an increase in the enthalpy for a
dynamic system. If the process is also reversible (in the thermodynamic definition), this change is also
isentropic.
Adiabatic Factor: The temperature change that occurs when all the limiting reactant is completely
consumed (normalized extent of reaction equals 1 when the reaction system is operated adiabatically).
This factor is useful for comparing exothermicity or endothermicity of several reactions.
Adiabatic Flame Temperature: The temperature of thermodynamic equilibrium in a reaction or in a set
of reactions that occurs in a process operating adiabatically.
Arrhenius Function: A mathematical model for defining the temperature dependency of an observed
macroscopic kinetic reaction rate. The rate is equal to Aexp(-E
a
/RT), where A is the preexponential
factor, E
a
is the apparent activation energy, R is the ideal gas law constant, and T is the absolute
temperature.
Authority Having Jurisdiction (AHJ): Organization, office, or individual responsible for approving
equipment, an installation, or a procedure. The designation is used in a broad manner because
jurisdiction and approval agencies vary, as do their responsibilities. Where public safety is primary, the
AHJ may be a federal, state, local, or other regional department or individual such as a fire chief, fire
marshal, chief of a fire prevention bureau, labor department, health department, building official, electrical
inspector, or others having statutory authority. In many circumstances, the AHJ is the property owner or
his designated departmental official. At government installations, the AHJ may be the commanding
officer or a designated department official. Approved is herein defined as being authorized by, or
acceptable, to the AHJ.
Autoignition Temperature: The lowest temperature at which a material will spontaneously ignite. No
additional ignition energy (ignition source) is required.
Burning Rate: The rate of liquid mass consumption per unit area (kg/(m
2
s)).
Burning Velocity: The velocity at which the liquid level decreases. It is the burning rate/density of the
liquid (m/s).
Catalyst: A chemical compound or chemical species that alters the rate of a chemical reaction. The
catalyst is not altered by the reaction.
Cell Size: Refers to the width of the characteristic fish scale-shaped cell pattern etched on a smoked foil
during a gas-phase detonation. The cell pattern is produced by the path of the triple-point, i.e.
intersection of a primary shock wave, a transverse shock wave, and a Mach stem wave. The cell width is
used as a parameter for characterizing detonations.
AIAA SP-084-1999
x
Chapman-Jouguet Detonation: Describes a stable detonation state that is consistent with most
experimental measurement. The steady-state solution of the conservation equations at which the Rankin
line is tangent to the Hugoniot curve.
Critical Diameter: The tube diameter necessary for a gaseous detonation to propagate from the tube to
an unconfined environment (Equation 3.6, d
c
= 13).
Critical Energy: The minimum energy required to initiate an unconfined detonation in a specified
mixture.
Deflagration: A flame moving through a flammable mixture in the form of a subsonic wave (with respect
to the unburned mixture).
Dermal Exposure: The penetration of a toxic chemical through skin and into the blood stream. Such
exposure can be rapid and difficult for the body to eliminate.
Detonable Mixture: The state of a specified mixture that, upon application of the critical energy for the
mixture, will initiate and sustain a detonation. This mixture is also flammable.
Detonation: Exothermic chemical reaction coupled to a shock wave that propagates through a
detonable mixture. The velocity of the shock wave is supersonic with respect to the unburned gases.
After initiation, the thermal energy of the reaction sustains the shock wave, and the shock wave
compresses the unreacted material to sustain the reaction. Initiation of a detonation can occur with a
deflagration to detonation transition (DDT).
Diluent: In a mixture, an inert material that reduces the concentration of the remaining materials.
Dynamic Parameters: Four parameters (detonability limits, initiation energy, cell size, and critical tube
diameter) used to describe the dynamic characteristics of a detonation. They are in contrast to the static
C-J detonation parameters; velocity, detonation temperature, detonation pressure, and detonation
density.
Equivalence Ratio: The ratio of the actual fuel-oxidizer ratio to the stoichiometric fuel-oxidizer ratio.
Exothermic: The production of thermal energy by a chemical reaction.
Explosion: The rapid equilibration of pressure between the system and the surroundings, such that a
shockwave is produced. Explosions may occur through mechanical failure of vessels containing high-
pressure fluids or through rapid chemical reactions producing large volumes of hot gases.
Explosion Potential: A parameter that is useful for comparing an unknown system (reaction) to well-
known reactions. It is the product of system volume (constant), adiabatic flame temperature, total moles
per unit of mass in the system at equilibrium, and the ideal gas law constant.
Fire: Sustained burning, as manifested by any or all of the following: light, flame, heat, and smoke
(ASTM E 176-97b).
*
Fire Point: The lowest temperature at which a flame continuously exists over a liquid surface upon
ignition by an open flame.

*
ASTM Fire Test Standards, 3rd Ed. 1990, ASTM, 1916 Race St., Philadelphia, PA 19103.
AIAA SP-084-1999
xi
Flame: A hot, usually luminous zone of gas, or particulate matter in gaseous suspension, or both, that is
undergoing exothermic chemical reaction. A flame may be stationary with the flammable mixture fed into
the reaction zone, or a flame may propagate through the flammable mixture, as in a deflagration.
Flame Speed: Refers to the velocity of propagation of the reaction zone through the flammable mixture,
as measured by a stationary observer. Usually measured at the front of the flame.
Flame Velocity: When coordinates are centered in the flame front it is the velocity at which unburned
gases move through the combustion zone in the direction normal to the flame front.
Flammability Limits: The lower (LFL) and upper (UFL) vapor concentrations (usually reported as
volume percent) of fuel in a flammable mixture that will ignite and propagate a flame. These limits are
functions of temperature, pressure, diluents, and ignition energy.
Flash Point: The lowest temperature, corrected to 101.3 kPa (14.7 psia) of pressure, of a material at
which application of an ignition source causes the vapor of the material to ignite momentarily under
specified conditions.
Froth: The froth is a medium in which gas bubbles are surrounded by a thin film of liquid hydrazine,
maximizing the surface area between the ullage gas and the liquid hydrazine.
Halocarbons: Organic compounds containing one or more of the elements fluorine, chlorine, bromine,
and iodine.
Hazard: A situation (or potential event) that may result in death or injury to personnel, or damage to
equipment. Includes the effect of fire, flash, explosion, shock, concussion, fragmentation, corrosion, or
toxicity.
Heat Generation Potential: The product of the temperature sensitivity and the adiabatic factor. A useful
dimensionless group for characterizing and comparing reactions.
Heat of Reaction: For a given temperature and pressure, the enthalpy of the products of a reaction
minus the enthalpy of the reactants.
Hypergolic: Spontaneous ignition of two materials upon contact (no additional ignition energy is
required).
Ignition: Introduction of sufficient energy into a flammable mixture or material to produce a flame.
Ingestion: The introduction of a toxic material into the body through the mouth or by breathing.
Material Compatibility: Materials are considered compatible with each other if their rate of degradation
when in contact is insignificant for the application.
Minimum Ignition Energy: The minimum energy required to ignite a flammable mixture under given
conditions (temperature, pressure, diluents).
Monopropellant: A liquid propellant that decomposes exothermically to produce hot gases; e.g.,
hydrogen peroxide, hydrazine.
Order of Reaction: A parameter used to define the concentration (or pressure) dependency of the
kinetic rate of reaction. For elementary reactions, the order coincides with the molecularity (for
AIAA SP-084-1999
xii
H
2
+ I
2
2HI the forward reaction rate is second order, the R
forward
{Conc. H
2
}
m
x {Conc. I
2
}
k
where
m=1 and k=1. The overall order of reaction n = m + k = 2 for this example.
Oxidizer: Primarily air, oxygen, halogens, the hypergolics (N
2
O
4
or fuming nitric acid) or any material that
undergoes a reduction in chemical terms (will readily accept electrons from the fuel).
Permissible Exposure Limit (PEL): The degree of exposure of workers to hazardous gases and vapors
is regulated by the Occupational Safety and Health Administration (OSHA). The regulation specifies
permissible exposure limits (PELs) that for hydrazine includes a ceiling limit (employee exposure not to be
exceeded during any part of the work day) of 1.0 ppm (1.3 mg/m
3
) and a skin designation that requires
the use of protective clothing and equipment. For more information, see Section 5.1.2.
Pool Fire: Used to describe the sustained burning of a pool of liquid fuel. The rate of burning, called
burn rate, is measured in terms of depth change/time or mass consumed/time.
Shock: A violent collision or impact and the subsequent transmission of energy through the system. The
energy moves as a wave at velocities greater than the speed of sound relative to the undisturbed
material.
Spacecraft Maximum Acceptable Concentration (SMAC): The maximum allowable concentration for
spacecraft applications, typically based on standards and requirements for ground applications. For
specific SMACs see Section 5.3.5.
Temperature Sensitivity: A measure of the response of reaction rate to temperature, E
a
/RT
2
(E
a
= activation energy; R = gas constant; T = temperature).
Thermal Runaway: Operation of a system that contains material which reacts exothermically, in a
manner that the rate of thermal energy generation exceeds the rate at which thermal energy is transferred
to the surroundings. The system temperature increase can lead to an increase in this imbalance and
then lead to autoignition of the reaction.
Threshold Limiting Value: The average concentration of toxic gas to which most workers can be
exposed during working hours (8 hours per day, 5 days per week) for prolonged periods without
adversely affecting their health.
Total Mass Burning Rate: The rate of total mass of reactant consumed (kg/s) in a burning system.
Toxic: Poisonous. A material that causes physiological damage to the body.
Ullage: The vapor space above the liquid surface within the system.
Vapor: A component in the gas phase that is in equilibrium with its corresponding liquid phase. The
temperature of the system must be below the critical temperature of the component so it can exist as a
liquid phase.
AIAA SP-084-1999
xiii
Trademarks
The following commercial products that require trademark designation are mentioned in this document.
This information is given for the convenience of users of this Special Report and does not constitute an
endorsement. Equivalent products may be used if they can be shown to lead to the same results.
Apiezon
Autospot
Cabot Alloy
Chemglaze
Chemturion
Ecolyzer
Freon
GASBADGE
Grafoil
Graphitar
Hastelloy
Haynes
Inconel X-750
Inconel
Interscan
Kalrez
Kel-F
Koropon
Krytox
Kynar
MACOR
Microseal
Mobil Jet Oil II
Monitox
Mykroy/Mycalex
Mylar
Permendur
Pyromet
Ryton
SCAPE
SCAN Kit
Splash uit
Stoody 6
Stycast
Teflon
Tefzel
Waspaloy
AIAA SP-084-1999
1
1 Introduction to hazard assessment
Hydrazine (HZ), N
2
H
4
, is used as an aerospace fuel, an antioxidant in industrial processes, and in the
production of pesticides and pharmaceuticals, to name just a few applications. The hazards associated
with the use of vapor and liquid hydrazine in aerospace systems are the focus of this special report.
1.1 About this special report
This special report presents information that designers, builders, and users of hydrazine systems can
use to avoid or resolve hydrazine hazards. Pertinent research is summarized, and the data is presented
herein as for concise quick-reference resource. Additional information can be found in the sources cited
throughout the special report. An example of a hazard analysis is provided in Annex A. The chemical,
physical, and thermodynamic properties of hydrazine are provided in Annex B.
Readers are cautioned that, although every reasonable effort has been made to
present accurate information, the authors and publisher make no warranty nor do they
assume legal responsibility for its validity. Readers are urged to assess each situation
carefully and to choose data that appear most appropriate.
Throughout this special report, the following conventions are used:
ambient pressure and temperature refer to a pressure of 101.3 kPa (14.7 psia) and temperature
of 298 K (77 F),
neat hydrazine is hydrazine that is free from adulteration,
high-purity hydrazine contains a minimum of 99.0 percent by weight hydrazine and less than or
equal to 0.005 percent by weight aniline,
monopropellant-grade hydrazine contains a minimum of 98.5 percent by weight hydrazine, and

data are given in SI (International System) units with US customary units in parentheses.
1.2 Approach to performing a hazard assessment
Section 1.2 presents guidelines for determining if a particular hazard exists, using the information
presented in this special report. A diagram depicting the overall hazards associated with hydrazine is
presented in Figure 1.
As shown in Figure 1, four categories of potential hazards are discussed in this special report, and
considered in assessing the overall hazards of hydrazine: fire, explosion, material compatibility, and
safety/exposure. For clarity, two of the potential hazards, explosion and material compatibility, have
been subdivided. Explosion has been divided into the three major processes that generate the pressure
that can lead to explosions in hydrazine systems. These processes are deflagration, detonation, and
thermal-chemical processes. Material Compatibility has been divided into two sections entitled "Effect of
Hydrazine" and "Effect of Material." The "Effect of Hydrazine" section describes how hydrazine degrades
materials by altering their physical properties. The "Effect of Material" section describes how materials
can accelerate the rate of hydrazine decomposition. Listed under each category of potential hazard in
Figure 1 are criteria that define when the potential hazard exists, and the effects produced by that
potential hazard.
AIAA SP-084-1999
2
The criteria for, and the effects produced, by each potential hazard are presented in more detail in
Sections 1.2.1 through 1.2.4. These sections should be read carefully to gain familiarity with the four
categories of potential hazards and the terminology that is used in assessing these hazards.
Sections 1.2.1 through 1.2.4 also specify information about the system or environment that is needed to
assess each potential hazard. With this information, the following approach can be utilized to establish
the overall hazard that exists when hydrazine is used.
Consider each of the four categories of potential hazards independently, using the
"Criteria" and the "Effects Produced" information contained in Sections 1.2.1 through
1.2.4.
Determine if the system state, environmental conditions, or both meet criteria resulting in a
hazardous situation.
The conditions that lead to a potential hazard are given under each Criteria. Further information on the
criteria of a potential hazard is listed under a corresponding section of this special report.
Figure 1 Overall hazards diagram
AIAA SP-084-1999
3
For example, a flammable mixture is one criterion of a fire. When assessing a potential fire
hazard, it is necessary to refer to the detailed information on flammability in Section 2.1.1 so that
the existence of a flammable mixture can be determined. Because a flammable mixture exists
does not necessarily signify that a hazardous condition, i.e. a fire will be produced.
Evaluate the potential effects in each appropriate hazard category. The effects of each
hazard are listed under the corresponding "Effects Produced" heading.
For example, once it has been shown that a potential fire hazard exists, the effect (heat,
in this case) produced by the fire must be evaluated. Section 2 presents information on
flame velocities and heats of reactions, which can be used to determine the rate and
duration of heat release.
Perform the overall assessment of all potential hazards by considering individual system
hazards and the hazards that occur through interaction between systems.
An example of a detailed hazard assessment is given in Annex A.
1.2.1 Fire hazards
To assess a possible fire hazard, the following information about the system and environment must be
known:
phase(s), pressure, and temperature of hydrazine,
environment (e.g., hydrazine, diluents, and oxidizers) and corresponding concentrations, and
ignition sources and amount of energy that can be released by each.
1.2.1.1 Criteria for a potential fire hazard
The criteria used to determine if there is a potential fire hazard from hydrazine vapor are: flammability
limits, minimum ignition energy, and autoignition temperature. The lower and upper flammability limits
(LFL and UFL) specify the minimum and maximum hydrazine vapor concentrations that will ignite and
propagate a flame. The flammability limits are a function of temperature, pressure, and other factors
(Section 2.1.1). The minimum ignition energy (MIE), the energy required to initiate a fire (Section 2.1.1),
is a function of temperature, pressure, and oxidizer content. The autoignition temperature (AIT) is the
temperature in a system at which the entire volume of gases or vapors spontaneously ignites
(Section 2.1.2). The AIT data for hydrazine is highly variable, system specific, and should be used with
caution. When the concentration of hydrazine vapor is within the flammability limits and the MIE is
present or when the temperature of the hydrazine vapor is at the AIT, a potential fire hazard exists.
When oxidizer vapors or diluent gases are present, the potential fire hazard from hydrazine liquid is
generally evaluated by using flash and fire points and MIE. The flash point of liquid hydrazine is the
temperature at which the vapor above the liquid forms an ignitable mixture with air. The fire point is the
temperature at which the vapor above the liquid can continuously support a flame. The fire point
normally occurs at a higher temperature than the flash point. Flash and fire points can change,
depending on the nature of the vapor above the liquid (e.g., the presence of diluents and oxidizers). If
the temperature of the liquid hydrazine is at or above the flash or fire point, and the MIE is present, a
potential fire hazard exists.
AIAA SP-084-1999
4
1.2.1.2 Effects produced by fire
The hazards from a fire are determined by assessing the effects produced by the fire on the system or
environment.
For the purposes of this special report, the primary effect of a fire is heat. A fire is characterized by the
rate and duration of the heat release. In the vapor phase, the rate and duration of the heat release can
be calculated using the flame velocity, the amount of hydrazine present, and the heat of reaction
(Section 2.1.3.4). In the liquid phase, the rate and duration can be calculated using the surface area of
the liquid pool, the heat of reaction, and the burning rate.
A fire hazard exists when an evaluation of the specific system or its environment shows the heat release
is sufficient to cause dangerous or undesirable conditions.
1.2.2 Explosion hazards
In this special report, it is assumed that an explosive event (see glossary) is produced either directly or
indirectly by the release of chemical energy. Explosions can arise from three processes that hydrazine
can undergo, and in some cases, generate high pressures. These are: deflagration, detonation, and
thermal-chemical processes.
1.2.2.1 Deflagration
A deflagration is a flame moving through a flammable mixture in the form of a subsonic wave (with
respect to the unburned mixture).
To assess an explosion hazard due to deflagration, the following information about the system and
environment must be known:
phase(s), pressure, and temperature of hydrazine,
environment (e.g., hydrazine, diluents, and oxidizers) and corresponding concentrations,
ignition sources and amount of energy released by the source,
mechanical properties of a confining system, and
presence of obstacles that can accelerate the deflagration.
1.2.2.1.1 Criteria for a potential explosion hazard from a deflagration
The flammability criteria for fire also apply to deflagrations. An explosion hazard exists when hydrazine
vapor is within the flammability limits, the MIE is present, and obstacles, turbulence, or confinement
necessary to accelerate the deflagration are present or when the deflagration is confined in a system
and the pressure due to the deflagration can exceed the burst pressure of the system.
When hydrazine liquid is present, a potential explosion hazard from a deflagration exists when the
temperature of the liquid hydrazine is at or above the fire point and the MIE is present.
AIAA SP-084-1999
5
1.2.2.1.2 Effects produced by a deflagration
The primary effects of a deflagration are fire and if the deflagration is confined, pressure. Calculation of
deflagration pressures in partially confined systems is very difficult. In a confined system the pressure
generated by the hydrazine (calculated from Equation 5) is compared to the burst pressure of the
system. If the maximum pressure which can be generated exceeds the burst pressure, an explosion
hazard exists. The venting rate of relief valves or burst disks must be great enough to prevent the burst
pressure of a system from being reached. The pressure increase rate can be calculated with
Equations 9 and 10.
The hazard associated with the heat generated by a deflagration is covered in Section 1.2.1. The hazard
resulting from a deflagration transition to a detonation is covered in Section 1.2.2.2.
1.2.2.2 Detonation
To assess hazards due to detonation (see Glossary), the following information about the system and
environment must be known:
pressure and temperature of hydrazine vapor,*
environment (e.g., hydrazine, diluents, and oxidizers) and the corresponding concentrations,
ignition sources and amount of energy released by the source, and
type of confinement.
1.2.2.2.1 Criteria for a potential explosion hazard from a detonation
The general criteria for detonation is that the combination of the system and the dynamics of the
deflagration (flame movement) lead to a deflagration to detonation transition (DDT) (Section 3.2). The
criteria that must be met for a detonation to occur are related to the composition and state of the
flammable material in the system, the size and shape of the vessel or enclosure, and the energy source
that initiates the detonation. The dynamic parameters, critical tube diameter, critical transition diameter,
and initiation energy are based on cell size which is determined from experiment or empirical correlation.
For a detonation to occur, there must be a sufficient volume for a detonation cell to form and sufficient
initiation energy or a flow path that will support a DDT (Equations 11 through 16). To use these
equations, the cell size, which varies with the composition, initial temperature, and initial pressure of the
hydrazine mixture, must be known.
1.2.2.2.2 Effects produced by a detonation
To determine if a detonation hazard exists, the resulting effects produced by the detonation must be
assessed.
The pressure produced by a detonation is described by the Chapman-Jouguet (C-J) condition
(Section 3.2) and is a characteristic of the reactant mixture. Computer programs have been developed
to calculate the C-J condition.

*Hydrazine vapor is readily detonated. Hydrazine liquid has not been detonated in cylindrical vessels of
diameters up to 10.2 cm (4 in.). The ignition energy was provided by approximately 887 g (1.96 lb) of
C4
initiated by a #8 blasters cap.
AIAA SP-084-1999
6
When a DDT occurs, the final detonation pressure (C-J pressure) is dependent on the extent of the
deflagration before transition. For example, consider neat hydrazine vapor contained in a system at
50 kPa (7.3 psia) and 300 K (80 F). If the vapor undergoes a DDT immediately after ignition, then the
detonation pressure (C-J pressure) will be approximately 28 times the initial pressure. If the DDT is
delayed, and the pressure in the system increases due to the deflagration, then the C-J pressure will be
28 times the pressure at the time of the DDT. A worst-case detonation hazard occurs when the
deflagration has proceeded almost to completion before the transition takes place. This case can be
estimated by using the adiabatic pressure (see Section 3.1.1) produced by the deflagration as the initial
system pressure in C-J calculations. Pressures produced when a detonation wave is reflected at
locations such as elbows and tees can be two to three times the incident pressure (i.e., the C-J
pressure).
1.2.2.3 Thermal-chemical processes
Any thermal-chemical reaction system is a potential explosion hazard if the reaction produces a net
increase in moles of gas or vapor (Section 3.1) when the system volume is confined, if the system
operates adiabatically (or nearly adiabatically), or when the rate of heat generated exceeds the rate of
heat exchange with the environment leading to a thermal runaway. Hydrazine is thermodynamically
unstable and while its degradation is slow under ambient conditions increasing temperature accelerates
this decomposition. In addition many materials act as catalysts for hydrazine decomposition and careful
selection of system materials must occur (Section 4).
Three cases that may cause explosion hazards when operating hydrazine systems are considered:
near-isothermal decomposition in a closed system; thermal runaway in an isolated system; rapid
compression in an open system.
To assess hazards due to a thermal-chemical reaction, the following information about the system and
environment must be known:
phase(s), pressure, and temperature of the hydrazine,
environment (e.g., hydrazine, diluents, and oxidizers) and corresponding concentrations,
materials in contact with hydrazine and corresponding surface areas,
system thermal capacity and heat transfer properties, and
pressure limitations of the system.
1.2.2.3.1 Near-isothermal hydrazine decomposition
If the enthalpy of reaction of hydrazine decomposition (Annex B) is dissipated to the environment as it is
generated, the system operates to maintain the temperature constant.
1.2.2.3.1.1 Criteria for hazard from near-isothermal hydrazine decomposition
If the system is a rigid volume and closed to addition or removal of mass, the pressure will increase as
the reaction proceeds to completion.
1.2.2.3.1.2 Effects produced by near-isothermal hydrazine decomposition
Pressure increases as the moles of decomposition product gases increase. For isothermal processes,
the pressure increase can be approximately calculated using the following:
AIAA SP-084-1999
7
Ideal Gas Law,
quantity of hydrazine and diluents,
kinetic parameters given in Table 17 or the figures of Section 4,
surface area of the material on which the hydrazine is decomposing if catalysis is occurring,
volume of system, and
initial conditions.
Assessment examples are given in Section 4.3.
1.2.2.3.2 Thermal runaway
If the system operates non-isothermally the temperature can reach a condition where heat is generated
faster than it can be dissipated to the environment and the system undergoes a thermal runaway.
1.2.2.3.2.1 Criteria for thermal runaway hazards
The system temperature increases continuously (unless limited by quantity of hydrazine) as long as the
rate of energy generation exceeds the rate of heat exchanged with the environment. If the system is
closed and a constant volume the pressure increases.
1.2.2.3.2.2 Effects produced by a thermal runaway
For a non-isothermal process, the pressure increase rate can be determined by solving a series of
nonlinear, differential equations. See Section 4.3 for examples. Under non-isothermal conditions, the
system will fail if sufficient hydrazine is present to decompose and produce pressures that exceed the
system burst pressure.
1.2.2.3.3 Rapid compression
Rapid compression occurs in a system when accelerating liquid compresses ullage gases or vapors in a
confined volume. The rapid temperature rise of the compressed material acts as a possible ignition
source for the hydrazine. If a froth is created at the liquid hydrazine ullage gas or vapor interface the
rate of thermal energy transfer into the hydrazine can increase producing a greater potential hazard.
Pressure increases as a result of compression, temperature rise and hydrazine decomposition product
gases. If system pressure exceeds design burst pressure, a failure resulting in potential hazard can
occur.
To assess an explosion hazard due to rapid compression, the following information about the system or
environment must be known:
initial temperatures of flowing hydrazine liquid and ullage gas,
pressure and volume of the ullage gas,
location of system dead-ends,
hydrodynamic surge pressure at the system dead-ends, and
burst pressure of the system.
AIAA SP-084-1999
8
1.2.2.3.3.1 Criteria for a hazard from rapid compression
The criteria for a hazard from rapid compression are the presence of liquid hydrazine, a dynamic
system,* a non-condensable gas ullage, a froth (see Section 3.3.3.1), and liquid confinement. The
initiation mechanism of this process is the enthalpy change of the ullage material during adiabatic
compression. The condition (i.e. temperature and pressure) of the froth is a function of the dynamic
surge pressure created when the moving liquid hydrazine impacts a dead-end. When the ullage
temperature starts at ambient and the fluid dynamic surge pressure is greater than 17.2 MPa (2500
psia), sufficient enthalpy is created to initiate exothermic decomposition of the hydrazine contained in the
froth. The froth and fluid dynamic surge pressure are both dependent upon the initial pressure and
volume of the ullage. The fluid dynamic surge pressure is also strongly dependent on the velocity of the
hydrazine liquid. Rapid compression which meets the above criteria will generate pressure, and thus a
potential explosion hazard will exist.
1.2.2.3.3.2 Effects produced by rapid compression
A hazard from rapid compression exists when the generated pressure exceeds the strength of the
confining system.
If the fluid dynamic surge pressure is known, the final generated pressure can be estimated from
Figure 21 and Tables 20 and 21. The fluid dynamic surge pressure can be estimated by applying
Equations 18 through 22, or by substituting water for hydrazine and performing experiments with the
system.
It should be noted that a hazard can also exist when the fluid dynamic surge pressure alone exceeds the
strength of the system.
1.2.3 Material compatibility
Hydrazine affects materials by degrading them and altering their physical and chemical properties.
Materials affect hydrazine by accelerating its decomposition. Section 4 of this special report provides
additional information on compatibility.
1.2.3.1 Effect of hydrazine
To assess a material compatibility hazard due to hydrazine contacting a material, the following
information about the system and environment must be known:
phase(s) and temperature of hydrazine,
exposure time,
surface condition and area exposed to hydrazine, and
contamination present.
1.2.3.1.1 Criteria for a potential material compatibility hazard from materials exposed to
hydrazine
The criteria for hydrazine's ability to affect a material (Section 4.1) are the length of time the material will

*
Explosive events caused by rapid compression have only been observed when the liquid hydrazine was
flowing.
AIAA SP-084-1999
9
be exposed to hydrazine, the temperature of the hydrazine in contact with the material, the surface
condition and area exposed to hydrazine, and the contamination present. Tables 14 and 15 for metals
and nonmetals, respectively, indicate the amount of corrosion and changes observed for various
materials exposed to hydrazine at a specified temperature. These corrosion rates may increase when
operating temperatures higher than those listed in the tables are used.
1.2.3.1.2 Effects produced on materials exposed to hydrazine
The physical and chemical properties of materials can be changed by exposure to hydrazine leading to
corrosion (loss of material), stress corrosion cracking (fracture), and embrittlement. A material
compatibility hazard exists when degradation of the material can lead to loss of system integrity or
component function.

1.2.3.2 Effect of material
Materials can act as a catalyst and accelerate the decomposition rate of hydrazine and thus the pressure
increase rate. Therefore, to assess the presence of a material compatibility hazard, the following
information about the system and environment must be known:
phase(s), quantity, and temperature of hydrazine,
environment (e.g., hydrazine, diluents, contaminants, and oxidizers) and corresponding
concentrations,
surface condition and area of the material,
system thermal capacity and heat transfer properties, and
pressure limitations of the system.
1.2.3.2.1 Criteria for a potential material compatibility hazard from hydrazine exposed to catalytic
materials
A potential material compatibility hazard exists when the material catalytically decomposes hydrazine.
Material compatibility data are presented in Section 4.
1.2.3.2.2 Effects produced in hydrazine exposed to catalytic materials
A material compatibility hazard exists when the hydrazine decomposition reaction is accelerated to the
point that pressures generated affect the operation of or damage the system or environment. (The
methods used to determine these pressures are presented in Section 1.2.2.3.) An actual material
compatibility hazard can lead to an explosion hazard if the generated pressures exceed the strength of
the confining system.
1.2.4 Exposure hazards
Exposure hazards can be divided into two categories: hazards to personnel (toxicity), and hazards to the
environment. Unlike the previous section, the criteria listed here do not apply to a potential hazard. If
these criteria are met, an actual exposure hazard exists.
To assess an exposure hazard, the following information about the environment must be known:
phase(s) of hydrazine,
AIAA SP-084-1999
10
temperature of the environment,
environment (e.g., hydrazine, diluents, and oxidizers) and corresponding concentrations, and
time of contact.
1.2.4.1 Criteria for an actual exposure hazard
The criteria for a toxicity hazard are exposure time and concentration. Personnel exposure to hydrazine
can be chronic (long-term exposure, Section 5.1.1) or acute (short-term exposure, Section 5.1.2.). A
toxicity hazard is present when personnel can be exposed to hydrazine concentrations which exceed the
Threshold Limit Value (TLV), Permissible Exposure Limits (PEL), or Recommended Exposure Limit
(REL) for chronic exposure (Section 5.1.1.2); or, Immediately Dangerous To Life or Health (IDLH) or
Spacecraft Maximum Acceptable Concentration (SMAC) for acute exposure (Section 5.1.2.2).
State and Federal regulations specify the levels of hydrazine that pose an exposure hazard to the
environment. For more information, refer to Section 5.
1.2.4.2 Resulting effects from exposure to hydrazine
The effects of a toxicity hazard and an environmental exposure hazard are listed in Section 5.
1.3 Overall hazard
The overall hazard to a system and environment may be intensified by a combination of several
individual events.
An example of the cumulative effect hazards can have was illustrated by an incident involving the ninth
Space Transportation System (STS-9). In this particular situation, a material compatibility hazard led to
stress corrosion of an injector tube constructed from Hasteloy B. The tube subsequently ruptured,
leaking hydrazine into the APU area, where it quickly froze. During re-entry, the hydrazine vaporized
and ignited. By itself, the heat from the burning hydrazine may have caused system damage; however,
the critical effect of the fire was to overheat the onboard APU's. When this happened, the hydrazine
underwent several explosive processes, damaging two of the three APU's.
Therefore, a proper hazard assessment considers not only each hazard and topic individually but the
potential effects to the system if several events occurred simultaneously.
AIAA SP-084-1999
11
2 Fire
Introduction
Fire is a rapid chemical reaction that produces heat and light (Lapedes 1974). Fire normally requires a
fuel, an ignition source, and an oxidizer, however, hydrazine is a monopropellant which decomposes
exothermically and does not require the presence of an oxidizer. A hydrazine fire can also begin without
the usual ignition source; hydrazine is hypergolic with oxidizing reagents and propellant oxidizers such
as inhibited red fuming nitric acid (IRFNA) and dinitrogen tetroxide.
A fire can occur with either vapor or liquid hydrazine, as well as in the mist, droplet, and spray forms of
the fuel. The fire hazard from the burning hydrazine is affected by system and environmental conditions,
i.e. the temperature, concentration, and pressure of the hydrazine, and the type of ignition source
involved. Hazard assessment varies depending on the form of hydrazine present.
In hydrazine vapor, the fire hazard can be quantified by considering the flammability limits, ignition
source, flame speed or flame velocity. In liquid hydrazine, fire points and burning rate can be used to
assess the degree of fire hazard. The fire hazard in hydrazine mists, droplets, or sprays can also be
measured by considering flash and fire points and burning rates. These factors vary with the purity of
hydrazine vapor (neat or completely free of contaminants, mixed with air, mixed with an inert diluent, or
mixed with an oxidizer) or liquid (high purity, monopropellant, or mixtures).
This section presents data on hydrazine fires. Readers are cautioned that, although every reasonable
effort has been made to present accurate information in this section, the authors and publisher make no
warranty nor do they assume legal responsibility for its correctness. Readers are urged to assess each
situation carefully and to choose data that appear most appropriate.
2.1 Hydrazine vapor
The reactants in a hydrazine fire are usually in the gaseous phase and must be present within a specific
concentration range to burn. The fire hazard of hydrazine vapor can be assessed by considering
flammability, ignition, flame speed and flame velocity in neat vapor, hydrazine-air mixtures, hydrazine-
diluent mixtures, and hydrazine-oxidizer mixtures.
2.1.1 Flammability
Flammability is a measure of the extent to which a vapor concentration of a fuel in a mixture will ignite
and propagate a flame. The limits of flammability of a gas or vapor are the minimum and maximum fuel
concentrations that can support flame propagation. The upper flammability limit (UFL) is the
concentration of the most concentrated mixture that is flammable; the lower flammability limit (LFL) is the
concentration of the most dilute fuel-air or fuel-diluent mixture that is flammable (see Glossary).
Flammability varies for neat hydrazine, hydrazine mixed with air, hydrazine mixed with an inert diluent,
and hydrazine mixed with an oxidizer. The flammability limits for each mixture are affected by pressure,
temperature, and other factors. The range of flammability is the range of concentrations between the
lower and upper flammability limits. In general for many flammable mixtures the flammability range is
widened by increasing temperature (Burgess and Wheeler 1911). There is limited data for hydrazine. In
general, decreased pressure (below ambient) narrows the flammability range by increasing the lower
flammability limit and decreasing the upper flammability limit (Coward and Jones 1952). As the pressure
decreases, the two limits approach each other. When the upper and lower limits are identical, the low
pressure limit (the minimum pressure required for ignition) is reached. The low pressure limit is affected
by ignition energy: as the ignition energy is increased, the low pressure limit decreases (Benz, Bishop,
and Pedley 1988).
AIAA SP-084-1999
12
Flammability limits also vary with the shape of the confining container and the direction of flame
propagation. Generally for many flammable mixtures, the flammability range widens with increasing tube
diameter (Coward and Jones 1952), and for closed tubes, the flammability range may narrow as the tube
is lengthened (Mullins and Penner 1959). The direction of flame propagation affects flammability limits
by widening the limits for upward propagation and narrowing the limits for downward propagation. The
limits for horizontal propagation are between those for upward and downward propagation (Benz,
Bishop, and Pedley 1988).
2.1.1.1 Neat hydrazine
2.1.1.1.1 Pressure effect
Neat hydrazine is flammable at pressures of 2.1 kPa (0.30 psia) and greater when ignited by a 1-J (9.5 x
10
-4
-Btu) spark (Benz, Guerrasio, and Rollins 1983). Increased ignition energy can lead to a decreased
low pressure limit (Section 2.1.1).
2.1.1.1.2 Temperature effect
The lower pressure limit of 2.1 kPa (0.30 psia) requires a minimum temperature of 300 K (81 F), based
on vapor pressure equations cited by Benz, Bishop, and Pedley (1988).
2.1.1.2 Hydrazine-air mixtures
The upper flammability limit of hydrazine-air mixtures is 100 percent hydrazine (Benz, Bishop, and
Pedley 1988). The most commonly quoted lower flammability limit for hydrazine/air mixtures is 4.7
percent (Scott, Burns, and Lewis 1949). A lower value, 2.9 percent, was reported by Olin Corporation
(National Fire Protection Association 1977). A summary of lower flammability limits obtained using
various methods is presented in Table 1.
Table 1 Lower flammability limits for hydrazine-air mixtures
Lower flammability Direction
of propagation Temperature Limit
K F percent v/v
Unknown
a
2.9
Upward
b
373 212 4.7
Omni-directional
c
366 200 5.0
a
National Fire Protection Association (1977)
b
Scott, Burns, and Lewis (1949)
c
Benz, Guerrasio, and Rollins (1983)
2.1.1.2.1 Pressure effect
The effect of reduced pressure on the flammability of hydrazine-air mixtures in a 2-L (122 in.
3
) spherical
chamber (omnidirectional propagation) at 311 K (100 F) and 366 K (200 F) is illustrated in Figure 2.
The lowest pressure that propagated a flame is 1.4 kPa (0.2 psia) at a hydrazine concentration of 35
percent by volume (twice stoichiometry). The ignition source in these tests was a 1 J (9.5 x 10
-4
Btu)
electrical spark. The low pressure limit is expected to decrease with increasing spark energy
(Section 2.1.1). Figure 2 shows a plot of hydrazine concentration (percent by volume) versus limiting
pressure and the corresponding altitude.
AIAA SP-084-1999
13
Figure 2 The effect of pressure on the flammability of hydrazine in air at 311 K (100 F) and 366 K (200 F) (Data
from Benz, Guerrasio, and Rollins 1983)
2.1.1.2.2 Temperature effect
The effect of temperature on the flammability of hydrazine-air mixtures at ambient pressure is small
(Figure 3). The lower flammability limit of hydrazine is expected to decrease as the temperature is
increased, but there are insufficient experimental data at present to confirm this.
The upper flammability limit is 100 percent for all temperatures above the flash point (Section 2.2.1).
Figure 3 Flammability limits for hydrazine vapor in air at ambient pressure as a function of temperature and
composition
The flammability limits for hydrazine vapor-air mixtures at various pressures and temperatures are
shown superimposed in heavy lines over curves of constant hydrazine concentration at changing total
pressure in Figure 4. The data in this figure indicate the flammable nature of the hydrazine vapor that
exists above liquid hydrazine at various conditions of temperature and pressure.
2.1.1.3 Hydrazine-diluent mixtures
The upper flammability limit of hydrazine-diluent mixtures is 100 percent (Benz, Bishop, and Pedley
1988). The lower flammability limits for various hydrazine-diluent mixtures are shown in Table 2. The
0 1 0 2 0 3 0 4 0 5 0 6 0 7 0 8 0 9 0 1 0 0
H yd r a z i n e C o n c e n t r a t i o n ( % v / v )
1
1 0
1 0 0
P
r
e
s
s
u
r
e

(
k
P
a
)
S
t
a
n
d
a
r
d

A
l
t
i
t
u
d
e

(
k
m
)
3 1 1 K
3 6 6 K
F l a m m a b l e R e g i o n
2 5
2 0
1 5
1 0
5
0
250 275 300 325 350 375 400 425
Temperature (K)
0
10
20
30
40
50
60
70
80
90
100
H
y
d
r
a
z
i
n
e

C
o
n
c
e
n
t
r
a
t
i
o
n

(
%

v
/
v
)

U.S. Air Force, 1973
Benz, Bishop, and Pedley 1988
Scott, Burns, and Lewis 1949
LFL
Flammable
Region
Saturated
Vapor
Mists
UFL
311
Closed Cup
Flash Point
AIAA SP-084-1999
14
lower flammability limits observed in these hydrazine-diluent mixtures are significantly greater than those
observed for hydrazine in air and other oxidizing atmospheres.
Figure 4 Flammability limits for hydrazine in air as a function of temperature and pressure (Data from Benz,
Bishop, and Pedley 1988)
Table 2 Lower flammability limits for hydrazine-diluent mixtures at a pressure of 101.3 kPa (14.7 psia)
Temperature Diluent
K F
Lower
flammability
limit
percent v/v
Argon
a
373-393 212-248 28.1
Benzene
a
373-393 212-248 72.1
Benzene
b
398 257 60.2
Butane
a
373-393 212-248 82.7
Cumene
b
398 257 76.2
Helium
c
378-391 221-244 37.0
n-Heptane
a
373-393 212-248 88.0
n-Heptane
b
398 257 79.0
n-Heptane
c
377-406 219-271 86.8
Hexane
a
373-393 212-248 87.2
Hydrogen
a
373-393 212-248 54.1
Nitrogen
a
373-393 212-248 35.2
Nitrogen
c
382-385 228-234 38.0
Toluene
a
373-393 212-248 76.0
Toluene
b
398 257 65.0
Water Vapor
c
403-408 265-275 30.9
Xylene
a
373-393 212-248 80.0
m-Xylene
b
398 257 72.7
a
Pannetier (1958)
b
Furno, Martindill, and Zabetakis (1962)
c
Scott, Burns, and Lewis (1949)
Diluents such as nitrogen, helium, and carbon dioxide generally reduce the flammability of a vapor. The
effectiveness of a diluent at quenching flames depends on the heat capacity and thermal conductivity of
270 280 290 300 310 320 330 340 350 360 370 380 390
Liquid Temperature (K)
0
10
20
30
40
50
60
70
80
90
100
T
o
t
a
l

P
r
e
s
s
u
r
e

(
k
P
a
)

[Benz, Guerrasso, and Rollins 1983]
Stoichiometric Mixture
Vapor Pressure 70% 50% 30% 10% 5%
LFL
UFL
Stoichiometry
Flammable Region
AIAA SP-084-1999
15
the diluent. Diluents with large heat capacities generally have greater quenching ability than those with
low heat capacities (Lewis and von Elbe 1961). Hydrocarbon vapors of benzene, xylene, butane, and
hexane mixed in hydrazine vapors were found to be effective at quenching the initiation of flames in the
mixture by electric sparks (Furno, Martindill, and Zabetakis 1962). Note that these hydrocarbons have
upper flammability limits below 10 percent by volume, much lower than the UFL for hydrazine, and large
heat capacities.
The flammability limits for mixtures of hydrazine, heptane, and air at ambient pressure and 398 K (257
F) are shown in Figure 5. Mixtures containing heptane in concentrations greater than 21 percent by
volume are nonflammable. In contrast, mixtures of hydrazine, oxygen, and nitrogen containing up to
86 percent nitrogen by volume are flammable (Figure 6).
2.1.1.3.1 Pressure effect
Although the low pressure limit is expected to increase with increasing concentration of diluent, no
supporting data are reported in the reviewed literature on the effect of pressure on the flammability of
hydrazine-diluent mixtures.
2.1.1.3.2 Temperature effect
The results in Table 2 (Pannetier 1958) indicate that the LFL for hydrazine does not change between
373 K (212 F) and 393 K (248 F) for certain diluents. No other data are reported in the reviewed
literature on the effect of temperature on the flammability of hydrazine-diluent mixtures.
Figure 5 Flammability limits as functions of composition for mixtures of hydrazine, heptane, and air at 398 K (257
F) and ambient pressure (Furno, Martindill, and Zabetakis 1962)
AIAA SP-084-1999
16
Figure 6 Flammability limits for ternary mixtures of hydrazine, oxygen, and nitrogen at 366 K (200 F) to 398 K
(257 F) and ambient pressure (Benz, Bishop, and Pedley 1988)
2.1.1.4 Hydrazine-oxidizer mixtures
The upper flammability limit for hydrazine-oxygen mixtures is 100 percent hydrazine (Benz, Bishop, and
Pedley 1988). The only reported value of the lower flammability limit for hydrazine-oxygen mixtures at
311 K (100 F) and 101.3 kPa (14.7 psia) is 4.9 (Benz, Bishop, and Pedley 1988). This value is similar
to the lower flammability limit, as determined by the same method, reported for hydrazine-air mixtures
(Section 2.1.1.2).
The effects of oxygen/nitrogen ratio are illustrated in Figure 6 which displays the flammability limits of
hydrazine-oxygen-nitrogen mixtures at ambient pressure and temperatures which varied from 366 K
(200 F) to 398 K (257 F). The experimental data used to generate Figure 6 are very limited, and this
figure should be considered only an approximation.
The effects of oxidizers other than oxygen were measured by Gray and Lee (1954). They found that the
lower flammability limit of hydrazine in both nitrous oxide and nitric oxide at ambient pressure and
temperature is 10 percent by volume. Hydrazine is hypergolic with dinitrogen tetroxide at ambient
pressure and temperature.
2.1.1.4.1 Pressure effect
No data are reported in the reviewed literature on the effect of pressure on the flammability of hydrazine-
oxidizer mixtures.
AIAA SP-084-1999
17
2.1.1.4.2 Temperature effect
No data are reported in the reviewed literature on the effect of temperature on the flammability of
hydrazine-oxidizer mixtures.
2.1.2 Ignition
Hydrazine vapor can be ignited with or without an ignition source. Ignition by a source requires that the
minimum ignition energy (MIE) be present; autoignition (ignition with no source) requires that the
autoignition temperature (AIT) be reached. In general, the minimum ignition energy decreases with
increasing oxygen content (Lewis and Von Elbe 1961), increasing temperature (Drell and Belles 1958),
and increasing pressure (Benz, Bishop, and Pedley 1988). The standard method for determining the
minimum ignition energy is ASTM E 582-88.
2.1.2.1 Ignition sources
Various sources have been used to ignite hydrazine: electric sparks (Overly 1976); a 15 kV, 60 mA
transformer (Furno, Martindill, and Zabetakis 1962); and shock waves (Michel and Wagner 1965). Other
common ignition sources include flames, heated or fused wires, incendiaries, hot surfaces, and rapid
compression. Many of these ignition sources, such as hot surfaces, can be affected by surface area,
surface temperature, and type or volume of container or extent of confinement.
For information on avoiding ignition hazards during storage, refer to Section 5.3.4.
2.1.2.2 Autoignition
Autoignition occurs when a mixture of gases or vapors ignites spontaneously with no external ignition
source after reaching a certain temperature, the autoignition temperature (AIT). The AIT is not an
intrinsic property of the gases or vapors (Kanury 1977) but is the lowest temperature in a system where
the rate of heat evolved from the gases or vapors increases beyond the rate of heat loss to the
surroundings resulting in ignition. The AIT of a mixture of gases or vapors is affected by pressure,
vessel shape and volume, surface activity, contaminants, flow rate, reaction rate, droplet and mist
formation, gravity, and reactant concentration (Benz, Bishop, and Pedley 1988).
In general, decreased pressure leads to an increased autoignition temperature (Benz and Pippen 1980;
Bodurtha 1980; Furno, Imhof, and Kuchta 1968); increased vessel size leads to a decreased autoignition
temperature (Setchkin 1954). For fuels in general the AIT is not very sensitive to fuel concentration
except at near-limiting concentrations (Furno, Imhof, and Kuchta 1968), but some studies with hydrazine
show off-stoichiometric mixtures lead to increased AITs (Miller and Schluter 1978). The effect of
catalytic surfaces on autoignition temperatures varies with the system. Some studies show that catalytic
surfaces in some systems increase the autoignition temperature (Lewis and Von Elbe 1961; Miller and
Schluter 1978); others indicate that the reaction between the fuel and the surface material leads to a
decreased autoignition temperature (Scott, Burns, and Lewis 1949; Stevens and Benz 1978).
One study found that as hydrazine is heated, decomposition increases, hydrazine concentration
changes, and the final result is an increased autoignition temperature. In a flowing system where
concentration is constant, the effect of materials appears to be determined by the catalytic properties of
the materials (Miller and Schluter 1978).
Hydrazine concentration and the presence of diluents affect the autoignition temperature. Mixtures near
the flammability limits have higher autoignition temperatures than those of intermediate composition
(Bodurtha 1980). Fuel-oxygen mixtures have slightly lower autoignition temperatures than similar
concentration fuel-air mixtures (Bodurtha 1980). Mixing inert gases with fuel generally increases the
AIAA SP-084-1999
18
autoignition temperature by diluting or altering the thermal conductivity, specific heat, or diffusivity of the
mixture (Mullins and Penner 1959).
The standard methods for determining the autoignition temperature are detailed by ASTM method E
659-78(1984), Mullins (1955), and Setchkin (1954). The AIT is recorded as the lowest temperature at
which autoignition occurs for a fuel. The ASTM method E 659-78(1984) (ASTM 1986a), notes that the
method is not designed for evaluating materials capable of exothermic decomposition. For this reason,
autoignition temperature data should be applied cautiously (Benz, Bishop, and Pedley 1988).
2.1.2.1 Neat hydrazine
2.1.2.1.1 Minimum ignition energy
Data on the minimum ignition energy that is required from an electrical spark in neat hydrazine are
limited, in part because of the ease of ignition. Overly (1976) found a minimum energy electrode
distance (d
m
) of about 2 mm (0.079 in.) which corresponds to a minimum ignition energy of 0.8 mJ (7.6 x
10
-7
Btu) (Figure 7) at 338 K (149 F) and 14.7 kPa (2.2 psia).
2.1.2.1.2 Pressure and temperature effect
Overly (1976) found that when the temperature of neat hydrazine vapor increases from 338 K (149 F) to
345 K (162 F) and the pressure increases from 14.7 kPa (2.2 psia) to 20.7 kPa (3.0 psia), the ignition
energy at an electrode distance of 4 mm (0.16 in.) decreases from 1.3 mJ (1.2 x10
-6
Btu) to a value that
is too low to measure. This value corresponds to a minimum ignition energy of less than 0.05 mJ
(4.7 x 10
-8
Btu).
Figure 7 The effect of electrode distance on the minimum ignition energy of pure hydrazine vapor at 338 K (149
F). (Data from Overly 1976)
2.1.2.1.3 Autoignition temperature
As noted previously (Section 2.1.2), autoignition temperature data should be applied cautiously.
AIAA SP-084-1999
19
2.1.2.1.3.1 Pressure effect
Gray, Lee, and Spencer (1963) found that autoignition temperatures of neat hydrazine vapor in a
borosilicate container decrease with increasing pressure, although they are significantly greater than
autoignition temperatures of hydrazine-air mixtures. The lowest measured autoignition temperature was
853 K (1080 F) at the highest test pressure of 8.4 kPa (1.2 psia). No data are reported in the reviewed
literature on the effects of pressures ranging from 8.4 kPa (1.2 psia) to ambient on the autoignition
temperature of neat hydrazine.
2.1.2.1.3.2 Material effects
No data are reported in the reviewed literature on the effects of materials on the autoignition temperature
of neat hydrazine vapor. Refer to Section 2.1.2 for a discussion of the effect of catalytic surfaces on
autoignition temperature.
2.1.2.2 Hydrazine-air mixtures
2.1.2.2.1 Minimum ignition energy
2.1.2.2.1.1 Pressure effect
No data are reported in the reviewed literature on the effect of pressure on the minimum ignition energy
of hydrazine-air mixtures.
2.1.2.2.1.2 Temperature effect
No data are reported in the reviewed literature on the effect of temperature on the minimum ignition
energy of hydrazine-air mixtures.
2.1.2.2.2 Autoignition temperature
Autoignition temperatures reported in Table 3 for hydrazine-air mixtures range from 438 K (328 F) to
673 K (752 F). Tests have been performed in an attempt to narrow this range; however, they did not
improved the data in Table 3 (Benz, Guerassio, and Rollins 1983). As noted previously (Section 2.1.2),
autoignition temperature data should be applied cautiously.
2.1.2.2.2.1 Concentration effect
A study of the effect of hydrazine concentration at reduced pressures using ASTM D2155-66 Standard
Method of Test for Autoignition Temperature of Liquid Petroleum Products procedures was conducted
with a 2.845 L (173.6 in
3
) 304 stainless steel chamber (Stevens and Benz 1978).
Table 3 Autoignition temperatures of hydrazine-air mixtures at 101.3 kPa (14.7 psia)
Autoignition Temperature Reference
K F
513-543 464-518 Van Dolah, Zabetakis, and Scott (1962)
543 518 Scott, Burns, and Lewis (1949)
673 752 Clark (1953)
438 328 Roth (1964)
543 518 Int. Tech. Inf. Inst., Tokyo Japan (1981)
AIAA SP-084-1999
20
The lowest AIT for a pressure of 101.3 kPa (14.7 psia) occurred at a hydrazine:oxygen ratio of 3.4:1;
similarly, at
41.3 kPa (6 psia) a ratio of 5.2:1,
20.7 kPa (3 psia) a ratio of 3.1:1, and
6.9 kPa (1 psia) a ratio of 4.9:1.
Hydrazine to oxygen ratios differing from stoichiometric combustion (1:1 or 2:1) were explained as
hydrazine being consumed by decomposition. A similar study using ASTM D2155-66 procedures but
with a 0.25 L (15.3 in
3
) borosilicate chamber containing metal specimens found that 304 stainless steel,
316 stainless steel, borosilicate glass, and copper displayed distinct minimum AITs while aluminum
(6061-T6), titanium (Ti-6Al-4V), and nickel did not (Miller and Schulter 1978).
2.1.2.2.2.2 Pressure effect
Autoignition temperatures for hydrazine-air mixtures were found to decrease with increasing pressure
(Figure 8). Some of these tests were performed in chambers made of stainless steel, and reaction
between the stainless steel and hydrazine may have lowered autoignition temperatures, as discussed in
the following section.
2.1.2.2.2.3 Material effects
The results in Figure 8 indicate that stainless steel, acting as a heterogenous catalyst for hydrazine
decomposition in air, lowers the autoignition temperature. This is supported by reports by Scott, Burns,
and Lewis (1949) and Stevens and Benz (1980). Tests conducted by Miller and Schluter (1978) indicate
that heterogenous catalysts increase the autoignition temperature of hydrazine-air mixtures. The effects
of various surface materials are highly dependent on system configuration. The reported data are given
in Table 4. For additional information on material effects, refer to Section 4.
Figure 8 The effect of pressure on the autoignition temperature of hydrazine in air (
a
Data from Furno, Imhof, and
Kuchta 1968,
b
Data from Stevens and Benz 1978a)
AIAA SP-084-1999
21
Table 4 Autoignition temperatures of hydrazine-air mixtures in contact with various materials
Autoignition contact
Pressure Temperature
kPa psia K F
Aluminum 6061-T6
a
85.3 12.4 613 644
Black Iron
b
101.3 14.7 405 269
Borosilicate Glass
b
101.3 14.7 543 518
Borosilicate Glass
a
85.3 12.4 548 526
Copper
a
85.3 12.4 538 509
Ferric Oxide
b
(powder) 101.3 14.7 <296 <73
Nickel
a
85.3 12.4 629 673
Platinum
b
101.3 14.7 499 439
18-8 Stainless Steel
b
101.3 14.7 429 313
304 Stainless Steel
c
101.3 14.7 433 320
304 Stainless Steel
a
85.3 12.4 549 529
316 Stainless Steel
a
85.3 12.4 559 547
Titanium (Ti-6Al-4V)
a
85.3 12.4 624 664
a
Miller and Schluter (1978)
b
Scott, Burns, and Lewis (1949)
c
Stevens and Benz (1978a)
2.1.2.3 Hydrazine-diluent mixtures
2.1.2.3.1 Minimum ignition energy
2.1.2.3.1.1 Pressure effect
No data are reported in the reviewed literature on the effect of pressure on the minimum ignition energy
of hydrazine-diluent mixtures.
2.1.2.3.1.2 Temperature effect
No data are reported in the reviewed literature on the effect of temperature on the minimum ignition
energy of hydrazine-diluent mixtures.
2.1.2.3.2 Autoignition temperature
The autoignition temperature reported for hydrazine in nitrogen at 101.3 kPa (14.7 psia) is shown in
Table 5. These values are higher than the autoignition temperature for hydrazine-air mixtures and are
dependent on the container material. As noted previously (Section 2.1.2), autoignition temperature data
should be applied cautiously.
Table 5 AIT's reported for hydrazine in nitrogen at 101.3 kPa (14.7 psia)
Container or surface AIT Reference
K F
Borosilicate Glass 802 984 Ladacki (1968)
304 Stainless Steel >644 700 Stevens & Benz (1978)
18-8 Stainless Steel >688 778 Scott, Burns, & Lewis (1949)
Black Iron 404 268 Scott, Burns, & Lewis (1949)
AIAA SP-084-1999
22
2.1.2.3.2.1 Concentration effect
No data are reported in the reviewed literature on the effect of concentration on the autoignition
temperature of hydrazine-diluent mixtures.
2.1.2.3.2.2 Pressure effect
No data are reported in the reviewed literature on the effect of pressure on the autoignition temperature
of hydrazine-diluent mixtures.
2.1.2.3.2.3 Material effects
The autoignition temperature for hydrazine-diluent mixtures is dependent on the container material.
Tests done in both stainless steel with water vapor as a diluent (Stevens and Benz 1980) and black iron
containers under nitrogen atmospheres (Scott, Burns, and Lewis 1949) resulted in decreased
autoignition temperatures.
2.1.2.4 Hydrazine-oxidizer mixtures
Ignition products and reaction mechanisms for hydrazine/nitrogen tetroxide systems at atmospheric and
2 MPa (300 psia) pressures have been studied and compared to thermodynamic equilibrium calculations
(Summers and McMullen 1966). A Time-of Flight Mass Spectrometer was used to identify the species
produced in a combustor during testing. The authors made the following observations:
the concentration of molecular nitrogen and oxygen increases with mixture ratio,

the concentration of nitric oxide decreases or remains relatively constant with increasing mixture
ratio,
the concentration of atomic nitrogen and oxygen is relatively constant at high mixture ratios, and

no NO
2
was observed.
The authors emphasized that these observations must be considered when developing reaction
mechanisms. In addition, they determined that the chemical species found in the combustion chamber
were not in equilibrium (thermodynamic equilibrium can not be assumed) and that high-temperature
kinetics may vary significantly from those at lower temperatures.
2.1.2.4.1 Minimum ignition energy
2.1.2.4.1.1 Pressure effect
No data are reported in the reviewed literature on the effect of pressure on the minimum ignition energy
of vapor hydrazine-oxidizer mixtures.
2.1.2.4.1.2 Temperature effect
No data are reported in the reviewed literature on the effect of temperature on the minimum ignition
energy of vapor hydrazine-oxidizer mixtures.
AIAA SP-084-1999
23
2.1.2.4.2 Autoignition Temperature
Measured autoignition temperatures for hydrazine-oxygen mixtures range from 423 K (302 F) (Furno,
Imhof, and Kuchta 1968) to 475 K (396 F) (Scott, Burns, and Lewis 1949). These values were
measured in borosilicate glass containers and are generally lower than autoignition temperatures for
hydrazine-air mixtures. As noted previously (Section 2.1.2), autoignition temperature data should be
applied cautiously.
Hydrazine combined with an equilibrium mixture of dinitrogen tetroxide and nitrogen dioxide is hypergolic
in the absence of diluents and at ambient temperature and pressure.
Hypergolic reaction between liquid hydrazine and gaseous dinitrogen tetroxide/nitrogen mixtures are
reported in Figure 9 as a function of the liquid temperature. Liquid hydrazine was flowed (at Reynolds
numbers of 29) into gaseous NO
2
*/N
2
mixtures (NO
2
* represents the equilibrium mixture of NO
2
and
N
2
O
4
). The temperature of the nitrogen tetroxide in the flow was maintained at 310 K (98 F). The curve
shown marks the boundary between ignition and nonignition regions.
Figure 9 Minimum spontaneous ignition temperatures for liquid hydrazine in contact with NO
2
* nitrogen mixtures
at 310 K (98 F) and ambient pressure (Data from Zung, Breen, and Kushida 1968)
2.1.2.4.2.1 Pressure effect
Autoignition temperatures for stoichiometric hydrazine-oxygen mixtures increase with decreasing
pressure (Figure 10).
The effect of reduced pressure on the ignition of hydrazine vapor-dinitrogen tetroxide mixtures was
measured by Martinkovic (1964). Vials of hydrazine and nitrogen tetroxide were simultaneously burst in
open and closed but not sealed compartments [12.9 L (787 in
3
)] located in a chamber (approximately
17000 L (600 ft
3
)) evacuated to simulate high altitudes. Ignition was observed in a closed compartment
at simulated high altitude pressures greater than 19.8 Pa (2.9 x 10
-3
psia). In a partially closed
compartment, ignition was only observed at simulated high altitude pressures greater than 1.1 kPa
(0.162 psia). In another study of the ignition of hypergols in simulated space environments (Chaun and
Wilber 1967) hydrazine and nitrogen tetroxide vapors were mixed in vessels of different sizes. For
AIAA SP-084-1999
24
mixtures at 300 K spontaneous ignitions occurred in a 10 L (610 in
3
) pyrex bell jar over a stainless steel
base for total pressures ranging from 267 Pa (3.8 x 10
-2
psia) to 533 Pa (7.7 x 10
-2
psia) and in a 1600 L
(98,000 in
3
) stainless steel sphere the spontaneous ignitions occurred at total pressures ranging from
20.0 Pa (2.9 x 10
-3
psia) to 32.0 Pa (4.7 x 10
-3
psia). The decrease in total pressure at which ignition
occurs is attributed by the authors to an inverse effect of the vessel volume to surface area. The values
measured by Martinkovic reflect a lower vessel volume to surface area ratio than the vessels tested by
Chaun and Wilber. Ignition was observed to be relatively insensitive to the mixture ratio.
Figure 10 The effect of reduced pressure on the autoignition temperature of hydrazine in oxygen
Zung, Breem, and Kushida (1968) performed hypergolic reaction experiments for liquid hydrazine and
NO
2
*/N
2
mixtures at reduced pressures. Their results are shown in Figure 11.
2.1.2.4.2.2 Concentration effect
Gray and Lee (1954) admitted known concentrations and pressures of hydrazine vapor and oxygen into
uniformly heated cylindrical borosilicate vessels. In a 4.6 cm (1.8 in.) inside diameter vessel (Figure 12)
ignitions at a vessel temperature of 809 K (997 F) occur at lower pressures than when the vessel
temperature is 731 K (865 F). The data show oxygen-rich mixtures igniting at lower pressures than
hydrazine rich mixtures. Other data show that when the vessel size is decreased from 4.6 cm (1.8 in.) to
a 2.3 cm (0.9 in.) inside diameter the minimum AIT for stoichiometric mixtures is observed to occur at
higher temperatures for a given pressure. AIT data for oxygen rich mixtures (1 N
2
H
4
+ 2 O
2
) in a 4.6 cm
(1.8 in.) inside diameter glass vessel show that for pressures less than 2.43 kPa (1.34 psia) a local
minimum AIT of 693 (788 F) is observed at a pressure of 0.27 kPa (0.039 psia). For these mixtures a
delayed ignition phenomenon of 5 to 10 s is observed rather than an explosion.
AIAA SP-084-1999
25
Figure 11 The effect of reduced pressure on the autoignition temperature of liquid hydrazine in dinitrogen
tetroxide (Data from Zung, Breen, and Kushida 1968)
Figure 12 The effect of concentration on the autoignition temperature of hydrazine in oxygen (Data from Gray and
Lee 1955)
Perlee, Imhof, and Zabetakis (1962) injected liquid hydrazine at different initial temperatures into heated
NO
2
*-air mixtures contained in a 250 cm
3
(15.3 in
3
) borosilicate glass flask.
AIAA SP-084-1999
26
Figure 13 shows the minimum temperature of the NO
2
*-air mixture required to spontaneously ignite liquid
hydrazine at varying temperatures as a function of NO
2
* concentration at pressures of approximately
100 kPa (14.5 psia). It was found that at any fixed concentration of NO
2
* in air, spontaneous ignition
temperature increased with decreasing initial liquid temperature. The AIT was found to be independent
of the injected fuel volume over the range of 0.01 cm
3
(6.1 x 10
-4
in
3
) to 0.07 cm
3
(4.3 x 10
-3
in
3
).
Additionally, when both the initial liquid hydrazine and the NO
2
*-air mixture were at or above ambient
temperature, the hydrazine ignited spontaneously in NO
2
*-air mixtures containing more than
approximately 15 percent NO
2
*. In all cases, the hydrazine and NO
2
* reacted on contact, but the
reaction did not always culminate in burning, as indicated by the nonignition region.
Figure 13 Minimum spontaneous ignition temperatures for liquid hydrazine in contact with NO
2
*-air mixtures at
ambient pressure (Data from Perlee, Imhof, and Zabetakis 1962)
The data in Figure 9 shows the effects of varying the mass fraction of nitrogen tetroxide in nitrogen gas
when it is mixed with hydrazine liquid.
2.1.2.4.2.3 Material effects
Figure 2.9 shows the difference in system response to glass and stainless steel containers. Potassium
chloride coatings on glass increase the autoignition temperature (Gray and Lee 1954).
2.1.2.4.2.3 Diluent effects
The effect of added inert gases on the autoignition temperature of stoichiometric mixtures of hydrazine
and oxygen was studied by Gray and Lee (1955). In Figure 14 argon and nitrogen reduce the AIT by as
much as 20 percent while the AIT increases with the addition of helium.
AIAA SP-084-1999
27
Figure 14 The effect of diluents on the autoignition temperature of hydrazine in oxygen (Data from Gray and Lee
1955)
2.1.3 Flame velocity
The movement of a flame is defined relative to a specific coordinate system. For coordinates centered
in the flame, the velocity at which unburned gases move through the combustion zone in the direction
normal to the flame front is defined as the flame velocity. For laboratory coordinates, the flame speed is
the velocity of the flame through the unburned gases.
The flame velocity (see Glossary) of hydrazine vapor can be calculated from the flame speed (see
Glossary), and the equation used depends on the nature of the confinement. When a flame propagates
throughout a tube, the flame velocity can be calculated by using Equation 1 (Kanury 1976):
U
v
= U
s
U
r
(1)
where:
U
v
= flame velocity, m/s (ft/s)
U
s
= flame speed, m/s (ft/s)
U
r
= fresh reactant mixture velocity, m/s (ft/s)
For a spherical flame, the flame velocity can be calculated by Equation 2 (Strehlow 1984):
(2)
U
v
=
V
1
V
2






U
s
=
r
1
r
2






3
U
s
AIAA SP-084-1999
28
where:
V
1
= initial volume of flame ball, m
3
(in
3
)
V
2
= final volume of flame ball, m
3
(in
3
)
r
1
= initial radius of flame ball, m (in.)
r
2
= final radius of flame ball, m (in.)
Combustion in a system may affect the boundary conditions of the combustion process causing the
flame velocities and flame speeds to vary. An average value may be used for the purpose of estimation.
In a flow system, the flame speed is dependent on whether the flow is laminar or turbulent. It is
important to consider flame speed for determining the rate that a fire can move through the flammable
mixture and subsequently distribute the fire throughout the system.
In evaluating the rate that heat is being generated throughout the system, the flame speed is used to
calculate the mass rate of consumption of the fuel.
2.1.3.1 Flame speed in laminar flow
The elementary combustion theory of Mallard and Le Chatelier (Glassman 1987) predicts that the
laminar-flow flame speed is proportional to the square root of the product of the thermal diffusivity and
the reaction rate (Equation 3).
U
L
( x RR) (3)
where
U
L
= laminar-flow flame speed, m/s (ft/s)
= thermal diffusivity, m
2
/s (ft
2
/s)
RR = reaction rate, (1/s)
2.1.3.2 Effect of temperature
The laminar flame speed varies with temperature through a change in unburned gas mixture properties
(heat capacity, thermal conductivity, density), a change in reaction rate with temperature, and the initial
temperature of the unburned gases, although the latter effect is small (Glassman 1987).
2.1.3.3 Effect of pressure
The reaction rate is a function of pressure, which translates to the following pressure dependency for
flame speed (Equation 4) (Glassman 1987)
U
L
p
n2
( )
1 / 2
(4)
where
n = overall order of the reaction (Kanury 1976)
p = pressure of the unburned gases
Gray and Lee (1955) have found that for neat hydrazine decomposition the order of reaction is n=2,
which would indicate that the flame speed would be approximately independent of pressure.
AIAA SP-084-1999
29
2.1.3.4 Effect of concentration
In general, flame speeds of mixtures are greatest for compositions close to stoichiometry and approach
zero at the limits of flammability (Kanury 1976).
2.1.3.5 Neat hydrazine
Flame velocities for hydrazine decomposition are shown in Table 6 for different experimental conditions.
The flame velocity for decomposition is a slowly changing function of pressure (Hall and Wolfhard 1956;
Gray and Lee 1959).
2.1.3.6 Hydrazine-air mixtures
No data are reported in the reviewed literature on the flame velocity of hydrazine-air mixtures.
Table 6 Flame velocity for neat hydrazine
Flame Velocity Pressure Temperature Composition
Measurement
Method
m/s ft/s kPa psia K F % v/v
1.12
a
3.7 5.3 0.77 335 143 99.9-100.2 Pyrex cylinder
1.17
a
3.8 6.7 0.97 335 143 99.9-100.2 Pyrex cylinder
1.79
b
5.9 20.0 2.90 393 248 97 Burner
1.85
b
6.1 9.3 1.35 393 248 97 Burner
1.96
b
6.4 4.0 0.58 393 248 97 Burner
1.85
c
6.1 101.3 14.7 423 302 97 Burner
a
Gray and Lee (1959)
b
Hall and Wolfhard (1956) Values represent average.
c
Murray and Hall (1951)
2.1.3.7 Hydrazine-diluent mixtures
The flame speed was found to decrease with increased concentration of diluents such as argon, helium,
nitrogen, hydrogen, and ammonia (Figure 15). Heat capacity and thermal conductivity are two factors
that contribute to the effect of a diluent on flame speed. Increasing the heat capacity of the mixture by
adding a diluent lowers the flame temperature and therefore the flame speed. Nitrogen and hydrogen,
with greater heat capacities than argon or helium cause a greater reduction in the flame speed. The
effect of thermal conductivity is seen with helium and argon which possess nearly the same heat
capacity. The system with a higher thermal conductivity will have a greater flame speed (Gray and Lee
1959).
2.1.3.8 Hydrazine-oxidizer mixtures
The flame velocities of various hydrazine-oxidizer systems in stoichiometic mixtures are given in Table 7.
The flame velocities of various concentrations of 97.2 percent aqueous hydrazine-oxygen mixtures at
423 K (302 F) and ambient pressure are displayed in Figure 16. These data suggest that the highest
burning velocities occur at concentrations near stoichiometry; velocities decrease dramatically as the
mixture becomes leaner or richer in hydrazine. The flame velocities of various hydrazine-water vapor
mixtures at 423 K (302 F) and ambient pressure are displayed in Figure 17.
AIAA SP-084-1999
30
Figure 15 Effects of diluents on the flame speed for hydrazine decomposition flames at an initial temperature of
335 K (144 F) and initial hydrazine pressure of 6.7 kPa (0.39 psia) (Data from Gray and Lee 1959)
Table 7 Heat of reaction, adiabatic flame temperature, and flame velocity for various hydrazine oxidizer systems
a
Reaction
Heat of
Reaction
H
298
Adiabatic
Flame Temp.
T
ad
Flame
Velocity
U
v
kJ/mol Btu/mol K F m/s ft/s
Decomposition (partial)
N
2
H
4
NH
3
+ 0.5N
2
+ 0.5H
2
141 134 1904 2968 1.10 3.61
Oxidation
N
2
H
4
+ O
2
N
2
+ 2H
2
O 579 549 2700 4390 2.80 9.19
N
2
H
4
+ 2N
2
O 3N
2
+ 2H
2
O 743 704 2655 4319 1.64 5.38
N
2
H
4
+ 2NO 2N
2
+ 2H
2
O 760 720 2745 4481 2.45 8.04
N
2
H
4
+ 2NO + N
2
3N
2
+ 2H
2
O 761 721 2665 4319 1.50 4.92
N
2
H
4
+ NO 1.5N
2
+ H
2
O + H
2
511 484 2645 4301 2.38 7.81
NOTE: Flame velocities measured at an initial pressure of 5.3 kPa (0.77 psia) and an initial
temperature of 335 K (144 F).
a
Gray and Lee (1959)
AIAA SP-084-1999
31
Figure 16 Flame velocities for 97.2% aqueous hydrazine-oxygen mixtures at 423 K (302 F) and ambient
pressure (Data from Murray and Hall 1951)
Figure 17 Flame velocities for hydrazine-water vapor mixtures at 423 K (302 F) and ambient pressure (Data
from Murray and Hall 1951)
AIAA SP-084-1999
32
2.2 Liquid hydrazine
Liquid hydrazine supports fire only if its temperature is above the fire point and an ignition source is
present.
In general, liquid pool fires are effected by temperature and wind. As the temperature exceeds the fire
point, the vaporization rate increases and the rate of flame spread across a liquid layer increases. At
temperatures well above the fire point, the flame spread rate does not increase with increasing
vaporization (Botteri et al. 1966; Kuchta 1973). A burning pool may be extinguished by a strong wind; as
the diameter of the pool increases, the velocity of the wind required to extinguish the fire also increases
(Atkinson and Eklund 1971).
2.2.1 Flash and fire points
Flash and fire points of liquid hydrazine vary depending on whether the hydrazine vapor above the liquid
surface is mixed with air, an inert gas, or another oxidizer.
The flash point of hydrazine is the lowest temperature at which liquid hydrazine gives off enough vapor
to form an ignitable mixture with air at ambient pressure near the surface of the liquid (Sax 1975). The
fire point, the lowest temperature at which liquid hydrazine continuously supports a flame (Jensen 1976),
is slightly higher than the flash point (as determined by the open-cup method).
Flash points of compounds and their mixtures are normally determined by the Tag closed-cup or
Cleveland open-cup test methods, with open-cup flash points generally measuring higher than closed-
cup flash points. The Cleveland open-cup method is described in the ASTM method D 92-85 (ASTM
1986b); the Tag method used to determine closed-cup flash points is detailed in ASTM method D 56-87
(ASTM 1986c). The methods used to determine fire points are detailed in Scott, Burns, and Lewis
(1949).
The hydrazine vapor concentration at the fire point is the same as the lower flammability limit of
hydrazine. Refer to Section 2.1.1 for information on hydrazine vapor flammability limits.
Mists can form when saturated vapor cools or when a liquid is mechanically sprayed and may ignite at
temperatures below the flash point (Botteri et al. 1966). Refer to Section 2.3 for additional information.
2.2.1.1 Neat hydrazine
Flash points are usually measured in air or a diluent gas at ambient pressure. Because no air or diluent
gas is present with neat hydrazine, flash points are not an appropriate measure of the fire hazard of neat
hydrazine. Neat hydrazine vapor is flammable above the low pressure limit of 2.1 kPa (0.30 psia)
(Section 2.1.1.1), and neat hydrazine vapor is flammable at 300 K (81 F), which is below the flash point
in air (Benz, Guerrasio, and Rollins 1983).
2.2.1.2 Hydrazine-air mixtures
Flash and fire point data of neat liquid hydrazine at ambient pressure are shown in Table 8.
At reduced pressures , hydrazine vapor may become flammable if its concentration relative to the
ambient atmosphere increases. This may be true even when the hydrazine temperature is lower than
ambient (Benz, Bishop, and Pedley 1988).
AIAA SP-084-1999
33
Table 8 Flash and fire points for neat liquid hydrazine at 101.3 kPa (14.7 psia)
Test Method Flash point Fire point
K F K F
Tag Closed Cup
a
311 100 ND
c
Tag Open Cup
a
325 126 ND
c
Tag Open Cup
b
324 124 324 124
a
U.S. Air Force (1973)
b
Olin Chemical Company (1967)
c
No data
The flash and fire points of hydrazine-water mixtures for different concentrations of hydrazine,
determined by Cleveland Open-Cup tester, are shown in Table 9 (Olin Chemical Company 1967). For
neat hydrazine the flash and fire points are identical (324 K, 123.5 F), but for hydrazine-water mixtures
the fire point may be up to 10 K higher than the flash point. The flash point of hydrazine is also
increased by addition of water. When the water concentration is 58 weight percent or higher, the mixture
can no longer be ignited under any open-air condition, not even with a very hot ignition source (Hshieh
1992).
Table 9 Flash and fire points of hydrazine-water mixtures
Hydrazine
Concentration
Flash point
a
Fire point
a
(Weight Percent) K F K F
100 324 124 324 124
90 329 133 329 133
80 335 143 335 143
70 342 156 343 158
60 353 176 357 183
50 371 208 381 226
42 NI
b
NI
b
a
The reported values were determined by a Cleveland Open-Cup tester. The
reported values have been corrected to a pressure of 101.3 kPa (14.7 psia).
b
NI indicates no ignition.
2.2.2 Burning rate and burning velocity
The burning rate for liquid hydrazine is the mass of fuel consumed per unit surface area per unit time for
the burning pool. A burning velocity can be defined as the rate that the level of the liquid pool
decreases. The burning velocity is the burning rate/density of the liquid.
After ignition (temperature exceeds the fire point, Section 2.2.1), the hydrazine liquid will continue to
burn only if the rate of vaporization is at least as great as the rate of consumption of the vapor (see
Glossary). The burning rate of a pool fire generally increases as the pool diameter increases and then
levels off at diameters greater than approximately 18 cm (7 in.) (Benz, Bishop, and Pedley 1988). Pool
depth is not reported to be a factor in the burning rate.
2.2.2.1 Neat or high purity
No data are reported in the reviewed literature on the burning rate or burning velocity of liquid pools of
neat or high purity hydrazine.
AIAA SP-084-1999
34
2.2.2.2 Monopropellant (MIL-P-26536D)
The mean burning rate of a 4.8-cm (1.9-in.) diameter pool of hydrazine in air at ambient pressure and
298 K (77 F) is 0.11 kg/(m
2
s) (1.3 lb/(ft
2
min)), and the mean burning velocity calculated from the
burning rate is 0.11 cm/s (0.04 in s). For a 15.2-cm (6-in.) diameter pool of hydrazine, the mean burning
rate is 0.22 kg/(m
2
s) (2.6 lb/(ft
2
min) (Burgess et al. 1969; Benz, Bishop, and Pedley 1988), and the
calculated mean burning velocity is 0.22 cm/s (0.09 in/s).
Data for hydrazine water mixtures at a concentration of 98.9 percent hydrazine (Table 10) approximate
the results expected for monopropellant hydrazine. The burning velocities reported in Table 10 were
measured in very narrow tubes.
2.2.2.3 Mixtures
The burning rate of hydrazine-water mixtures is reduced with increased amounts of water. Burgess et
al. (1969) found that a 15.2-cm (6-in.) diameter pool containing water in a concentration of 20 percent by
weight has a burning rate of 0.034 kg/(m
2
s) (0.4 lb/(ft
2
min)); this corresponds to a calculated burning
velocity of 0.034 cm/s (0.013 in./s). A pool containing 50 percent water does not ignite. This contrasts
with the data of Table 9 obtained with a Cleveland Open Cup Tester. There data indicate that ignition
can occur for a 50 weight percent concentration and the mixture can no longer be ignited until the water
concentration is 58 weight percent or higher. The data are significant and are of practical use in
designing procedures for fighting hydrazine fires as discussed in Section 5.
The burning velocities of several concentrations of liquid hydrazine-water mixtures in nitrogen at varying
pressures and tube diameters are in Table 10 Adams and Stocks (1953) show in this table that a
maximum nitrogen pressure (P
Ext
) exists above which liquid hydrazine does not burn.
2.3 Hydrazine mists, droplets, and sprays
Mists can form when saturated vapor cools or when liquid fuel is mechanically sprayed. If an ignition
source persists, vaporization from the mist may result in ignition at lower temperatures than vapor
ignition. In addition, the rapid vapor formation that occurs in hydrazine mists, droplets, or sprays can
support fire if the amount of vapor that forms is equal to or greater than the amount of vapor that is
consumed (Benz, Bishop, and Pedley 1988).
2.3.1 Flash and fire points
Mists or foams have flash points and/or fire points that range from less than to greater than that of
corresponding vapor systems.
2.3.2 Burning rates
The mass burning rate of a single liquid hydrazine droplet increases with increasing droplet diameter,
oxygen concentration, and surrounding temperature (Allison and Faeth 1972). The burning rate of
monopropellant-grade hydrazine droplets in the absence of oxygen is slower than the burning rate of
monopropellant-grade hydrazine droplets in oxygen (Figure 18).
Table 10 Burning velocities for liquid hydrazine-water mixtures pressurized by nitrogen gas
a
Applied
N
2
pressure
101kPa 203 kPa 405 kPa 608 kPa 811 kPa 1115 kPa 1621 kPa 2128 kPa 2634 kPa
HZ Conc.
% weight
Burning Velocity
cm/s
P
Ext
b
kPa
3 mm tube diameter
94-95 0.026 0.047 0.067 0.093 0.092 2685
94-95 0.023 0.057 0.076 0.107 c
5 mm tube diameter
98.9 0.028 0.036 0.041 0.077 0.077
98.9 0.029 0.038 0.048 0.078 0.078
97.0 0.028 0.084 0.060 4256
97.0 0.027 0.061 0.125 4458
94.5 0.021 0.046 0.055 0.065 0.077 3445
94.5 0.023 0.047 0.057 0.065 0.061 3546
89.8 0.011 0.026 0.021 1419
89.8 0.011 0.025 0.023 1469
88.8 c 0.019 c 1064
88.8 c 0.016
88.8 c 0.017 c
a
Adams and Stocks (1953)
b
Pressure above which liquid did not burn
c
Failed to maintain burning after ignition
3
5
A
I
A
A

S
P
-
0
1
6
-
2
-
1
9
9
9
AIAA SP-084-1999
36
Figure 18 Hydrazine burning rates as a function of drop diameter for oxidation and decomposition at ambient
pressure (Data from Allison and Faeth 1972)
AIAA SP-084-1999
37
3 Explosion
Introduction
The potential for an explosion always exists when a flammable mixture of hydrazine is present in a
system (see Glossary). While it is possible to design pressure relief into a system, the rate of pressure
increase during a detonation usually cannot be dissipated through such relief devices, and failure of the
mechanical containment may result. The subsequent explosion may injure personnel and/or may
damage surrounding components.
In assessing the explosion potential of a hydrazine system, three separate processes must be
considered: deflagration (Section 3.1), detonation (Section 3.2) and exothermic reaction (Section 3.3).
Two things should be remembered: (1) deflagrations and exothermic reactions can under go transitions
to detonations, which is potentially the most hazardous process, and (2) deflagrations and exothermic
reactions in isolated systems build up pressure that can lead to explosions.
3.1 Deflagration
A deflagration is defined as a flame moving through a flammable mixture at subsonic velocity (Kanury
1976). The propagating flame is a zone of exothermic reaction in which heat and mass transfer sustain
the reaction. In other words, the heat produced by the reaction must be transferred into the unburned fuel
so that the rate of reaction is sufficiently fast to sustain the flame, and in a similar manner, the fuel mixture
must be available at the flame zone to sustain the reaction. The zone may be a fraction of a millimeter to
more than a centimeter thick (Zeldovich et al. 1985). When the system is filled with a flammable mixture,
the usual occurrence is that ignition takes place in a small region, and the flame propagates throughout
the system. Deflagrations have less destructive power and propagate more slowly than detonations
(Section 3.2); however deflagrations are still considered very rapid explosion events (Sax and Lewis
1987).
The assessment of the potential for a deflagration is similar to that followed for fires. The following criteria
are necessary for a deflagration to exist:
1. A flammable mixture exists (see Section 2).
2. An ignition source is present (see Section 2).
3. If ignition occurs, the ignition source is located where a flame can be propagated throughout the
system.
The propagation of a free flame depends to a large degree on the gas dynamic flow structure (Lee, Chan,
and Knystautas 1982). Laminar flames have been described in Section 2. Obstacles in the path of the
flame can produce turbulence, which can accelerate the flame. This acceleration can occur over a short
distance.
Turbulence can also quench the flame. If the mixing that occurs during a turbulent process reduces the
flame-zone temperature, the rate of reaction will decrease, and the generation of thermal energy
decreases. This could also occur through interaction with the walls (surroundings). Deflagration is very
dependent on the system configuration, and each system is a unique consideration.
The effects produced by a deflagration come from an increase in pressure caused by the release of
thermal energy and in the case of hydrazine, an increase in the number of moles. The deflagration, if in
an open system, could create a blast wave in the surrounding medium (Kinney and Graham 1985). If a
AIAA SP-084-1999
38
deflagration transforms to detonation, effects associated with supersonic detonation waves can result
(Section 3.2).
The critical concerns are the over pressurization of closed systems and the transition of the deflagration
into a detonation; this can occur if a mechanism for flame acceleration exists in the path of the
propagating flame (Section 3.2).
3.1.1 Hydrazine vapor
Hydrazine vapor deflagration can generate significant pressures. Experimentally determined pressure
rises and pressure generation rates due to deflagration depend on experimental conditions and the type
of data collected which can vary significantly. Thus the remaining discussion is restricted to a summary of
factors in real, non-adiabatic systems that effect the pressure rise and pressure generation rate due to
deflagration, and the theoretical expressions to estimate them.
The effects of several factors on the pressure and pressure generation rate due to deflagrations in real,
non-adiabatic systems can be summarized briefly:
increased flame velocity leads to increased pressure change (Zabetakis 1965);
the addition of inert diluents or the presence of excess fuel or oxidizer reduces the deflagration
speed and narrows the flammability range. An exception may arise with the addition of helium or
argon; if the ratio of thermal conductivity to specific heat is increased, then the deflagration speed
can increase (Kanury 1976). This can lead to increased pressure change;
stoichiometric oxidant-fuel mixtures result in the greatest pressures (Benz, Bishop, and Pedley
1988);

given a constant initial pressure, an increased initial temperature leads to a decreased maximum
pressure. This occurs because the density of the fuel-oxidizer mixture is lower and less chemical
energy is released in the deflagration (Benz, Bishop, and Pedley 1988);

increased system pressure leads to increased pressure due to deflagration (Benz, Bishop, and
Pedley 1988);
upward propagation of the explosion leads to greater flame velocity (and subsequent pressure
from the deflagration products) than downward propagation (Coward and Jones 1952);
turbulence caused by obstacles in the flow path can increase the rate of energy release of the
deflagration and in nonadiabatic systems lead to increased pressure, especially near the lower
and upper flammability limits (Harris 1967);
vessel volume or shape does not usually affect the pressure from deflagration products unless
the surface area of the vessel is large; in spherical chambers, the pressure generation rate
decreases as the vessel volume increases (Harris 1967; Nagy et al. 1971); and
ignition source alone has little effect on pressure resulting from a deflagration (Nagy et al. 1971),
although increased ignition energy and multiple ignition sources lead to increased pressure
generation rate (Lee and Guirao 1981).
AIAA SP-084-1999
39
3.1.1.1 Temperature of a Deflagration
Given an appropriate mathematical model of the system, the temperature distribution both in time and
space could be obtained (Bird, Stewart, and Lightfoot 1960). However, an estimate of the flame
temperature can be obtained assuming adiabatic conditions (Equation 5):
T
ad
= T
o
+
Q - n
j
j

j
n
j
C
pj
j

(5)
where
T
ad
= adiabatic flame temperature (K)
T
o
= initial temperature (K)
Q = the total enthalpy of reaction for hydrazine mixture (kJ)
n
j
= moles of j
th
product (mol)
j
= latent heat of phase change for the j
th
product for temperatures between T
o
and T
ad
(kJ/mol)
pj
C = mean heat capacity of j
th
product (kJ/(molK))
j = number of products
Table B-2 in Annex B gives T
ad
for various mixtures of hydrazine and air.
3.1.1.2 Pressure of a deflagration
Under the conditions that the deflagration proceeds throughout the entire system (all available hydrazine
is reacted), and the system is closed (rigid vessel) and adiabatic (no heat loss to surroundings), a
deflagration pressure can be calculated as an adiabatic pressure (Equation 6):
ad P =
o P
ad
T
o T


_
,
f
n
o n


_
,
(6)
where
P
ad
= adiabatic pressure (kPa)
P
o
= initial pressure (kPa)
T
ad
= adiabatic flame temperature (Equation 5) (K)
T
o
= initial temperature (K)
n
f
= moles after complete reaction (mol)
n
o
= initial moles total (mol)
Table B-2 in Annex B gives P
ad
for various mixtures of hydrazine and air.
3.1.1.3 Deflagration transition to detonation
The transition of a deflagration to a detonation is described in Section 3.2.1.
3.1.1.4 Time dependent pressure
The ability to predict the total pressure generated at a given time and the rate of pressure generation is as
important as the ability to predict the maximum pressure. Protective devices such as burst disks and
relief valves, vital to the protection of pressure vessels, must be effective not only at a particular pressure
but must also accommodate a rapidly increasing pressure.
AIAA SP-084-1999
40
The total pressure at a given time, although primarily affected by the flame velocity (Sections 2.1.3 and
2.2.2), is also affected by ignition point and vessel geometry. The following examples illustrate the
pressure generated at a given time and the pressure generation rate in adiabatic systems for two
commonly used geometrical configurations. These examples assume that the pressure changes linearly
as the volume fraction burned and ignores differences between the burned portion and the unburned
portion. Similar equations can be derived for other geometrical configurations. For ignition starting at one
end of a closed cylinder and propagating axially, the time-dependent pressure (Equation 7) is
P
AU
b
t P
ad
P
o
( )
V
+ P
o
(7)
where
P = total pressure as a function of t (kPa)
A = cross-sectional area of the cylinder (m
2
)
U
b
= flame velocity (m/s)
t = time (s)
P
ad
= adiabatic pressure (kPa)
P
o
= initial pressure (kPa)
V = volume of container (m
3
)
In Equation 7, it is assumed that there are no effects from the walls as the flame propagates down the
cylinder. For central ignition in a spherical container a similar expression for the theoretical total pressure
is given by Equation 8
P P P
o

kP
o
U
b
3
t
3
3V
P
ad
P
o
(8)
where k is a constant (Zabetakis 1965).
In both Equations 7 and 8, P = P
ad
when the flame has propagated through the entire container volume.
These equations yield a calculated ideal gas pressure. Real systems are affected by factors such as
non-central ignition, non-uniform burning, and wall effects. Zabetakis (1965) has shown that these factors
reduce the total pressure to less than P
ad
.
The rate of pressure increase can be found by differentiating Equations 7 and 8 with respect to time:
dP
dt

AU
b
(P
ad
P
o
)
V
(9)
and
dP
dt

kP
o
U
b
3
t
2
V
(10)
Equations 9 and 10 are ideal models. The pressure generation rate for real systems is lower.
3.1.2 Liquid hydrazine
Liquid hydrazine is flammable above the closed-cup flash point of 311 K (100 F). In a closed vessel, fire
in the vapor above the liquid can lead to explosion. Refer to Section 2.1 for additional information.
AIAA SP-084-1999
41
3.2 Detonation
The rapid and violent form of reaction called detonation differs from other forms in that all the important
energy transfer is in strong shock (compression) waves, with negligible contribution from other processes
like heat conduction, which are important in deflagration and thermal runaway (Fickett and Davis 1979).
After initiation, the thermal energy of the reaction sustains the shock wave, and the shock wave
compresses the unreacted material to sustain the reaction. The velocity of the shock wave is supersonic
with respect to the unburned gases. As shown in Table B-1 in Annex B, in hydrazine/air mixtures the
stable shock wave for propagation varies from velocities of 1043 m/s (318 ft/s) to 2511 m/s (765 ft/s). It is
the overpressure (the difference between product pressure and initial pressure) that is potentially
destructive and constitutes the hazard of detonation. The pressure ratio for hydrazine or hydrazine/air
mixtures, product pressure to initial pressure, shown in Table B-1 gives values between 5.8 and 27.6 for
detonations. The distribution of products changes and mole numbers increase leads to an increased
pressure for higher equivalence ratios.
The criteria that must be met for a detonation to occur are related to the composition and state of the
flammable material in the system, the size and shape of the vessel or enclosure, and the energy source
that can ignite the flammable mixture. The dynamic parameters, based on cell width (Section 3.2.1) that
must be determined from experiment or empirical correlation, are critical tube diameter, and initiation
energy. As in all systems involving flammable materials, catalytic reaction with structural components can
accelerate reactions, which can initiate a detonation.
An additional consideration is the transition of a deflagration into a detonation. Although there are
general criteria for such a transition (Section 3.1), it is a combination of the system and the dynamics of
the deflagration (flame movement) that lead to this transition, and therefore are specific to a given system
and environment.
The effects of a detonation are the result of the pressures associated with the primary shock wave and/or
reflected shocks. This pressure is a function of the initial state of the mixture. This section considers
detonation in hydrazine vapor and in liquid hydrazine.
3.2.1 Detonation theory
The classic Chapman-Jouguet (C-J) theory is a thermodynamic equilibrium calculation for the detonation
states (the detonation velocity, pressure, temperature, and density ratio across the compression wave)
(Kanury 1976). When C-J conditions occur, the parameters are known as C-J parameters. These static,
thermodynamic equilibrium calculations generally agree with experimental values (Zeldovich et al. 1985;
Glassman 1987). The Gordon and McBride code performs C-J calculations (Gordon and McBride 1976).
The dynamic parameters of detonations are far more complicated concepts, and at present there is no
acceptable theory comparable to the C-J calculations (Beeson et al. 1989). These parameters require a
knowledge of the structure and hence the chemical reaction rates, coupled to the shock (compression)
wave. The parameters are detonability limits, the initiation energy, the critical tube diameter, and the
thickness of the reaction zone (Lee 1984). A one-dimensional model for the detonation structure, the
Zeldovic-Von Neuman-Doring (ZND) model provides a mathematical description of the one-dimensional
detonation structure. Although there have been many variations of the ZND model and many attempts
have been made to describe the three-dimensional detonation structure, none have been adequate.
3.2.2 Hydrazine vapor
Detonations are characterized by a structure, which under appropriate experimental measurements show
up as a fish-scale pattern often tracked on smoked foils. This pattern results from the triple point
interaction of the incident shock, the Mach stem, and the transverse shock. The dimensions of the
AIAA SP-084-1999
42
detonation cell are obtained from such a pattern, and the dynamic parameters have been correlated in
terms of the cell width.
3.2.2.1 Cell-dimension parameters for gases and vapors
Detonation susceptibility is inversely proportional to cell width. The smaller the cell width, the more easily
a detonation can occur. Cell width is affected by the system as follows:
Temperature may increase or decrease cell width (Lee et al. 1982). For hydrazine, cell width
increases as temperature decreases (Benz, Bishop, and Pedley 1988).
In fuel-oxidizer mixtures, stoichiometric ratios usually produce the smallest cells. A change from
stoichiometric ratios increases cell size until detonation is not possible (Lee et al. 1982).
Diluents, such as nitrogen, increase cell width (Bull, Ellsworth, and Shiff 1982).
Pressure increase usually decreases cell width (Bull, Ellsworth, and Shiff 1982).
Critical vessel size
For a detonation to occur, there must be sufficient volume for a detonation cell to form.
For totally confined detonation , the tube diameter must be greater than or equal to the width of
one cell divided by (Dupre, Knystautas, and Lee 1986) (Equation 11):
tube d

(11)
where
d
tube
= diameter of the tube, cm (in.)
= detonation cell width, cm (in.)
For transition of a detonation from a tube to an unconfined volume, the minimum diameter of the
tube is empirically determined to be 13 times the width of one cell (Lee 1984) (Equation 12):
critical tube d = 13 (12)
For transmission of a detonation through a slot (or channel) with square dimensions into an
unconfined volume, the minimum channel width of the vapor cloud is empirically determined to be
10 times the width of one cell (Lee 1984) (Equation 13):

W
c
= critical channel width = 10 (13)
For a totally unconfined detonation of a cloud of gaseous fuel-air mixture, the minimum diameter
of the vapor cloud is empirically determined to be 6.5 times the width of one cell (Lee et al. 1982)
(Equation 14):
d
min
= 6.5 =
critical tube
d
2
(14)
For detonation of a cloud of gaseous fuel-air mixture partially confined against a planar surface,
the minimum diameter of the vapor cloud is empirically determined to be five times the width of
one cell (Lee et al. 1982) (Equation 15):
AIAA SP-084-1999
43
min d = 5 =
W
c
2
(15)
3.2.2.3 Detonation energy
An expression indicating the relationship between the detonation cell width and the minimum ignition
energy for direct detonation of an unconfined gas has been developed: (Guirao et al. 1982) (Equation 16):

E
min
= 431
0
C-J
2
V I
3
(16)
where
E
min
= minimum energy (J)

o
= density of unburned gas (kg/m
3
)
V
C-J
= velocity of stable C-J detonation wave (m/s)
I = a dimensionless energy integral; for rough estimates with hydrazine, I 1
= detonation cell width (m)
3.2.2.4 Deflagration transition to detonation
Detonations can be initiated indirectly by a deflagration that accelerates into a detonation. Turbulence is
the main cause of transition (Lee and Moen 1980). Turbulence can cause flame acceleration (Moen et al.
1980) and start the deflagration wave spinning, consequently causing transition to detonation.
Turbulence created by obstacles in the flow path of a deflagration has also been found to lead to
detonation (Sherman et al. 1986). Detonation may also occur when the flame front associated with a
deflagration is reflected off a surface.
A high-pressure reflected wave can enter into a precompressed, preheated medium and detonate. The
overpressure that results is, initially, significantly higher than the stable C-J pressure, because the
detonation occurs in unburned gas that has been precompressed by the compression or pressure wave
generated by and moving ahead of the deflagration flame. The initially high overpressure eventually
decays to the level of the stable C-J pressure.
3.2.2.5 Catalytic materials
Some materials of construction can cause catalytic acceleration of chemical reactions. Hydrazine and
material compatibility are discussed in Section 4. These reactions can cause ignition and deflagration,
which may lead to detonation.
3.2.2.6 Detonation effects
Effects of some of the variables on detonation pressure are as follows:
near-stoichiometric fuel-oxidant mixtures generally result in the greatest detonation pressures,

increased initial temperature at constant initial pressure decreases the detonation pressure for a
constant volume (ideal gas law) (Benz, Bishop, and Pedley 1988); and the reduced mixture
density results in less chemical energy being released in the detonation,

increased initial pressure increases final pressure for a constant volume (ideal gas law), and

increased diluents normally decrease overpressure (Kanury 1976).
AIAA SP-084-1999
44
3.2.2.7 Reflection of shock (compression) wave
In confined systems, including piping (or tubing), the constructive superposition of reflected compression
waves can cause increased pressures. These pressures can reach two or more times the incident
detonation pressure (Bodurtha 1980; Kinney and Graham 1985).
3.2.2.8 Neat hydrazine
3.2.2.8.1 Pressure and velocity
Stable detonations of pure hydrazine vapor were observed by Pedley et al. (1988) at pressures greater
than 0.5 kPa (0.07 psia) with velocities close to the C-J velocity of approximately 2500 m/s (8200 ft/s).
Earlier research by W. Jost (1962), A. Jost et al. (1963), and Heinrich (1964) generated similar results.
Calculation of the detonation properties at C-J conditions for neat hydrazine at an initial pressure of
101.3 kPa (14.7 psia) and 387 K (237 F) indicates that the C-J pressure is 2.80 MPa (406 psia), the
velocity is 2510 m/s (8170 ft/s), and the temperature is 2050 K (3230 F).
3.2.2.8.2 Cell width
The detonation cell width for neat hydrazine vapor at ambient pressure and 393 K (248 F) was reported
by Pedley et al. (1988) to be 1.8 mm (0.067 in.). As the initial pressure and temperature decrease,
detonation cell width increases (Figure 19). Kinetic modeling was done to develop the theoretical results
shown in the Figure 19.
3.2.2.9 Hydrazine-air mixtures
3.2.2.9.1 Concentration
Hydrazine vapor-air mixtures formed at 400 K (261 F) in a 15-cm (6-in.) internal diameter horizontal
detonation tube were initiated with a an E2B blasting cap (approximately 1735 J (1.6 Btu)). In Table 11
the number of tests for which detonations did or did not occur are indicated over 10 percent increments
for different total pressures. The data indicate that for the initiating energy applied, greater concentrations
of air are required to prevent detonation at subatmospheric pressures as follows: 50 percent at 100 kPa
(14.5 psia), 60 percent at 50 kPa (7.25 psia), and 90 percent at 10 kPa (1.45 psia).
3.2.2.9.2 Pressure and velocity
Velocity measurements for the detonation events shown in Table 11 were within 24 percent of values
predicted by equilibrium code C-J calculations.
Calculation of the detonation properties at C-J conditions of a stoichiometric hydrazine-air mixture at an
initial pressure of 101.3 kPa (14.7 psia) indicates that the C-J pressure is 1.96 MPa (284 psia), the
velocity is 2000 m/s (6560 ft/s), and the temperature is 3060 K (5050 F). The C-J pressure, as well as
velocity and temperature for several hydrazine-air mixtures are given in Annex B, Table B-1.
AIAA SP-084-1999
45
Figure 19 Detonation cell width of neat hydrazine at varying pressure and temperature (Pedley et al. 1988)
Table 11 Detonation occurrence as a function of hydrazine vapor concentration in air at 400 K for different total
pressures
a
Concentration Total pressure
% N
2
H
4
100 kPa (14.5 psia) 50 kPa (7.3 psia) 10 kPa (1.5 psia)
100 NT NT NT
90 NT NT NT
80 NT NT NT
70 NT NT NT
60 3Y 1Y NT
50 3N 3Y 1Y
40 NT 3N 2Y/1N
30 1N NT 1Y
20 NT NT 3Y/2N
10 NT NT 1Y/3N
0 NT NT NT
NOTE: It is important to note that the percent air dilution required to prevent a detonation of the
hydrazine vapor can change with a change in the initiation energy.
NT = No test.
#Y = Indicates the number of tests in which a detonation occurred.
#N = Indicates the number of tests in which no detonation occurred.
a
Rathgeber 1997a.
AIAA SP-084-1999
46
3.2.2.9.3 Cell width
Cell width measurements for the detonation events shown in Table 11 fell within the following ranges: 15
to 30 mm (0.59 to 1.18 in.) at 40 percent air dilution, 8 to 40 mm (0.31 to 1.57 in.) at 50 percent air
dilution, and 3 to 20 mm (0.12 to 0.79 in.) at 80 percent air dilution.
3.2.2.10 Hydrazine-diluent mixtures
3.2.2.10.1 Concentration
Hydrazine vapor-diluent mixtures formed at 400 K (261 F) in a 15-cm (6-in.) internal diameter horizontal
detonation tube were initiated with a an E2B blasting cap (approximately 1735 Joules (1.6 Btu)). The
threshold concentration for detonation as a function of pressure is shown in Figure 20 for nitrogen,
hydrogen, helium, and ammonia diluents.
0 10 20 30 40 50 60 70 80 90 100
Pressure (kPa)
0
10
20
30
40
50
60
70
80
90
100
H
y
d
r
a
z
i
n
e

V
a
p
o
r

C
o
n
c
e
n
t
r
a
t
i
o
n

(
%
)
LIMITING CONCENTRATION FOR DETONATION
IN HYDRAZINE VAPOR DILUENT MIXTURES
Detonable Region
Detonable Region
Detonable Region
Non-detonable Region
Non-detonable Region
Non-detonable Region
Nitrogen
a
Hydrogen
b
Helium
c
Ammonia
d
Neat Hydrazine
e
Figure 20 Non-detonable concentrations of hydrazine vapor mixed with diluent gases at 400 K shown below the
indicated curve (Data from
a
Rathgeber 1995;
b
Rathgeber 1997a,
c
1997b,
d
1997c; and
e
Pedley et al.
1988)
Testing with 1 percent hydrazine vapor in helium at 0.55 MPa (80 psia) and 320 K (117 F) does not
detonate when initiated with a 1735 Joule (1.6 Btu) blasting cap.
3.2.2.10.2 Cell width
Cell width measurements were taken for the detonation events shown in Figure 20. Typically the
detonations are characterized by a range of cell sizes. The relative numbers of cell sizes were not
measured.
AIAA SP-084-1999
47
Nitrogen: At 100 kPa (14.5 psia) and 50 percent nitrogen dilution the cell sizes were within the range 2 to
15 mm (0.08 to 0.59 in.). At 50 kPa (7.25 psia) and 60 percent nitrogen dilution cell sizes were within the
range 8 to 20 mm (0.31 to 0.79 in.). At 10 kPa (1.45 psia) and 60 percent nitrogen dilution cell sizes were
within the range 2 to 80 mm (0.08 to 3.15 in.).
Hydrogen: At 100 kPa (14.5 psia) and 50 percent hydrogen dilution cell sizes were within the range 10 to
25 mm (0.39 to 0.98 in.). At 50 kPa (7.25 psia) and 50 percent hydrogen dilution cell sizes were within
the range 8 to 20 mm (0.31 to 0.79 in.). At 10 kPa (1.45 psia) and 60 percent hydrogen dilution cell sizes
were within the range 3 to 55 mm (0.12 to 2.16 in.).
Helium: At 100 kPa (14.5 psia) and 80 percent helium dilution cell sizes were within the range 8 to
20 mm (0.31 to 0.79 in.). At 50 kPa (7.25 psia) and 80 percent helium dilution cell sizes were within the
range 2 to 25 mm (0.08 to 0.98 in.). At 10 kPa (1.45 psia) and 60 percent helium dilution cell sizes were
within the range 3 to 37 mm (0.12 to 1.46 in.).
Ammonia: At 100 kPa (14.5 psia) and 20 percent ammonia dilution cell sizes were within the range 4 to
15 mm (0.16 to 0.59 in.). At 50 kPa (7.25 psia) and 30 percent ammonia dilution cell sizes were within
the range 5 to 15 mm (0.20 to 0.59 in.). At 10 kPa (1.45 psia) and 30 percent ammonia dilution cell sizes
were within the range 2 to 40 mm (0.08 to 1.57 in.).
3.2.2.11 Hydrazine-oxidizer mixtures
3.2.2.11.1 Pressure and velocity
The measured detonation velocities for hydrazine-oxygen mixtures show considerable scatter. The
mixture composition at the time of initiation of the detonation was thought not to be known due to rapid
hydrazine oxidation changes in the test system (W. Jost 1962).
Fairly stable detonations of hydrazine-nitrous oxide mixtures were observed at hydrazine concentrations
from 8 to 100 percent by volume (A. Jost et al. 1963). Using the initial conditions of 101.3 kPa (14.7 psia)
and 367 K (201 F), the calculated CJ parameters of a stoichiometric hydrazine-NO
2
mixture are 3.83
MPa (555 psia), 2500 m/s (8200 ft/s), and 3740 K (6270 F).
3.2.2.11.2 Cell width
No data are reported in the reviewed literature on the detonation cell widths of hydrazine-oxidizer
mixtures.
3.2.3 Liquid hydrazine
The detonation of neat liquid hydrazine has not been observed. Liquid hydrazine of unknown water
content did not detonate in the following standard propellant tests (Tannenbaum and Beardell 1969):
in the Trauzl test, liquid hydrazine did not expand at all whereas nitroglycerin expanded 27 cm
3
/g
(750 in
3
/lb);

in the Drop Weight test, 2 kg (4.4 lb) of liquid hydrazine did not explode at up to 30.5 m (100 ft)
whereas 2 kg (4.4 lb) of nitroglycerin exploded 50 percent of the time at 0.305 m (1 ft);

in the Card Gap test, a sample of hydrazine in a 2.54- cm (1-in.) diameter tube did not explode
with all cards removed whereas a similarly sized sample of nitromethane exploded with 20 cards.
AIAA SP-084-1999
48
Condensed phase testing (Rathgeber 1990) subjected neat hydrazine in a 10.16 cm-inside diameter by
25.4 cm-long (4-in.-inside diameter by 10-in.-long) steel pipe sealed with a 0.64 cm (0.25-in.) thick base
plate to a plane shockwave generated by 887 grams of Composition C4 high explosive. The liquid
hydrazine did not detonate. Results of other tests using driver explosives indicate that liquid hydrazine
does not support stable detonations (Forshey, Cooper, and Doyak 1969; Schmidt 1984; Scott, Burns, and
Lewis 1949; Vander Wall et al. 1971).
Although there have been reported explosions of liquid hydrazine, at present none of these observations
is thought to be the result of liquid detonation but rather they are attributed to the presence of multiple
phases in the liquid hydrazine or other mechanisms, i.e. the result of rapid runaway reaction (thermal
runaway) (Fritchman and Benz 1980; Section 3.3.2), water-hammer effects, rapid compression (Vander
Wall et al. 1971; Briles and Hollenbaugh 1978; Section 3.4), vapor detonation (Fritchman and Benz 1980;
Section 3.2.1), or flashback from a vapor detonation (Benz and Pippen 1980; Section 3.2.1). If they
occur, explosions of liquid hydrazine can lead to very large explosion pressures and severe system
damage.
Transition to detonation may occur in liquid hydrazine, especially when the liquid is in motion, but no
conclusive studies have been performed. No data are reported in the reviewed literature on the DDT of
liquid hydrazine.
3.2.3.1 High purity
No data are reported in the reviewed literature on the critical energy and minimum dimensions required
for detonation of high purity liquid hydrazine.
3.2.3.2 Monopropellant
Negative results for the detonation of monopropellant liquid hydrazine were obtained by Scott, Burns, and
Lewis (1949). Subsequent testing performed with monopropellant grade hydrazine (MIL-P-26536D) in
1.3 and 5.08 cm (0.5 and 2.0 in.) diameter stainless steel tubes (Beeson et. al. 1988) and 10.16 cm
(4.0 in.) diameter stainless steel tubes (Rathgeber 1990) did not produce evidence of a C-J detonation.
These results indicate the critical diameter for liquid hydrazine is greater than 10.16 cm (4.0 in.).
3.2.3.3 Mixtures
Negative results were obtained for liquid mixtures of hydrazine, monomethylhydrazine, and hydrazine
nitrate at low hydrazine nitrate concentrations (Tannenbaum 1962). As the hydrazine nitrate
concentration increases above 45 percent by weight, the mixture becomes detonable.
3.3 Thermochemical reaction
Any thermochemical reaction can lead to a potential explosion hazard if one or more of three conditions
are met: (1) the reacting system is confined to a rigid volume; (2) the reacting system is adiabatic or
nearly adiabatic; or (3) the rate of heat generated exceeds the rate of heat exchange with the
environment (Ayeni 1978). Hydrazine alone or in the presence of oxidizers is such a system. This
section describes three cases that could lead to an explosion, resulting from the operation of an hydrazine
system:
near-isothermal hydrazine decomposition,
thermal runaway, and
rapid compression.
AIAA SP-084-1999
49
3.3.1 Thermodynamic instability
Hydrazine is thermodynamically unstable. Decomposition of a hydrazine molecule produces an increase
in the number of gaseous molecules (Annex B). In a closed, rigid system, under isothermal conditions,
pressure will increase. In non-isothermal systems, estimates of maximum final temperature and pressure
can be calculated using Equations 5 and 6, respectively.
3.3.2 Thermal runaway
If the hydrazine system temperature increases because more thermal energy is retained than is
dissipated during hydrazine reaction, thermal runaway can result.
The rate of thermal decomposition or oxidation of hydrazine (except with hypergols) is small under
ambient conditions. However, the presence of catalytic materials can lead to rapid acceleration of these
reactions. Hydrazine and material compatibility is covered in Section 4.0.
The pressure generation rate observed with a typical thermal runaway confined in a system is shown in
Figure 21. The pressure-time curve shows three distinct pressure areas: the initial pressure rise, the
high pressure area, and the region of pressure decay. During the initial pressure rise, the reaction self-
accelerates at a rate governed by the chemical kinetics of the reaction. The pressure maximum results
when further pressure increase is limited by amounts of reactants or chemical dissociation. Pressure
decay occurs as cooling from contact with the containment walls takes effect and as possible pressure
relief devices or leaks reduce the pressure (Kinney and Graham 1985).
The maximum rate of pressure generation for a given composition or mixture with similar vessel shape,
vessel length-to-diameter ratio, degree of turbulence, and point of ignition, is suggested by Bodurtha
(1980) (Equation 17):
dP
dt

1
V
1
3
(17)
where
dt
dP
= pressure generation rate (kPa/s)
V = volume of vessel (m
3
)
Central ignition or multiple ignition points lead to the greatest maximum rate of pressure
dt
dP
increase
(Bodurtha 1980).
AIAA SP-084-1999
50
Figure 21 Typical pressure versus time profile for confined explosions (a confined thermal runaway will produce
the same result.)
The effects of several factors on explosion pressure and pressure generation rate in closed, non-adiabatic
systems can be summarized briefly:
given a constant initial pressure, an increased initial temperature leads to a decreased maximum
reaction pressure due to the reduction in mixture density (Benz, Bishop, and Pedley 1988),
increased initial system pressure leads to increased final reaction pressure (Benz, Bishop, and
Pedley 1988),

turbulence increases the rate of pressurization and leads to increased final system pressure,
especially near the lower and upper flammability limits (Harris 1967),

for fixed initial temperature and pressure inert diluents normally decrease final reaction pressure
(Kanury 1976), and

vessel volume or shape does not usually affect explosion final reaction pressure unless the
surface area of the vessel is large; in spherical chambers, the pressure generation rate
decreases as the vessel volume increases (Harris 1967; Nagy et al. 1971).
Final reaction pressures for adiabatic processes can be calculated using Equations 5 and 6.
When hydrazine is present in a system, hazard assessment must always include the effects of hydrazine
decomposition leading to an explosion. The effects can be approximately evaluated by calculating the
maximum pressure produced by thermodynamic equilibrium, i.e. maximum heat generation and maximum
temperature. The thermal runaway reaction is relatively slow (as compared to detonation) and systems
fail when the burst pressure of the weakest component of the system is reached. Therefore pressure-
relief devices can usually provide protection from thermal runaway.
AIAA SP-084-1999
51
3.3.2.1 Hydrazine vapor
Thermal runaway of hydrazine can occur in vapor, but because of the limited amount of fuel available in
vapor, runaway reactions of hydrazine vapor mixtures produce less pressure and heat than runaway
reactions of liquid hydrazine (Benz, Bishop, and Pedley 1988).
Although hydrazine is hypergolic with dinitrogen tetroxide vapor at ambient temperature, hydrazine vapor
reacts slowly with air and oxygen at ambient temperature unless catalysts are present. Concentrations of
60-100 ppm of hydrazine in air were found to drop at an approximate rate of 25% in 24 h in the absence
of catalysts (Martin, Davis, Kilduff, Mahone 1989). Vapor hydrazine exposed to catalytic materials causes
localized surface heating (see Section 4) that may be sufficient to cause ignition (Miller and Schluter
1978). This ignition and the subsequent deflagration appear to be the major hazard associated with
hydrazine vapor. For information on the effects of vapor hydrazine exposed to various materials, refer to
Section 4.
3.3.2.2 Liquid hydrazine
At high temperature or in the presence of a catalyst liquid hydrazine can undergo rapid decomposition
and produce violent thermal explosions. The conditions for a thermal explosion depend on the rate of
heat generated by hydrazine decomposition and the rate of heat lost from the system.
The heat generation rate is described by Barrow (1973) in Equation 18:
Q
rx
HSP
n
V
RT


_
,
Aexp
E
a
RT


1
]
(18)
where
Q
rx
= heat generation rate (J/s)
A = preexponential factor (kPa
1-n
/(scm
2
), depending on the reaction order n
E
a
= activation energy (J/mol)
R = gas constant (J/(molK)
T = temperature (K)
S = surface area (cm
2
)
V = volume (m
3
)
H = heat of reaction (J/mol)
P = pressure (kPa)
n = reaction order (dimensionless)
The terms needed to solve this equation are obtained from kinetic analysis of data from tests under near-
adiabatic conditions. Because liquid hydrazine is highly susceptible to surface catalysis, the heat
generation rate varies with exposure to different materials, and the kinetic parameters in Equation 18
must be determined for each material. Plots of heat generation rate versus temperature for various
materials are given in Section 4.2.
As with vapor systems, thermal runaway in a liquid hydrazine system occurs when the rate of generation
of thermal energy exceeds the rate that thermal energy is dissipated to the surroundings. In a closed
system even with isothermal operation, the decomposition of hydrazine can lead to pressure buildup.
The larger quantities can produce very high pressures. This process can be modeled, including the case
where catalytic decomposition is occurring, if the appropriate Arrhenius rate parameters are available
(See Section 4).
AIAA SP-084-1999
52
3.3.3 Rapid compression
Hydrazine systems under flow can result in potential hot spots developing by the rapid compression of a
noncondensable ullage gas, the rapid compression of residual hydrazine vapor, or the rapid compression
of a froth generated within the system during flow. These hot spots can lead to accelerated hydrazine
decomposition or reaction.
3.3.3.1 Adiabatic heating of an ullage gas
When an ullage gas or vapor is compressed, the temperature of the gas or vapor and the surroundings
increases. If the compression process is adiabatic, compression occurs without heat flow in or out of the
substance being compressed (Parker 1984). The change in internal energy and the resulting
temperature increase of the gas can be determined from Equations 19 and 20 (Smith and Van Ness
1975):
W U
RT
i
1
1
P
f
P
i



_
,

1 ( )





1
]
1
1
(19)
T
f
T
i
P
f
P
i



_
,

1 ( )
(20)
where
-W = work performed by the compression process (J/mol)
U = internal energy (J/mol)
R = gas constant (J/(molK))
T
i
= initial temperature (K)
T
f
= final temperature (K)
P
i
= initial pressure (kPa)
P
f
= final pressure (kPa)
= the ratio of the specific heats = C
p
/C
v
3.3.3.2 Criteria for initiation of rapid compression
Recent studies in transparent glass and stainless steel chambers indicate that the explosion process may
be an exothermic reaction occurring in a froth formed at the liquid-gas interface between hydrazine and
ullage gases (Baker et al. 1988, Bunker et al. 1990). This is based on the following. Operating
temperatures typically found in liquid hydrazine flow systems do not approach the autoignition
temperature for hydrazine. While hydrazine vapor burns (see Section 2) and detonates (Pedley et al.
1988), bulk hydrazine liquid does not detonate (Beeson, Plaster, and Bunker 1988). Some researchers
have described rapid compression events as detonations (Hagemann, Briles, and Farkas 1983) but the
damage to steel lines reported does not indicate detonation of hydrazine vapor as a likely cause. Comer
(1977) has shown with hydrazine monopropellants used as gun propellants that ignition is critically
dependent upon the initiation, growth and propagation of gas cavities, i.e. froth, formed by Taylor
instabilities at the liquid-gas interface and Helmholtz mixing of gas and liquid. No explosive
AIAA SP-084-1999
53
decomposition has been observed when a stationary column of liquid hydrazine without froth was rapidly
compressed (Briles et al. 1985). Therefore when the froth is compressed sufficiently and rapidly enough
such that heat transfer to the hydrazine liquid in the film is large compared to heat loss to the system,
rapid heating will cause decomposition of the hydrazine throughout the froth, generating more thermal
energy and rapidly accelerating the temperature rise, which can lead to thermal run away.
3.3.3.3 Surge pressures
Water hammer is a potential problem in any system involving pressurized liquid flow. Equipment rupture
may result because of water hammer. In systems in which the fluid media can undergo exothermic
decomposition, adiabatic heating can lead to accelerated exothermic reactions and, possibly, explosive
events.
Rapid compression hazards were observed early in the development phase of the Space Shuttle Auxiliary
Power Unit (APU) and have been duplicated in experimental hydrazine systems (Benz, Long, and Weary
1984; Briles et al. 1985; and Vander Wall et al. 1971).
When hydrazine liquid (pressurized) is suddenly released into a system with ullage gas present, such as
a line with a dead-end, compression of the ullage gas causes the gas to heat adiabatically. This hot gas
in contact with hydrazine liquid can accelerate hydrazine decomposition (Baker et al. 1988, Bunker et al.
1990). The surge pressures are large compared to the pressurization (push) pressures that accelerate
the liquid through the line.
An estimate of surge pressures can be obtained by Equations 21 through 23 developed for water-
hammer effects in lines (Hwang 1981).
P V
o
C (21)
where
P = water hammer pressure Pa
= density of the liquid (kg/m
3
)
V
o
= velocity of the column of liquid (m/s)
C = velocity of the wave traveling along the pipe, celerity (m/s)
Celerity can be determined by
C
E
c

(22)
and E
c
can be determined from the elasticity of the pipe walls, E
p
, and the elasticity of the fluid, E
f
, as
follows:
1
E
c

1
E
f
+
Dk
E
p
t
(23)
where
D = diameter of the pipe (m)
k = constant that depends on the method of anchoring the pipe (k is usually = 1)
E
f
= modulus of elasticity of the fluid (Pa)
E
p
= modulus of elasticity of the pipe (Pa)
t = thickness of the pipe wall (m)
The dependence of hydrazine decomposition on hydrodynamic surge pressure has been studied (Baker
et al. 1988, Bunker et. al. 1990). Tests were performed in which a column of liquid propellant-grade
AIAA SP-084-1999
54
hydrazine, aqueous hydrazine mixtures, or water was accelerated into a chamber filled with nitrogen at
ambient temperatures. In these tests, pressurization pressure was increased and the resulting pressure
was measured at the dead-end. Results of these tests indicate that as the surge pressures, Equation 20,
exceeded 17 MPa (2500 psia), the resulting pressures generated at the dead-end were much larger than
those obtained with water (Figure 22). This excess pressure increase was attributed to hydrazine
decomposition. When water was added to the hydrazine, larger hydrodynamic surge pressures were
required to initiate hydrazine decomposition.
Note: Readers are cautioned that when estimating the internal energy and adiabatic temperature of the
compressed ullage gas (Equations 18 and 19) the surge pressure and not push pressure is used as the
final pressure.
3.3.3.4 Effects of ullage gas on hydrodynamic surge pressure and hydrazine decomposition
The effects of ullage gas pressure on hydrodynamic surge pressure and hydrazine decomposition using
water and monopropellant-grade hydrazine were studied by Baker et al. (1988). In these tests at ambient
temperature, water or hydrazine was accelerated into a chamber filled with nitrogen at various pressures.
The hydrodynamic surge or resulting pressure was measured at the chamber dead-end (Table 12).
Figure 22 Measured pressures when 20-cm (8-in.) liquid column is accelerated into 38-cm (15 -in.) long, 1.27-cm
(0.5-in.) diameter chamber at various hydrodynamic surge pressures at ambient temperature
In tests with water, the surge pressure decreased as the ullage pressure increased. These results
indicate that the pressure drop driving fluid flow, i. e. the difference between the push pressure and the
ullage pressure, decreases as ullage pressure increases. This causes the surge pressure to decrease.
Similar results were reported by Knowles (1981) and Prickett, Mayer, and Hermel (1988).
AIAA SP-084-1999
55
Table 12 Effects of gas ullage pressure on surge pressure and hydrazine decomposition
a
Surge pressure
Ullage pressure Water Hydrazine
kPa psia MPa psia MPa psia
0.005 7.3E-4 131 19,005 131
b
19,005
b
7 1.02 90
b
13,057
b
203
b
29,451
b
14 2.03 96
b
13,927
b
183
b
26,549
b
21 3.05 86
b
12,477
b
141
b
20,456
b
28 4.06 79
b
11,461
b
196
b
28,435
b
41 5.9 76 11,026 210
b
30,446
b
55 8.0 69 10,010 169
b
24,518
b
69 10.0 69 10,010
c c
85 12.3 69 10,010
c c
99 14.7 62 8995
c c
113 16.4 55 7979
c c
127 18.4 45 6529
c c
134 19.4 28 4062 179
b
25,969
b
141 20.5 24 3482 48 6964
a
20-cm (8-in.) liquid column accelerated into a 38-cm (15-in.) long, 1.3-cm (0.5-in.) diameter
chamber; push pressure at 2 MPa (300 psia)
b
Chamber yielded, no rupture
c
Chamber ruptured, pressure increase was greater than 276 MPa (40,000 psia), the burst
strength of the tubing.
Resulting pressures with hydrazine show a different trend. At ullage pressures less than 0.005 kPa
(0.0007 psia), the resulting pressures were similar to hydrodynamic surge pressures seen in water, and
hydrazine decomposition was not initiated. As the ullage gas pressure was increased to 134 kPa
(19.4 psia), the resulting pressures increased to values large enough to cause chamber ruptures. As the
ullage pressure was increased above 134 kPa (19.4 psia), resulting pressures decreased and
approached hydrodynamic surge pressures observed with water. These results suggest adiabatic
compression of the ullage gas caused the ullage gas to reach a sufficient temperature (Equation 19) to
initiate the exothermic decomposition of the hydrazine (froth). This occurred until the initial ullage gas
pressure reached a value sufficiently high that the final temperature would not initiate exothermic
decomposition.
The effects of ullage gas volume on hydrazine decomposition were also studied by Baker et al. (1988),
and the results are shown in Table 13 for water and monopropellant-grade hydrazine. In these tests,
water or hydrazine was accelerated into variable-length chambers filled with nitrogen at ambient
temperature. The push pressure was kept constant, and the surge or resulting pressures were measured
at the chamber dead-end. In tests with water, the surge pressures generally decreased as the length or
volume of the bottom chamber increased. In tests with hydrazine, chamber lengths less than 76 cm
(30 in.) caused the chambers either to yield or rupture. As the chamber length was increased above
76 cm (30 in.), however, the resulting pressures decreased and approached the hydrodynamic surge
pressure observed with water.
AIAA SP-084-1999
56
Table 13 Effects of gas ullage volume on surge pressure and hydrazine decomposition
a
Tube
length
Tube
volume
Surge pressure
water
Resulting pressure
hydrazine
cm in. cm
3
in
3
MPa psia MPa psia
13 33 12 197 62 8995 230
b
33,368
b
25 64 24 393 62 8995
c c
38 97 36 590 69 10,010
c c
51 130 48 787 55 7979
c c
64 163 60 983 48 6964
c c
76 193 72 1118 45 6529
c c
89 226 84 1377 31 4497 96
b
13,927
b
102 259 96 1573 28 4062 38 5513
a
20-cm (8-in.) liquid column accelerated into 1.3-cm (0.5-in.) diameter chamber; ullage pressure
at 85 kPa (12.4 psia); push pressure at 2 MPa (300 psia)
b
Chamber yielded, no rupture
c
Chamber ruptured, pressure increased beyond 276 MPa (40,000 psia), the burst strength of the
tubing.
3.3.3.5 Design criteria
Decomposition of hydrazine is initiated by adiabatic heating of the ullage gas when compressed with a
liquid hydrazine column and is dependent on the hydrodynamic surge pressure generated when a liquid
column impacts at a dead-end. Surge pressures depend on the configuration and operational conditions
of the system. Based on results by Baker et al. (1988), systems containing gas ullages should be
designed to avoid surge pressures greater than 17 MPa (2500 psia). Although careful design of
manifolds, orifices, and line lengths, and proper selection of ullage gas pressures are important in
avoiding large surge pressures, estimation of surge pressure in systems is very difficult. Therefore,
configuration tests should be performed to ensure that surge pressures are below the recommended
value. These tests can be performed with water to predict hydrazine surge pressures in the system. The
measured surge pressures must be adjusted for the difference in bulk modulus between water and
hydrazine. Systematic studies of the effect of tubing configuration and components on surge pressure
have been conducted (Rathgeber 1992, Webb 1992, Webb (2) 1993, Webb, Poe, and Peterson 1992).
Systems that operate at greater than ambient temperatures should be tested with hydrazine because
surge pressures required to initiate hydrazine decomposition have not been fully determined at elevated
temperatures.
Design issues associated with surge in hydrazine are summarized as follows:
decomposition correlates with dynamic surge pressure,

the greater the difference between supply and ullage pressure the greater the surge,

decomposition is more likely to occur for higher temperatures for a given dynamic surge,

lighter push gases produce greater surges, i.e. He>N
2
,

Surge increases with line length for a given push pressure,

elbows and turns in the line do not reduce surge in evacuated ullages,

AIAA SP-084-1999
57
orifices can be used to reduce surge so long as the pressure drop across the orifice does not
produce cavitation and adiabatic heating of collapsing bubbles sufficient to initiate a thermal
runaway,

venturi devices can be used to reduce surge so long as the collapse of the bubbles produced by
cavitation does not cause sufficient adiabatic heating to initiate a thermal runaway,

fast valve openings can contribute to higher surge pressures, and

greater tubing elasticity contributes to an increase in surge pressure.
AIAA SP-084-1999
58
4 Hydrazine and material compatibility
Introduction
This section considers the effect of hydrazine contacting materials of construction normally used in
hydrazine systems and inadvertently contacting other materials used in aerospace operations. The
materials considered have been identified by Tauber and Lee.*
The chemical reaction of hydrazine with materials can result in loss of the material's integrity, i.e.
corrosion, loss of ductility, soluble, etc. or in decomposition of the hydrazine, i.e. catalytic reaction. The
loss of a materials integrity can result in loss of function for the system component. The decomposition
of hydrazine can lead to either system pressurization (Section 3.1) or to thermal runaway (Section 3.3) or
both.
When assessing a material's compatibility with hydrazine, it is important to carefully detail the conditions
of the system and its environment. Important considerations include the phases present, the
concentration and purity of the hydrazine, the surface area of the material and the temperature and
pressure of the system. The condition of the material surface is also important, including the presence of
oxide or other coatings and the possibility of surface contamination. The extent of the hydrazine contact
with the material assists the evaluation of the potential hazard, e.g. a material may have adequate
compatibility for short or inadvertent contact but be unacceptable for routine immersion in hydrazine. Also
materials may be acceptable for low temperature applications but unacceptable for higher temperature
applications.
This section presents available information for assessing hydrazine and material compatibility. The
reader is advised that while the authors have attempted to present accurate information, extrapolation to
systems other than those described should proceed cautiously.
4.1 Material degradation
Materials are categorized as either metals or nonmetals. The effect of hydrazine on metals is usually
measured in terms of corrosion (loss of mass) although some metals exhibit stress corrosion cracking.
The effect of hydrazine on nonmetals is reported in terms of a change in the mechanical or physical
properties of the material. In both cases, any visual change in the hydrazine is also reported.
4.1.1 Test method
The gross compatibility of materials with liquid hydrazine as a function of exposure time and temperature
is determined by the immersion test, Reactivity of Materials in Aerospace Fluids (Test 15), from NASA
(1991). In Test 15, the total pressure and test temperature must simulate the worst-case use
environment for degradation of material or fluid. The material is immersed in hydrazine liquid, under its
own vapor pressure plus an ambient nitrogen blanket, elevated to a minimum temperature of 344.1 K
(160 F), and held for 48 h or until sufficient hydrazine decomposition occurs to end the test as shown by
an excessive pressure increase.
4.1.2 Test conditions
The major controlled variables for compatibility tests are contact time, temperature and contaminants.

* Taeuber, R. and L. Lee, NASA Johnson Space Center, Houston, TX; private communication, undated.
AIAA SP-084-1999
59
4.1.2.1 Contact time
Experiments designed to test the long-term hydrazine-material compatibility have been performed. Initial
efforts are reported by Toth et al. (1976). A more extensive test was initiated by Moran and Bjorklund
(1982). Several attempts have been made to evaluate long-term compatibility on the basis of short-term
tests. No consistent data was obtained.
4.1.2.2 Temperature
In general, increased temperature decreases material compatibility with hydrazine. Specifically, many
nonmetals that have been accepted at ambient temperature degrade at elevated temperatures.
4.1.2.3 Contaminants
The presence of contaminants generally decreases the compatibility of materials with hydrazine. In
particular, the presence of CO
2
, acidic impurities, and halogenated solvents were reported to cause
material failure. Because of this potential for embrittlement, chlorofluorocarbon solvents, such as Freons,
must not be used to clean stainless steel lines that will be exposed to hydrazine.* Carbazic acid forms
when CO
2
is present in hydrazine, increasing the corrosion of metals (Green, Pullen, and Stebbins 1974;
Green et al. 1973) and causing failure of nonmetals (Schmidt 1984). The effects of CO
2
contamination on
the corrosion rates of alloys were studied using electrochemical impedance spectroscopy (Bhardwaj
1996). A summary of these results is given in Section 4.1.4. The compatibility of aluminum with
hydrazine decreases as the water content of the hydrazine increases (Schmidt 1984).
4.1.3 Material compatibility data
Table 14, Metals, and Table 15, Nonmetals, give the available compatibility data for hydrazine and
materials. The testing to obtain material compatibility data very seldom can separate the effect of
hydrazine on materials and the effect of materials on hydrazine; therefore, the user should examine
available data in Section 4.2 when making a material selection decision.
Table 14 includes the following data about metals:
the identity of the material indicated, when possible, by a chemical formula or chemical name,

the temperature at which the material was tested,

the extent to which corrosion occurred given in mil/year (1 mil= 0.001 in. = 0.00254 cm) or
percent hydrazine decomposition by N
2
per year,
testing results, and

the source of the data (in footnotes at the end of Table 14).
Table 15 includes the following data about nonmetals:
the identity of the material indicated, when possible, by a chemical formula or chemical name,
the temperature at which the material was tested,

*
NASA Pre-Alert E4-70-03A (Form 863). W. H. Lockyear, Alert Coordinator, Jet Propulsion Laboratory, CA,
Aug. 16, 1971.
AIAA SP-084-1999
60
physical changes seen in the material,
mechanical property changes seen in the material,
compatibility test results, and
the source of the data (in footnotes at the end of Table 15).
4.1.4 Alloy corrosion studies
Electrochemical Impedance Spectroscopy (EIS) has been used to study the effect of CO
2
concentrations
on alloy corrosion rates (Bhardwaj 1996). Iron (Fe), nickel (Ni), cobalt (Co) and titanium (Ti) alloys were
used. The corrosion study used monopropellant grade hydrazine (MIL-P-26536D) and dry ice (CO
2
)
dissolved in hydrazine as the contaminated test fluid. Testing conditions were performed at 298 t 3 K, at
ambient pressure and with CO
2
concentrations of 0, 25, 50, 100, 200, and 414 ppm. Testing conditions
were performed in an electrochemical cell. By utilizing the Stern Geary equation (Stern and Roth 1957)
and experimentally determining polarization resistance values the corrosion current density was
calculated. The corrosion rates in mils per year were then calculated using the Faraday law in terms of
penetration rate. The accelerated corrosion of the alloys in the presence of CO
2
may be attributed to the
action of carbazic acid (N
2
H
3
CO
2
H).
The EIS study found the following: nickel and cobalt alloys are more prone to corrosion than iron and
titanium alloys; the corrosion rates as a function of time and CO
2
concentration for the iron and titanium
alloys increased while the corrosion rates decreased for the nickel and cobalt alloys in CO
2
-contaminated
hydrazine; when tested in hydrazine, the titanium alloy exhibited the least amount of corrosion while the
nickel alloys exhibited the most amount of corrosion. Corrosion rates are shown Table 16.
4.2 Material effects on hydrazine
This section evaluates hydrazine-material compatibility using the decomposition of the hydrazine as the
criterion for assessment of potential hazards. Hydrazine system applications span a wide-range of
operating periods, e.g. some applications have a ten-year operating period, and while decomposition at
ambient conditions for a selected material may be slow, these long periods may lead to appreciable
degradation potential. Also, the material compatibility data must be available for liquid hydrazine, vapor
hydrazine or both.
4.2.1 Test methods
Although a standardized test method has not been established for determining rate of decomposition of
hydrazine on various materials, four methods have been consistently selected. They are:
1. Isothermal rate of gas evolution, as measured by volume (Sayer 1968; Sayer 1971; Taylor et al.
1969; Sayer 1972; Sayer 1974).

2. Isothermal rate of gas evolution, as measured by pressure (Axworthy et al. 1969; Chang and
Gokcen, 1976; Bennett et al., 1980).
Table 14 Effects of hydrazine on metals
Temperature Corrosion rate,
Material K F mil/year, unless
otherwise indicated
Test period Testing
results
Reference
Aluminum Compatible a
Aluminum 1100 333 140 <1 b
Aluminum 2014 300 80 <1 b
Aluminum 2014 11 years Not good c
Aluminum 2014-T4 300 80 <1 d
Aluminum 2021 10 years Excellent c
Aluminum 2024 294 70 <1 b
Aluminum 2219 10 years Excellent c
Aluminum 3003-0 300 80 <1 b
Aluminum 356-T6 316 109 <1 b
Aluminum 4043 344 160 <1 b
Aluminum 6061-0 344 160 <1 b
Aluminum 6061-T6 316 109 <1 b
Chromium Coating, CVD 294 70 <1 b
Chromium-Plate Coating 294 70 <1 b
Cobalt NC e
Cobalt, Pure >50 d
Copper >50 d
Gold 297 75 <1 b
Hastelloy C 325 125 <1 b
Hastelloy C NC e
Inconel X-750 300 80 1 b
Iron, Pure >50 b
Molybdenum NC e
Molybdenum, Pure >50 d
Nickel, Pure <1 d
Niobium NC e
6
1
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 14 Effects of hydrazine on metals (continued)
Temperature Corrosion rate,
Material K F mil/year, unless
otherwise indicated
Test period Testing
results
Reference
Steel, 17-7 PH 10 years Very good c
Steel, A-286 10 years Excellent c
Steel, AISI 301SS (aged) 11 years Excellent c
Steel, AISI 301SS (cryostretched) 14 years Good c
Steel, AISI 301SS (unaged) 11 years Excellent c
Steel, CRES 17-4 PH 333 140 <1 b, d
Steel, CRES 17-4 PH 344 160 >50 d
Steel, CRES 17-7 PH 297 75 <1 b
Steel, CRES 302 300 80 <1 b
Steel, CRES 303 >50 d
Steel, 304 Compatible a
Steel, 304 NC e
Steel, CRES 304 333 140 <1 b
Steel, CRES 304L <1 d
Steel, CRES 316 366 200 <1 b
Steel, CRES 316 366 200 >50 d
Steel, CRES 321 333 140 <1 b
Steel, CRES 321 333 140 1-5 d
Steel, CRES 347 366 200 <1 b
Steel, CRES 410 300 80 <1 b
Steel, CRES 410 300 80 1-5 d
Steel, CRES 416 366 200 <1 b, f
Steel, CRES 416 366 200 >50 d
Steel, 430 344 160 NC 94-28381
g
Steel, CRES 430 293 68 <1 b
Steel, CRES 430 293 68 1-5 d
Steel, CRES 440C 300 80 <1 b
Steel, CRES AM350 344 160 <1 b
Steel, CRES AM350 10 years Very good c
Steel, CRES AM350 316 109 0.11 %/year 174 days h
Steel, CRES AM350 316 109 0.25 %/year 328 days h
Steel, CRES AM350 333 140 0.32 %/year 174 days h
6
2
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 14 Effects of hydrazine on metals (continued)
Temperature Corrosion rate,
Material K F mil/year, unless
otherwise indicated
Test period Testing
results
Reference
Steel, CRES AM350 333 140 0.35 %/year 328 days h
Steel, CRES AM350 HZ+CO
2
316 109 0.40 %/year 182 days h
Steel, CRES AM350 HZ+CO
2
316 109 0.45 %/year 293 days h
Steel, CRES AM350 HZ+CO
2
333 140 1.38 %/year 293 days h
Steel, CRES AM355 344 160 <1 b
Steel, CRES AM355 316 109 0.34 %/year 174 days h
STEEL, CRES AM355 316 109 1.34 %/year 328 days h
Steel, CRES AM355 316 109 0.019 %/year 4732 days h
Steel, CRES AM355 333 140 6.7 %/year 174 days h
Steel, CRES AM355 HZ+CO
2
316 109 2.05 %/year 182 days h
Steel, CRES AM355 HZ+CO
2
316 109 4.78 %/year 293 days h
Steel, CRES AM355 HZ+CO
2
333 140 3.8 %/year 182 days h
Steel, M-2 Tool 394 250 i
Tantalum, Pure 300 80 <1 b
Titanium, 13Nb-13Zr Diffusion Hardened 344 160 NC 95-29434
g
Titanium, 6Al-4V HZ+CO
2
316 109 0.123 %/year 293 days j
Titanium, 6Al-4V HZ+CO
2
316 109 0.07 %/year 182 days j
Titanium, 6Al-4V HZ+CO
2
333 140 0.343 %/year 293 days j
Titanium, 6Al-4V HZ+CO
2
333 140 0.14 %/year 182 days j
Titanium, 6Al-4V 344 160 <1 b
Titanium, 6Al-4V 316 109 0.020 %/year 293 days j
Titanium, 6Al-4V 316 109 0.020 %/year 139 days j
Titanium, 6Al-4V 333 140 0.046 %/year 293 days j
Titanium, 6Al-4V 333 140 0.06 %/year 139 days j
Zirconium, Pure 297 75 <1 b
6
3
A
I
A
A

A
P
-
0
1
6
-
2
-
1
9
9
9
Table 14 Effects of hydrazine on metals (continued)
NOTES: NC denotes no change in posttest.
Mil per year = 0.001 in. per year
a
Kilduff (1990); Those materials which exceed two or more of the criteria for the posttest material and fluid analyses are considered
incompatible by the MCT procedure and not recommended for use with that fluid.
b
Salvinski, Howell, and Lee (1967)
c
Gill (1986); Compatibility rating based on measurements in change in pressure in each tank, analysis on propellant
impurity and metallurgical examination of the interior tank surfaces and welds.
d
Marsh and Knox (1970)
e
Davis (1993); no visual changes in the metal samples were observed in the posttest material and fluid analysis.
f
Pitting was observed.
g
Tests performed at NASA Johnson Space Center White Sands Test Facility (Contact: NASA Laboratories Office, NASA
JSC WSTF, P. O. Box 20, Las Cruces, NM 88004.)
h
Lawton (1985); Corrosion is given in hydrazine decomposition by N
2
in percent per year. Those entries with the term HZ + CO
2

indicate tests done with carbon dioxide contamination.


i
Worthington (1986); pitting was observed, no corrosion rate was reported. NA Not available
j
Lawton (1985): Those entries with the term HZ + CO
2
indicate tests done with carbon dioxide contamination.
6
4
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
ADHESIVES
Epoxi-Patch, Kit 1C 300 80 Yes c
Epoxylite 5403 300 80 Yes c
COATINGS
Akzo 663-3-21/X-310A
Polyurethane Topcoat
344 160 1.4 %
Weight loss
Coating changed from
rigid coating to
softened, flexible and
wrinkled.
48 h 94-28373
d
Epoxy Polyamide 300 80 Yes c
Koropon, Super High Temp
(Superkoropon 515-700)
298 77 No 78-9978
d
LUBRICANTS AND OILS
Brayco 815Z Oil 344 160 Yes 75-5134
d
DuoSeal
344 160 NC 48 h 94-28333
d
Fomblin Y25 344 160 Changed from
transparent colorless,
adherent liquid to
translucent milky.
48 h 94-28332
d
Fyrquel 220 344 160 Changed from
transparent,
yellow/green adherent
liquid to yellow.
48 h 94-28331
d
Houghto-Safe 1055 344 160 Material was partly
dissolved.
48 h 94-28344
d
Houghto-Safe 1120 344 160 Changed from light
green to translucent
cloudy white.
48 h 94-28345
d
6
5
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals (continued)
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
LUBRICANTS AND OILS (continued)
Krytox 143 AC Oil 344 160 No 76-5782
d
Mobil Jet Oil II 298 77 Yes 78-9775
d
Zonyl FSN Fluorosurfactant 298 77 Not Compatible 2 h 95-28920
d
PLASTICS AND ELASTOMERS
AFE-332 Yes e
AFE-332 344 160 No Varies f
AF-E-332 316 109 No serious
degradation of
polymer observed.
3 years g
AF-E-411 450 350 Slight 76-8590
d
Butyl Rubber B591-80 316 110 No Yes 77-9212
d
Butyl Rubber 344 160 No e
Butyl Rubber B318 Surface crazing and
some blistering in
posttest
k
EPR 344 160 No e
EPR E515-80 344 160 No Varies 81-13810
d
EPR E692
Not Compatible
l
EPR E740 Not Compatible l
White EPR Not Compatible l
COMPOSITES
COPV Brunswick Corp. Model
No. 220088-1
No significant
effect on burst
strength.
96-30013
d
COPV Graphite/Epoxy SCI-
REZ 100 Matrix with Toray-
T 10006B Carbon Fiber
None TR-804-001
d
COPV Lincoln Composites
Model
None TR-804-002
m
6
6
A
I
A
A
S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals (continued)
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
COPV SCI Model AC 5229 No significant
effect on burst
strength.
TR-804-003
n
COPV SCI Model AC 5229 No significant
effect on burst
strength.
96-29786
d
EPDM, Polymerized Yes 8 weeks
immersion,
200 h stress
j
Glass, Borosilicate Plate e
Graphite/Epoxy Laminate SCI 344 160 1.6% Weight
gain
None Posttest material
was delaminated
48 h 94-28426
d
Halar R-CTFE 344 160 No
76-8674
d
Kadel 1230 (polyketone) 296 73 40% increase in
elongation; 6%
decrease in yield
strength
12 days j
Kadel E 100 (polyketone) 296 73 79% decrease in
elongation; 5.2%
decrease in yield
strength
12 days i
Kadel E 1140 (polyketone) 296 73 27% decrease in
elongation; 0%
change in yield
strength
12 days i
Kadel EP 3140 (polyketone) 296 73 17% decrease in
elongation; 1.7%
increase in yield
strength
12 days i
Kalrez 1045 344 160 No 76-8758
d
Kel-F 81 Slightly darker in
posttest fluid
analysis
k
6
7
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals (continued)
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
Kynar 460 Compatible;
Slightly darker in
posttest fluid
analysis; Not
Compatible
k
k,l
376 218 Yes Not Recommended
79-11761
d
MFX 7008 (polypropylene) 296 73 33% decrease in
elongation; 13%
decrease in yield
strength
12 days i
Mylar, Polyester Type A Film Yes d
Nylon 6/6 Not Compatible l
OF-700-07 (polyphenylene
sulfide)
296 73 200% increase in
elongation; no
information on
yield strength.
12 days i
PEEK 450G (polyether
Etherketone)
296 73 0.22% decrease in
elongation; 10%
decrease in yield
strength.
12 days i
PES 4100G (poly-
ethersulfone)
296 73 91% decrease in
elongation; 33%
decrease in yield
strength
12 days i
Poly-chlorotrifluoro-ethylene,
Kel-F81
>344 >160 Yes NA Not Recommended 76-8778
d
Polypropylene Compatible l
Radel AG 230
(polyarylsulfone)
296 73 25% decrease in
elongation; 20%
decrease in yield
strength
12 days i
Roulon J Film Not Compatible i
6
8
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals (continued)
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
Roulon J Film 344 160 14.6%
Weight loss.
Changed from
medium to light
brown.
48 h 95-29067
d
Ryton BR90A (polyphenylene
sulfide)
296 73 67% increase in
elongation; 16%
decrease in yield
strength
12 days i
Ryton R402 (polyphenylene
sulfide sulfone)
296 73 0% change in
elongation; 2%
decrease in yield
strength
12 days i
Stackpole 2331 Ceramic
Permanent Magnet
344 160 0.28 %
Weight gain
NC** 48 h 96-29676
d
Teflon AF 1600 344 160 0.051%
Weight gain
Changed from
colorless to cloudy
94-28139
d
Teflon FEP 344 160 No Varies e
Teflon TFE 344 160 No Varies e
Teflon Compatible l
Tefzel 2004 296 73 15% decrease in
elongation; 4%
decrease in yield
strength
12 days i
Tygon S-54-HL PVC Tubing 298 77 No k
Udel GF 130

(polysulfone) 296 73 29% increase in
elongation; 14%
decrease in yield
strength
12 days i
NOTES: All test results were based on immersion testing.
NC denotes no posttest material change.
COPV = Composite Overwrapped Pressure Vessel
6
9
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 15 Effects of hydrazine on nonmetals (continued)
Temperature Physical Mechanical Compatibility Test Reference
Material K F changes
a
changes
b
test results period
a
Physical changes include changes in weight, color, dimension, etc.
b
Mechanical changes include changes in hardness, tensile strength, viscosity, etc.
c
Marsh and Knox (1970)
d
Tests performed at NASA Johnson Space Center White Sands Test Facility (NASA Laboratories Office, P. O. Box 20, Las Cruces, NM NM 88004.)
e
Salvinski, Howell, and Lee (1967)
f
Coulbert, Cuddihy, and Fedors (1973)
g
Piper (1985)
h
Boyd and White (1960)
i
Arah et al. (1991)
j
Chew et al. (1990)
k
Davis (1993)
l
Kilduff, Davis, and Linley (1990); Those materials which exceed two or more of the criteria for the posttest material and fluid analyses are
considered incompatible by the MCT procedure and not recommended for use.
m
Results Subtask 3.6 TR-804-002 (Delgado et al. 1997a), White Sands Test Facility, Las Cruces, NM. This COPV is a spherical COPV with a 5086
Al alloy liner overwrapped with T-40 graphite fiber epoxy resin. The COPV underwent immersion/wet burst tests to determine COPV integrity.
n
Results Subtask 3.6 TR-804-003 (Delgado et al. 1997b) This is a subscale cylindrical COPV with a seamless Al liner and epoxy overwrap. The
COPV underwent an immersion wet/burst test to determine COPV integrity.
o
Results Subtask 3.6 WSTF # 96-29786, White Sands Test Facility, Las Cruces, NM. This is a subscale cylindrical COPV with a seamless
aluminum liner and graphite epoxy overwrap. The COPV underwent an immersion/wet burst test followed by an immersion/dry burst test.
p
Results Subtask 3.6 WSTF # 96-30013, White Sands Test Facility, Las Cruces, NM. This is a spherical COPV of 5086 aluminum alloy
overwrapped with T-40 graphite fiber and epoxy resin. The COPV underwent an immersion/wet burst test.
7
0
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
AIAA SP-084-1999
71
Table 16 Corrosion rate of alloys in hydrazine and CO2-contaminated hydrazine at 48 hours
Corrosion rate in mils per year
Material CO
2
Concentration
ppm
HZ
a
25 50 100 200 414
SS 304L 0.054 0.240 0.280 0.384 0.524 0.720
MP35N 0.056 2.308 2.780 3.232 3.424 3.810
SS 17-7Ph 0.074 0.090 0.140 0.165 0.210 0.240
TiGR5 0.013 0.024 0.035 0.041 0.046 0.060
Udimet 720 0.917 ND ND ND ND 4.500
NOTES: 1 mil = 0.001 in. = 0.00254 cm
HZ denotes hydrazine
ND means Not Done
a
<3 ppm CO
2

3. Exothermic reaction (Benz 1981; Pedley 1987).

4. Calorimetric methods - accelerated rate (ARC) (Wedlich, Davis, and Peters 1988; Davis and
Wedlich 1988; Davis and Wedlich 1991) and microcalorimetry (Davis et al. 1993).
The first two methods were used predominantly for developing the catalytic material used in the
monopropellant thrusters and power systems. The more recent work at WSTF has concentrated on the
calorimetric methods - first with ARC and most recently with the microcalorimetry system (Davis,
Hornung, and Bunker 1993; Davis et al., publication in process).
The first step in reduction of decomposition rate data to standard kinetic parameters is the determination
of the appropriate stoichiometry of the reaction occurring. Although they do not describe the exact kinetic
mechanism, most data has been represented based on one of the following stoichiometric reactions for
liquid hydrazine (Equations 24 and 25):
3N
2
H
4
( ) N
2
(g) + 4NH
3
(g) 123,300 J/mol (116 Btu/mol) at 515 K (468 F) (24)
and
2N
2
H
4
( ) N
2
(g) + H
2
(g) + 2NH
3
(g) 97,000 J/mol (92 Btu/mol) at 298 K (77 F) (25)
where
= liquid
g = gas
The application of kinetic data should verify the appropriate stoichiometry (See assessment examples,
Section 4.3).
The rate of decomposition of hydrazine is very temperature sensitive (Section 2.1.2, Autoignition
Temperatures and Section 3.3, Exothermic Reaction). Appropriate test methods such as ARC and
microcalorimetry are used to establish the temperature dependency of the catalyzed reaction. Kinetic
temperature data should be used carefully when applied outside of the temperature range in which it was
acquired (See Assessment examples, Section 4.3).
4.2.2 Test conditions
Monopropellant hydrazine (Military Specification: MIl-P-26536D) is the most commonly studied grade of
hydrazine, followed by standard grade. Liquid hydrazine in equilibrium with its vapor and in contact with a
AIAA SP-084-1999
solid surface is the most frequently studied hydrazine system. In a study by Lucien (1961), hydrazine
was confined to the liquid phase by excluding the ullage space.
In all these studies, the experimental conditions were controlled to include the following:
the test vessel is constructed of either a relatively inert material or the test material itself,
the level of hydrazine contamination caused by the materials used to construct, clean, or
passivate the test vessel and sample material is minimized,
the appropriate surface area of the test vessel and the test material are measured,
the condition of the test material surface is determined, i.e. oxide coatings or other surface
contamination must be known, and
the temperature and pressure must be continuously monitored.
Application of kinetic data for hazard assessment from hydrazine decomposition should include the
above. If not available, the data should be used with great care.
The majority of aerospace applications for hydrazine will exclude air and therefore rate of decomposition
testing should exclude air. There have been several studies reported for the hydrazine-air system
(Kilduff, Davis, and Koontz 1988; Miller and Schluter 1978). Schmidt (1984) has reported on hydrazine-
nitrogen systems.
4.2.3 Hydrazine decomposition data
The decomposition of hydrazine catalyzed by a solid surface is a heterogeneous reaction. Appropriate
reaction rate equations would be some form of the Langmuir-Hinshelwood-Hougen-Watson model
(Froment and Bischoff 1979), with the temperature dependency represented in terms of Arrhenius
parameters (A, E
a
). Where this level of detail is not available, hazard assessment should at a minimum
include consideration of the thermal runaway temperature, i.e. that temperature at which the rate of heat
generation (exothermic reaction, Section 2.12) exceeds the rate of heat removal from the system. Such
an evaluation is system dependent and care must be exercised when such a determination is made (See
assessment examples, Section 4.3).
The summary of the ARC testing on hydrazine degradation by materials is given in Table 17. These data
are based on the solid surface being saturated with hydrazine at all times, i.e. complete surface coverage
and the stoichiometry of the decomposition occurring according to Equation 24. The initial temperature
and initial rate given in the table correspond to the ARC instrument sensitivity of 0.02 K/min, i.e. when the
amount of reaction causes the system temperature to increase at this rate. The Arrhenius kinetic
parameters correspond to the given temperature range and are a least squares fit of a set of experimental
runs. The use of this data in hazard assessment is demonstrated in Section 4.3.
The kinetic parameters A and E
a
, of Table 17 can be used to produce a curve showing the rate that
thermal energy is generated for a specified material. An example of such a plot and its use is given in
Example 6, Section 4.3.
Fewer data are available on the heterogeneous decomposition of hydrazine by non-metallic materials.
Data available are summarized in Table 18 in which the following are included:
material description,
Table 17 Hydrazine decomposition data
Onset Onset Kinetic parameters Temp. Pressure
Material WSTF# Form temp. rate A A
a
E
a
E
a
a
range range
K mol/(sm
2
) mol/(sm
2
) mol/(sm
2
) J/mol J/mol K MPa
Al 1100 90-23710 Turning 433 1.040E-06 8.600E+00 3.230E+00 57316 1495 433-493 0.87-13.7
Al 2014 89-22829 Powder 468 1.100E-06 1.790E+05 3.300E+04 103845 781 453-513 1.6-11.9
Al 6061 89-22825 Powder 472 1.260E-06 7.780E+01 4.190E+01 72489 2241 463-513 1.8-11.9
Al/Ni(50/50) 89-23452 Powder 427 1.350E-05 3.410E+13 8.220E+13 146793 9161 423-463 0.62-12.6
Au 90-23611 Powder 428 1.270E-05 3.080E+03 1.100E+03 69316 1461 433-503 0.71-13.1
Au/Ni Braze Pieces 383 1.234E-04 3.870E+05 2.200E+05 69959 2088 383-453 0.12-10.9
Be/Cu Turning 434 2.840E-06 6.390E+08 2.300E+08 124505 1150 453-493 0.75-12.5
Chromium Powder 430 4.570E-07 7.030E+06 9.580E+06 109720 5269 433-473 0.56-12.5
Chromium
b
95-28907 Powder 3.57E-05 59,800 1000 298-328
Copper 93-27848 Powder 420 8.210E-07 1.110E+01 4.870E+00 58086 1750 423-493 0.57-12.9
Copper
b
96-30056 Powder 4.01E+03 78,400 1080 298-328
E-Brite 26-1 89-23467 Powder 428 1.250E-05 3.260E+05 1.140E+05 84760 1379 443-483 0.52-12.6
E-Brite 29-4 89-23466 Turning 440 1.730E-05 6.550E+06 1.520E+06 96470 919 453-483 0.77-14
Ferrotic HT-6A 92-26281 Turning 407 7.930E-07 9.540E+09 1.150E+09 120747 457 413-463 0.5-13.2
Hastelloy C-22 89-23473 Powder 414 1.110E-05 2.870E+05 2.230E+05 83212 2994 423-473 0.35-13.3
Hastelloy C276 91-25322 Rods 413 1.860E-04 6.700E+02 4.900E+02 51017 2802 433-473 0.5-11.7
Incoloy 909 90-23602 Turning 422 5.670E-06 2.780E+07 1.950E+07 100629 2717 433-473 0.54-12.6
Inconel 600 91-25320 Rods 420 8.330E-05 7.210E+01 3.210E+01 48345 1728 423-483 0.57-12.8
Inconel 718 Powder 382 1.660E-06 1.200E+05 6.470E+04 76477 1952 383-443 0.14-10.9
Inconel X-750 90-25908 Turning 403 4.410E-07 3.820E+08 1.940E+08 112820 1882 413-453 0.51-11.9
Iron (Fe) 92-25982 Powder 412 7.610E-05 5.000E+09 1.480E+09 108382 1118 423-463 0.5-11
Iron
b
95-28990 Powder 3.21E+17 146,800 1020 298-328
Lead (Pb) 90-23876 Powder 476 2.660E-06 7.860E+06 2.910E+07 117677 15852 483-523 1.9-12.7
M-2 Tool Steel Turning 387 7.430E-07 4.520E+08 1.460E+08 108066 1139 393-430 0.2-12
7
3
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 17 Hydrazine decomposition data (continued)
Onset Onset Kinetic parameters Temp. Pressure
Material WSTF# Form temp. rate A A
a
E
a E
a
a
range range
K mol/(sm
2
) mol/(sm
2
) mol/(sm
2
) J/mol J/mol K MPa
Mallory 2000 90-23554 Powder 379 3.150E-06 1.350E+03 2.940E+02 63100 766 383-433 0.11-10.9
Manganese 90-23786 Powder 421 5.430E-07 5.000E+04 1.150E+04 88864 904 423-483 0.6-13
Manganese
b
95-28988 Powder 1.64E+08 10,670 100 298-328
Mo-30W 90-23560 Turning 412 4.560E-05 1.380E+02 2.960E+01 51638 834 413-483 0.35-13.3
Molybdenum Powder 389 1.390E-05 5.600E+04 4.090E+04 72060 2626 393-443 0.16-12
Molybdenum
b
95-28991 Powder 2.85E+04 75,500 1000 298-328
Monel 400 89-23456 Powder 397 2.700E-05 4.040E+06 3.980E+06 81918 3798 413-473 0.4-11.5
MP35N 89-23507 Turning 417 2.880E-05 1.990E+03 7.150E+02 62718 1378 423-473 0.42-13.3
Nickel Foil 364 2.740E-05 3.160E+07 5.800E+07 85767 6334 383-423 0.07-9.1
Nickel Rene Turning 415 5.490E-06 5.880E+07 1.010E+07 62459 714 393-423 0.37-13.3
Nitronic 60 89-23499 Turning 419 1.450E-06 8.080E+04 2.640E+04 80029 1256 433-473 0.44-13.3
Rhenium (Re) 90-23722 Powder 357 7.400E-06 1.000E+04 7.620E+03 63163 2606 353-423 0.2-10
SS 201 89-22933 Turning 445 1.030E-06 9.320E+02 5.770E+02 78428 2530 453-503 1-13.3
SS 301 87-21167 Turning 422 1.190E-05 7.960E+10 3.800E+10 127761 1775 413-453 0.48-12.6
SS 302 Turning 422 4.180E-06 9.010E+06 4.320E+06 98998 1812 423-463 0.45-12.6
SS 303L 89-23453 Powder 409 9.480E-07 7.750E+05 1.350E+05 93862 687 413-483 0.5-13
SS 304 92-25985 Powder 422 7.960E-07 1.580E+06 1.420E+06 99631 345 413-473 0.6-11
SS 310 89-23501 Turning 429 2.190E-06 9.170E+05 1.690E+05 90711 726 443-483 0.63-12.6
SS 316 89-23459 Powder 436 4.810E-06 1.220E+05 6.230E+04 87465 2017 433-483 0.7-12.5
SS 316L Powder 416 3.800E-06 2.590E+07 1.280E+07 104729 1876 423-463 0.48-13.3
SS 321 90-23603 Turning 437 1.010E-05 2.910E+06 4.430E+04 96110 58 433-493 0.9-12.3
SS 347 90-23601 Turning 450 7.010E-05 7.180E+04 4.500E+04 78807 2563 453-503 1.1-13.2
SS 410L 92-25916 Powder 423 1.900E-06 8.420E+07 1.440E+07 111296 675 423-483 0.6-12.7
SS 430L Powder 411 8.080E-06 9.350E+08 6.400E+08 119134 2653 423-473 0.43-13.3
SS 440C 89-23503 Turning 400 6.520E-07 4.990E+06 2.710E+06 98889 2054 403-463 0.4-11.8
SS 446 89-23502 Turning 420 3.960E-04 2.300E+05 1.630E+05 86568 2730 423-473 0.42-12.6
AISI 4130 90-23740 Turning 420 3.060E-06 5.860E+10 3.270E+10 133537 2162 433-473 0.6-12.6
Stellite 25 89-23474 Powder 439 2.410E-06 1.050E+07 8.470E+06 108517 3233 443-493 1-13
7
4
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 17 Hydrazine decomposition data (continued)
Onset Onset Kinetic parameters Temp. Pressure
Material WSTF# Form temp. rate A A
a
E
a E
a
a
range range
K mol/(sm
2
) mol/(sm
2
) mol/(sm
2
) J/mol J/mol K MPa
Tungsten
b
95-28989 Powder 2.75E+00 52,770 1020 298-328
Tungsten
CarbideK96
90-24353 Powder 331 7.360E-06 4.710E+01 1.640E+01 43217 1074 333-383 0.02-9.1
Udimet 700 89-23484 Turning 390 1.960E-06 1.030E+04 8.260E+02 73217 288 393-443 0.3-12
Udimet 720 89-23485 Turning 372 2.080E-06 8.790E+01 1.110E+01 54872 433 373-423 0.2-11.3
Vanadium Powder 373 2.780E-06 6.810E+05 4.170E+05 86340 2209 373-443 0.11-10.9
NOTE: WSTF # = White Sands Test Facility number
a
The values of A and E
a
are to be used in sensitivity analysis for comparing values of the observed rate constant, k. The equation is:
k/k = A/A + E
a
/RT (Davis 1994) where E
a
is the activation energy, A is the pre-exponential factor, k is observed rate constant, R is the molar gas
constant, T is the temperature, and denotes the change in the parameters A, E
a
, and k.
b
Data obtained from microcalorimetry tests. All others were obtained from accelerated rate calorimetry (ARC) tests.
The values of E
a
for the microcalorimetry studies are the standard deviation of the E
a
value.
7
5
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 18 Effects of nonmetals on hydrazine
Surface
Material
Heat generation
rate
a
per unit area Temperature
area/
volume Phase
b
Posttest change
results Test
period
Method/
references
J/(sm
2
) K F cm
-1
COATINGS
Akzo 663-3-2/X-310A
Polyurethane Topcoat
344 160 Liquid Test fluid changed
to yellow.
48 h Immersion test
94-28373
Koropon Super High Temp
(Superkoropon 515-700)
< 55 480 405 1.25 Liquid Exothermic test
c
LUBRICANTS
Apiezon L Grease < 55 479 403 Liquid
Liquid,
vapor
Exothermic test
c
Accelerated rate
calorimetry
f
DuoSeal 344 160 Liquid Test fluid changed
to yellow
48 h Immersion test
94-28333
Fomblin Y25 344 160 Liquid Test fluid was
unchanged
48 h Immersion test
94-28332
Fyrquel 220 344 160 Liquid Test fluid changed
to yellow
48 h Immersion test
94-28331
Houghto-Safe 1055 344 160 Liquid Test fluid changed
to a light green color
48 h Immersion test
94-28344
Houghto-Safe 1120 344 160 Liquid 48 h Immersion test
94-28345
Microseal 100-1, Dry Film 620 480 404 1.25 Liquid Exothermic test
c
Microseal 100-1, Dry Film 560 431 316 1.25 Liquid Exothermic test
c
Microseal 100-1, Dry Film 220 408 274 1.25 Liquid Exothermic test
c
Zonyl FSN Fluorosurfactant 297 75 Liquid Test fluid was
unchanged
2 h Immersion test
95-28920
7
6
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 18 Effects of nonmetals on hydrazine (continued)
Surface
Material
Heat generation
rate
a
per unit area Temperature
area/
volume Phase
b
Posttest change
results Test
period
Method/
references
J/(sm
2
) K F cm
-1
PLASTICS AND ELASTOMERS
AF-E-332 (EPR) Liquid,
vapor
Accelerated rate
calorimetry
f
AF-E-411 (EPR)
h
55 480 405 1.25 Liquid Exothermic test
c
AF-E-411 (EPR)
h
< 55 430 315 1.25 Liquid Exothermic test
c
Butyl rubber B318 Liquid Chloride detected Immersion test
i
EPR 692-75 Liquid Slight yellowing of
fluid
Immersion test
i
EPR 740-75 Liquid Nonvolatile
residue detected
Immersion test
i
EPR E 740-75, Parker
Teflon FEP
d,e
7500 478 400 0.03 Liquid Exothermic test
c
EPR E 740-75, Parker
Teflon FEP
d,e
5000 450 350 0.03 Liquid Exothermic test
c
EPR E 740-75, Parker
Teflon FEP
d,e
2000 422 300 0.03 Liquid Exothermic test
c
EPR E515-80 93 478 400 1.25 Liquid Exothermic test
c
EPR E515-80 < 55 429 313 1.25 Liquid Exothermic test
c
Glass, Borosilicate Plate -
Inorganic
Liquid,
vapor
Accelerated rate
calorimetry
f
Graphitar-86/Carbon 120 480 405 1.25 Liquid Exothermic test
c
Graphitar-86/Carbon 78 470 386 1.25 Liquid Exothermic test
c
Halar ECTFE Liquid Slight yellowing of
fluid
Immersion test
i
Kalrez 1045
e,f
1200 450 350 0.25 Liquid Exothermic test
c
Kalrez 1045
e,f
430 436 325 0.25 Liquid Exothermic test
c
Kalrez 1045
e,f
210 422 300 0.25 Liquid Exothermic test
c
7
7
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
Table 18 Effects of nonmetals on hydrazine (continued)
Surface
Material
Heat generation
rate
a
per unit area Temperature
area/
volume Phase
b
Posttest change
results Test
period
Method/
references
J/(sm
2
) K F cm
-1
Kel-F 81 Liquid Chloride and
Fluoride detected
Immersion test
i
Kynar 460 Liquid Chloride and
Fluoride detected.
Immersion test
i
Kynar Grade 460
h
144 394 250 1.25 Liquid Exothermic test
c
Kynar Grade 460
h
115 389 240 1.25 Liquid Exothermic test
c
Kynar Grade 460
h
78 367 200 1.25 Liquid Exothermic test
c
MACOR Ceramic-
Inorganic
< 55 478 400 1.25 Liquid Exothermic test
c
Rulon J Film 344 160 Liquid Test fluid changed
to brown
48 h Immersion test
95-29067
Teflon PTFE Liquid No changes Immersion test
i
Teflon AF 1600 344 160 Liquid Test fluid changed
to yellow
48 h Immersion test
94-28139
Teftzel ETFE Liquid Nonvolatile
residue detected;
Fluoride detected
Immersion test
i
Tefzel ETFE
f
8300 450 350 0.05 Liquid Exothermic test
c
Tefzel ETFE
f
3400 422 300 0.05 Liquid Exothermic test
d
Tungsten Carbide K96 393 478 400 1.25 Liquid Exothermic test
c
Tungsten Carbide K96 196 455 359 1.25 Liquid Exothermic test
c
Tungsten Carbide K96 88 437 327 1.25 Liquid Exothermic test
c
Tungsten Carbide K96 < 55 396 253 1.25 Liquid Exothermic test
c
Tungsten Carbide KZ96 346 478 400 1.25 Liquid Exothermic test
c
Tungsten Carbide KZ96 196 450 350 1.25 Liquid Exothermic test
c
Tungsten Carbide KZ96 < 55 427 308 1.25 Liquid Exothermic test
c
7
8
A
I
A
A
S
P
-
0
8
4
-
1
9
9
9
Table 18 Effects of nonmetals on hydrazine (continued)
Surface
Material
Heat generation
rate
a
per unit area Temperature
area/
volume Phase
b
Posttest change
results Test
period
Method/
references
J/(sm
2
) K F cm
-1
COMPOSITE
Graphite/Epoxy
Laminate, SCI
344 160 Liquid Test fluid was
unchanged
48 h Immersion test
94-28426
Stackpole 2331 Ceramic
Permanent Magnet
344 160 Liquid Test fluid was
unchanged
48 h Immersion test
96-29676
a
For the exothermic test, heat generation rate for nonmetals is measured as (energy)/(time)(unit area of

catalyst); unit area is the initial
surface area of the material. For accelerated rate calorimetry, heat generation rate for nonmetals is measured as (energy)/(time)(gram of
catalyst).
b
Primary phase of hydrazine present during test
c
Fritchman and Benz (1980); initial (pad) pressure of 2.76 MPa (400 psig), GN
2
d
Mancus and Benz (1981)
e
The pre- and posttest surface areas differed greatly. If the material surface area was underestimated, the reported rate may be too high.
f
Accelerated rate calorimetry test method is described in Pedley (1987).
g
Benz (1980)
h
Benz (1981)
i
Davis (1993)
7
9
A
I
A
A

S
P
-
0
8
4
-
1
9
9
9
AIAA SP-084-1999
80
decomposition rate calculated as the thermal energy generated per unit surface area,
temperature of the decomposition rate measurement,
sample surface area,
phase, and
method of test.
The sources of this data are given as footnotes at the end of the tables. These test results are very
system dependent and extrapolation of these results to different test systems or to other temperature
conditions should be done cautiously.
4.2.4 Chemical reactivity of hydrazine in air
Metal surfaces also catalyze the combustion of hydrazine. Data for hydrazine/metal reactivity in air are
given in Table 19, (Miller and Schluter 1978). The authors assert that the thermal runaway temperatures
represent the point at which the surface of the material registered a runaway temperature increase.
These temperatures are totally configuration-dependent and should not be treated as absolute
temperatures, below which hydrazine vapor systems are safe. The values of the thermal runaway
temperatures, however, do provide a relative ranking of the compatibility of these materials with hydrazine
vapor in air at elevated temperatures.
Table 19 Thermal runaway temperatures for hydrazine vapor in contact with metallic materials in air
a
Material Thermal runaway temperature
K F
Aluminum, 6061-T6 465 378
Copper 432 317
Gold 411 280
Inconel-600 452 354
NICKEL 437 328
Silver 398 256
Steel, CRES 303 430 315
Steel, CRES 304 420 297
Steel, CRES 316 452 354
Steel, CRES 321 434 322
Titanium-3A1-2.5V 446 343
Titanium-6A1-4V 490 422
a
Miller and Schluter (1978)
4.3 Assessment examples
The chemical equation used for these examples is Equation 24, rewritten as Equation 26:
N
2
H
4
( ) 1/3 N
2
(g) + 4/3 NH
3
(g) 123,300 J/mol (116 Btu/mol) at 515 K (468 F) (26)
The source of kinetic information is accelerating rate calorimetry (ARC) as given in Table 18. The heats
of formation for hydrazine, N
2
H
4
( ), and ammonia, NH
3
(g), are 50.37 kJ/mol and -45.6 kJ/mol at 298 K
(77 F), respectively (Benz, Bishop, and Pedley 1988). It is assumed that the surfaces of the material are
completely saturated at all times with hydrazine. The rate of decomposition based on the Arrhenius rate
expression (Equation 27) is:
AIAA SP-084-1999
81
dn
dt
SAexp
E
a
RT


_
,
(27)
where:
dn/dt = rate of change of moles hydrazine (mol/s)
E
a
= activation energy (J/mol)
R = gas constant = 8.314 (J/(moleK))
A = pre-exponential value (mol/(sm
2
))
S = surface area (m
2
)
T = Temperature (K)
For example calculations, we will use the sizes and volumes for two actual fuel tanks as used by Bennett
et al. 1980, in development of acceptance test procedures for hydrazine tankage, and fuel system tubing.
The data used in the examples are:
Gas generator (GG) tank 304 SS diaphragm tank
Volume: 1.358 E-3 m
3
(82.87 in
3
)
Hydrazine Weight: 1212 g (2.67 lb)
Ullage: 0.161 E-3 m
3
(9.82 in
3
)
Surface Area: 0.0615 m
2
(95.32 in
2
)
Weight of Tank: 1500 g (estimated) (3.31 lb)
Attitude control system (ACS) tank 304 SS diaphragm tank
Volume: 6.480 E-3 m
3
(395.4 in
3
)
Hydrazine Weight: 5382 g (11.87 lb)
Ullage: 1.142 E-3 m
3
(69.69 in
3
)
Surface Area: 0.1889 m
2
(292.8 in
2
)
Weight of Tank: 4600 g (estimated) (10.14 lb)
321 SS Tubing
Diameter: 6.35 E-3 m (0.25 in.)
Wall Thickness: 5.08 E-4 m (0.020 in.)
Thermal Conductivity: 16.1 W/(ms) (9.3 Btu/(hft
2
))
Heat capacity data
304 SS: 0.5 J/(gK) (0.12 Btu/lbF)
321 SS: 0.5 J/(gK) (0.12 Btu/lbF)
Hydrazine: 3.08 J/(gK) (0.74 Btu/lbF)
4.3.1 Estimation of relative decomposition rates for material applications for which the material
response to hydrazine is well known
If the reactivity of Material A is less than Material B, then Material A can be substituted for Material B, all
other factors, such as surface area, use temperature, and general suitability for the application under
consideration being equal.
If the reactivity of Material A is greater than Material B, but the surface area of Material A is less than
Material B then a comparison of the products of the reactivity times the surface area is appropriate. Thus,
if R
A
S
A
< R
B
S
B
then Material A can be used in place of Material B other factors, such as use
temperature and general suitability for the application under consideration being equal. If the reverse,
R
A
S
A
>R
B
S
B
is true, then additional assessment may be required to determine if the material could still
be used.
AIAA SP-084-1999
82
Example 1: Ranking of materials
Given: The gas generator tank, GG, is made of 304 stainless steel. It has a surface area of 0.0615 m
2
.
Wanted: Consideration is being given to replacing this material with either 321 stainless steel or
Hastelloy C22. Are these acceptable replacements if the operating temperature is 430 K?
Solution: Using the available ARC data of Table 17, the following analysis can be made:
Compare ARC onset temperatures and rates.
Material ARC onset Onset rate Range of
temp K mol/(sm
2
) temp K
SS 304 422 7.96E-7 413-473
SS 321 437 1.01E-5 433-493
Hastelloy C-22 414 1.11E-5 423-473
These temperatures indicate that the kinetic parameters of Table 17 were obtained over similar
temperature ranges. Therefore Equation 27 can be used to calculate decomposition rates for
comparison. The comparison is made at a temperature of 430 K.
Material
Decomposition Rate, k
at T=430 K
S
1 dn
dt
mol/(sm
2
)
k/k
Range of
k k
SS 304 1.246E-6 0.996 0 to 2.5E-6
SS 321 6.144E-6 0.031 5.9E-6 to 6.3E-6
Hastelloy C-22 2.235E-5 1.61 0 to 5.8E-5
Conclusion: At the ARC onset temperature for a given material, both SS 321 and Hastelloy C-22 have
greater decomposition rates than SS 304, Table 4.4. Similarly at 430 K, SS 321 and Hastelloy C-22 have
greater decomposition rates than SS 304. However, the experimental error associated with the kinetic
parameters (k/k = A/A + E
a
/RT, see Table 4.5 for the specified temperature range for these materials)
give limits for the rates of decomposition at 430 K for SS 304 and Hastelloy C-22 which overlap.
This example illustrates the care which must be exercised when using these data for design engineering.
The mean values for decomposition rates suggest that SS 304 is the preferred choice among these
materials. However, when one considers the error bands around these mean values, the preferred
choice becomes less clear.
Material Mean
k -k k +k
% error
SS 304 1.246 E-6 0 2.5 E-6 t 100 %
SS 321 6.144 E-6 5.9 E-6 6.3 E-6 t 3 %
Hastelloy C-22 22.35 E-6 58 E-6 58 E-6 -100 to +161 %
AIAA SP-084-1999
83
4.3.2 Effects of materials on the heat generation rate
The effects of a material or materials on the heat generation rate can be evaluated by assuming an
adiabatic environment. Real system effects such as thermal conductivity of the containment and heat
loss to the surroundings are neglected. The rate of heating is proportional to the decomposition rate.
4.3.2.1 Determine the heat generation rate
For a given material, the heat generation rate (dQ/dt) due to decomposition can be calculated using
Equation 28:
dQ
dt
C
s
r H S A exp
E
a
RT


_
,
W ( ) (28)
where:
H = heat of reaction = 123,300 J/mol
C
s
= system heat capacity, m
i
C
pi
(J/K)
i = material component contribution
r = self heating rate (K/s)
S = surface area (m
2
)
A = pre-exponential factor (mole/(s m
2
))
T = Temperature (K)
R = gas constant = 8.314 J/(molK)
For a material in contact with hydrazine, the heat generation rate can be approximated as the sum of the
individual heat generation rates.
4.3.2.2 Determine the temperature that will produce a given rate of thermal decomposition due to
a single material
Under adiabatic conditions, Equation 29 can be used to determine the temperature T for a specified self-
heating rate (r) due to thermal decomposition:
T
E
a
R ln A S H/(C
s
r) [ ]
(K) (29)
where:
C
s
= system heat capacity, m
i
C
pi
(J/K)
E
a
= activation energy (J/mol)
R = gas constant = 8.314 J/(molK)
A = pre-exponential value (mole/(sm
2
))
S = surface area (m
2
)
H = heat or reaction = 123,300 J/mol
C
s
, the system heat capacity (J/K), is calculated by summing the contributions of the components of the
system.
Example 2: Consider a Gas Generator (GG) tank made of Al 2014. At what temperature will the self
heating rate equal 0.02 K/min during adiabatic operation?
C
s
= C
p
(Al) wt (Al) + C
p
(N
2
H
4
) wt (N
2
H
4
)
AIAA SP-084-1999
84
Thus; C
s
= 0.5 1500 + 3.08 1212 = 4483 J/K.
And by substituting from Table 17 into Equation 29 the following is obtained:
T = 103845 J/mol/{(8.314 J/(molK) ln[1.790E+05 mol/(sm
2
) 0.0615 m
2
123,300 J/mol 60 s/min/
(4483 J/Kx 0.02 K/min)]}
T = 608 K
For an Al 2014 GG tank the temperature of 608 K is the temperature at which the rate of hydrazine
decomposition is equivalent to 0.02 K/min (3.33 x 10
-4
K/s) during adiabatic operation of the system.
Note: This calculation is based on a lumped parameter model where the entire thermal capacity of the
system is changing at the specified rate.
4.3.2.3 Determine the temperature that will produce a given rate of thermal decomposition due to
multiple materials
For multiple materials the process is more complicated but can be approximated by a trial and error
calculation for T. If materials have large differences in their reactivity and similar surface areas, then a
calculation for only the most reactive should be made. Equation 30 can be used to determine the
temperature resulting from a specified self-heating rate for multiple materials whose reactivities are
approximately equal:
C
s

r
H


_
,
S
i
A
i
exp
E
a
RT


_
,


_
,
0

(30)
where:
C
s
= system heat capacity, m
i
C
pi
(J/k)
i = component contribution
R = self heating rate (K/s)
H = heat or reaction J/mol
S
i
= surface area of component i
A
i
= pre-exponential factor of material i (mole/(sm
2
))
E
a
= activation energy (J/mol)
Several trial and error iterations ofT will yield a solution to the expression.
Example 3: The Altitude Control System (ACS) tank
Consider the ACS tank (see data) with 50% of the surface area being 321 SS and the other 50% being
304 SS. For the ACS tank 50% of the surface area is 0.09445 m
2
. Assume the weight of the ACS tank is
split 50% between the 304 SS and the 321 SS. The specific heat for the system is:
C
s
= C
p
(304 SS) wt(304 SS) + C
p
(321 SS) wt(321 SS) + C
p
(N
2
H
4
) wt(N
2
H
4
).
Thus; C
s
=0.5 J/(gK) 2300 g + 0.5 J/(gK)

2300 g + 3.08 J/(gK) 5382 g =18877 J/K.
Substituting C
s
and values from Table 17 into Equation 30 yields:
0 =18877 J/K 0.02 K/min/60 s/min/111000 J/mol - 0.09445 m
2
1.580E6 mol/(sm) exp(-99631 J/mol
8.314 J/(molK)T) - 0.09445 m
2
2.920E6 mol/(sm
2
) exp(-96110 J/mol/(8.314 J/(molK) T))
AIAA SP-084-1999
85
Simplifying terms to a dimensionless expression gives:
f(T) = 1 - 2.6325E9 exp(-1.1984E4/T)- 4.8485E9 exp(-1.156E4/T) = 0
Trial and error evaluation of the function f(T) for different values of T yields the values shown in the
following table:
T (K) f (T)
500 0.456
550 -3.517
510 0.1415
512 0.0614
513 0.0188
514 -0.0255
Thus, the approximate solution for the transcendental equation f(T) = 0 is T 513 K. This temperature
corresponds to a rate of hydrazine decomposition under adiabatic conditions equivalent to 0.02 K/min in
the modified ACS tank.
4.3.3 Hazard analyses
Assessment example covering the pressurization rate in a system, the temperature for onset of thermal
runaway in a system, and general unsteady state analysis are given.
4.3.3.1 Calculate pressurization rate
To evaluate possible pressure hazards from the decomposition reaction, the hydrazine decomposition
rate and the Ideal Gas Law (P=nRT/V) are used to calculate the rate of pressure increase in a vessel
under isothermal conditions.
Use the ideal gas equation (31) for the vapor in the ullage volume:
dP
dt

dn
dt

R T (1 + U/ V) 3600
U 3 1000
(31)
where:
dP/dt = pressure rate (kPa/h)
dn/dt = HZ decomposition rate (mol/s)
R = gas constant = 8.314 J/(molK)
T = temperature (K)
V = system volume (m
3
)
U = ullage volume (m
3
)
n = quantity of hydrazine (mol)
and 3600 converts seconds to hours, 1000 converts Pa to kPa and the factor of (1 + U/V)/3 is an
approximate correction for ammonia solubility in hydrazine (Bennett et al. 1980). At ullage-to-volume
ratios less than approximately 5, the ammonia produced is largely soluble in the liquid hydrazine.
Example 4: Calculate the pressurization in the GG tank at 422 K over a one year period.
AIAA SP-084-1999
86
From Table 18, a temperature of 422 K corresponds to the ARC onset temperature for SS304; therefore
the hydrazine decomposition rate can be read directly at 422 K;
k
1
S
dn SS304 ( )
dt
7.96E 7
mol
2
m s
Now evaluate Equation 30 with the value for dn/dt and other variables:
dP
dt
= (7.96E - 7 /(mol/(m
2
s))) (0.0615 m
2
) 8.314 J/(mol K) 422 K
1+ 0.161E - 3
3
m /1.358E - 3
3
m ( )
[ ]
3600s /h/[ 0.161E - 3 ( ) 3
1000Pa] 1.432E- 3 kPa h 2.08E - 4 psi h ( )
For 1 year: 1.432 kPa/h 24 h 365 days = 1.254E4 kPa (1820 psi)
Clearly, this pressure build up is a hazard and would require alternative materials if the temperature of
422 K were the operational temperature. Note: At 422 K the initial tank pressure would be 600 kPa
(87.1 psia).
Example 5: For multiple materials, calculate the pressure rate for each component and sum the
individual values for the total pressure rate.
Consider a sealed tube containing hydrazine that possesses an internal surface area of 204 cm
2
of
stainless steel (SS304) control material and 0.2 cm
2
of AuNi braze. The system volume is 0.042 L and
the system ullage is 0.034 L. If the system is at 421 K what is the pressurization rate?
Calculate dn/dt
AuNi
+ dn/dt
SS304
using Equation 27 and the ARC kinetic data:
dn/dt
AuNi
= 2E-5 m
2
3.870E5 mol/(m
2
s) exp[-69959 J/mol /(8.314 (J/(molK)) 421 K)]
=1.616E-8 mol/s
dn/dt
SS304
= 0.0204 m
2
1.580E+6 mol/(m
2
s) exp[-99631 J/mol / 8.314 J/(molK) 421 K)]
=1.401E-8 mol/s
dn/dt(total) = 3.02E-8 mol/s
The pressure rate calculated using dn/dt(total) in Equation 30 is:
dP/dt = 3.02E-8 mol/s 8.314 J/(molK) 421 K (1+0.034E-3 m
3
/0.042E-3 m
3
) 3600 s/h/(0.034E-5 m
3
3 1000 Pa/kPa)

= 6.74 kPa/h (0.98 psi/h)
4.3.3.2 Determination of onset temperature for the non-isothermal decomposition of hydrazine:
steady state analysis
A thermal runaway hazard is evaluated in the context of a system. If during the operation of a system
containing materials which react exothermically the rate of thermal energy generation exceeds the rate at
which thermal energy is transferred to the surroundings, then system temperature increase can lead to an
increase in this imbalance and lead to a thermal runaway.
The following steady state analysis indicates whether the rate of heat generation is greater than or less
than the rate of heat loss. The system considered is a flat segment of material with hydrazine
AIAA SP-084-1999
87
decomposition occurring on the inner surface. Heat is lost both to the bulk liquid hydrazine and through
the segment of wall to the surroundings.
The procedure for this calculation is as follows:
a. Select the material, obtain the thermal conductivity, k
w
, and wall thickness, x.
b. Set the film coefficients, h
i
and h
o
.
c. Set the inner surface temperature, T
s
.
d. Set the ambient (outside system) temperature, T
a
.
e. Calculate q
(T
a
T
s
)
1
h
o
+
x
k
w
f. Calculate T
f
T
s

q
h
i
g. Check if Equation 32 is satisfied. If yes, the solution has been obtained. If no, choose different
T
s
and repeat.
Equation 32 describes the steady state balance of the enthalpy generated by hydrazine decomposition
and the heat exchanged to the fluid and surroundings for a material surface:
Q
GEN
= Q
HXG
or
H
R
o
S A exp
E
a
RT
s



_
,

S
T
a
T
f ( )
1
h
f
+
x
k
o
+
1
h
o



_
,






1
]
1
1
1
(32)
where Q
GEN
is calculated using Equation 28 and Q
HXG
is calculated using Fouriers law for heat transfer
through several bodies in series.
Example 6: Storage Tank
This example concerns the time-rate of change of the temperature of a hydrazine storage tank for two
cases: (1) where the tank contents initially are above ambient temperature, and (2) where the tank
contents are below ambient temperature. The worst case calculation is for an ambient temperature of
317 K (110 F). Note: Because of the extreme extrapolation of the kinetic rate data, the uncertainty
analysis is applied for both cases.
Single spherical fuel storage vessel: Gas Generator Tank (SS 316L) at sea level on summer day.
Radius =6.870E-2 m
Volume =1.358E-3 m
3
Weight N
2
H
4
( ) =1212 g
Ullage =0.151E-3 m
3
Surface Area (S
rx
) =0.0615 m
2
(for reaction)
Surface Area (S
ves
) =0.0699 m
2
(for heat transfer)
Weight metal (SS 316L, 304) =1500 g
Wall thickness, x =0.0012 m
Case 1. T
TANK
initially above ambient.
AIAA SP-084-1999
88
1. Assume entire external surface area available for heat transfer to atmosphere.
2. Assume only 85% of surface area available for decomposition reaction.
3. Assume T
surr
= 110 F (317 K)
P
atm
= 14.7 psia (101.325 kPa)
4. First trial - no flow; stagnant atmosphere
In order to calculate for the film (heat-transfer) coefficient, the Nusselt number is used (Equation 33):
N
nu

hD
k
(33)
The Nusselt number (N
nu
) is the dimensionless group which relates the film (heat-transfer) coefficient, h,
the thermal conductivity, k, of the fluid, and a characteristic dimension, D (Geankoplis 1983). For a single
sphere being heated or cooled by a fluid flowing past it, the following equation can be used (Equation 34):
N
nu
2.0 + 0.60N
Re
0.5
N
Pr
1 / 3
(34)
where N
Re
is the Reynolds number and N
Pr
is the Prandtl number. For a no flow, stagnant atmosphere,
the Reynolds number is zero and the Nusselt number is equal to 2.0.
h = 2.0 k
air
/D
sphere
k
air
= 0.21 k
O2
+ 0.79 k
N2
= 0.21(0.028) + 0.79(0.027) = 0.027 W/(mK)
h = 2.0 0.027/(20.0699) = 0.39 W/(m
2
K)
5. T
initial
= 350 K
Lumped parameter model/neglect mass change from reaction Enthalpy Balance:
Accumulation of
enthalpy of tank
and contents




_
,

Rate of enthalpy
due to reaction on
tank surface




_
,

Rate of heat exchange with


surroundings.



_
,
The enthalpy balance can be written in equation form as (Equation 35):

d
dt
M
l
C
vl
+ M
tk
C
ptk
( )
T t ( ) T
ref
( )
S
rx
A

exp
E
a
RT


_
,
hS
ves
T t ( ) T
surr
[ ] (35)
where:
C
v( )
= 3.358 J/(gK)
C
ptk
= 0.502 J/(gK)

A = (AH
rx
) = 3.186E12 J/(m
2
s)
E
a
= 104729 J/mol
S
rx
= 0.85 S
ves
(m
2
)
h = 0.39 W/(m
2
K)
R = 8.314 J/mol K
Solve Equation 35 for dT(t)/dt to obtain (Equation 36):
AIAA SP-084-1999
89

d T t
( )
dt

1+
k
k


_
,
AS
rx
exp
E
a
RT


_
,
l ( ) M
C
v l ( )
+ M
tk
C
ptk
( )

h S
ves
T t
( ) T
surr
( )
l ( ) M
v l ( ) C
+M
tk
C
ptk
( )


(36)
where:
k/k = 1.14 (Sensitivity factor for the observed rate constant, see Table 18)
Evaluation of the terms in Equation 36 gives (Equation 37)

d T t
( )
dt
8.694E7 exp(
-1.26E4
T(t)
) 5.6524E6 T t ( )317
( )

(37)
The storage tank and hydrazine will cool to the ambient temperature (317 K) although they initially were
at 350 K, as shown in Figure 23.
For this example, the enthalpy of heat exchange term in Equation 37 dominates the enthalpy of
generation term causing a negative slope for the temperatures-time curve. Eventually the temperature
will reach ambient. However, if a larger T
initial
is used, the slope could be positive and a thermal runaway
will occur. The following table gives the magnitudes for each term on the right hand side of Equation 37.
Between 500 and 515 K, the slope changes from negative to positive. Thus, if the initial temperature is
greater than 515 K, the temperature will continue to rise until a system failure occurs.
Temperature
K
Enthalpy of
generation
J/s
Enthalpy of heat
exchange
J/s
250 1.14E-14 -3.8E-4
300 5.0E-11 -9.6E-5
400 1.82E-6 4.69E-4
500 9.9E-4 10.34E-4
515 2.06E-3 1.12E-3
525 3.28E-3 1.18E-3
Case 2. If the hydrazine liquid is initially below ambient and the system conditions are such that when the
hydrazine liquid is above ambient the system will cool to ambient. The end result for case 2 is an ambient
steady state.
AIAA SP-084-1999
90
Figure 23 Temperature profile of storage tank and hydrazine cooling to ambient temperature
AIAA SP-084-1999
91
5 Safety
Introduction
Exposure of personnel to trace amounts of hydrazine can cause serious health consequences. Because
of its high toxicity, hydrazine is hazardous not only to humans and other animal life, but also to the
environment (soil and vegetation). This section presents information on the health effects of exposure to
hydrazine, reviews regulatory guidelines for hydrazine use, presents current published hydrazine
minimum exposure limits, and presents recommended procedures for hydrazine handling. The
assessment of exposure hazards should be performed according to the statutory regulations that apply in
a given situation and with the assistance of a toxicologist or an industrial hygienist. Failure to do so can
result in monetary penalties and even civil and criminal prosecution.
Readers are cautioned that, although every reasonable effort has been made to present accurate
information in this section, the authors and publisher make no warranty nor do they assume legal
responsibility for its correctness. Readers are urged to contact state and local regulatory
agencies for complete information on the applicable safety regulations.
In practice, handling hydrazine in a given situation will always entail some degree of potential exposure.
The health effects of the hydrazine family of compounds are still being actively investigated and are not
yet fully understood (Schmidt 1984). Consequently, the guidelines for personnel exposure to hydrazine
are frequently changed. For example, the threshold limit value - time weighted average (TLV-TWA) for
hydrazine, recommended by the American Conference of Governmental and Industrial Hygienists
(ACGIH), was recently reduced by a factor of 10 (ACGIH 1996). Current values are 0.01 ppm and a
carcinogen status of A3. The circumstances of exposure are also important. As an example, the
exposure limit developed for workers who are at risk for known periods during the work day will be
different from the exposure limit for the general public from an industrial accident. The recommendations
for such different aspects of personnel exposure often come from different organizations with different
charters. The exposure limits stipulated by the Occupational Safety and Health Administration (OSHA)
and the Environmental Protection Agency (EPA) carry the force of law, while organizations such as the
ACGIH and the National Institute of Occupational Safety and Health (NIOSH) can only recommend limits.
Organizations such as the ACGIH continually review health and safety data, and therefore provide
recommendations that may be more up-to-date than statutory regulations. No published exposure limit in
itself assures protection; only proper control provides protection. A safe working environment is
established through the proper use of administrative policies, engineering controls, and personal
protective equipment.
5.1 Hydrazine toxicity
The term toxicity, as used in the professional literature, is slightly inaccurate in that it fails to recognize the
complex relationship that personnel and environmental exposures often have. An acceptable definition
for toxicity is, The ability of a substance to cause damage to living tissue, impairment of the central
nervous system, severe illness or, in extreme cases, death when ingested, inhaled, or absorbed by the
skin (Hawleys 1994 ). To evaluate the toxicity of hydrazine, the following must be considered: (1) the
route of absorption, (2) the concentration of exposure, and (3) the exposure duration.
Hydrazine can enter the body through the lungs, skin, eyes, and gastrointestinal (GI) tract. These four
forms of absorption can also be categorized as inhalation, dermal exposure, ocular exposure, and
ingestion. Dosage is the total amount of a substance taken into a body. It may well be the single most
important factor in determining whether a given chemical, such as hydrazine, will produce a toxic effect.
It is an axiom that the dose makes a poison.
AIAA SP-084-1999
92
The duration of exposure can be described as acute or chronic (Sax 1994). Acute exposure refers to
short durations (less than 24 h). When absorption is through ingestion, an acute exposure refers to one
dose. Typically, acute exposures will produce noticeable effects, often immediately. Chronic exposure
refers to longer durations, usually measured in days or more. Chronic exposure to hydrazine can
damage internal organs, especially the liver, kidneys, and blood. When chronic absorption is through
ingestion, it refers to repeated doses of the same chemical; for all other means of absorption, chronic
exposure refers to repeated or prolonged exposure. As a general rule of thumb, the following delineation
is made: acute exposures generally affect the individual quickly and noticeably. Chronic exposures are
very often insidious; they may offer little or no clue to the damaging effects on the body.
5.1.1 Results of hydrazine exposure
All forms of hydrazine are extremely hazardous to unprotected personnel. Hydrazine is a strongly
reducing, basic chemical. Regardless of the way in which it enters the body, even small amounts of
hydrazine can be fast-acting and destructive on tissues and fluids. A single acute dosage is highly toxic.
The results of animal studies show that at sufficiently high dosages (four times the median lethal dosage)
death occurs without warning indications within a minimum of 7 min (Schmidt 1984). The specific effects
of personnel exposure to hydrazine can vary greatly from individual to individual, but the warning cannot
be overemphasized: the best practice is to avoid any contact with hydrazine. Do not breathe hydrazine
vapor, and never allow hydrazine to contact your eyes, skin, or clothing.
General symptoms for human exposure to hydrazine include the following:
eye irritation,
dizziness,
nausea,
skin burns,
nasal lesions,
bronchitis,
abdominal pain, and
convulsions.
General conditions arising from human exposure to hydrazine include the following:
corneal damage,
methemoglobinemia,
cyanosis,
dermatitis,
respiratory edema,
kidney damage,
liver damage,
hemolysis,
central nervous system depression,
hematuria,
AIAA SP-084-1999
93
Heinz bodies in red blood cells, and
reproductive effects (observed in experimental animals).
There have been few reported cases of hydrazine inducing death in humans. Two cases of intoxication
demonstrate the deadly effect of hydrazine (Reinhardt and Britteli 1981a):
One person accidentally drank `between a mouthful and a cupful' [of hydrazine.] [He]
immediately vomited and lost consciousness. Hospitalized, he was flushed but afebrile,
unconscious, and vomiting; pupils were dilated but central and light reactive. Within 12 h
vomiting ceased, the pupils were smaller but diverged to the right, and he was
sporadically violent. He was treated with pyridoxine. Later, his memory and voluntary
movements were normal but he was ataxic and unable to write. There was lateral
nystagmus to the right and his ability to sense vibration was lost. He improved but his
final condition was not reported.
[Another]...worker handled hydrazine hydrate once a week for 6 months. Early signs of
toxicity were lethargy, conjunctivitis, and tremors. On the last day of exposure he
developed fever, vomiting, and diarrhea. Later, he developed abdominal pains and black
feces and became incoherent. His abdomen was enlarged and his liver was palpable
and tender. There was fluid in the chest cavity and lung shadowing. Bilirubin and
creatinine levels were elevated. Urine volume was low and contained protein and
erythrocytes. Treatment to correct fluid imbalance was given and he temporarily
improved but died 20 days after the last exposure to hydrazine. Autopsy revealed severe
tracheitis, bronchitis, lungs filled with exudate, enlarged kidneys with severe tubular
necrosis, interstitial hemorrhages and inflammation, liver with microscopic focal areas of
necrosis and granular cytoplasmic degeneration, and an enlarged heart with muscle fiber
degeneration and hyperemia. Conditions of exposure were simulated and hydrazine
concentration in the air was found to be 0.071 mg/m
3
(0.053 ppm). Skin absorption could
have been a significant factor. However, duration of exposure and the actual
concentration he was exposed to are not known, but the investigators considered the
lung, liver, and kidney damage to be from hydrazine poisoning.
Data based upon the response in animals indicate that hydrazine is a strong skin and mucous
membrane irritant, a convulsant, a hepatotoxin, and a moderate hemolytic agent (Reinhardt and Britteli
1981).
5.1.2 Inhalation
By far the most likely form of hydrazine exposure in the workplace is that from inadvertent inhalation. The
average worker breathes approximately 10 m
3
(353 ft
3
) of air during an 8-hour working day (Sax 1984).
At ambient temperatures hydrazine evaporates readily with a vapor pressure approximately 60 percent
that of water and can persist for hours in air (see Section 5.2.1). There is little data correlating long term
effects in humans. It is known that acute exposure resulting from spills can result in pulmonary irritation
even dyspnea and choking hundreds of yards from the source (Hall and Rumack, 1996). Pulmonary
edema has been observed for acute exposure. Other symptoms include rhinitis in the nose and in the
mouth salivation, sore throat, and coughing (Hall and Rumack, 1996).
Animal studies at high concentrations, for example the 4-hour inhalation LC
50
* for rats with hydrazine is
570 ppm, show toxicity characterized by respiratory tract irritation, pathological damage to lung, liver, and
kidney (Reinhardt and Brittelli 1981b).

* Lethal concentration that causes death in 50 percent of the tested population.
AIAA SP-084-1999
94
5.1.3 Exposure to skin, eyes, and mucous membranes
Primarily because of its ability to break down the skin's protective barrier of oil, all forms of hydrazine can
gain rapid access to the bloodstream through the skin, and therefore can enter many of the internal
organs as well. In some cases, hydrazine entering the body in this way may be even more dangerous
than inhaled hydrazine. Substances that gain entrance to the body through the skin can be harder for the
body to eliminate, because only a small portion is filtered by the liver and kidneys.
Hydrazine is a strong skin irritant producing discoloration, rashes, blisters and edema (Schmidt 1984). It
is readily absorbed through the skin (Schmidt 1984). Dermal exposure tests in canines with
monomethylhydrazine (MMH), a compound related to hydrazine, found that MMH made its way to the
dogs' bloodstreams within 30 s. As a general warning of the dermal exposure hazard, the following
example is offered:
These studies of absorption through skin clearly demonstrate that although some
decrease in toxicity may occur by this route, death may result from cutaneous application
of hydrazine or its methyl derivatives. The surprisingly small volume required is
illustrated by the calculation that only 5 mL (0.305 in
3
) hydrazine spilled on a 10-kg (22-lb)
dog would undoubtedly cause serious, if not fatal, intoxication. (Smith and Clark 1969).
The study recommended that MMH should be washed from the affected skin immediately. The skin may
afford some protection from MMH, allowing time to wash the MMH off the skin, especially when compared
to MMH entering the body through inhalation. The level of safety added, however, in no way diminishes
the hazard. Of particular importance, if the MMH was not washed from the skin within 2 min, the
concentrations in the bloodstream were as high as if no washing had occurred (Smith and Clark 1969).
Liquid hydrazine introduced into the eyes of test animals causes permanent corneal damage
(Schmidt 1984).
The skin does not readily absorb vapor or gaseous hydrazine, but the ACGIH does give hydrazine a
skin designation (Table 22).
5.1.4 Ingestion
Compared to inhalation or dermal exposure, ingestion of hydrazine is unlikely. It is rare, yet it is perhaps
the most difficult route of exposure to prevent. Ingestion can result from the inadvertent drinking of liquid
hydrazine; however, a far more likely cause of ingestion is from eating in contaminated areas or with
contaminated hands (Dinman 1979). A dramatic example of the effects of ingestion is given in
Section 5.1.1.
5.1.5 Carcinogenicity
There is no conclusive data demonstrating hydrazine to be a cancer causing agent in man. High
exposure concentrations have been shown to cause cancer in laboratory animals. This has led NIOSH
(see section 5.3.3) to declare hydrazine a carcinogen while the ACGIH lists hydrazine as an Animal
Carcinogen, status A3 (see Table 5.3).
AIAA SP-084-1999
95
5.2 Environmental fate of hydrazine
Hydrazine introduced to the environment is considered a pollutant. Depending upon the method of
release the hydrazine may contaminate some combination of air, water, or soil. In assessing the effects
of a hydrazine release on the environment factors of interest include the length of time that hydrazine
persists, the toxicity to animals and plants, and the potential for bioaccumulation.
Although hydrazine can be volatile and reactive in the environment, it can also be relatively stable in the
absence of the right degradation catalysts. For these situations treatments to neutralize hydrazine can be
used.
5.2.1 Air
A variety of mechanisms lead to the rapid degradation of hydrazine released into the atmosphere.
Hydroxyl radicals, ozone, and nitrogen oxides all act to degrade vapor phase hydrazine. Estimates of the
half life of hydrazine for different reactions in the lower troposphere are given in Table 20 based upon
environmental data and kinetic assumptions.
Table 20 Estimated half life for hydrazine in the atmosphere
a
Reactant/Process Reactant concentrations
Hydrazine half life
h
Hydroxyl Radical 1 E6

molecules/cm
3
3.0
Ozone 7 E11 molecules/cm
3
9.2
Ozone In natural Troposphere <2.0
Ozone At Pollution Levels <0.2
No Ozone At 15 C & 16% Rel. Humidity 5.0
No Ozone At 14 C & 85% Rel. Humidity 1.8
Nitrogen oxides Not Specified 2.0
b
Direct Photolysis None Large
c
a
Hall A.H. & Rumack B.H. (Eds): TOMES(R) Information System. Micromedex, Inc., Englewood,
Colorado (Edition expires 4/30/96).
b
Half life applies in light or dark.
c
Not expected to significantly effect hydrazine concentration.
5.2.2 Water
Hydrazine released into water will degrade. The rate of volatilization depends upon the concentrations of
organic matter, oxygen, water hardness, the concentration of hydrazine, and the position of the hydrazine
relative to the water surface. High concentrations of organic matter and oxygen will result in a greater
volatilization of hydrazine. At the waters surface the rate of volatilization is greater. Water hardness
increases the volatilization rate. The results of a comparison study showing the effect of organic matter
on a 5 mg/L hydrazine concentration in water are shown in Table 21. The pond water and river water
were maintained at the same temperature and adjusted to contain identical dissolved oxygen levels. The
river water differed from the pond water by containing substantial amounts of organic matter.
5.2.3 Soil
Soils are thought to either degrade, absorb, or permit hydrazine to leach through. The amount and rate of
degradation will be greatest for soils containing large amounts of organic carbon. Soils high in clay
content will absorb hydrazine. In sandy soils hydrazine may leach through to ground water. Volatilization
and biodegradation are slow processes that will only have a significant role in the removal of hydrazine
AIAA SP-084-1999
96
when the concentrations of hydrazine are low. High concentrations of hydrazine that occur during a spill
are not expected to be affected by volatilization or biodegradation (Hall and Rumack 1996). Additional
information concerning soil interactions with respect to pH, organic content, available transition metals,
colloidal properties of soils, and soil sorption is available (Street 1988).
Table 21 Degradation of 5 mg/L hydrazine in water
a
Time Pond Water Percent Degradation River Water
b
Percent Degradation
1 h 20 22.6
1 day 74 96
2 days 80 100
3 days 81.6 ---
a
Hall and Rumack (1996)
b
River water possesses substantial quantities of organic matter.
5.2.4 Bioaccumulation and biodegradation
Other mechanisms that cause the removal of hydrazine from the environment include bioaccumulation
and biodegradation. Guppies exposed to 0.5 g/g hydrazine in water developed hydrazine
concentrations of 144 g/g after 96 h. This is described as a moderate tendency to bioconcentrate but
the rapid degradation of hydrazine in the environment is likely to prevent significant bioconcentration
(Hall and Rumack 1996). Biodegradation is possible with strains of bacteria able to metabolize hydrazine
(Kuch 1996).
5.2.5 Remediation
Treatments found effective for the neutralization of hydrazine released in the environment include sodium
hypochlorite oxidation, other chlorination procedures, or ozonation. Any remediation procedure should be
specified by and directed by the cognizant scientist in the organization responsible for compliance with all
federal and state environmental regulations.
5.3 Hydrazine exposure guidelines
A variety of guidelines have been created to help protect the worker and the public from exposure to
hazardous chemicals. The following generalizations can be made about the scope of regulation of the
advisory and regulatory agencies. The exposure limits published by OSHA are designed to protect the
worker, and they carry the force of law. The regulations developed by OSHA are often based upon the
recommendations of organizations such as the ACGIH and NIOSH. The recommendations coming from
the ACGIH and NIOSH, while not mandatory, represent the most up-to-date consensus on exposure
limits required for worker safety. The public must also be protected from hazardous substances used in
the work place. Title III of the Superfund Amendments and Reauthorization Act (SARA) administered by
the EPA directs communities and industry to inventory hazardous substances, track toxic chemical
releases, plan for chemical accidents, and provide public access to information on hazardous substances.
NASA has implemented a policy of using the most stringent Special Report lines.
*
The more widely used
hydrazine exposure guidelines are discussed below.

*
Policy letter from Dr. Harry C. Holloway MD, Agency Representative to OSHA Basic Safety and Health
Program Elements for Federal Employees, NASA Headquarters, Washington D.C. (1995).
AIAA SP-084-1999
97
5.3.1 Threshold limit values of the American Conference of Governmental Industrial Hygienists
The most widely used measure of human tolerance to toxins is the TLV of the ACGIH. The TLV's are
used as guidelines for airborne contaminants and are acknowledged by federal regulations although they
do not in and of themselves carry the weight of a legal regulation. Acceptance of the ACGIH guidelines,
now the most stringent, as limits not to be exceeded is recommended. The ACGIH reviews health data
on chemicals annually.
*
An estimation based on laboratory research and occupational experience, the TLV, refers to airborne
concentrations of substances and represent conditions under which it is believed that nearly all workers
may be repeatedly exposed day after day without adverse effects, (ACGIH 1996).
Table 22 shows the TLV-TWA for hydrazine.
**
For hydrazine the ACGIH specifies a threshold limit
valuetime weighted average (TLV-TWA). This is an average concentration, assuming a normal 8-h
workday and a 40-h workweek. The allowed variation in worker exposure above the TLV-TWA is
specified by the excursion limits which may exceed 3 times the TLV-TWA for no more than a total of
30 min during a work-day, and under no circumstances should they exceed 5 times the TLV-TWA,
provided that the TLV-TWA is not exceeded (ACGIH 1996). The ACGIH also notes that absorption of
hydrazine through the skin can result in systemic toxicity. Strict adherence to the ACGIH guidelines is
recommended.
Table 22 ACGIH threshold limit value guidelines for hydrazine
a
TLV-TWA
ppm mg/m
3
Designations
0.01 0.013 Skin
b
Animal
c
Carcinogen
a
ACGIH 1996
b
Contact through the skin, mucous membranes, and/or eyes to hydrazine vapor or liquid can lead
to a potential significant contribution to the overall exposure. This does not refer to the effects
upon the skin itself.
c
Designation A3: Carcinogenic in experimental animals at relatively high dose, by route(s) of
administration, at site(s), of histologic type(s), or by mechanism(s) that are not considered relevant
to worker exposure. Available epidemiological studies do not confirm an increased risk of cancer
in exposed humans. Available evidence suggests that the agent is not likely to cause cancer in
humans except under uncommon or unlikely routes or levels of exposure.
5.3.2 Guidelines of the Occupational Safety and Health Administration
OSHA has established permissible exposure limits (PEL's) of workers to hydrazine.* These exposure
limits carry the full force of federal regulations and are binding unless a waiver is obtained. This

*
New guidelines proposed by the ACGIH stand as recommendations under evaluation for a period of
2 years. During this probationary period, the members of the health research community can seek to
dissuade the ACGIH from formalizing the recommendation. If there are no successful challenges to the
proposed limits, the recommendation is adopted after 2 years.
**
At the time of this writing there are no portable personal instrumentation or area monitoring devices that
can register these low concentrations on a continuous basis. Also, given the propensity of hydrazine to
be catalyzed by surrounding surfaces, currently available area monitoring devices [require off-line
analysis] may not adequately detect transient levels of hydrazine. Consult a hygienist for directions on
how to meet ACGIH guidelines in the work place.
* OSHA PEL's are given in 29 CFR 1900.1000 Table Z-1, Air Contaminants. As of July 7, 1992, the U.S.
Court of Appeals for the Eleventh Circuit struck down OSHA's 1989 PEL standard for general industry
AIAA SP-084-1999
98
information is published annually in the Federal Register, which is the source for the most up-to-date
OSHA regulations. Again, it is important to note that PEL's differ from the other exposure limits in that
they are regulations, not recommended guidelines.
OSHA is in the process of revising existing PEL's to improve the protection against sensory irritation and
catastrophic effects to the cardiovascular system, liver, kidneys, lungs, and central nervous system.
The new PEL's employ a TWA, an STEL, a ceiling limit, and a notation for possible skin hazard. The
TWA is described by OSHA as the employee's average airborne exposure in any 8-hour shift of a
40-hour workweek which shall not be exceeded. The STEL is defined by OSHA as the employee's
15-min (or other specified period) time weighted average exposure which shall not be exceeded at any
time during a work day. The ceiling is defined by OSHA to be the employee's exposure which shall not
be exceeded during any part of the work day. It is further stipulated that when instantaneous monitoring
is not possible, the STEL limit shall not be exceeded at any time over a working day. The skin
designation, when listed for a substance, indicates that the substance shall be prevented or reduced to
the extent necessary in the circumstances through the use of gloves, coveralls, goggles, or other
appropriate personal protective equipment, engineering controls or work practices. OSHA only lists a
time weight average and gives a skin designation for hydrazine (Table 23). It is recommended that
operations meet the more stringent ACGIH guidelines (see 5.2.1).
Table 23 OSHA permissible exposure limit for hydrazine
a
Time weighted average Skin designation
1.0 ppm (1.3 mg/m
3
) Protective clothing and equipment required
a
29CFR1910.1000 Table Z-1 Limits for Air Contaminants (June 1996).
5.3.3 The National Institute Of Occupational Safety And Health
NIOSH provides recommendations to OSHA. Guidelines published by NIOSH include the recommended
exposure limits (REL's) and the immediately dangerous to life or health (IDLH) limit. Because NIOSH
considers hydrazine a carcinogen no REL is published.
**
The IDLH limit is designed to aid with respirator
selection. The IDLH limit is a maximum concentration from which escape within 30 min is possible
without escape-impairing symptoms or irreversible health effects. For hydrazine the IDLH limit is 50 ppm.
NIOSH has compiled OSHA PEL's, REL's, IDLH limits, and ACGIH TLV's, as well as suggestions on
corrective actions, preventive measures, protective clothing, and general emergency procedures in an
easy-to-use handbook.**

5.3.4 Emergency planning requirements
The Emergency Planning and Community Right to Know Act, also known as SARA Title III, directs each
state to appoint a State Emergency Response Commission (SERC) and apportion emergency planning
districts, each with a Local Emergency Planning Committee (LEPC). The SARA Tier I and Tier II forms
used to comply with these regulations allow local community agencies to develop emergency response
plans and compile an inventory of all chemicals in their vicinity. In addition, the regulations allow the
general public access to information about hazardous chemicals that are stored and used in their
communities.

(29 CFR 1910.100) and ordered a return to levels adopted by OSHA in 1971. The 1989 PEL's are still
in force and OSHA is appealing the ruling (Occupational Hazards 1992).
**
Appendix A, NIOSH POCKET Special Report, U.S. Department of Health and Human Services, 1994.
AIAA SP-084-1999
99
SARA Title III requires facilities that possess one or more of 366 extremely hazardous substances in
quantities above limits set by the EPA to make available information on these substances
*
to the public
and to work with their LEPC to develop comprehensive plans to cover accidents. Hydrazine is one of the
366 substances.
Emergency planning provisions come into effect when the threshold planning quantity (TPQ) is exceeded.
The EPA has established a TPQ based upon an index that accounts for both the toxicity of the chemical
and its potential to become airborne; however, the TPQ's are not intended to be predictive of risk. The
TPQ for hydrazine is 227 kg (500 lb).
5.3.5 Spacecraft maximum acceptable concentrations
**
Personnel exposure guidelines have been developed by NASA for setting levels of containment needed
for payload and utility chemicals used or produced on manned flights and to determine the health impact
should these materials enter the cabin atmosphere. The NASA policy for ground-based operations shall
comply with standards developed for the workplace established by nationally recognized groups such as
the ACGIH (Section 5.2.1).
The continuous exposure for periods of up to 1 week, and perhaps 6 months or longer on future missions,
that the astronauts will experience is very different from occupational exposure based upon an 8 h per
day for 5 days per week. For the conditions of spaceflight NASA toxicologists have developed spacecraft
maximum acceptable concentration (SMAC) values (Garcia, James, and Limero 1991). Given different
durations for flight and flight operations SMAC's have been developed for 1 h, 24 h, 7 days, 30 days, and
180 days. These values are given in Table 24.
Table 24 Spacecraft maximum allowable concentrations for hydrazine
a
Time Concentration
ppm
b
Concentration
mg/m
3
1 h 3.800 5.000
24 h 0.310 0.400
7 days 0.038 0.050
30 days 0.023 0.030
180 days 0.004 0.005
a
Private communication with Dr. John T. James, JSC Toxicology, Houston TX, June 27, 1996.
b
Concentration computed from mg/m
3
values according to the formula (ACGIH, 1995-96)
5.3.6 Environmental regulations
Controls over the release of hydrazine specified in the Code of Federal Regulations cover the release of
hydrazine above a fixed quantity, while in transport, into the air, or as waste. Compliance with these
regulations is mandatory. Operations involving the use of hydrazine should be coordinated with
environmental and industrial hygiene specialists as appropriate.
CERCLA. The comprehensive Environmental response, Compensation, and Liability Act (CERCLA)
requires that releases of hazardous substances into the environment above a threshold level be

*
This is done with the Toxic Chemical release Inventory Reporting Form (EPA Form R) required by
section 313 of the Emergency Planning and Community Right-To-Know Act.
**
This section is based on a conversation with Dr. J. James, JSC Toxicology, Houston TX, September 15,
1992.
AIAA SP-084-1999
100
documented. For hydrazine the Reportable Quantity (RQ) is 453.6 g (1 lb). The RQ is also applied in 49
CFR, subpart A, Hazardous Materials Transportation by the Department of Transportation (DOT).
Federal Clean Air Act. Hydrazine is regulated under Title III of the Federal Clean Air Act as amended.
Depending on the source category delineated by the EPA, or approved State Implementation Plans, new
or modified emission sources emitting hydrazine above established regulatory levels may require an
Operating Permit for compliance.
Federal Clean Water Act. These regulations and EPA authorized state water pollution agencies will
regulate discharges of hydrazine to surface waters. Process waters released to holding ponds or other
collection areas may require discharge plans. Discharge plans can require sampling for constituents of
interest, leak detection and control systems and groundwater monitoring to determine potential impacts to
underlying aquifers. In addition, discharges to navigable waters and storm water runoff that may be
contaminated could require permitting and compliance monitoring under the NPDES (National Pollutant
Discharge Elimination System) regulatory requirements.
Food, Drug, and Cosmetics Act (FFCA). The primary concern of these regulations is steam used in the
processing of food. The regulations require that steam used in contact with food does not contain
hydrazine. Many boilers that produce steam use hydrazine as an oxygen scavenger in process water.
RCRA. Discarded commercial chemical product, off-specification materials, container residue in excess
of 2.5 cm (1 in.), spill residue, and liquids or materials containing hydrazine constituents are designated
hazardous waste (waste code U133) by the Resource conservation and Recovery Act (RCRA) Solid
Waste Regulations (40 CFR, Part 261.33). Large quantity generators of hydrazine waste must be
permitted pursuant to 40 CFR 270. Both small and large quantity generators of hydrazine wastes must
utilize permitted transport, disposal and storage facilities, whether such facilities are on-site or off-site.
SARA Title III. Organizations/corporations that maintain an inventory of hydrazine (or any other regulated
hazardous material) above a Threshold Planning Quantity must prepare reports that disclose chemical
information (see Section 5.3.4). This information includes inventory quantities, locations of stored
material, types of storage containers, and storage conditions such as pressure and temperature. This
regulation is part of the Community Right-to-Know requirements. The Toxic Release Inventory (TRI)
Form R is required by organizations that release toxic chemicals as part of normal operations. This form
provides communities and the general public a nationwide inventory of hazardous chemical releases.
5.4 Hydrazine exposure remediation and control
The toxic nature of hydrazine exposure has been outlined in the previous parts of Section 5. Proper
engineering design, adherence to appropriate operating procedures, and the cooperation of all personnel
in good safety practices can mitigate hazards from operations with hydrazine. However, if personnel
hazards from hydrazine occur, proper procedures should be followed: this section details such
procedures.
5.4.1 First aid
In Case of Emergency, a physician should be notified while first aid is being performed. If no one is
available to send for a physician, treat the individual first and then get help. In less severe situations, the
affected individual should seek a physician's opinion. These guidelines are general and must be carefully
AIAA SP-084-1999
101
evaluated for their appropriateness to the specific situation (Marsh and Knox 1970; CPIA 1984; Olin
Corporation 1992).*
It is important to note that in many cases, incorrect first-aid may be worse than no first aid at all
(Sax 1994).
All personnel, first responders, emergency medical technicians, nurses, doctors, etc., responding
to an exposure emergency involving hydrazine must have adequate personal protection. Caution
must be exercised at the accident site to avoid additional exposure of rescue personnel. Remove and
isolate contaminated clothing and shoes at the site. Secondary exposure is possible from contact with
exposure sites on victims, handling contaminated articles of clothing, and coming in contact with
contaminated exudate from the victim. Secondary exposure can take any form: inhalatory, dermal,
ocular, or even accidental ingestion.
5.4.1.1 Inhalation
Send for qualified personnel. The following actions should be performed:
(1) Immediately remove the individual from the exposure area. If the exposed individual is
conscious, have the person hold his or her breath or breathe shallowly while moving into fresh air
or until protective breathing equipment is available.
(2) If breathing has stopped, perform artificial resuscitation.
(3) Keep the individual quiet until a physician arrives. A physician may treat convulsions with
barbiturates, taking into consideration respiratory complications.
5.4.1.2 Skin Splash
Send for qualified personnel. The following actions should be performed by the exposed person. If it is
not possible for the exposed person to perform the following actions, the actions should be performed by
personnel with adequate personal protective equipment.
(1) Move the individual from the contaminated area and remove the person's contaminated clothing.
(2) Immediately wash the affected area with soap (mild detergent), and flush skin with large amounts
of water until no evidence of hydrazine remains (approximately 15 to 20 min); it is very important
to use large amounts of water to prevent skin blistering. If available, use an emergency shower.
(3) Seek treatment from a physician for alkali-type burns.
Water washing and tannic acid solution can help arrest hydrazine absorption through skin (Smith and
Clark 1971).
5.4.1.3 Eye Splash
Send for qualified personnel. The following actions should be performed by qualified personnel:
(1) Move the individual from the contaminated area.

*See also the database of Occupational Health Services, Inc. 11 West 42nd Street, 12th Floor, New York,
New York 10036. Database listing January 1991
AIAA SP-084-1999
102
(2) Flush the eyes with large amounts of water or normal saline solution, occasionally lifting upper
and lower lids, until no evidence of hydrazine remains (approximately 15 to 20 min). Do not put
anything in the eyes except water or normal saline solution. If available, use an emergency eye
wash.
(3) If no one is available to send for a physician, flush eyes first for 10 min, seek help, then resume
flushing. Treatment by a physician (preferably an ophthalmologist) should begin as soon as
possible.
5.4.1.4 Ingestion
Send for qualified personnel. The following actions should be performed by qualified personnel:
(1) Have the exposed individual drink large amounts of water immediately; then induce vomiting by
putting a finger down the person's throat. Do not use ipecac to induce vomiting. Do not use milk
as a diluent. Do not make an unconscious person vomit: get medical attention immediately.
(2) Keep the individual quiet until a physician arrives. A physician may treat convulsions with
barbiturates, taking into consideration respiratory complications.
NOTE: The effects from hydrazine exposure may be delayed. Keep the victim under observation.
Instructions for Qualified Medical Personnel
The following general treatments can only be undertaken by qualified medical personnel.
Ingestion of hydrazine: Administer charcoal slurry, aqueous or mixed with saline cathartic or sorbitol.
Dosage:
Adults 30 to 100 g
Children 15 to 30 g
Infants 1 to 2 g/kg
One dose of the cathartic may be mixed with the charcoal or given separately (Hall and Rumack 1996).
Seizures: Administer diazepam IV bolus; Adult - 5 to 10 mg initially, repeated every 15 min PRN up to
30 mg. Child - 0.25 to 0.4 mg/kg/dose up to 10 mg/dose. For uncontrollable or recurring seizures
administer phenytoin or phenobarbital (Hall and Rumack 1996).
Convulsions: Give vitamin B6 to offset the convulsant action of hydrazine poisoning (Back, Carter, and
Thomas 1978). Perform blood analysis for glucose, calcium, urea nitrogen, and carbon dioxide
(Occupational Health Services 1987).
Pulmonary Edema: Maintain ventilation and oxygenation with close arterial blood gas monitoring. If the
P02 remains less than 50 mmHg, PEEP, or CPAP may be necessary. Avoid net positive fluid balance;
monitoring through central line or Swan-Ganz catheter (Hall and Rumack 1996).
Methemoglobinemia: Administer 1 to 2 mg/kg of 1 percent methylene blue slowly by IV if the patient is
cyanotic and symptomatic, or the methemoglobin level is greater than 30 percent in an asymptomatic
patient. Additional doses may be required (Hall and Rumack 1996).
AIAA SP-084-1999
103
Antidotes: Pyridoxine may be antidotal. The dose of pyridoxine is 25 mg/kg, 1/3 given IM and 2/3 given
by IV over 3 h. Increase the dose by 25 mg/kg every 5 to 10 min to a maximum of 300 mg/kg/dose for
continuing symptoms (Hall and Rumack 1996).
Monitoring: Monitor the following to track the progress of an hydrazine exposure patient (Occupational
Health Services 1987):
liver enzymes,
hypoglycemia (decrease in glucose),
lung, liver, and renal function,
hemolytic anemias,
cardiac monitor, hypotension, and cardiac pressure. (Slowing of the heart has been observed in
test animals exposed to hydrazine.),
pulmonary edema after 48 to 72 h (lung damage),
convulsions,
serum hydrazine, and
methemoglobin concentration.
5.4.2 Personnel protection
The following general guidelines apply to protect personnel from exposure to hydrazine:
avoid breathing dust, vapors, or fumes from a contaminated area,
avoid skin contact with hydrazine or hydrazine-contaminated materials,
do not handle broken packages or containers of hydrazine without protective equipment,
wear full protective clothing in the presence of hydrazine (fireman's gear is inadequate), and
wear a self-contained breathing apparatus when fighting fires involving hydrazine or cleaning up
hydrazine spills where volatilization can occur.
The following is required by law:
*
do not eat or smoke in areas where hydrazine is handled, processed, or stored,
locate an eyewash fountain in the immediate vicinity of areas where exposure is possible, and
locate an emergency shower in the immediate vicinity of areas where exposure is possible.
5.4.3 Smell
The common-sense relationship of inhalation and olfaction (sense of smell) raises the question of using
the nose to detect airborne hydrazine. Because of its sharp ammoniacal odor (it is often described as
fishy or ammonia-like), hydrazine is easy to identify; however, concentrations high enough to detect
through this means [3 - 4 ppm, (Hall and Rumack 1996) are greater than the TLV-TWA. If hydrazine is

*
Occupational Health Services, Inc., 11 West 42nd Street, 12th Floor, New York, New York 10036.
Database listing January 1991
AIAA SP-084-1999
104
detected by smell, the area should be immediately vacated. Commercial detectors are available and are
required for safety (Section 5.4.5).
5.4.4 Medical surveillance
Manufacturers and certain processors of chemical substances that can cause significant adverse
reactions to health or the environment are required to keep general medical histories for 30 years on
personnel who are at risk (40 CFR 717 and the Toxic Substances Control Act (TSCA) Section 8(c)).
Emphasis is on monitoring the following:
central nervous system tests, peripheral neuropathy,
respiratory history,
pulmonary functions,
blood chemistry,
renal and liver functions,
kidney function,
skin exam,
by 17 chest P.A. X-ray,
differential blood cell morphology, and
ACGIH biological exposure indices for methemoglobin inducers: 1.5 percent of hemoglobin
methemoglobin in blood at any time during or at the end of a work shift.
5.4.5 Evacuation procedures
In the event of an uncontrollable fire involving hydrazine or a container of hydrazine exposed to direct
flame, evacuate personnel beyond a radius of 1524 m (5000 ft).*
In the event of spilled or leaking hydrazine, evacuation of downwind personnel must be considered.
5.4.6 Protective apparel
Protective clothing and equipment necessary to prevent any possibility of skin contact with hydrazine are
required of employers by the CFR (29 CFR 1910.133 (1991), Sections (a)(2), (a)(4), (a)(5), and (a)(6)).
Employers are required to ensure that clothing wet with hydrazine is kept in sealed containers until its
removal or decontamination and to inform those performing decontamination operations of the hazardous
properties of hydrazine. Goggles must comply with 29 CFR 1910.133 (1991), Sections (a)(2) through
(a)(6). If detectable concentrations are present, respirators with the following characteristics are required:
a self-contained breathing apparatus with a full face piece operated in a pressure-demand or positive-
pressure mode, or a supplied-air respirator with a full face piece operated in a pressure-demand or other
positive-pressure mode.
The protective clothing chosen depends on the type of operation being performed. Several sources
recommend specific protective apparel to be worn when working with or near hydrazine (CPIA 1984; Olin
Corporation 1992; Marsh and Knox 1970; USAF 1973). These recommendations cover gloves, shoes,
headgear, respirators, and suits. Because many materials (especially synthetic materials) react violently
with hydrazine, apparel must be carefully assessed for its compatibility with hydrazine and must be
chosen by knowledgeable personnel. Additional information about the compatibility of materials is
available in Section 4.
AIAA SP-084-1999
105
Special suits for use near hydrazine are commercially available: for example, the ILC Dover Model 12
Chemturion suit, the Self-Contained Atmospheric Protection Ensemble (SCAPE) manufactured by
Standard Safety Equipment Company, and the standard splash suit manufactured by Fab-Ohio
Company.
The chlorinated polyethylene ILC Dover suit encapsulates the wearer and requires connection to an
external breathing source. The SCAPE suit is made of polyvinyl chloride on cotton. This suit also
encapsulates the wearer, and an air supply is carried inside the suit on the wearer's back. The splash
suit is made of polyvinyl chloride and does not provide respiratory protection (Smith 1981). Additional
information on protective apparel is available from the U.S. Department of Health, Education, and Welfare
(HEW) (1978).
5.4.7 Fire fighting
When hydrazine is involved in a fire, extinguish the fire only if the flow of hydrazine can be stopped. In a
shallow spill or deep pool, supported by air or by another oxidizer (flare combustion), hydrazine fires are
most effectively doused with a coarse water spray. Water applied as a solid stream may be ineffective.
Extinguishers such as vaporizing liquids, powders, water fog, and foams are less effective because of the
wide flammability limits and re-ignition hazard of hydrazine. Halogenated hydrocarbon extinguishers
create additional hazards, and on large spills or deep pools, dry sodium bicarbonate extinguishes a flame
but does not cool the hydrazine, and re-ignition is possible (Marsh and Knox 1970; CPIA 1984). Use
flooding quantities of water to cool all affected containers. Water should be applied from as far a distance
as possible.
When hydrazine is not on fire but has been accidentally released, keep away sparks, flames, and other
ignition sources. Prevent hydrazine from entering water sources or sewers; control flow with dikes as
necessary. Stop leaks if it is possible to do so without hazard. Vapors and pools should be dispersed
with a water spray.
A self-contained breathing apparatus and protective apparel are recommended when dealing with burning
hydrazine (Olin Corporation 1992; HEW 1978). For additional information on fighting hydrazine fires,
refer to Shaver and Berkowitz (1984).
5.4.8 Spill cleanup
Spilled hydrazine liquid quickly leads to vapor concentrations in the air that exceed safe limits. Therefore,
prompt cleanup of spilled hydrazine is very important, and notification of the EPA may be necessary.
Hydrazine is a regulated hazardous substance under the Superfund regulations of 29 CFR 302.4 (1986a).
These regulations state that spills of more than 4.54 kg (10 lb) must be reported to the EPA's National
Response Center ((800) 424-8802). Guidelines on how to handle spills into surface water can be
obtained from the U.S. Department of Transportation (DOT) through the U.S. Coast Guard Chemical
Response Information System.
Spilled hydrazine is not only a toxicity hazard, but also presents fire, explosion, and compatibility hazards.
Because hydrazine readily reacts with many substances, these hazards must be carefully assessed. For
additional information, see Sections 2, 3, and 4.
The following points, adapted from Olin Corporation (1992), briefly summarize the procedures for
cleaning up hydrazine spills. These procedures are general and must be carefully evaluated for their
appropriateness to the specific situation:
AIAA SP-084-1999
106
(1) Wear Mine Safety and Health Administration/NIOSH-approved self-contained breathing
apparatus. Follow OSHA regulations for respirator use specified in the Code of Federal
Regulations, 29 CFR 1910 (1986b).
(2) Wear goggles, coveralls, butyl rubber gloves,* boots, and an apron. In cases of large spills,
exposed skin surfaces must be protected, and full protective suits with supplied air respiratory
systems may be necessary. See Section 5.3.2 for additional information on protective apparel.
(3) Depending on the quantity spilled, either absorb spilled hydrazine with a special absorbent
material (Delgado and Davis 1995) or flush spilled hydrazine with large amounts of water and
drain into a catch basin; transfer diluted hydrazine from the catch basin into an approved DOT
container, and seal the container.
(4) Neutralize any remaining material with dilute hypochlorite solution and flush with water.
(5) Dispose of all contaminated material (spill cleanup supplies, residue, contaminated soil, and
clothing) by environmentally approved methods (Section 5.3.6).
Additional information on cleaning up hydrazine spills is available from Shaver and Berkowitz (1984).
5.5 Hydrazine handling
Use of hydrazine requires special consideration be given to the design and construction of process
equipment and material storage areas. The toxicity of hydrazine requires that monitoring equipment,
sensitive to the mandated exposure limits, be available and operational. The EPA has designated
hydrazine as a hazardous material, and disposal of waste material containing hydrazine must adhere to
strict guidelines. This section details current information on these topics.
5.5.1 Engineering design
The purpose of engineering design is to control levels of hydrazine at or below the recommended
environmental limits and to minimize personnel contact. The following examples provide some general
design guidelines for hydrazine fuels (HEW 1978):
Storage tanks: employ sump and dip leg design.
Joints and connections: use welded connections.
Components: choose commonly used relief valves and control valves; minimize overhead
valves; and use material selection guidelines in Section 4.
Water supply: ensure that adequate water is available (Section 2.4.2).
Explosion-proofing: follow electrical and fire codes.
Emergency equipment: provide emergency shower, eye wash, breathing-air station, and
sprinkler system.
Nitrogen padding: maintain an inert atmosphere.

* Butyl rubber gloves have been found to offer excellent permeation protection against hydrazine (Smith
1980). Caution: surface degradation of the butyl rubber gloves has been observed upon direct
exposure to hydrazines, especially in conjunction with pump oil. Gloves must be discarded when
degradation is observed.
AIAA SP-084-1999
107
Good housekeeping: keep combustible materials separate.
Material compatibility: consider metals, soft goods, and lubricants for compatibility (additional
information about compatibility is available in Section 4).
Maintenance: ensure that components can be removed and repaired as necessary.
5.5.2 Storage containers
Improper storage and handling can quickly lead to other types of hydrazine hazards, including exposure,
fire, explosion, and material incompatibility.
Hydrazine is highly reactive with oxygen in the air, oxidizing agents, organic matter, and oxides of metals
such as iron, copper, lead, manganese, and molybdenum (Olin Corporation 1992). Because hydrazine is
so reactive, storage conditions must be chosen carefully. For additional information on hydrazine
reactivity, consult Sections 2, 3, and 4.
Although it is not susceptible to decomposition through normal impact or friction (hydrodynamic shear),
hydrazine reacts with oxygen and carbon dioxide in the air; and therefore, should be stored under a
nitrogen blanket.*
Guidelines for designing safe storage and processing equipment for hydrazine fuels are available (Cloyd
and Murphy 1965):
1. use a minimum of mechanical joints,
2. consider maximum system pressure,
3. avoid low-lying liquid traps,
4. provide a purge system for inert gases, and
5. vent with a high stack or scrubber.
The following recommendations apply to containers used for storing and transferring hydrazine (Olin
Corporation 1992):
store hydrazine in a DOT-approved shipping container; keep container tightly closed and vent
carefully when opening,
ground the container electrically and protect it from electrical sparks, open flames, and heat
sources, and
maintain hydrazine in a nitrogen atmosphere.
For additional information on storage equipment requirements, consult the CPIA (1984). Instructions for
cleaning hydrazine systems are detailed by the USAF (1973).\

* WSTF internal communication with D. Davis, March 1993.
AIAA SP-084-1999
108
5.5.3 Storage areas
The following recommendations apply to buildings and areas where hydrazine is stored (Olin Corporation
1992; USAF 1973):
store only in well-ventilated areas,
locate away from occupied areas,
separate from oxidizing materials,
maintain at a temperature below 322 K (120 F); hydrazine contracts on freezing, so there is no
need to worry about vessel damage (Hannum 1985),
store in a fire-resistant building with an automatic sprinkling system over storage tanks,
provide adequate water supply for fighting fires and controlling spills,
ground electrically and protect from lightning,
surround by a dike system that drains into a collection basin,
provide access by at least two roads that are wide enough for turning a motor vehicle around,
ventilate well, either naturally or mechanically,
keep clean and free of organic materials and oxidizers,
keep clean and free of metal oxides; hydrazine decomposition is catalyzed by metal oxides
(Section 4.2), and
store with a nitrogen blanket to reduce absorbing moisture and thus fuel contamination; hydrazine
is hygroscopic.
5.5.4 Transportation
Transportation of hydrazine on public thoroughfares is covered by federal and state transportation
standards and guidelines (see 5.5.7). Where conditions and requirements of use on private thoroughfare
are similar to those of public thoroughfares, federal and state transportation standards and guidelines
must be used.
Hydrazine can be transported by means that vary from tanks on barges, railroad cars, and trucks to small
cylinders. Standard terminology* for transport containers includes:
cargo tanks: transport containers designed for highway service, such as over-the-road trailers
and tank motor vehicles, and
cylinders: pressure vessels with a circular cross section designed for pressures greater than
275.7 kPa (40 psia).

5.5.4.1 Transport on public thoroughfares
While most commerce on public thoroughfares involves commercial carriers, the responsibility for
complying with federal and state transportation laws rests not only with them but also with the
organizations that handle and receive hydrazine.

* Definitions have been developed by the Department of Transportation (DOT) in 49 CFR 171.8 (1992).
AIAA SP-084-1999
109
5.5.4.1.1 Training
Personnel involved in handling, receiving, shipping, and transport of a hazardous material must receive
Hazardous Materials (HAZMAT) training (49 CFR 172.700 1996).
5.5.4.1.2 Emergency response
The information required to adequately respond to all phases of transport emergency is required at
facilities where hazardous materials are either loaded, stored, or handled (49 CFR 174.600 1992).
Emergency actions for first responders arriving at a transportation incident involving hydrazine are
outlined in SPECIAL REPORT 28 of the Emergency Response Special Reportbook published by the
Research and Special Programs Administration of the DOT (DOT, 1993). The guidelines direct how to
secure and isolate the accident scene, tell who to contact, give fire fighting directions, give handling
directions for a spill or leak, and list first aid procedures.
5.5.4.1.3 Transport requirements for hydrazine
The Department of Transportation (DOT) designates for the purpose of transportation, hazardous
materials and prescribes the requirements for shipping, shipping papers, package marking, labeling, and
placarding. The requirements for shipping anhydrous hydrazine or aqueous solutions containing greater
than 64 percent hydrazine by mass are shown in Table 25.
Table 25 Department of Transportation shipping requirements
a
Category Requirement
Hazard Class Flammable liquid
ID Number UN2029
Labels Flammable Liquid and Poison
Packaging (Exceptions) None
Packaging (Specific Requirements) 49 CFR 173.276
Passenger aircraft or railcar Forbidden
Cargo aircraft only 2.4 L (5 pints)
Water Shipment Requirements May be stowed on deck of a cargo ship but must
be segregated per requirements for corrosives.
Forbidden on passenger vessels.
a
49 CFR 172.101 Hazardous Materials Table 1996.
Federal shipping requirements are subject to change so the latest issue of the CFR should be consulted.
5.5.4.1.4 Transport design requirements
Cylinder and tankage design must follow accepted design practice (ASME 1995) for cargo tankers to
transport hypergolic propellants over public thoroughfares (Pohlhammer and Drown 1991).
AIAA SP-084-1999
110
5.5.4.2 Transport on privately controlled thoroughfares
Federal and state transportation guidelines can be applied in lieu of special requirements on privately-
controlled sites where conditions and requirements of use are similar to public thoroughfares.
Special equipment or operations used for the transport of hydrazine must meet federal and state labor
requirements (29 CFR 1996) as well as additional requirements of the Authority Having Jurisdiction
(AHJ).
5.5.5 Monitoring equipment
Any operation involving hydrazine has the potential of exposure of personnel to the vapors. Monitors
should be used to assess the safety of an area before unprotected personnel are allowed to enter. Once
personnel enter into an area where exposure is possible, personnel monitors or area monitors should be
used continuously.
The selection of a detector or laboratory analysis technique requires a thorough understanding of the
intended application and personnel monitoring procedure required by regulations. The monitoring
method chosen will depend on the reason for monitoring. Monitors are typically used for real-time
hydrazine measurement applications, while laboratory analysis techniques are required by health
regulations.
Monitors can be conveniently classified by application into four types; personnel exposure or TLV
monitors, leak or spill monitors, flammability limit monitors, and monitors for vacuum service. Personnel
exposure monitors, which are employed to prevent exposure of unprotected personnel to more than the
TLV concentrations of toxic gases or vapors, are used in two distinctly different manners; area monitors
and personnel dosimeters. Area monitors are used to assure that the concentration of the toxic
compound is below the TLV or, very often, below some action level such as half of the TLV, before
personnel are allowed to enter and work in an area without breathing protection. Area monitors may
either be hand-held portable units or statically mounted units. Area monitors provide real-time
measurement of the concentration of the toxic compound in the air and are equipped with an alarm.
Personnel dosimeters are attached directly to the clothing of personnel near the breathing zone.
Generally, the dosimeters employ some sort of an absorbent tube and a sampling pump which
continuously pulls a portion of the atmosphere through the absorbent tube. Passive dosimeters are also
possible. Personnel dosimeters do not provide real-time measurement of the concentration of the toxic
compound. Instead they provide the time-weighted average concentration to which the person was
actually exposed.
Personnel exposure monitors have the lowest detection range of any of the monitors. Generally they
need to be capable of detecting hydrazine at 10 percent of the TLV. With the proposed ACGIH TLV of 10
ppb, manufacturers of hydrazine area monitors have been challenged to lower the detection limits of their
monitors by at least an order of magnitude. Currently, no commercially available monitor can meet this
requirement although there are some promising development efforts underway (Leavitt et al. 1991;
Leavitt, Mattson, and Young 1992). Although these monitors possess extreme sensitivity, care must be
taken in matching the monitor to the application since some of the monitors possess a limited
measurement range and may not measure more than 2 to 10 times above the TLV concentration.
Dosimeters are available that can provide detection and measurement over the 8-h work day that is well
below the 10-ppb time-weighted average.
Leak or spill monitors are intended to be used by personnel already in protective breathing gear to locate
leaks or spills and to aid in the clean-up or decontamination effort. They may also be statically mounted
in unattended areas as a warning device for a spill or leak problem. Because the possible concentration
range resulting from leaks or spills is large, these monitors need to have a large dynamic range and be
AIAA SP-084-1999
111
relatively tolerant of exposure to very high concentrations of the toxic compound. Typically, the lower
detectable limit of these devices is in the ppm range and the upper detectable limit is in the 1000-ppm
range.
Flammability limit monitors are usually statically mounted area monitors used for continuous surveillance
of storage areas or equipment operational areas. The measurement range is from 5 to 200 percent of the
lower flammable limit for the compound of interest, although some monitors have a much larger range.
Monitors for vacuum service are a relatively new requirement. These are used to provide information
about a toxic compound inside of a vacuum test chamber prior to repressurizing to atmospheric pressure.
These may be used in any of the three modes discussed above: as TLV monitors, as leak or spill
detectors, and for prevention of flammable mixtures resulting from back-filling a chamber with air.
Various methods have been developed to measure hydrazine in air, including electrochemistry,
calorimetry, chemiluminescence, chemiresistors, ion mobility spectrometry, infrared spectroscopy, gas
chromatography, solid-state semiconduction, CO
2
laser photoacoustics, and photoionization. In one
laboratory monitoring method based upon OSHA Method # 20, a known quantity of contaminated air is
passed through a sampling media such as a glass tube filled with acidified firebrick. The sampling
medium is then analyzed at a specially equipped laboratory to determine the amount of hydrazine
collected (Johnson and Baker 1992). Immediate measurement of hydrazine generally employs a direct-
reading method or a passive dosimeter that indicates a level of exposure. The principles governing many
of these methods are summarized by Saunders and Larkins (1976).
If personnel are likely to be exposed to hydrazine vapors or aerosols, exposure monitoring is
recommended by NIOSH. The monitoring procedure should quantify personnel exposure during the
workday, and the results must be comparable to the OSHA PEL or the ACGIH TLV. According to NIOSH
recommendations, each regulated area should be sampled at least every 6 months while hydrazines are
in use. For operations lasting less than 6 months, sampling is suggested at least once during the
operation. If samples indicate exposure to elevated levels of hydrazines, NIOSH recommends that the
contamination be controlled and that exposed employees be monitored at least once a week until two
consecutive measurements, at least one week apart, are within safe limits (HEW 1978). Personnel
exposure to hydrazine is usually monitored by an industrial hygienist.
Determining the best choice among the many different types of detection devices is difficult. The best
advice is to survey the literature (some excellent material is cited in the References) and carefully
evaluate the detector for the needs of your application. In any application, it is important to abide by the
information given in the product literature. Install and use each unit only in strict accordance with the
manufacturer's instructions. Young, Helms, and Travis provide an up-to-date analysis of competing
detectors and theories (1990).
5.5.5.1 Monitors commercially available and under development (Proposed ACGIH guidelines)
The following list of commercially available monitors and prototype systems under development, and the
literature in which they are reviewed, may aid in selecting an appropriate instrument for monitoring
hydrazine levels. The list is organized by detector purpose or level of detection with an annotation to the
basic technology used. While many systems have been investigated, the list reflects systems that either
have been successfully operated or positively evaluated and it is not meant to be exhaustive.*
Technologies that have been investigated but at the present time do not meet current requirements
include chemiluminescence, gas chromatography, solid-state semiconduction, and photoionization. This
list is not an endorsement for these products. Additional monitors are reviewed in Schmidt (1984).

* Personal communication with R. Young, NASA Kennedy Space Center, August 1992.
AIAA SP-084-1999
112
(1) Detector (CO
2
Photoacoustics) Aerospace Corp., reviewed by the CPIA (1981) and Loper (1981).
Comment: laboratory prototype.
(2) Interscan (electrochemistry), reviewed by the CPIA (1981); Loper (1981); Rose
and Holtzclaw (1984); Miller (1987); and Young, Helms, and Travis (1990).
(3) MK-2 Portable Vapor Detector, GMD Systems, Inc. (colorimetry), reviewed by Leavitt et al.
(1991).
(4) Dosimeter badge, GMD Systems, Inc. (colorimetry), reviewed by Brown and Nunez (1991).
(5) Dosimeter badge, GEO-CENTERS, Inc. (colorimetry), reviewed by Taffe, Travis, and Rose-
Pehrsson (1989) and Rose-Pehrsson and Crossman (1992).
Comment: planned for lapel or wall attachment.
(6) Interim Active Vanillin Sampler (colorimetry), reviewed by Blaies, Leavitt, and Young (1991) and
Young et al. (1992).
(7) Airborne Vapor Monitor, Graseby Analytical, Ltd, Watford, UK (ion mobility spectrometry),
reviewed by Eiceman et al. (1990), Leavitt et al. (1991), Cross et al. (1992a), and Cross et al.
(1992b).
Comment: possesses limited dynamic range.
(8) Amperometric technique (liquid chromatography), reviewed by Johnson and Baker (1992).
Comment: laboratory technique.
5.5.5.2 Monitors for leak detection (Meet current ACGIH guidelines)
(1) Ecolyzer Series 6000 and 7000 (electrochemistry), Energetics Science, Inc., reviewed by the
CPIA (1981); Energetics Science, Inc. (1977); Loper (1981); Luskus and Kilian (1986); Rose and
Holtzclaw (1984); Saunders et al. (1978); Young and Travis (1987); Young, Helms, and Travis
(1990).
Comments: series 6000 - very portable, designed for exposure monitoring 0 to 2 ppm; series
7000 - designed for leak detection; does not meet current ACGIH guidelines.
(2) Model 4000, Interscan (electrochemistry), reviewed by the CPIA (1981); Loper (1981); Rose and
Holtzclaw (1984); Miller (1987); and Young, Helms, and Travis (1990).
Comments: very portable; designed for exposure monitoring 0 to 2 ppm.
(3) Model LD-18, Interscan (electrochemistry), reviewed by Young and Travis (1987).
Comment: wall-mount version of Model 4000 monitor.
(4) Bruel & Kjaer 1302, Bruel & Kjaer, Denmark (photo-acoustic infrared spectroscopy), reviewed by
Young, Helms, and Travis (1990) and KSC-DL-3384.
Comments: 4 orders of magnitude detection range; does not meet current ACGIH guidelines.
AIAA SP-084-1999
113
(5) Toxic Level Detector 1 (TLD-1), MDA Scientific, Inc. (colorimetry), reviewed by Rose and
Holtzclaw (1984); paper tape for TLD-1, MDA Scientific, Inc., reviewed by the CPIA (1981), and
Loper (1981).
Comments: can change application by changing tape; humidity reduces tape reading.
(6) Model 7100, MDA Scientific, Inc., reviewed by Johnson (1986); Paper tape for TLD-1, MDA
Scientific, Inc., reviewed by the CPIA (1981), and Loper (1981).
Comments: area personnel exposure monitor for use in area with controlled humidity; can be
calibrated at low concentrations.
5.5.5.3 Flammable concentration monitor
Bruel & Kjaer 1302, Bruel & Kjaer, Denmark (photo-acoustic infrared spectroscopy), reviewed by
Young, Helms, and Travis (1990).
5.5.5.4 Vacuum detection
(1) Sensors (chemiresistor), Naval Research Laboratory, Washington D. C., reviewed by Rose-
Pehrsson, Grate, and Klusty (1989)
Comments: under development; not commercially available
(2) Dosimeter based on conductive polymer technology (chemiresistor), reviewed by Zakin,
Bernstein, and Moody (1990)
Comments: under development; not commercially available
(3) Real-time monitoring based on conductive polymer technology (chemiresistor), reviewed by
Zakin, Bernstein, and Ellis (1992)
Comment: not commercially available
(4) Time of flight mass spectrometer detector, A. D. Little Co.
Comment: not commercially available
5.5.6 Waste disposal
Hydrazine waste is created in two ways:
(1) hydrazine used in a process (process waste), and
(2) off-specification or spilled hydrazine.
The disposal of any contaminated product or exposed materials must be done in a manner approved by
the appropriate federal, state, and local regulatory agencies. Hydrazine spill residues have been given
the hazardous waste number U133 by the EPA. Hydrazine can also meet the definition of a hazardous
waste defined by the Resource Conservation and Recovery Act (RCRA) (40 CFR 260 (1991)).
AIAA SP-084-1999
114
Exposure to hydrazine can affect wildlife, soil, water, and air, although hydrazine degrades relatively
quickly when exposed to the environment. Information on the environmental effects is beyond the scope
of this document. For more information, contact the EPA.
To avoid environmental contamination, proper procedures for disposal of hydrazine should be followed.
In general, process waste constituting a reactivity or toxicity hazard is a regulated waste, and must be
disposed of at a permitted treatment site that is regulated by the EPA. Specification product that has
degraded or spilled is a regulated hazardous waste and must also be disposed of at a permitted
treatment site. In addition, any residue or contaminated soil, water, or debris resulting from a spill is
regulated as a hazardous waste per 40 CFR 261 (1991).
Although process waste that does not pose a reactivity or toxicity hazard is not regulated by the EPA, the
hydrazine it contains is considered a hazardous constituent, and as such must be treated before being
discharged into the environment. Such waste may be treated on-site by evaporation or other standard
methods.
Hydrazine can be destroyed by chlorination, but this process creates significant amounts of methyl
chloride, lesser amounts of nitrogen trichloride NCl
3
, and detectable amounts of N,N-
dichloromethylamine, all of which can create a potential environmental hazard. Ozone is not cost-
effective for decontaminating hydrazine in aqueous solution. For more information on EPA requirements
for hydrazine disposal, refer to 40 CFR 261 (1991).
5.5.7 Current regulatory enforcement
A list of current regulatory enforcement is provided below:*
OSHA Standard 29 CFR 1910.1200 - Hazard Communication
OSHA Standard 29 CFR 1910.1000 - Air Contaminants Table Z-1
OSHA Standard 29 CFR 1910.94 - Ventilation
OSHA Standard 29 CFR 1910.134 - Respiratory Protection
OSHA Standard 29 CFR 1910.20 - Access to Employee Exposure and Medical Records
OSHA Standard 29 CFR 1910.132 - Personal Protective Equipment
OSHA Standard 29 CFR 1910.141 - Sanitation
OSHA Standard 29 CFR 1910.133 - Eye and Face Protection
OSHA Standard 29 CFR 1910.151 - Medical Services and First Aid
OSHA Standard 29 CFR 1910.1450 - Occupational Exposure to Hazardous Chemicals in
Laboratories
OSHA Standard 29 CFR 1910.120 - Hazardous Waste Operations
OSHA Standard 29 CFR 1910.119 - Process Safety Management of Highly Hazardous
Chemicals Standard
40 CFR 717 - Records and reports of allegations that chemical
substances cause significant adverse reactions to health
or the environment Section 8(c) of the Toxic Substances
Control Act (TSCA)
40 CFR 116 - Designation of Hazardous Substances - Section
311(b)(2)(a) of the Clean Water Act; Isomers and hydrates
of hydrazine and solutions and mixtures containing these
materials

* Occupational Health Services, Inc. 11 West 42nd Street, 12th Floor, New York, New York 10036. Data
base listing January 1991.
AIAA SP-084-1999
115
40 CFR 172.101 - Tables of hazardous materials, their description, shipping
name, class, label, packaging, and other requirements;
Designated as hazardous material for the purpose of
transportation. Also see 49 CFR 172.102 and 49 CFR
171.12 for international shipping requirements.
40 CFR 261.33(e) - Discarded commercial chemical products, off-specification
species, containers, and spill residues of commercial
chemical products or manufacturing chemical
intermediates identified as acute hazardous waste
45FR33084 5/19/80
5.5.8 Additional information
It is beyond the scope of this Special Report to offer full coverage of the industrial hygiene practices. Nor
could this Special Report ever provide a comprehensive plan for establishing safe working practices in
every conceivable scenario. For those readers needing such broad-based information, consult Clark
et al. (1968), and the more recent Patty's Industrial Hygiene and Toxicity (Reinhardt and Britteli 1981); the
latter offers the most thorough coverage in a three-volume series. Back, Carter, and Thomas (1978)
provide a decent compendium of toxicity studies published between 1970 and 1978. Data base services
now offer excellent up-to-date information on all aspects of hydrazine usage.*
5.6 Assessment example
The example is to complete the information about hydrazine found on a Material Safety Data Sheet
(MSDS).**
Section I - Product Identification
Chemical Name & Synonyms: Anhydrous Hydrazine
Chemical Family: Formula: Product Name:
Hydrazine N
2
H
4
Anhydrous Hydrazine
Description: Propellant and aerospace fuel
OSHA Hazard Classification: Carcinogen, corrosive, highly toxic, sensitizer, skin and eye hazard; liver,
kidney, nervous system, blood, reproductive and lung toxin, flammable liquid.
Section II - Component Data
Product Composition:
CAS or Chemical Name: Hydrazine
CAS No.: 302-01-2
Percentage Range: 98 - 100%
Hazardous Per 29 CFR 1910.1200: Yes
Exposure Standards:
OSHA (PEL) TWA: 0.1 ppm (0.1 mg/m
3
) Ceiling: none STEL: none

* Much of the data in this section was researched from the TOMES system (Hall and Rumack 1996).
**This MSDS is based upon material provided by Olin Corporation for their hydrazine MSDS. More up-to-
date information (in brackets) has been provided where available.
AIAA SP-084-1999
116
ACGIH(TLV) TWA: 0.01 ppm (0.013 mg/m
3
) Ceiling: none STEL: none
CAS or Chemical Name: Water
CAS No.: 7732-18-5
Percentage Range: 0.1-1.0%
Hazardous per 29 CFR 1910.1200: No
Exposure Standards: None established
Section III - Precautions for Safe Handling and Storage (see Section 5.4 of this special report)
Handling Precautions: Avoid contact with eyes, skin or clothing. Do not take internally. Upon contact
with skin or eyes, wash off with water. Avoid breathing mist or vapor.
Storage Conditions: Store in a well ventilated area under nitrogen atmosphere away from heat, sparks,
open flames, and oxidants. Do not contaminate. Keep container closed when not in use. Drums should
be vented carefully when opening. All containers should be electrically grounded. An inert atmosphere
must be maintained over anhydrous hydrazine at all times. Nitrogen has been adopted as the padding
material for anhydrous hydrazine storage and transfer.
Product Stability And Compatibility:
Shelf Life Limitations: One year
Incompatible Materials For Packaging: Package only in Teflon or 304L or 347 stainless steels
containing less than 0.5% molybdenum.
Incompatible Materials For Storage Or Transport: Oxidizing agents; acids; metal oxides; metals
other than low-molybdenum (less than 0.5%) stainless steel, Inconel, titanium, selected aluminum alloys
and chromium; and organic materials with high surface area such as rags, cotton waste, sawdust, etc.
Section IV - Physical Data (see Annex A of this special report)
Appearance: Odor: Molecular Weight:
Clear, colorless liquid Ammonia 32.04
Freezing Point: Vapor Pressure: Volatiles:
1.5 C (35F) 14.9 mmHg at 25 C 100% by volume
Boiling Point: Solubility In Water: Evaporation Rate:
113.5 C (236 F) miscible No data
Specific Gravity: Bulk Density: pH @ 25 C:
1.004 1.004 (g/cm
3
) 10.1 - 10.7(1% solution, in
neutral distilled water)
Vapor Density: Decomposition Temperature:
1.1 (Air=1) > 250 C (482 F)
Coefficient Of Oil/Water Distribution: No Data
Section V - Personal Protective Equipment Requirements
Protective Equipment: Ventilation Requirements:
Respiratory: Wear a NIOSH/ As required to keep airborne concentrations
MSHA approved respirator if any below the TLV.
AIAA SP-084-1999
117
exposure is possible
Skin: Wear gloves, boots, apron and a face shield with safety glasses. A full impermeable suit is
recommended if exposure is possible.
Equipment Specifications:
Respirator Type: Wear NIOSH/MSHA approved positive-pressure air-supplied respirator
Glove Type: Butyl rubber
Boot Type: Butyl rubber
Apron Type: Butyl rubber
Protective Suit: Butyl rubber
Section VI - Fire and Explosion Hazard Data (see Sections 2 and 3 of this special report)
Flammability Data:
Flammable Yes
Combustible: No
Pyrophoric: No
Flash Point: 51C (124 F) Method: Open Cup
Explosive Limit: Upper: 98% Lower: 2.5%
Autoignition Temperature: 270C (518F)
Method: U.S. Bureau of Mines Bulletin 627.
Extinguishing Media: Alcohol foam, carbon dioxide, water spray.
Special Fire Hazard & Fire Fighting Procedures Flood with water to prevent reignition and to keep fire-
exposed containers cool. See Section XI for protective equipment for fire fighting.
Section VII - Reactivity Information
Conditions Under Which This Product May Be Unstable:
Temperatures Above: 270 C (518 F)
Mechanical Shock or Impact: No
Electrical (Static) Discharge: Yes
Hazardous Polymerization: Will not occur
Incompatible Materials: Avoid contact between anhydrous hydrazine and strong oxidizers such as
hydrogen peroxide, nitrogen tetroxide, fluorine, halogen fluorides, and fuming nitric acid. Such
contact will result in immediate ignition or explosion. Avoid contact with metal oxides such as those
of iron, copper, lead, manganese, and molybdenum. Contact with such metallic oxides may lead to
flaming decomposition. Avoid contact with organic materials having large surface areas or porous
surfaces. Adsorption of anhydrous hydrazine by rags, cotton waste or similar organic materials will
eventually result in spontaneous combustion.
Hazardous Decomposition Products: Ammonia, hydrogen
Summary of Reactivity:
Oxidize: No
Pyrophoric: No
Organic Peroxide: No
Water Reactive: No
Corrosive: Yes
AIAA SP-084-1999
118
Other: Reducing agent
Section VIII - First Aid
Eyes: Immediately flush with large amounts of water for at least 15 min, occasionally lifting the upper
and lower eyelids. Call a physician at once.
Skin: Immediately flush with water for at least 15 min Call a physician. If clothing comes in contact with
the product, the clothing should be removed immediately and should be laundered before re-use.
Ingestion: Immediately drink large quantities of water. Induce vomiting. Call a physician at once. DO
NOT give anything by mouth if the person is unconscious or if having convulsions.
Inhalation: If a person experiences nausea, headache or dizziness, the person should stop work
immediately and move to fresh air until these symptoms disappear. If breathing is difficult,
administer oxygen, keep the person warm and at rest. Call a physician. In the event that an
individual inhales enough vapor to lose conscious, the person should be moved to fresh air at
once and a physician should be called immediately. If breathing has stopped, artificial respiration
should be given immediately. In all cases, ensure adequate ventilation and provide respiratory
protection before the person returns to work.
Section IX - Toxicology and Health Information (see Sections 5.1 and 5.2 of this special report)
Routes of Absorption: Inhalation, dermal, oral, eye contact
Warning Statements And Warning Properties:
May be fatal if inhaled, absorbed through skin or ingested. Harmful if exposed to eyes.
Human Threshold Response Data:
Odor Threshold: 3.7 ppm
Irritation Threshold: This value has not been established
Immediately Dangerous To Life Or Health: 80 ppm
Signs, Symptoms, and Effects of Exposure
Inhalation:
Acute: Hydrazine is highly irritating to the nose, throat, upper respiratory tract and lungs. Inflammation
of the respiratory tract may lead to bronchitis. Vapor can also cause eye irritation. Pulmonary
edema and lung damage may occur. Damage may also result to liver, kidneys and blood. High
exposure may give rise to hemolysis of the blood cells. Vomiting, diarrhea, nausea, dizziness,
cyanosis, and convulsions may also occur. A single exposure would not likely produce fetal toxicity
and malformations, but several exposures may cause these effects to the fetus.
Chronic: The repeated inhalation of hydrazine may produce inflammation of the nasal, tracheal and
bronchial tissue. Chronic bronchitis can result. Damage to the liver, kidneys, and blood may also
occur. Damage to blood may be characterized by hemolysis and reduction of packed cell volume.
Fetal toxicity and malformations can also result. Repeated inhalation may produce cancer.
Animal Toxicity:
Acute Oral LD 50: 60 mg/kg (rat)
AIAA SP-084-1999
119
Acute Dermal LD 50: 93 mg/kg (rabbit)
Acute Inhalation LC 50: 570 ppm for 4 h (rat)
Irritation: Corrosive to the skin and eyes
Aquatic Toxicity: Hydrazine was fatal to rainbow trout at a concentration of 146 mg/L within one
hour of exposure.
Acute Target Organ Toxicity: Damage to lungs, liver, kidneys, blood and central nervous system.
Chronic Target Organ Toxicity: Repeated exposure to hydrazine has produced damage to the
lungs, liver, kidneys and blood. Cancer has also been observed in laboratory animals
Reproductive And Developmental Toxicity Hydrazine has been shown to produce embryolethality
and fetal malformations in laboratory animals.
Carcinogenicity Hydrazine is recognized as a cancer-causing agent in animals by IARC and OSHA.
It is considered a suspect carcinogen in humans by these two organizations.
Mutagenicity Hydrazine has been shown to cause DNA and chromosomal damage in a number of
test systems. It is considered mutagenic.
Section X - Transportation Information
This material is regulated as a DOT hazardous material.
DOT Description From The Hazardous Materials Table 49 CFR 172.101: Anhydrous hydrazine -
Flammable Liquid UN 2029 Poison - Inhalation Hazard
Reportable Quantity: 1 lb (per 49 CFR 172.101)
The material described above is subject to the U.S. DOT Hazardous MATERIALS REGULATIONS via the
modes and packaging quantities indicated below with the letter x
Mode: Packaging Quantities
x rail x Bulk x Non-Bulk
x motor x Bulk x Non-Bulk
x water x Bulk x Non-Bulk
x air x Bulk x Non-Bulk
The applicable packaging sections in 49 CFR is 173.276.
Section XI - Spill and Leakage Procedures (see Sections 5.3 and 5.4 of this special report)
For all transportation accidents contact Chemtrec 800-424-9300.
Spill Mitigation Procedures: Hazardous concentrations in air may be found in the local spill area and
immediately downwind. Remove all sources of ignition. Stop the source of spill as soon as possible and
notify appropriate personnel.
Air Release: Vapors may be suppressed by the use of a water fog.
Water Release: This material is soluble in water. Contain all water for treatment and/or disposal.
Land Spill: Dike spill area immediately. Dilute material with water on a ratio of at least 3 to 1. Remove
liquid and place in proper containers for the material. Begin neutralization or disposal procedures as
soon as possible.
Spill Residues: Dispose of per guidelines under Section XII, WASTE DISPOSAL. This material may be
neutralized for disposal; you are requested to contact OCEAN at 800-OLIN-911 before beginning any
such operation.
AIAA SP-084-1999
120
Personal Protection For Emergency Spill And Fire-Fighting Situations: In case of fire, use normal
fire fighting equipment. Response to this material requires the use of a full encapsulated suit and self-
contained breathing apparatus (SCBA).
Additional protective clothing must be worn to prevent personal contact with this material. Those items
include but are not limited to: boots, gloves (butyl rubber), hard hat, splash-proof goggles, full face
shield and impervious clothing, i.e., chemically impermeable suit.
Section XII - Waste Disposal
If this product becomes a waste, it meets the criteria of a hazardous waste as defined under 40 CFR 261
and would have the following EPA hazardous waste number: U133.
If this product becomes a waste, it will be a hazardous waste which is subject to the Land Disposal
Restrictions under 40 CFR 268 and must be managed accordingly. As a hazardous waste, it must be
disposed of in accordance with local, state and federal regulations in a permitted hazardous waste
treatment, storage and disposal facility by incineration.
CARE MUST BE TAKEN TO PREVENT ENVIRONMENTAL CONTAMINATION FROM THE USE OF
THIS MATERIAL. THE USER OF THIS MATERIAL HAS THE RESPONSIBILITY TO DISPOSE OF
UNUSED MATERIAL, RESIDUES AND CONTAINERS IN COMPLIANCE WITH ALL RELEVANT
LOCAL, STATE AND FEDERAL LAWS AND REGULATIONS REGARDING TREATMENT, STORAGE
AND DISPOSAL FOR HAZARDOUS AND NONHAZARDOUS WASTES.
Section XIII - Additional Regulatory Status Information
Toxic Substances Control Act: The components of this product are listed on the Toxic Substance
Control Act Inventory.
Superfund Amendments and Reauthorization Act Title III: Hazard Categories, per 40 CFR 370.2
Emergency Planning and Community Right to Know: (40 CFR 355, APP. A)
Extremely hazardous substance - threshold planning quantity 454 kg (1000 lbs).
AIAA SP-084-1999
121
Annex A: Hazard assessment example
This section presents a hazard assessment example using a typical hydrazine propulsion system that is
stored and operated in a test cell. A hazard is created when the propulsion system cannot contain the
effects of a hazardous event occurring inside the system. If the propulsion system fails, a hazard
assessment must also be performed on the test cell. The propulsion system and test cell are
characterized in the specifications (Table A.1) and illustrated in Figure A.1. Assumptions made in
performing this assessment are stated in the text.
A.1 Assessing system hazards
A.1.1 Fire hazard
A.1.1.1 Criteria
Assume that after operating the propulsion system, the line between the isolation valve and thruster
valves contains hydrazine vapor in concentrations within the flammability limits. To have a fire hazard,
ignition energies that exceed the minimum ignition energy or temperatures at which the hydrazine vapor
could autoignite must be present. Assume that neither the minimum ignition energy nor the autoignition
temperature is present in our example.
After operating, the propulsion system may still contain liquid hydrazine. When neat liquid hydrazine is at
300 K (80 F), the hydrazine vapor above the liquid is flammable. As specified, the liquid hydrazine in this
system is at 315 K (108 F). The potential for a fire hazard from the liquid is still low, because ignition
sources are not present.
A.1.1.2 Effects
A propagating fire in the propulsion system is unlikely. Nevertheless, if the heater failed in the on
position, the temperature in the system could reach the autoignition point. The amount of heat generated
would depend on the quantity of hydrazine vapor present. In any case, damage to softgoods in the
valves could occur, releasing liquid hydrazine into the test cell. The hazards resulting from such a spill
are considered in Section 2 of this example.
A.1.2 Explosion hazard
A.1.2.1Deflagration
A.1.2.1.1 Criteria
Assume an ignition source is present and that the lines contain an ignitable hydrazine vapor mixture after
the propulsion system has been operated.
A.1.2.2 Effects
In the unlikely event of a deflagration, the system can contain the adiabatic pressure of a neat hydrazine
deflagration (approximately 1.05 MPa or 152 psia).
AIAA SP-084-1999
122
Table A.1 Specifications for hazards assessment example
Component Parameter(s) Specification
Propellant Tank Operational temperature 315 K (108 F)
Maximum operating pressure 2.1 MPa (300 psia)
Max. allowable working pressure 3.4 MPa (500 psia)
Volume of hydrazine 0.028 m
3
(1 ft
3
)
Tank material 304L stainless steel
Diaphragm material EPR-AF-E-332 rubber
Burst pressure 13.8 MPa (2000 psia)
Tubing Diameter 0.64 cm (0.25 in.)
Wall thickness 0.051 cm (0.020 in.)
Material 304L stainless steel
Burst pressure 52 MPa (7500 psia)
Fill Line Disconnect Body material 304L stainless steel
Softgood material Kynar
Filter Size 25 m absolute
Material 304L stainless steel
Line Heater Coil Normal operating temperature 333 K (140 F)
Maximum fail open temperature 589 K (600 F)
Isolation Valve Type Solenoid
Body material 304L stainless steel
Softgood material AFE-4-11 rubber
Maximum operating temperature 366 K (200 F)
Maximum operating pressure 20.7 MPa (3000 psia)
Maximum fail open temperature 477 K (400 F)
Thruster Valve Type Solenoid
Body material 304L stainless steel
Softgood material AFE-4-11 rubber
Maximum operating temperature 366 K (200 F)
Thruster Valve Maximum operating pressure 20.7 MPa (3000 psia)
Maximum fail open temperature 477 K (400 F)
Test Cell Material 304L stainless steel plate wall
liner with structural concrete
support, reinforced
Environment ambient temperature and
pressure
Exhaust venting ports 61 cm (2 ft) diameter
Chamber volume 28.3 m
3
(1000 ft
3
)
Overpressure rating 1.4 MPa (200 psia)
Static rating 345 kPa (50 psia)
AIAA SP-084-1999
123
Figure A.1 Typical propulsion system and test cell
A.1.2.2 Detonation
A.1.2.2.1 Criteria
Assume that the lines are filled with hydrazine vapor at 101.3 kPa (14.7 psia); thus, the corresponding cell
size is 2 mm (0.08 in.). From Equation 15, the tube diameter in the propulsion system is sufficient to
support a vapor detonation. In this case, assume that the required E
m
(minimum energy, Section 3.2.1) of
approximately 21 J (0.02 Btu) is not present; thus, detonation can only be initiated by a deflagration-to-
detonation transition (DDT). Because liquid hydrazine has not detonated in 10.2-cm (4-in.) lines with
large energies, a liquid hydrazine detonation is very unlikely.
A.1.2.2.2 Effects
The potential for a detonation is low. If a detonation did occur, the system could withstand the C-J
pressure of neat hydrazine, approximately 2.80 MPa (406 psia).
AIAA SP-084-1999
124
A.1.2.3Thermal-chemical process
A.1.2.3.1 Criteria
To evaluate an explosion hazard from a thermal-chemical process, assume three different situations: one
situation under normal conditions and two situations under failure conditions. The first situation considers
the time required to fail the tank given the pressure generation rate produced under normal conditions,
with materials exposed to liquid hydrazine at 315 K (108 F).
In the second situation, the time required to fail the tank is assessed, given the pressure generation rate
produced under a failure condition, such as a continually powered coil in the isolation valve. Assume that
the coil heats the valve and 0.3 m (1 ft) of line to an average temperature of 477 K (400 F).
The third situation considers the time required to fail the tank given the pressure generation rate produced
under another failure condition. In this situation, assume that one heater fails in the on position and
heats liquid hydrazine contained in 0.3 m (1 ft) of line upstream of the thruster valve to an initial
temperature of 589 K (600 F).
A.1.2.3.2 Effects
In the first situation (under normal conditions), approximately 900 moles of hydrazine, contained in 0.03
m
3
(1 ft
3
) of hydrazine, can decompose. If this hydrazine decomposes into 0.1 m
3
(3.5 ft
3
) at 315 K
(108 F), the final pressure will be approximately 162 MPa (23,500 psia). To determine the rate at which
the pressure is generated, the rate at which the hydrazine is decomposing must be determined from the
heat generation rate, the surface area of the material on which hydrazine is decomposing, and the heat of
reaction.
At 315 K (108 F), materials that affect the decomposition rate of hydrazine are AF-E-411 rubber and
Kynar. For the diaphragm material (EPR AF-E-332 rubber), an exothermic reaction is not detected until
430 K (315 F) is reached; therefore, it can be assumed that this material does not significantly accelerate
the hydrazine decomposition reaction. Similarly, the rate of hydrazine decomposition on 304L stainless
steel at 315 K (108 F) is negligible.
Assume that before the system is operated, approximately 6.5 cm
2
(1 in.
2
) surface area of both Kynar and
AF-E-411 rubber are exposed to hydrazine. The heat generation rates from Table 18 in the main text are
78 W/m
2
for Kynar and 55 W/m
2
for AF-E-411 rubber. From this information, it can be seen that
approximately 1.6 x 10
-6
moles/s of hydrazine are decomposing. Note that although the data in Table 18
in the main text for AF-E-411 rubber and Kynar are for higher temperatures, these data were used
because they are the only data available. Calculating a dP/dt of 2.9 x 10
-4
kPa/s (4.2 x 10
-5
psia/s), the
tank will not fail for approximately 1.5 years. After the propulsion system is operated, the AF-E-411
rubber present in the thruster valves increases the hydrazine decomposition rate slightly but not enough
to produce a hazard.
In the second situation, the final pressure generated is again 162 MPa (23,500 psia). The effect of the
increased temperature on the decomposition rate must be determined for 0.3 m (1 ft) of 304L stainless
steel line and AF-E-411 rubber exposed at 477 K (400 F). The effects of Kynar at 315 K (108 F) must
also be considered. From Table 18, the heat generation rate is 1 W/m
2
(9.48 x 10
-4
Btu/(sft
2
) for 304L
stainless steel and 55 W/m
2
(5.21 x 10
-2
Btu/(sft
2
) for AF-E-411 rubber. The total decomposition rate for
the hydrazine exposed to Kynar, AF-E-411 rubber and 304L steel is approximately 1.7 x 10
-6
moles/s.
This shows that the decomposition reaction is not significantly accelerated by the continual supply of
power to the valve coil. The time required to fail the tank is again about 1.5 years.
AIAA SP-084-1999
125
Assume that in the third situation, the hydrazine temperature initially starts at 589 K (600 F), but the
temperature rises because there are insufficient heat sinks. To determine the pressure generation rate, a
series of nonlinear differential equations must be solved; therefore, the time to tank failure is difficult to
determine. To prevent tank failure, the heater must be prevented from failing in the on position or a
cooling system must be provided.
If the line is maintained at 589 K (600 F), calculations similar to those performed for the first two
situations can be performed. Using these calculations, and assuming that liquid hydrazine is in contact
with 304L stainless steel and AF-E-411 rubber at 589 K (600 F) and in contact with 304L stainless steel,
Kynar, and AF-E-411 rubber (isolation and remaining thruster valve) at 315 K (108 F), the time to tank
failure is approximately 1 month. The shorter time determined for this situation is caused mainly by the
heat generation rate of 304L stainless steel (265 W/m
2
) at 589 K (600 F).
A.1.2.4 Rapid Compression
A.1.2.4.1 Criteria
Because a dynamic system is created when the isolation valve is opened and the lines are filled with a
noncondensable ullage, there is potential for an explosion produced by a rapid compression process. An
evaluation of the hydrodynamic surge pressure can determine whether sufficient froth is formed or
sufficient heat is released during compression to produce explosive decomposition of hydrazine.
The determination of the hydrazine decomposition hazard to the operation of the system requires
accurate information about surge pressures. If accurate estimates of system hydrodynamic surge
pressures are not available, it is recommended that actual hydrodynamic measurements with water be
performed on a simulated system.
A.1.2.4.2 Effects
Assume a maximum velocity of 40 m/s (131 ft/s). Using Equation 19, the estimated surge pressure is
calculated to be over 19 MPa (2756 psia). A surge pressure of 19 MPa (2756 psia) is dangerously close
to the operating limits of the isolation and thruster valves. The surge pressure calculation assumes the
contained fluid behaves like water and does not account for rapid hydrazine decomposition effects which
appear for pressures greater than 17 MPa (2466 psia). The additional pressure generated from rapid
hydrazine decomposition could result in severe damage to the propulsion system and the release of liquid
hydrazine into the test cell. The hazards resulting from this type of spill are assessed in Section A.2.4.
A.1.3 Compatibility hazard
A.1.3.1 Effect of hydrazine
A.1.3.1.1 Criteria
Assume that the propulsion system is exposed to hydrazine at a maximum temperature of 315 K (108 F)
for approximately one week before the test is conducted. From Table 4.1 in the main text, it can be seen
that 304L stainless steel has a corrosion rate of less than 1 mil/year (1 mil=0.001 in.). From Table 4.2 in
the main text, it can be seen that Kynar is not recommended for use in hydrazine systems, there is only a
slight change in AF-E-411 rubber at 450 K (351 F), and AF-E-332 rubber shows no changes at 344 K
(160 F).
AIAA SP-084-1999
126
A.1.3.1.2 Effect
The Kynar will degrade, possibly leaking hydrazine into the test cell. Hazards to the test cell caused by
this failure are assessed in Section A.2. The remaining materials will not cause a hazard during the week
before testing.
A.1.3.2 Effect of material
This assessment is discussed in Section A.1.2.3.
A.1.4 Exposure hazard
Neither toxicity hazard nor environmental hazard exists inside the propulsion system.
A.2 Assessing test cell hazards
A hazardous condition in the test cell may be created when a failure occurs in the propulsion system.
The hazard in this case is the possibility of the explosive event damaging the equipment within the cell. If
the test cell cannot contain the effects of the hazardous event, the integrity of the test cell will also be lost.
A.2.1 Fire hazard
A.2.1.1 Criteria
Assume that the temperature of the test cell is 315 K (108 F). The vapor pressure of hydrazine at 315 K
(108 F) is 4.9 kPa (0.71 psia). This corresponds to a hydrazine concentration of 4.84 percent by volume,
which is above the conservative lower flammability limit of 2.9 percent. Given that ignition sources are
present, a fire hazard from a propagating fire exists for this scenario. Assume that the autoignition
temperature of 438 to 673 K (328 to 752 F) is not reached. A fire hazard may exist from the liquid
hydrazine, because the temperature of the test cell is 315 K (108 F) and the conservative flash point of
liquid hydrazine is 311 K (100 F). Note that a flammable mixture can form inside the test cell if as little
as 0.5 percent of the available liquid hydrazine spills into the test cell.
A.2.1.2 Effects
The burning velocity for hydrazine-air mixtures is not known, but the burning velocity of hydrazine-oxygen
mixtures (Figure 13) is approximately 2.0 m/s (6.6 ft/s). Hydrazine-air velocities are lower than hydrazine-
oxygen velocities because of the diluent effect of nitrogen. The heat of reaction from hydrazine-oxygen
mixtures (577 kJ/mole or 548 Btu/mole) can also be used for a worst-case calculation, keeping in mind
that nitrogen will act as a heat sink. From the vapor pressure data, approximately 1.7 kg (3.75 lb) of
hydrazine is contained in a volume of 28 m
3
(1000 ft
3
).
It can be calculated from these data that approximately 30,600 kJ (29,070 Btu) of heat can be released
into the test cell in 1.5 s if the vapor ignites and a fire propagates from one end of the test cell to the
other. Because the liquid hydrazine is at the flash point temperature, additional energy could be released
if the vapor above the liquid burns. If all the liquid hydrazine contained in the propulsion system spills into
the test cell, approximately 27 kg (59 lb) of hydrazine will remain in the liquid state.
Approximately 4.81 x 10
5
kJ (4.6 x 10
5
Btu) would be released from burning this quantity of hydrazine.
The burning rate of a 15.2 cm (6-in.) diameter pool is 0.22 kg/(m
2
s) (2.6 lb/(ft
2
min)) (as stated in
Section 2.2.2.2 in the main text). Using this burning rate, 4.81 x 10
5
kJ (4.6 x 10
5
Btu) is released in
AIAA SP-084-1999
127
14 s; however, the pool diameter is considerably larger than 15 cm (6 in.) and is expected to burn at a
much faster rate.
A.2.2 Explosion hazard
A.2.2.1 Deflagration
A.2.2.1.1 Criteria
Assume as before that an ignitable mixture and ignition energies are present; therefore, an explosion
hazard from a deflagration also exists.
A.2.2.1.2 Effects
Table B.2 in Annex B gives the adiabatic explosion pressure for a 5 percent hydrazine-air mixture as
0.39 MPa (24.2 psia). Because the test cell is rated to withstand a static pressure of 0.34 MPa (50 psia),
the integrity of the cell and the equipment in the cell may be compromised.
A.2.2.2 Detonation
A.2.2.2.1 Criteria
Cell size data are not available for hydrazine-air mixtures, but cell size data for neat hydrazine can be
used. Neat hydrazine data approximate a worst-case situation because the cell size for hydrazine-air
mixtures is larger than the cell size for neat hydrazine. At the vapor pressure in the test cell, 5 kPa
(0.711 psia), the cell size for neat hydrazine is 20 mm (0.78 in.); thus a 130-mm (5.12-in.) diameter
unconfined sphere contains enough cells to support a detonation.
Because the test cell is many times larger than the sphere, there is potential for a detonation in the test
cell. The minimum energy required to directly initiate a detonation is 6625 J (6.28 Btu), which is on the
order of the energy released from 1.4 g (0.0031 lb) of TNT. If this amount of energy is not available in the
test cell, the detonation cannot be initiated directly. Instead, because a deflagration hazard exists, the
detonation may be initiated indirectly through a DDT. The potential of a hazard from a liquid detonation is
unlikely because liquid hydrazine has not been observed to detonate.
A.2.2.2.2 Effects
A fuel concentration of 5 percent (Table B-1 in Annex B) results in a detonation pressure of approximately
0.92 MPa (133 psia). The test cell was designed to withstand 1.4 MPa (200 psia) overpressure and 0.34
MPa (50 psia) static pressure. The test cell will not be damaged if a detonation occurs in the cell;
however, the propulsion system will probably be severely damaged.
A.2.2.3 Thermal-chemical processes
A.2.2.3.1 Criteria
Assume a thermal-chemical process causes the confining vessel to fail. Further assume that materials
contained within the test cell are not significantly catalytic at 315 K (108 F) and that large heat sinks are
present, the rate of hydrazine decomposition will be very slow. Furthermore, in this scenario a vent is
used to maintain ambient pressure in the test cell.
AIAA SP-084-1999
128
A.2.2.3.2 Effects
No hazard exists from a thermal-chemical process.
A.2.2.4 Rapid compression
A rapid compression hazard is not present in the test cell.
A.2.3 Compatibility hazard
A.2.3.1 Effect of hydrazine
Assume that materials contained in the test cell are compatible with hydrazine; thus, a compatibility
hazard does not exist.
A.2.3.2 Effect of material
This assessment is discussed in Section A.1.2.3.
A.2.4 Exposure hazard
A.2.4.1 Criteria
Assume that approximately 1.7 kg (3.75 lb) of hydrazine evaporate into the air. This corresponds to a
concentration of 51,000 ppm if the vapor pressure of hydrazine is reached. A toxicity hazard exists in the
test cell, and may extend beyond the confines of the cell.
Assume that no environmental hazard is present.
A.2.4.2 Effects
The effects of toxicity hazards are addressed in Sections 5.1.1.1 and 5.1.2.1.
A.3 Summary of hazard assessment
A.3.1 System hazard assessment
Under the conditions assumed, the potential for a fire, deflagration, or detonation occurring in the
propulsion system is minimal. Because the propulsion system is stored in the test cell for such a short
time, a thermal-chemical hazard will only exist if the heater fails in the on position and a thermal
runaway occurs.
On the other hand, the potential for an explosion hazard caused by rapid compression is great, and a
single system failure can leak hydrazine into the test cell. A material compatibility hazard exists because
Kynar is used in the propulsion system. A toxicity hazard does not exist in the propulsion system.
A.3.2 Test cell hazard assessment
A fire, deflagration, or detonation hazard may exist in the test cell if liquid or vapor hydrazine is present. If
any of these events occur, equipment contained in the cell will be damaged, but the integrity of the test
cell will not be lost. There is no explosion hazard expected from a rapid compression event and the
AIAA SP-084-1999
129
potential for a thermal-chemical hazard or material compatibility hazard is minimal. A toxicity hazard
exists when hydrazine is in the test cell, and this hazard may extend beyond the boundaries of the cell.
AIAA SP-084-1999
130
Annex B: Chemical, physical, and thermodynamic properties of
hydrazine
B.1 General Information
Common Name: Hydrazine, Anhydrous hydrazine
Chemical Name: Hydrazine
Chemical Formula: N
2
H
4
(H
2
NNH
2
)
Molecular Weight: 32.04516 (Chase 1986)
Appearance: Colorless, water-like liquid; fuming in air
B.2 Chemical properties
The following properties are at ambient temperature and pressure unless otherwise specified.
B.2.1 Physical constants
Melting Point 274.68 K (34.75 F) (Marsh 1970)
Boiling Point 387.4 K (237.6 F) (Marsh 1970)
Liquid Density 1.004 Mg/m
3
(62.68 lb/ft
3
) (Marsh 1970), at 298.15 K (77 F)
Critical Temperature 653 K (716 F) (Marsh 1970)
Critical Pressure 14700 kPa (2131 psia) (Yaws 1977)
Critical Density 0.231 Mg/m
3
(14.4 lb/ft
3
) (Marsh 1970)
Triple Point Temperature 274 t 0.5 K (33.5 t 0.2 F) (Benz, Bishop, and Pedley 1988)
Triple Point Pressure 0.319 t 0.004 kPa (0.0463 t 0.0006 psia) (Benz, Bishop, and Pedley
1988)
Thermal conductivity
Gas* 1.377 x 10
-2
W/(m

K) (9.56 x 10
-2
Btuin/(ft
2
hF))
Liquid 0.89 W/(m

K) (6.18 x 10
-1
Btuin/(ft
2
hF)) (Marsh 1970)
Viscosity
Gas* 8.013 x 10
-6
Ns/m
2
(8.013 x 10
-3
cP)
Liquid 9.13 x 10
-4
Ns/m
2
(9.13 x 10
-1
cP) (Marsh 1970)
Vapor Pressure 1.89 kPa (0.274 psia) (Marsh 1970)
Liquid compressibility
Adiabatic 0.232 GPa
-1
(1.59 x 10
-6
psi
-1
) (Marsh 1970)
Isothermal 0.251 GPa
-1
(1.72 x 10
-6
psi
-1
) (Marsh 1970)
Surface tension 6.645 x 10
-2
N/m (4.554 x 10
-5
lb
f
/ft) (Marsh 1970)
Speed of sound
Vapor** 346.4 m/s (1120 ft/s) (at 387.3 K (237.5 F))
Liquid 2074 m/s (6803.7 ft/s) (Schmidt 1980)

* Value calculated from temperature dependency constants (Yaws 1977).
** Ideal gas law calculation based on JANAF data (Chase 1986).
AIAA SP-084-1999
131
B.2.2 Thermodynamic data
Heat capacity
Gas C
p
50.813 J/(molK) (0.0267 Btu/(molF)) (Chase 1986)
C
v
*
42.499 J/(molK) (0.0224 Btu/(molF))
Liquid 98.840 J/(molK) (0.0519 Btu/(molF)) (Chase 1986)
Heat of fusion 12.657 kJ/mol (11.996 Btu/mol) (Chase 1986)
Heat of vaporization 44.73 kJ/mol (42.396 Btu/mol) (Chase 1986)
Heat of formation
Vapor 95.353 kJ/mol (90.377 Btu/mol) (Chase 1986)
Liquid 50.626 kJ/mol (47.984 Btu/mol) (Chase 1986)
Heat of combustion (Chase 1986)
Vapor Higher Heating Value:
**
N
2
H
4
(v) + O
2
(g) N
2
(g) + 2H
2
O( )
-667.01 kJ/mol (-632.2 Btu/mol)
Lower Heating Value:
**
N
2
H
4
(v) + O
2
(g) N
2(
(g)+ 2H
2
O(g)
-579.01 kJ/mol (-548.8 Btu/mole)
Liquid Higher Heating Value:
**
N
2
H
4
( ) + O
2(
(g) N
2
(g) + 2H
2
O

( )
-622.29 kJ/mol (-589.8 Btu/mol)
Lower Heating Value:
**
N
2
H
4
( ) + O
2
(g) N
2
(g) + 2H
2
O(g)
-534.28 kJ/mol (-506.4 Btu/mol)
Heat of decomposition
N
2
H
4
N
2
+ 2H
2
(Benz, Bishop, and Pedley 1988)
Vapor
***
-95.353 kJ/mol (-90.377 Btu/mol)
Liquid
*
-50.626 kJ/mol (-47.984 Btu/mol)
N
2
H
4
NH
3
(g) + 1/2 N
2
+ 1/2 H
2
(Benz, Bishop, and Pedley 1988)
Vapor -140 kJ/mol (-132.69 Btu/mol)
Liquid -97 kJ/mol (-91.94 Btu/mol)
N
2
H
4
4/3 NH
3
(g) + 1/3 N
2
(Benz, Bishop, and Pedley 1988)
Vapor -157 kJ/mol (-148.81 Btu/mol) at 298 K (77 F)
Liquid -112 kJ/mol (-106.16 Btu/mol) at 298 K (77 F)
Liquid -123 kJ/mol (-116.58 Btu/mol) at 515 K (468 F)
B.2.3 Thermodynamic properties of hydrazine (Barragan, Woods, and
Wilson 1998)

*
Calculated from JANAF value(s) (Chase 1986).
**
Ideal gas low calculation based on JANAF data (Chase 1986).
***
Value calculated from temperature dependent constants (Yaws 1977).
AIAA SP-084-1999
132
A partial Mollier diagram for hydrazine is given in Figure B.1. Some thermodynamic properties of
hydrazine are given in Table B.1. Some calculated Chapman-Jouguet (C-J) parameters for hydrazine-air
detonations are given in Table B.2
B.2.4 Temperature dependency constants
Liquid density Range: 275 K (35 F) to 653 K (716 F)

L
(Mg/m
3
) = 0.3171 x 0.2538
-(1 - T/653.2)
(Yaws 1977)
Thermal conductivity
Gas Range: 273 K (32 F) to 1673 K (2552 F)

g
(W/(mK) = -2.255 x 10
-2
+1.192 x 10
-4
T + 8.4 x 10
-9
T
2
+ 7.9

x 10
-13
T
3
(Yaws 1977)
Liquid Range: 275 K (35 F) to 591 K (604 F)

L
(W/(mK))= 1.19662-7.3316 x 10
-4
T-1.011 x 10
-6
T
2
(Yaws 1977)
Viscosity Range: 273 K (32 F) to 1673 K (2552 F)
Gas
g
(Ns/m) = -1.705 x 10
-6
+ 3.401 x 10
-8
T - 4.751 x 10
-12
T
2
(Yaws 1977)
Viscosity Range: 273 K (32 F) to 450 K (350 F)
Liquid
L
(Ns/m
2
) = -0.8126 - 0.015384T + 1.5395 x 10
-5
T
2
(Marsh 1970)
Vapor Pressure Range: 288 K (59 F) to 653 K (716 F)
log P (kPa) = 60.003-(3880.3/T)-20.575 log T+1.5585 x 10
-2
T-5.0525 x
10
-6
T
2
(Yaws 1977)
Heat of vaporization Range: 275 K (35 F ) to 653 K (716 F)
H
v
(kJ/mol) = 40.4 ((653-T)/266.5)
0.38
(Yaws 1977)
Heat capacity
Gas Range: 298 K (77 F) to 1500 K (2240 F)
C
p
(J/(molK)) = 16.19 + 0.1494T - 9.62 x 10
-4
T
2
- 2.5 x 10
-8
T
3
(Yaws 1977)
Liquid 275 K (35 F) to 473 K (392 F)
C
p
(J/(molK)) = -620.4 + 6.561T - 2.010 x 10
-2
T
2
+ 2.101 x 10
-5
T
3
(Yaws 1977)
B.2.5 Explosion pressures
The data in Tables B-3 and B-4 were calculated using the Gordon-McBride computer program (Gordon
and McBride 1976).
B.3 Military specifications
Military specification MIL-P-26536D (Military Specification 1986) covers the requirements for standard,
monopropellant, and high-purity grades of hydrazine. Purity requirements and maximum allowable
amounts of impurities are summarized in Table B-5. MIL-P-26536D also provides information on test
methods to be used to determine the content of hydrazine.
AIAA SP-084-1999
133
0
1
10
100
1,000
10,000
100,000
-300 200 700 1,200 1,700 2,200
Enthalpy (J/g)
P
r
e
s
s
u
r
e

(
k
P
a
)
H (Vap)
H (Liq)
Figure B.1 Partial Mollier diagram of hydrazine
AIAA SP-084-1999
134
Table B.1 Thermodynamic properties of hydrazine
T
K
P
kPa
VL
m
3
/g
Vv
m
3
/g
HL(T,P)
Sat. liquid
J/g
H
v
(T,P)
Sat. vapor
J/g
SL(T,P)
Sat. liquid
J/g K
S
V
(T,P) Sat.
vapor
J/gK
273 1.50 1.26E-06 3.29E-02 -78 844 -0.0075 3.11
283 2.88 1.27E-06 1.77E-02 -46 859 -0.0048 3.05
293 5.26 1.28E-06 1.00E-02 -15 874 0.0045 2.99
298 7.19 1.29E-06 7.46E-03 0 882 0.0000 2.96
303 9.13 1.29E-06 5.98E-03 15 890 0.0194 2.94
313 15.17 1.31E-06 3.71E-03 46 905 0.0391 2.90
323 24.26 1.32E-06 2.39E-03 76 921 0.0631 2.87
333 37.50 1.33E-06 1.59E-03 106 938 0.0908 2.84
343 56.18 1.35E-06 1.09E-03 137 954 0.1217 2.82
353 81.86 1.36E-06 7.66E-04 168 971 0.1555 2.80
363 108.27 1.38E-06 5.93E-04 198 988 0.2048 2.79
373 149.28 1.40E-06 4.40E-04 230 1,005 0.2447 2.78
383 201.96 1.42E-06 3.31E-04 262 1,023 0.2863 2.77
393 268.49 1.44E-06 2.54E-04 294 1,040 0.3295 2.77
403 351.29 1.46E-06 1.97E-04 328 1,058 0.3742 2.76
413 452.92 1.49E-06 1.55E-04 362 1,075 0.4201 2.76
423 576.09 1.51E-06 1.23E-04 397 1,093 0.4673 2.76
433 723.67 1.54E-06 9.90E-05 432 1,110 0.5157 2.76
443 898.62 1.57E-06 8.03E-05 469 1,128 0.5651 2.76
453 1103.96 1.60E-06 6.56E-05 507 1,145 0.6157 2.76
463 1342.82 1.64E-06 5.40E-05 546 1,162 0.6673 2.77
473 1618.32 1.68E-06 4.47E-05 585 1,179 0.7199 2.77
483 1933.63 1.73E-06 3.73E-05 626 1,195 0.7736 2.77
493 2291.87 1.78E-06 3.12E-05 669 1,211 0.8285 2.77
503 2696.16 1.84E-06 2.62E-05 712 1,226 0.8846 2.78
513 3149.56 1.91E-06 2.20E-05 757 1,240 0.9420 2.78
523 3655.06 1.99E-06 1.86E-05 803 1,254 1.0009 2.78
533 4215.56 2.09E-06 1.57E-05 852 1,266 1.0617 2.78
543 4833.87 2.21E-06 1.32E-05 902 1,277 1.1249 2.77
553 5512.68 2.36E-06 1.10E-05 955 1,285 1.1912 2.76
563 6254.57 2.57E-06 9.15E-06 1,013 1,290 1.2625 2.75
573 7061.96 2.91E-06 7.46E-06 1,079 1,290 1.3439 2.73
580 7667.36 3.51E-06 6.36E-06 1,144 1,285 1.4274 2.70
T = temperature
P = pressure
VL = specific volume, liquid
Vv = specific volume, vapor
HL = enthalpy, saturated liquid
Hv = enthalpy, saturated vapor
SL = entropy, saturated liquid
Sv = entropy, saturated vapor
AIAA SP-084-1999
135
Table B.2 Calculated Chapman-Jouguet (C-J) parameters for hydrazine-air detonations
Adiabatic
Concentration Initial Equivalence flame C-J
by volume
a
temperature ratio temperature velocity pressure
% K
b
K
b
m/s
c
MPa
d
2.5 303.3 0.122 1056.3 1043.2 0.588
5.0 316.0 0.250 1619.6 1339.6 0.923
10.0 329.5 0.529 2449.1 1708.8 1.462
15.0 338.1 0.840 2945.8 1933.1 1.861
17.36
e
341.3 1.000 3060.8 2002.9 1.963
20.0 344.5 1.190 3104.8 2058.7 2.085
25.0 349.8 1.587 3070.8 2127.7 2.184
30.0 354.3 2.040 2991.2 2177.6 2.252
35.0 358.0 2.564 2903.3 2219.3 2.313
40.0 361.4 3.174 2816.5 2255.9 2.366
45.0 364.5 3.896 2732.8 2288.7 2.416
50.0 367.3 4.761 2653.0 2318.3 2.461
55.0 369.8 5.820 2577.0 2345.2 2.505
60.0 372.3 7.142 2505.1 2369.8 2.544
65.0 374.4 8.843 2436.9 2392.4 2.582
70.0 376.5 11.111 2372.5 2413.3 2.618
75.0 378.4 14.285 2311.5 2432.5 2.652
80.0 380.3 19.047 2253.9 2450.4 2.684
85.0 382.1 26.984 2199.5 2467.1 2.715
90.0 383.7 42.857 2147.9 2482.7 2.745
95.0 385.3 90.476 2099.2 2497.3 2.774
100.0 387.0 Decomposition 2053.2 2511.2 2.801
a
101.325 kPa (14.696 psia) initial pressure
b
To convert to F, T (F) = 1.8 (T(K))-459.67
c
To convert to ft/s, V (ft/s) = 3.281 V (m/s)
d
To convert to lb/in.
2
, P (psi) = 145.038 P (MPa)
e
Stoichiometric hydrazine concentration
AIAA SP-084-1999
136
Table B.3 Isolated system pressures based on thermodynamic equilibriu of hydrazine and air
Adiabatic
Concentration Initial Equivalence flame C-J
by volume
a
temperature ratio temperature velocity pressure
% K
b
K
b
m/s
c
MPa
d
2.5 303.3 0.122 762.1 551.0 0.261
5.0 316.0 0.250 1158.5 676.8 0.390
10.0 329.5 0.529 1814.5 847.8 0.614
15.0 338.1 0.840 2299.9 952.7 0.796
17.36
e
341.3 1.000 2433.5 980.7 0.858
20.0 344.5 1.190 2441.7 1009.5 0.889
25.0 349.8 1.587 2338.4 1040.7 0.912
30.0 354.3 2.040 2229.6 1061.3 0.928
35.0 358.0 2.564 2127.9 1078.3 0.943
40.0 361.4 3.174 2034.4 1093.0 0.956
45.0 364.5 3.896 1948.2 1105.7 0.967
50.0 367.3 4.761 1869.2 1116.9 0.978
55.0 369.8 5.820 1796.3 1126.8 0.988
60.0 372.3 7.142 1729.3 1135.6 0.997
65.0 374.4 8.843 1667.4 1143.6 1.006
70.0 376.5 11.111 1610.0 1150.8 1.013
75.0 378.4 14.285 1556.9 1157.4 1.021
80.0 380.3 19.047 1507.4 1163.5 1.028
85.0 382.1 26.984 1461.4 1169.1 1.035
90.0 383.7 42.857 1418.3 1174.3 1.042
95.0 385.3 90.476 1378.0 1179.2 1.048
100.0 387.0 Decomposition 1340.2 1183.7 1.054
a
101.325 kPa (14.696 psia) initial pressure
b
To convert to F, T (F) = 1.8 (T(K))-459.67
c
To convert to ft/s, V (ft/s) = 3.281 V (m/s)
d
To convert to lb/in.
2
, P (psi) = 145.038 P (MPa)
e
Stoichiometric hydrazine concentration
AIAA SP-084-1999
137
Table B.4 Purity requirements for hydrazine
a
Constituent Grade
(maximum allowable percent by
weight)
Standard Monopropellant High Purity
Hydrazine
b
98.0 98.5 99.0
Ammonia 0.4
Aniline 0.50 0.005
Carbon Dioxide 0.003 0.003
Chloride 0.0005 0.0005
Iron 0.002 0.004
Nonvolatile Residue 0.005 0.001
Other Volatile Carbonaceous Material
c
0.02 0.005
Particulate (mg/L) 10.0 1.00 1.00
Water 1.5 1.00 1.00
a
Military Specification, 1986
b
Minimum required percent by weight
c
Total of Monomethylhydrazine (MMH), Unsymmetrical dimethylhydrazine (UDMH), and alcohol
AIAA SP-084-1999
138
Annex C: References
ACGIH. Threshold Limit Values and Biological Exposure Indices for 1996. Cincinnati, Ohio: American
Conference of Governmental Industrial Hygienists, 1996, p. 10.
Adams, G. K. and G. W. Stocks. "The Combustion of Hydrazine." In Fourth Symposium (International)
Combustion. Baltimore, MD, Williams and Wilkins (1953): pp. 239-248.
Allison, C. B. and G. M. Faeth. "Decomposition and Hybrid Combustion of Hydrazine, MMH, and UDMH
as Droplets in a Combustion Gas Environment. Combust. Flame, 19 (October 1972): pp. 213-
226.
Amenta, J. S., and A. M. Dominguez. The Effect of Hydrazine and Congeners on 14 CO
2
Respiratory
Pattern of Various Metabolic Substrates. Applied Pharmacology, Vol. 4 (1965), pp. 236-246.
Arah, C. O., H.M. Hand, D. K. McNamara, J. A. S. Green, and W. J. Arbegast Jr. The Effect of Rocket
Propellants on High-Performance Engineering Thermoplastics. 36
th
Annual International
SAMPE Symposium, pp. 1545-1560. Denver, CO: Martin Marietta Astronautics Group, 1991.
ASME. Pressure Vessels ASME Boiler and Pressure Vessel Code, Sect. VIII, Div. 1, American Society
for Mechanical Engineering, New York, NY (1995, or latest revision).
ASTM. Test Method for Autoignition Temperature of Liquid Chemicals. In Annual Book of ASTM
Standards, Vol. 05.03 and 14.02, Method E 659-78 (1984), American Society for Testing
Materials, Philadelphia, PA, 1986a.
ASTM. Test Method for Flash and Fire Points by Cleveland Open Cup. In Annual Book of ASTM
Standards, Vol. 04.04, 05.01, and 10.03, Method D 92-85, American Society for Testing
Materials, Philadelphia, PA, 1986b.
ASTM. Test Method for Flash Point by Tag Closed Tester. In Annual Book of ASTM Standards, Vol.
05.01 and 06.03, Method D 56-82, American Society for Testing Materials, Philadelphia, PA,
1986c.
ASTM. Test Method for Minimum Ignition Energy and Quenching Distance in Gaseous Mixtures. In
Annual Book of ASTM Standards, Vol. 14.02, Method E 582-88, American Society for Testing
Materials, Philadelphia, PA, 1988.
ASTM. Standard Terminology of Fire Standards. In Annual Book of ASTM Standards, Vol. 4.07,
American Society for Testing Materials, West Conshohocken, PA, 1997.
Atkinson, A. J. and T. I. Eklund. Crash Fire Hazard Evaluation of Jet Fuels. FAA-RD-70-72, Naval Air
Propulsion Test Center, Philadelphia, PA, 1971.
Axworthy, A. E., J. M. Sullivan, S. Cohz, and E. Welz. Research on Hydrazine Decomposition. AFRPL-
TR-69-146, Edwards AFB, CA, July 1969.
Ayeni, R. O. Thermal Runaway. Ph.D. Dissertation, Cornell University, 1978.
Back, K. C., V. L. Carter, and A. A. Thomas. Occupational Hazards of Missile Operations with Special
Regard to the Hydrazine Propellants. Aviation, Space, and Environmental Medicine, April 1978,
pp. 591-598.
Baker, D., H. Beeson, D. Fernandez, and M. Plaster. Explosive Decomposition of Hydrazine by Rapid
Compression of Gas Ullages. In 1988 JANNAF Safety and Environmental Protection
Subcommittee Meeting. pp. 15 - 26. Johns Hopkins University, Laurel, MD: Chemical
Propulsion Information Agency, May 1988.
Barragan, M., S. Woods, and D. B. Wilson. An Equation of State for Hydrazine and
Monomethylhydrazine, Its Validation and Use for Calculation of Thermodynamic Properties.
Paper presented at the 1998 JANNAF Propellant Development & Characterization Subcommittee
and Safety and Environmental Protection Subcommittee Joint Meeting, Houston, TX, April 20-24,
1998.
Barrow, G. M. Physical Chemistry. 3rd ed. New York: McGraw-Hill, 1973. p. 9
Beeson, H. D., R. D. McClenagan, C. V. Bishop, F. J. Benz, W. J. Pitz, C. K. Westbrook, and J. H. S. Lee.
Detonability of Hydrocarbon Fuels in Air. Progress in Astronautics and Aeronautics, Dynamics
of Detonations and Explosions: Detonations, Vol. 133, pp. 19-36, July 1989.
AIAA SP-084-1999
139
Beeson, H., M. Plaster, and R. Bunker. Investigation of Detonations in Liquid Hydrazine. In 1988
JANNAF Safety and Environmental Protection Subcommittee Meeting. pp. 49 - 56. Johns
Hopkins University, Laurel, MD: Chemical Propulsion Information Agency, May 1988.
Bennett, C. R., D. R. B. Saw, and D. Sutton. Laboratory Tests at Elevated Temperatures for the
Prediction of the Rates of Pressure Rise in Hydrazine Tanks at Normal Storage Temperatures.
Journal of Hazardous Materials, Vol. 4, (1980): pp. 23-44. Elsevier Scientific Publishing
Company, Amsterdam, The Netherlands.
Benz, F. J. Compatibility of AFE-411 with Hydrazine at Elevated Temperatures. TR-291-001, NASA
Johnson Space Center White Sands Test Facility, Las Cruces, NM, May 1981.
Benz, F. J. The Effect of Type 321 Stainless Steel Welded to Type 304 Stainless Steel on the
Exothermic Decomposition of Hydrazine at Elevated Temperature. TR-361-001, NASA White
Sands Test Facility, Las Cruces, NM, Oct. 1984.
Benz, F. J. The Exothermicity of Hydrazine and Kynar 460. NASA-TR-283-001, NASA Johnson Space
Center White Sands Test Facility, Las Cruces, NM, Nov. 1980.
Benz, F. J. and D. L. Pippen. "Autoignition, Flammability and Explosion Properties of Hydrazine and
Monomethylhydrazine." In JANNAF Safety and Environmental Protection Specialist Session. J.
A. E. Hannum, Ed., pp. 603-616. Laurel, MD: Chemical Propulsion Information Agency, 1980.
Benz, F. J., C. V. Bishop, and M. D. Pedley. Ignition and Thermal Hazards of Selected Aerospace Fluids:
Overview,
Data, and Procedures. RD-WSTF-0001, NASA Johnson Space Center White Sands Test
Facility, Las Cruces, NM, Oct. 1988.
Benz, F. J., D. Guerrasio, and F. Rollins. "New Observations in Flammability and Autoignition
Temperatures of Hydrazine, Methylhydrazine, and Aerozine-50. In JANNAF Safety and
Environmental Protection Subcommittee Meeting, L. D. Simmons, Ed., pp. 61-69. Laurel, MD:
Chemical Propulsion Information Agency, 1983.
Benz, F. J., T. L. Long, and D. P. Weary. Explosive Hydrazine Decomposition Due to Rapid Gas
Compression (Adiabatic Heating). In JANNAF Safety and Environmental Protection
Subcommittee Meeting, p. 21. Johns Hopkins University, Laurel, MD.: Chemical Propulsion
Information Agency, May 1984.
Bhardwaj, R. C., D. D. Davis, and D. L. Baker. An EIS Study of the Corrosive Behavior of Propellant
Hydrazine. Corrosion 96 (The NACE International Annual Conference and Corrosion Show),
Denver, CO, March 24-29, 1996.
Bird, R. B., W. E. Stewart, and E. N. Lightfoot. Transport Phenomena. Department of Chemical
Engineering, University of Wisconsin. Madison, WI: John Wiley & Sons, Inc., 1960.
Bitter, H. L., D. A. Clark, and W. W. Lackey. Metabolic Effects of Hydrazine. Aerospace Medicine, Vol.
38 (1967), pp. 1214-1219.
Blaies, D. M., N. A. Leavitt, and R. C. Young. Development of an Interim Active Vanillin Sampler Used to
Detect 10 Parts Per Billion of Hydrazine and Monomethylhydrazine in Air. In 1991 JANNAF
Safety and Environmental Protection Subcommittee Meeting. NASA Kennedy Space Center, FL:
Chemical Propulsion Information Agency, (1991): pp. 203-208.
Bodurtha, F. A. Industrial Explosion Prevention and Protection. New York, NY: McGraw-Hill, 1980.
Botteri, B. P., R. E. Cretcher, J. R. Fultz, and H. R. Lander. A Review and Analysis of the Safety of Jet
Fuel. AFAPL-TR-66-9, Wright-Patterson AFB, OH, 1966.
Boyd, W. K. and E. L. White. Compatibility of Rocket Propellants with Materials of Construction. DMIC
Memorandum 65, Battelle Memorial Institute, Columbus, OH, Sept. 1960.
Briles, O. M. and R. P. Hollenbaugh, Sr. Adiabatic Compression Testing of Hydrazine. Paper presented
at the Fourteenth Joint Propulsion Conference of the American Institute of Aeronautics and
Astronautics and the Society of Automotive Engineers, Las Vegas, NV, 1978.
Briles, O., D. Hagemann, F. J. Benz, and T. Farkas. Explosive Decomposition of Hydrazine Due to
Rapid Gas Compression. In 1985 JANNAF Propulsion Meeting. K. L. Strange, ed., pages 519 -
525. Johns Hopkins University, Laurel, MD: Chemical Propulsion Information Agency, April
1985.
AIAA SP-084-1999
140
Brinkley, S. R. and B. Lewis. On the Transition from Deflagration to Detonation. In Seventh Symp.
(Intl.) Combust. London: Butterworth, 1959.
Brown, S. W. and C. R. Nunez. Hydrazine Exposure Levels at the Kennedy Space Center. In 1991
JANNAF Safety and Environmental Protection Subcommittee Meeting, Publication 569. NASA
Kennedy Space Center, FL: Chemical Propulsion Information Agency, (July 1991) pp. 21-31.
Bull, D. C., J. E. Ellsworth, and P. J. Shiff. Detonation Cell Structures in Fuel/Air Mixtures. Combust.
Flame, 45 (1982): pp. 7 - 22.
Bunker, Radel L., D. Baker and J. H. Lee. Explosive Decomposition of Hydrazine by Rapid Compression
of a Gas Volume. In Dynamics of Explosives, American Institute of Aeronautics and
Astronautics, Inc 1990.
Burgess, D. S., G. H. Martindill, J. N. Murphy, F. J. Perzak, J. M. Singer, and R. W. Van Dolah. "Inerting
and Extinguishing of Aerozine-50 with Water and/or CF
3
Br." J. Spacecraft, 6 (1969): pages 1259-
1268.
Burgess, M. J. and R. V. Wheeler. "Lower Limit of Inflammation of Mixtures of Paraffin Hydrocarbons
with Air." J. Chem. Soc., 99 (1911): pp. 2013-2030.
Cadwallader, E. and L. B. Piper. Hydrazine Compatibility Survey. Silver Springs, MD: Chemical
Propulsion Information Agency, June 1973.
CFR Title 29. Labor. Code of Federal Regulations, Parts 1910.20, 1910.94, 1910.119, 1910.120
1910.132, 1910.133, 1910.134, 1910.141, 1910.151, 1910.1000, 1910.1200, and 1910.1450
(1991 or latest revision).
CFR Title 40. Protection of Environment. Code of Federal Regulations, Parts 116, 172.101, 261.33(e),
268, 355 Annex A, 370.2, and 717 (1991 or latest revision).
CFR Title 49. Transportation. Code of Federal Regulations, Parts 171.8, 172.102, 173.276 (1992 or
latest revision.
Chang, E. T. and N. A. Gokcen. Compatibility of Alloys with Hydrazine Containing Freon. TR 76-34,
Aerospace Corp., El Segundo, CA, Feb. 1976.
Chase, M. W. Jr. et. al. JANAF Thermochemical Tables Third Edition. 335 East 45th Street, New York,
New York 10017: American Chemical Society and the American Institute of Physics, 1986.
Chew, W. M., B. D. Allan, C. L. Greer, D. L. May and L. Whitman. Compatibility of Elastomeric Materials
with Selected Liquid Propellants. Critical Technology, Redstone Arsenal, AL, October 1990.
Chuan, R. L. and P. C. Wilber. "Ignition of Hypergolic Propellants in a Simulated Space Environment." J.
Spacecr. Rockets, 4 (1967): pp. 282-284.
Clark, C. C. Hydrazine. Baltimore, MD: Mathieson Chem. Corp., 1953.
Clark, D. A., J. D. Bairrington, H. L. Bitter, F. L. Coe, M. A. Medina, J. H. Merritt, and W. N. Scott.
Pharmacology and Toxicology of Propellant Hydrazines. USAF School of Aerospace Medicine,
Brooks AFB, TX, 1968.
Cloyd. D. R., and W. J. Murphy. Handling Hazardous Materials. Washington, D.C.: NASA, 1965.
Comer, R. H. Ignition and Combustion of Liquid Monopropellants at High Pressures. In Sixteenth
Symposium (International) on Combustion, pp. 1211 - 1219. Pittsburgh, PA: The Combustion
Institute, 1977.
Coulbert, C. D., E. F. Cuddihy, and R. F. Fedors. Long-Time Dynamic Compatibility of Elastomeric
Materials With Hydrazine. NASA TM 33-650, Jet Propulsion Laboratory, Pasadena, CA, Sept.
1973.
Coward, H. F. and G. W. Jones. Limits of Flammability of Gases and Vapors. Bulletin 503, Bureau of
Mines, Washington, D.C., 1952.
Coward, H. F. and G. W. Jones. Limits of Flammability of Gases and Vapors. Bulletin 503, Bureau of
Mines, Washington, DC, 1952.
CPIA. Air Force/NASA Workshop on Hypergolic Vapor Detectors. Chemical Propulsion Information
Agency. Laurel, MD: Johns Hopkins University Applied Physics Laboratory, June 1981.
CPIA. Hazards of Chemical Rockets and Propellant, Volume III, Liquid Propellants. Chemical Propulsion
Information Agency. Laurel, MD: Johns Hopkins University Applied Physics Laboratory, 1984.
AIAA SP-084-1999
141
Cross, J. H., S. W. Beck and T. F. Limero, J. T. James. Hydrazine Monitoring in Spacecraft. In Fifth
Annual Workshop on Space Operations Applications and Research (SOAR '91), NASA
Conference Publication 3127, Volume 2, Feb. 1992b, pp. 627-636.
Cross, J. H., T. F. Limero, S. W. Beck, J. T. James, N. B. Martin, H. T. Johnson, R. C. Young, and C. B.
Mattson. Monitoring Hydrazines in the Space Shuttle Airlock with a Proof-of-Concept Ion
Mobility Spectrometer. In 1992 JANNAF Safety and Environmental Protection Subcommittee
Meeting, Naval Postgraduate School, Monterey, CA: Chemical Propulsion Information Agency,
August 10-14, 1992a.
Davis, D. D. and R. C. Wedlich. "Kinetics and Mechanism of the Thermal Decomposition of Hydrazine,
Hydrazine Hydrate and Monomethylhydrazine Using Accelerating Rate Calorimetry". JANNAF
Propellant Characterization Subcommittee, Nov 1-3, 1988, Las Cruces, NM.
Davis, D. D. and Richard Wedlich. Kinetics and Mechanism of the Thermal Decomposition of
Monomethylhydrazine by Accelerating Rate Calorimetry. Thermochimica Acta, 175 (1991)
pp.189-198.
Davis, D. D. and S. D. Hornung. Microcalorimetry and Long-term Immersion Testing for the Evaluation of
Metal-Hydrazine Interactions. WSTF-IR-98-0103, NASA Johnson Space Center White Sands
Test Facility, Las Cruces, NM, publication in process.
Davis, D. D., S. D. Hornung, and R. L. Bunker. Evaluation of Microcalorimety for Determination of
Material Reactivity with Hydrazine. JANNAF Propellant Development and Characterization
Subcommittee, April 28-30, 1993, Livermore CA.
Davis, D. D., S. D. Hornung, and R. L. Bunker. Evaluation of Microcalorimetry for Determination of
Materials Reactivity with Hydrazine. 1993 JANNAF Propellant Development and
Characterization Subcommittee Meeting, pp. 23-27. Livermore, CA: CPIA Publication 597, April
1993.
Davis, D. D., S. Hornung, and H. T. Johnson. NHB 8060.1C Test 15 Reactivity of Materials in Aerospace
Fluids Methodology and Results. NAS9-95619 White Sands Test Facility, Las Cruces, NM, 1993.
Delgado, R. H. and D. D. Davis. Application Test of a Hydrazine Spill Control System: Copper II Oxide
Pillows. TR-587-002. Johnson Space Center White Sands Test Facility, Las Cruces, NM 88004.
June 12, 1995
Delgado, R. H., D. D. Davis, W. L. Ross, Sr., H. D. Beeson, and H. T. Johnson. Enhanced Technology for
Composite Overwrapped Pressure Vessels Subtask 3.6: Material Compatibility Testing -
Exposure/Burst Tests of Lincoln Composites Vessels Summary Report, TR-804-002, August 4,
1997a.
Delgado, R. H., D. D. Davis, W. L. Ross, Sr., H. D. Beeson, and H. T. Johnson. Enhanced Technology for
Composite Overwrapped Pressure Vessels Subtask 3.6: Material Compatibility Testing -
Exposure/Burst Tests of Structural Composites Vessels Summary Report, TR-804-003, August 4,
1997b.
Dinman, B. D. The Mode of Entry and Action of Toxic Materials. Patty's Industrial Hygiene and
Toxicology. New York: John Wiley and Sons, (1979).
DOT. Emergency Response Guidebook, United States Department of Transportation, Washington, DC
(1993, or latest revision).
Drell, I. L. and F. E. Belles. Survey of Hydrogen Combustion Properties. NACA Report 1383, Lewis
Flight Propulsion Lab, Cleveland, OH, 1958.
Dupre, G., R. Knystautas, and J. H. Lee. Near-Limit Propagation of Detonation in Tubes. In Dynamics
of Explosions, J. R. Bowen, J. C. Leyer, and R. I. Soloukhin, ed. New York: American Institute of
Aeronautics and Astronautics, Inc., 1986.
Eiceman, G. A., R. C. Young and J. C. Travis, and Z. Laney. "Ion Mobility Spectrometry in Monitoring
Hydrazines and Hazardous Organic Compounds In Air and Water." In 1990 JANNAF Safety and
Environmental Protection Subcommittee Meeting, Publication 543. Lawrence Livermore National
Laboratory, Livermore, CA: Chemical Propulsion Information Agency, (June 1990): pp. 115-120.
Energetics Science, Inc. A Study for Hypergolic Vapor Sensor Development. NASA-CR-158048,
Energetics Science, Inc., Elmsford, NY, April 1977.
AIAA SP-084-1999
142
Forshey, D. R., J. C. Cooper, and W. J. Doyak. Detonability of the Nitromethane-Hydrazine-Methanol
System. Explosivstoffe, 17 (1969): pp. 125 - 129.
Fritchman, T. T. and F. J. Benz. The Exothermicity of Liquid Hydrazine Exposed to Various Auxiliary
Power Unit Materials. TR-226-001, NASA Johnson Space Center White Sands Test Facility, Las
Cruces, NM, May 1980.
Fritchman, T. T. and F. J. Benz. The Exothermicity of Liquid Hydrazine Exposed to Various Auxiliary
Power Unit Materials. TR-226-001, NASA Johnson Space Center White Sands Test Facility, Las
Cruces, NM, May 1980.
Froment, G. F. and K. B. Bischoff. Chemical Reactor Analysis and Design. John Wiley and Sons, New
York 1979. pp. 90.
Furno, A. L., A. C. Imhof, and J. M. Kuchta. "Effect of Pressure and Oxidant Concentration on
Autoignition Temperatures of Selected Combustibles in Various Oxygen and Dinitrogen Tetroxide
Atmospheres." J. Chem. Eng. Data, 13 (1968): pp. 243-249.
Furno, A. L., G. H. Martindill, and M. G. Zabetakis. "Limits of Flammability of Hydrazine-Hydrocarbon
Vapor Mixtures." J. Chem. Eng. Data, 7 (July 1962): pp. 375-376.
Garcia, H. D., J. T. James, and T. F. Limero. Human Exposure Limits to Hypergolic Fuels. In Fifth
Annual Workshop on Space Operations Applications and Research (SOAR '91), NASA
Conference Publication 3127, Volume 2 (1992): pp. 620-626.
Geankoplis, C. J. Transport Processes and Unit Operations 2nd Ed. Prentice Hall, Englewood Cliffs, NJ,
1983
Gill, D. D. Storability and Expulsion Test Program. AD-B106 491 Edwards Air Force Base, CA,
September 1986.
Glassman, I. Combustion. 2nd ed. Orlando, FL: Academic Press, Inc., 1987.
Gordon, S. and B. J. McBride. Computer Program for Calculation of Complex Chemical Equilibrium
Compositions, Rocket Performance, Incident and Reflected Shocks, and Chapman-Jouguet
Detonations. NASA-SP-273, NASA Lewis Research Center, Cleveland, Ohio, March 1976.
Gray, P. and J. C. Lee. "Explosive Decomposition and Combustion of Hydrazine." In Fifth Symp. (Intl.)
Combust., pp. 692 - 699. New York, NY: Reinhold, 1955.
Gray, P. and J. C. Lee. "Recent Studies of the Oxidation and Decomposition Flames of Hydrazine." In
Seventh Symp. (Intl.) Combust. London: Butterworth, 1959.
Gray, P. and J. C. Lee. "The Combustion of Gaseous Hydrazine." Trans. Faraday Soc., 50 (1954):
pages 719 -728.
Gray, P., J. C. Lee, and M. Spencer. "Combustion, Flame and Explosion of Hydrazine and Ammonia:
The Spontaneous Ignition of Pure Gaseous Hydrazine." Combust. Flame, 7 (December 1963):
pp. 315 - 321.
Green, R. L., J. P. Stebbins, A. W. Smith, and K. E. Pullen. Advanced Techniques for Determining Long-
Term Compatibility of Materials with Propellants, D180-14839-2, Boeing Aerospace Corp.,
Seattle, WA, Dec. 1973.
Green, R. L., K. E. Pullen, and J. P. Stebbins. "Long-Term Propellant Material Compatibility Assessment
from Short-Term Data." In 1974 JANNAF Propulsion Meeting, Vol. I, Part 2, pp. 519 - 541.
Laurel, MD: Chemical Propulsion Information Agency, Dec. 1974.
Grelecki, C. Fundamentals of Fire and Explosion Hazards Evaluation. AIChE Today Series Seminar,
American Institute of Chemical Engineers, New York, NY, 1976.
Guirao, C. M., R. Knystautas, J. H. Lee, W. Benedick, and M. Berman. Hydrogen-Air Detonations. In
Nineteenth Symp. (Intl.) Combust., Pittsburgh, PA: Combustion Institute, 1982.
Hagemann, D., O. Briles, and T. Farkas. The Compression Ignition Characteristics of Hydrazine. In
1983 JANNAF Propulsion Meeting, pp. 77 - 86. Johns Hopkins University, Laurel, MD: Chemical
Propulsion Information Agency, 1983.
Hall, A. H. and B. H. Rumack (editors). TOMES

Information System. Micromedex, Inc., Englewood,


Colorado (1996).
Hall, A. R. and H. G. Wolfhard. "Hydrazine Decomposition Flame at Sub-Atmospheric Pressures." Trans.
Faraday Soc., 52 (1956): pp. 1520 - 1525.
AIAA SP-084-1999
143
Hannum, E. A. (editor). Hazards of Chemical Rockets and Propellants, Volumes I, II, III. Chemical
Propulsion Information Agency, Laurel, Md.: The Johns Hopkins University, (1985).
Harris, G. F. P. The Effect of Vessel Size and Degree of Turbulence on Gas Phase Explosion Pressures
in Closed Vessels. Combust. Flame, 11 (Feb. 1967): pp. 17 - 25.
Hawleys Condensed Chemical Dictionary, Twelfth Editions, Van Nostrand Reinhold, 1994
Heinrich, H. J. Propagation of Detonation Vapor. In Z. Phys. Chem. N.F., 42 (1964): pp. 149 - 165.
HEW (U.S. Department of Health, Education, and Welfare). NIOSH Criteria for a Recommended
Standard: Occupational Exposure to Hydrazine, DC: Government. Printing Office, June 1978.
Hoffman, B. J. Olfactory Response to Monomethylhydrazine. TR-140-001, NASA White Sands Test
Facility, Las Cruces, NM, 1976.
Hshieh, Fu-Yu Flash and Fire Points of Monomethylhydrazine-Water Mixtures. TR-704-001, NASA
Johnson Space Center White Sands Test Facility, Las Cruces, NM, January 13, 1992.
Hwang, N. H. Fundamentals of Hydraulic Engineering Systems. Englewood Cliffs, NJ: Prentice-Hall,
1981.
Internation Technical Information Institute. Toxic and Hazardous Industrial Chemicals Safety Manual.
Tokyo, Japan, 1981.
Jacobson, K. H., J. H. Clem, H. J. Wheelwright Jr., W. E. Reinhart, and N. Mayes. The Acute Toxicity of
the Vapors of Some Methylated Hydrazine Derivatives. AMA Archives of Industrial Health,
Volume 12, April 15, 1955, pp. 609-616.
Jensen, A. V., ed. Chemical Rocket/Propellant Hazards. Vol. 1: General Safety Engineering Design
Criteria. Silver Spring, MD: Johns Hopkins University, 1976.
Johnson D. L. and D. L. Baker. Determination of Hydrazine, hydrazine, and UDMH in Aqueous Solution
Using Cation Exchange HPLC with Amperometric Detection. In 1992 JANNAF Safety and
Environmental Protection Subcommittee Meeting, Naval Postgraduate School, Monterey, CA:
Chemical Propulsion Information Agency, August 10-14, 1992.
Johnson, R. P. Evaluation of MDA Scientific Models 7080 and 7100 Hydrazine Monitors for the Detection
of Monomethylhydrazine. TR-346-001, NASA White Sands Test Facility, Las Cruces, NM,
January 1986.
Jost, A., K. W. Michel, J. Troe, and H. Gg. Wagner. Detonation and Shock Tube Studies of Hydrazine
and Nitrous Oxide. AR 63-157, Wright-Patterson AFB, OH, 1963.
Jost, W. Investigation of Gaseous Detonations and Shock Wave Experiments with Hydrazine. ARL 62-
330, Wright-Patterson AFB, OH, April 1962.
Kanury, A.M. Introduction to Combustion Phenomena. New York, NY: Gordon and Breach, 1976.
Kilduff, J. E., D. D. Davis, and S. L. Koontz. "Surface-Catalyzed Air Oxidation Reactions of Hydrazines:
Tubular Reactor Studies." In the Third Conference on the Environmental Chemistry of Hydrazine
Fuels, D. A. Stone and F. L. Wiseman, eds., pp. 128 - 137. Tyndall AFB, FL: AFESC, 1988.
Kilduff, J. E., D. D. Davis, and S. L. Koontz. Surface-Catalyzed Air Oxidation of Hydrazines:
Environmental Chamber Studies. In Third Conference on the Environmental Chemistry of
Hydrazine Fuels, D. A. Stone and F. L. Wiseman, ed., pp. 38 - 49. Air Force Engineering and
Services Center, Tyndall AFB, FL, Jan. 1988b.
Kilduff, J., D. D. Davis, and L. Linley. The Evaluation of Materials of Compatibility with Aerospace Fluids.
1990 JANNAF Propellant Development and Characterization Subcommittee, Western Space and
Missile Center, Vandenberg AFB, CA , November 13-15, 1990.
Kinney, G. G. and K. J. Graham. Explosive Shocks in Air. New York: Springer-Verlag, 1985.
Knowles, P. J. Hydrazine Pressure Surges Into an Evacuated Manifold Assembly. In 1981 JANNAF
Propulsion Meeting, pp. 607 - 627. Johns Hopkins University, Laurel, MD.: Chemical Propulsion
Information Agency, May 1981.
Kuch, David J. Bioremediation of Hydrazine: A Literature Review. AL/EQ-TR-1994-0055, Armstrong
Laboratory, Environics Directorate, Tyndall Air Force Base, Florida, April 1996.
Kuchta, J. M. Fires and Explosion Manual for Aircraft Accident Investigations. AFAPL-TR-73-74, Wright-
Patterson AFB, OH, 1973.
Ladacki, M. Surface Activity of Metals with Hydrazine Fuels at Elevated Temperatures. Memo RR-68-6,
Rocketdyne, Canoga Park, CA, personal correspondence.
AIAA SP-084-1999
144
Lapedes, D. N. Dictionary of Scientific and Technical Terms. New York, NY: McGraw-Hill, 1974.
Lawton, E. A., C. M. Moran, and S. DiStefano. Corrosion of Metals by Hydrazine. 1985 JANNAF
Propulsion Meeting, Vol. 1 CPIA AD-A161084 San Diego, CA April 9-12.
Leavitt, N. A. and C. B. Mattson and R. C. Young. Evaluation of Portable Vapor Detectors for hydrazine,
N
2
H
4
, and UDMH at a 10-PPB Concentration. In 1992 JANNAF Safety and Environmental
Protection Subcommittee Meeting, Naval Postgraduate School, Monterey, CA: Chemical
Propulsion Information Agency, August 10-14, 1992.
Leavitt, N. A., C. B. Mattson, Z. Laney and R. C. Young. Evaluation of GMD MK-2 Portable Vapor
Detector for hydrazine, N
2
H
4
, and UDMH. In 1991 JANNAF Safety and Environmental Protection
Subcommittee Meeting, Publication 569. NASA Kennedy Space Center, FL: Chemical Propulsion
Information Agency, July 1991, pp. 219-223..
Lee, J. H. Dynamic Parmeters of Gaseous Detonations. Ann. Rev. Fluid Mech., Volume 16 (1984), pp.
311-336.
Lee, J. H. and C. M. Guirao. Factors that Influence the Pressure Development in Closed and Vented
Vessels. Paper presented at the Fifteenth Loss Prevention Symposium of the American Institute
of Chemical Engineers, Ottawa, Canada, August 1981.
Lee, J. H. and I. O. Moen. The Mechanism of Transition from Deflagration to Detonation in Vapor Cloud
Explosions. Prog. Energy Combust. Sci., 6 (1980): pp. 359 - 389.
Lee, J. H., C. Chan, and R. Knystautas. Hydrogen-Air Deflagrations: Recent Results. Proceedings of
the Second International Workshop on the Impact of Hydrogen on Water Reactor Safety, M.
Berman, ed., SAND82-2456, Sandia National Laboratories, Albuquerque, NM, October 1982.
Lee, J. H., R. Knystautas, C. M. Guirao, W. A. Benedick, and J. E. Shepherd. Hydrogen-Air
Detonations. In Proceedings of the Second Intl. Workshop on the Impact of Hydrogen on Water
Reactor Safety. M. Berman, ed. Albuquerque, NM: Sandia National Labs, 1982.
Lewis, B. and G. von Elbe. Combustion, Flames, and Explosions of Gases. 2nd ed. New York, NY:
Academic Press, 1961.
Litchfield, E. L., M. H. Hay, and D. R. Forshey. Direct Electrical Initiation of Freely Expanding Gaseous
Detonation Waves. In Ninth Symp. (Intl.) Combust., pp. 282 - 286. New York: Academic Press,
1963.
Loper, G. L. Hypergol Vapor Detector (HVD) Technical Interchange. TOR-0081(6447)-1, Aerospace
Corp., El Segundo, CA, August 1981.
Lucien, H. W. "Thermal Decomposition of Hydrazine." J. Chem. Eng. Data, 6 (1961): pp. 584 - 586.
Luskus, L. J. and H. J. Kilian. Test and Evaluation of an Energetics Science Incorporated Model 7660
Electrochemical Hydrazines Analyzer. SAM-TR-80-10, USAF School of Aerospace Medicine,
Brooks AFB, TX, June 1986.
MacEwen, J. D., and E. H. Vernot. Toxic Hazards Research Unit Annual Technical Report: 1971.
Aerospace Medical Research Laboratory, Aerospace Medical Division, Air Force Systems
Command, WPAFB, Ohio, 1971.
MacEwen, J. D., and E. H. Vernot. Toxic Hazards Research Unit Annual Technical Report: 1979.
Aerospace Medical Research Laboratory, Aerospace Medical Division, Air Force Systems
Command, WPAFB, Ohio, 1979.
MacEwen, J. D., J. Theodore, and E. H. Vernot. Human Exposure to EEL Concentrations of
Monomethylhydrazine. Proceedings of the 1st Annual Conference on Environmental Toxicology.
AMRL-TR-70-102, Paper no. 23, Aerospace Medical Research Laboratory, WPAFB, Ohio, 1970,
pp. 363.
Mallinckrodt. Material Safety Data: Methyl hydrazine. Paris, Ky.: Mallinckrodt, Inc., 1987.
Mancus, H. V. and F. J. Benz. Thermal Compatibility of Kalrez 1045, FEP Teflon, Dyed and Undyed
Tefzel in Hydrazine at Elevated Temperatures. TR-293-001, NASA White Sands Test Facility,
Las Cruces, NM, July 1981.
Manson, N. and F. Ferrie. Contribution to the Study of Spherical Detonation Waves. In Fourth Symp.
(Intl.) Combust., pp. 486 - 494. Baltimore, MD: Williams and Wilkins, 1953.
AIAA SP-084-1999
145
Manson, N., C. H. Brochet, J. Brossard, and Y. Pujol. Vibratory Phenomena and Instability of Self-
sustained Detonations in Gases. In Ninth Symp. (Intl.) Combust., pp. 461 - 469. New York:
Academic Press, 1963.
Marsh, W. R. and B. P. Knox. US Air Force Propellant Handbook: Vol. 1, Hydrazine Fuels. Buffalo, NY:
Bell Aerospace CO, 1970.
Marsh, W. R., and B. P. Knox. U.S. Air Force Propellant Handbook: Vol. 1, Hydrazine Fuels. Buffalo,
NY: Bell Aerospace Co., 1970.
Marsh, Walter R. and Bruce P. Knox USAF Propellant Handbooks Hydrazine Fuels. Volume 1. Air Force
Systems Command, Edwards, California: Air Force Rocket Propulsion Laboratory Research and
Technology Division, March 1970.
Martin, N. B., D. D. Davis, J. E. Kilduff, and W. C. Mahone. Environmental Fate of Hydrazines: AFESC
Final Report. NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM, June
1989.
Martinkovic, P. J. Space Vehicle Propulsion Compartment Fire Hazard Investigation. Vol. 1: Fire Hazard
Investigation of Combustible Materials in Spacecraft Propulsion Compartment. RPL-TDR-64-
103, Edwards AFB, CA, 1964.
Michel, K. W. and H. Gg. Wagner. "The Pyrolysis and Oxidation of Hydrazine Behind Shock Waves." In
Tenth Symp. (Intl.) Combust. pp. 353 - 364. Pittsburgh, PA: Combustion Institute, 1965.
Military Specification. Propellant, Hydrazine. MIL-P-26536D, November 1, 1986.
Miller, E. L. Evaluation of the Interscan Model 4186 Compact Portable Analyzer as a Monitor for
Threshold Limit Value (TLV) Concentrations of Hydrazine, Monomethylhydrazine (hydrazine), and
Unsymmetrical Dimethylhydrazine (UDMH) Vapors in Air. TR384-001, NASA White Sands Test
Facility, Las Cruces, NM, Nov. 1987.
Miller, E. L. and L. A. Schluter. Thermal Regeneration Temperatures of Materials Exposed to Hydrazine
Vapor and Air Mixtures. TR-225-001, NASA Johnson Space Center White Sands Test Facility,
Las Cruces, NM, June 1978.
Miller, E. L. and L. A. Schluter. Thermal Regeneration Temperatures of Materials Exposed to Hydrazine
Vapor and Air Mixtures. TR-225-001, NASA Johnson Space Center White Sands Test Facility,
Las Cruces, NM, June 1978.
Moen, I. O., D. Bjerketvedt, A. Jenssen, and P. A. Thibault. Transition to Detonation in a Large Fuel-Air
Cloud. Combustion and Flame, 61 (1985): pp. 285 - 291.
Moen, I. O., M. Donato, R. Knystautas, and J. H. Lee. Flame Acceleration Due to Turbulence Produced
by Obstacles. Combustion and Flame, Vol. 39 (1980): pp. 21-32.
Moen, I. O., R. Knystautas, J. H. Lee, H. Gg. Wagner, and M. Donato. Turbulent Flame Propagation and
Acceleration in the Presence of Obstacles. In Seventh (Intl.) Colloquium on Gas Dynamics of
Explosions and Reactive Systems. Gottingen, Germany, August 1979.
Moran, C. and R. Bjorklund. Propellant/Material Compatibility Program and Results: Ten-Year Milestone.
Report 82-62, Jet Propulsion Laboratory, Pasadena, CA, July 1982.
Moran, C.M. and R. A. Bjorklund, Test Program to demonstrate the Stability of Hydrazine in Propellant
Tanks - Final Report, NASA, JPL Publication 83-37, Pasadena, CA. April 1983.
Mullins, B. P. Spontaneous Ignition of Liquid Fuels. London: Butterworth, 1955.
Mullins, B. P. and S. S. Penner. Explosions, Detonations, Flammability, and Ignition. London, UK:
Pergamon Press, 1959.
Murray, R. C. and A. R. Hall. "Flame Speeds in Hydrazine Vapor and in Mixtures of Hydrazine and
Ammonia with Oxygen." Trans. Faraday Soc. 47 (1951): pp. 743 - 751.
Nagy, J., E. C. Seiler, J. W. Conn, and H. C. Verakis. Explosion Development in Closed Vessels. Report
of Investigation 7505, Bureau of Mines, Pittsburgh, PA, 1971.
NASA. Flammability, Odor, and Offgassing Requirements and Test Procedures for Materials in
Environments that Support Combustion. NASA-STD-6001 (formerly NHB 8060.1C). Office of
Space Transportation Systems, National Aeronautics and Space Administration, Washington,
DC, February 9, 1998.
National Fire Protection Association. Fire Hazard Properties of Flammable Liquids, Gases, and Volatile
Solids. Boston, MA, 1977.
AIAA SP-084-1999
146
NRC. Emergency and Continuous Exposure Guidance Levels for Selected Airborne Contaminants,
Volume 5. National Research Council, 1985.
Occupational Hazards. OSHA Court Strikes Down OSHA PELs. Occupational Hazards. August 1992,
pp. 11-12.
Occupational Health Services, Inc. Hazardline. New York, New York: Occupational Health Services,
Inc., 1987.
Olin Chemical Company. Anhydrous Hydrazine Handling and Storage. Product Data AD-1010-367, Olin
Chemical Division, Olin Chemical Company, 1967.
Olin Corporation. Monomethyl Hydrazine (hydrazine) Material Safety Data Sheet. Stamford, CT, 1992.
Olin Corporation. Monomethyl Hydrazine (Hydrazine) Product Data Sheet. Stamford, CT, 1981.
Overly, J. Establishing Criteria for the Ignition: An Experimental and Theoretical Study of the Effect of
Spark Geometry on the Minimum Ignition Energy of Hydrazine. Ph.D. thesis, Vanderbilt
University, Nashville, TN, 1976.
Pannetier, G. Bultingair-Laborde. Madeleine, Bull. Soc. Chim., France, 1958.
Parker, S. P., ed. Dictionary of Scientific and Technical Terms. 3rd ed. New York: McGraw-Hill, 1984.
Pedley, M. D. Accelerated Rate Calorimetry Analysis of Various 300-Series Stainless Steels Found in the
Rockwell RCS Low-Pressure Solenoid Valve. TR-515-001, NASA White Sands Test Facility, Las
Cruces, NM, April 1987.
Pedley, M. D., C. V. Bishop, F. J. Benz, C. A. Bennett, R. D. McClenagan, D. L. Fenton, R. Knystautas, J.
H. Lee, O. Peraldi, G. Dupre, and J. E. Shepherd. Hydrazine Vapor Detonations. In Dynamics
of Explosions, A. L. Kuhl, J. R. Bowen, J.-C. Leyer, and A. Borisov, eds., pages 45 - 63.
Washington, D.C.: AIAA, 1988.
Perlee, H. E., A. C. Imhof, and M. G. Zabetakis. "Flammability Characteristics of N
2
H
4
Fuels in N
2
O
4
Atmospheres." J. Chem. Eng. Data, 7 (July 1962): pp. 377 - 379.
Pinkerton, M. K., E. A. Hagan, and K. C. Back. Distribution and Excretion of 14C-Monomethylhydrazine.
AMRL-TR-67-175, Aerospace Med. Res. Lab., WPAFB, Ohio, 1967.
Piper, L. B. Hydrazine Compatibility. CPTR-85-32, Chemical Propulsion Information Agency, Laurel,
Md., August 1985.
Pohlhammer, M and E. J. Drown. Design of Cargo Tankers for Transport of Hypergolic Propellants. In
JANNAF Interagency Propulsion Committee Conference, Chemical Propulsion Information Agency
(1991).
Prickett, R. P., E. Mayer, and J. Hermel. Water Hammer in a Spacecraft Propellant Feed System. In
24th Joint Propulsion Conference. Washington, D.C.: AIAA, 1988.
Rathgeber, K. Hydrazine Vapor/Air Detonation Tests Special Test Data Report. WSTF-95-28962,
NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM, May 13, 1997a.
Rathgeber, K. Space Station Rapid Compression Special Test Data Report. WSTF-92-26694, NASA
Johnson Space Center White Sands Test Facility, Las Cruces, NM, Nov. 1992.
Rathgeber, K. Condensed Phase Detonation Studies Special Test Data Report. WSTF-90-24354, NASA
Johnson Space Center White Sands Test Facility, Las Cruces, NM, Sept. 1990.
Rathgeber, K. Hydrazine Vapor/Ammonia Vapor Detonation Tests Special Test Data Report. WSTF-95-
29316, NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM. May 13,
1997d.
Rathgeber, K. Hydrazine Vapor/Helium Gas Detonation Tests Special Test Data Report. WSTF-95-
28785, NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM. May 13,
1997c.
Rathgeber, K. Hydrazine Vapor/Hydrogen Gas Detonation Tests Special Test Data Report. WSTF-94-
28600, NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM. May 13,
1997b.
Rathgeber, K. Hydrazine Vapor/Nitrogen Gas Detonation Tests Special Test Data Report. WSTF-94-
28578, NASA Johnson Space Center White Sands Test Facility, Las Cruces, NM. 1995.
Reinhardt, C. F., and M. R. Britteli. Heterocyclic and Miscellaneous Nitrogen Compounds. Patty's
Industrial Hygiene and Toxicology, 3rd Edition, G. D. Clayton and F. E. Clayton, editors, New
York: John Wiley & Sons, 1981, pp. 2671-2823.
AIAA SP-084-1999
147
Reinhardt, C. F., and M. R. Britteli. Hydrazines. Patty's Industrial Hygiene and Toxicology, 3rd Edition,
G. D. Clayton and F. E. Clayton, editors, New York: John Wiley & Sons, 1981, pp. 2791-2805
Rose, S. L., and J. R. Holtzclaw. The Critical Comparison of Commercially Available Hydrazine
Detectors. In JANNAF Safety and Environmental Protection Subcommittee Meeting. Laurel,
MD, Chemical Propulsion Information Agency (1984): pp. 171-183.
Rose-Pehrsson, S. L. and K. P. Crossman. Development of a Dosimeter System for UDMH, hydrazine,
and Hydrazine. In 1992 JANNAF Safety and Environmental Protection Subcommittee Meeting,
Naval Postgraduate School, Monterey, CA: Chemical Propulsion Information Agency, August 10-
14, 1992.
Rose-Pehrsson, S. L., J. W. Grate, and M. Klusty. Hydrazine Detection Using Chemiresistors. In 1989
JANNAF Safety and Environmental Protection Subcommittee Meeting, Publication 508, Brooks
AFB, San Antonio, TX: Chemical Propulsion Information Agency (April 1989) pp. 157-166.
Roth, E. M. Space-Cabin Atmospheres. Part II: Fire and Blast Hazards. NASA-SP-48 Wash., D.C.:
NASA, 1964. N64-20744
Saad, M. A., M. B. Detweiler, and M. A. Sweeney. "Analysis of Reaction Products of Nitrogen Tetroxide
with Hydrazines Under Nonignition Conditions." AIAA J., 10 (1972): pp. 1073 - 1078.
Salvinski, R. J., G. W. Howell, and D. H. Lee. Advanced Valve Technology: Materials Compatibility and
Liquid Propellant Study, Vol. 2. Interim report, 06641-6041-R000, TRW Systems, Redondo
Beach, CA, Nov. 1967.
Saunders, R. A. and J. T. Larkins. Detection and Monitoring of Hydrazine, Monomethylhydrazine, and
Their Decomposition Products. NRL 3313, Naval Research Laboratory, DC, June 1976.
Saunders, R. A., J. J. DeCorpo, B. J. Stammerjohn, and R. J. Kautter. Evaluation of an Electrochemical
Detector for Trace Concentrations of Hydrazine Compounds in Air. NRL-8199, Naval Research
Laboratory, DC, April 1978.
Sawyer, R. F. and I. Glassman. Gas Phase Reactions of Hydrazine with Nitrogen Dioxide, Nitric Oxide,
and Oxygen. In Eleventh Symp. (Intl.) Combust, pages 861 -869. Pittsburgh, PA: Combustion
Inst., 1967.
Sax, I. N. Dangerous Properties of Industrial Materials. New York: Van Nostrand, 1984.
Sax, N. I. Dangerous Properties of Industrial Materials. 4th ed. New York, NY: Van Nostrand Reinhold,
1975.
Sax, N. I. and R. J. Lewis Sr., ed. Hawley's Condensed Chemical Dictionary. 11th ed. New York: Van
Nostrand Reinhold Company, 1987.
Sayer, C. F. The Heterogeneous Decomposition of Hydrazine: Part 1 Kinetics of the Liquid Phase
Decomposition on a Supported Indium Catalyst. Technical Report 68/8, Rocket Propulsion
Establishment, Wescott, England, Oct. 1968.
Sayer, C. F. The Heterogeneous Decomposition of Hydrazine: Part 4 Kinetics of the Decomposition of
Liquid Hydrazine on a Supported Rodium Catalyst. Technical Report 71/16, Rocket Propulsion
Establishment, Wescott, England, Nov. 1971.
Sayer, C. F. The Heterogeneous Decomposition of Hydrazine: Part 5 Kinetics of the Liquid Hydrazine
on a Supported Ruthenium Catalyst. Technical Report 72/1, Rocket Propulsion Establishment,
Wescott, England, Oct. 1968.
Schmidt, E. W. Hydrazine and Its Derivatives: Preparation, Properties, Applications. New York: John
Wiley & Sons, 1984.
Scott, F. E., J. J. Burns, and B. Lewis. Explosive Properties of Hydrazine. BM-RI-4460, Bureau of Mines,
Pittsburgh, PA, 1949.
Setchkin, N. P. "Self-Ignition Temperatures of Combustible Liquids." J. Res. NBS, 53 (1954): pp. 49 -
66.
Shaver, D. K. and R. L. Berkowitz. Pollution Technology Review No. 109: Post-Accident Procedures for
Chemicals and Propellants. Park Ridge, NJ: Noyes Publications, 1984.
Sherman, M. P., S. R. Tieszen, W. B. Benedick, J. W. Fisk, and M. Carcassi. The Effect of Transverse
Venting on Flame Acceleration and Transition to Detonation in a Large Channel. In Dynamics of
Explosions, J. R. Bowen, J. C. Leyer, and R. I. Soloukhin, ed. New York, NY: American Institute
of Aeronautics and Astronautics, 1986.
AIAA SP-084-1999
148
Smith, E. B., and D. A. Clark. Nomograms Correlating Dose of hydrazine with Blood Levels. Aerospace
Medicine, Vol. 40 (1969), pp. 1373-1377.
Smith, E. B., and D. A. Clark. Evaluation of Tannic Acid and Water Washes in Prevention of Absorption
of hydrazine Through Skin. Aerospace Medicine, Vol. 42 (1971), pp. 662.
Smith, I. D. Evaluation of Personnel Protection Systems Utilizing Compressed Air.
Smith, I. D. Permeation of Personnel Protective Equipment by Nitrogen Tetroxide, Monomethylhydrazine,
and Hydrazine. TR-265-001, NASA White Sands Test Facility, Las Cruces, NM, March 1980.
Smith, M. M. and H. C. Van Ness. Introduction to Chemical Engineering Thermodynamics, 3rd ed. New
York: McGraw-Hill, 1975.
Stevens, B. D. and F. J. Benz. Autoignition Characteristics of Hydrazine at Reduced Pressure. TR-205-
004, NASA JSC White Sands Test Facility, NM, 1978a.
Stevens, B. D. and F. J. Benz. Flammability of Vapor Hydrazine and Monomethylhydrazine in Air at
Reduced Pressures and Elevated Temperatures. NASA-TR-282-001, NASA JSC White Sands
Test Facility, Las Cruces, NM, 1980.
Street, Jimmy, Clifford Johnston, Li Ou, Robert Mansell, and Steve Bloom. Environmental Interactions of
Hydrazine Fuels in Soil/Water Systems. ESL-TR-88-24, University of Florida, Soil Science
Department, Gainsville, Florida, October 1988.
Strehlow, R. A. Combustion Fundamentals. New York, NY: McGraw-Hill, 1984.
Stull, D. R. Monograph Series: Fundamentals of Fire and Explosions. New York: American Institute of
Chemical Engineers, 1977.
Summers, W. H., and E. T. McMullen. "Combustion of the N
2
H
4
-N
2
O
4
Propellant System." Paper
presented at the Second Propulsion Joint Specialist Conference, Colorado Springs, CO, 1966.
Taffe, P. A., J. C. Travis, and S. L. Rose-Pehrsson. Field Evaluation of the Passive Hydrazine
Dosimeters. In 1989 JANNAF Safety and Environmental Protection Subcommittee Meeting,
Publication 508. Brooks AFB, San Antonio, TX: Chemical Propulsion Information Agency (April
1989) pp. 167-172..
Tannenbaum, S. and A. J. Beardell. Characterization, Chemical Compatibility, Storability, and Hazard
Testing of Liquid Propellants. In Propellant Manufacturing, Hazards, and Testing Symp. D.
Boyars and K. Klager, ed., pages 344 - 368. Washington, D. C.: Amer. Chem. Soc., 1969.
Tannenbaum, S. Packaged Liquid Propellants. RMD 5005-F, Thiokol Chemical Corp., Denville, NJ,
1962.
Tarver, C. M. Chemical Energy Release in the Cellular Structure of Gaseous Detonation Waves.
Combustion and Flame, 46 (1982): pp. 135 - 156.
Taylor, W. F., M. Lieberman, and M. S. Cohen. Development of a Low Cost Catalyst for Hydrazine.
ESSO Report # GR-4-DCH-69, ESSO Research and Engr. Div. Linden, N. J. March 1969.
Toth, B. Actual New Cancer-Causing Hydrazines, Hydrazides, and Hydrazones. Journal of Cancer
Research and Clinical Oncology. Vol. 97 (1980), pp. 97-108.
Toth, B. Hydrazine, Methylhydrazine, and Methylhydrazine Sulfate Carcinogenesis in Swiss Mice.
International Journal of Cancer, Vol. 9 (1972), pp. 109-118.
Toth, B., and H. Shimizu. Methylhydrazine Tumorigenesis in Syrian Golden Hamsters and the
Morphology of Malignant Histiocytomas. Cancer Research, Vol. 33 (1973), pp. 2744-2753.
Toth, L. R., W. A. Cannon, C. D. Coulbert, H. R. Long, "Propellant/Material Compatibility-Program and
Results." NASA Tech. Memorandum 33-779, JPL, Pasadena, CA. August 1976.
TR-278-001, NASA White Sands Test Facility, Las Cruces, NM, June 1981.
U. S. Air Force. Chemical Rocket/Propellant Hazards Manual Vol. 2: Liquid Propellants. Washington,
D.C.: 1973.
U.S. Department of Health and Human Services. NIOSH Pocket Guide to Chemical Hazards. U.S.
Government Printing Office, 1985.
USAF. Chemical Rocket/Propellant Hazards. Vol. II, Liquid Propellants. DC: Government Printing
Office. April 1973.
Van Dolah, R. W., Zabetakis, M. G., and Scott, G. S. Review of Fire and Explosion Hazards of Flight
Vehicle Combustibles. Bureau of Mines Information Circular 8137, 1962.
AIAA SP-084-1999
149
Vander Wall, E. M., J. K. Suder, R. L. Beegle, Jr., and J. A. Cabeal. Propellant-Material Compatibility
Study. AFRPL-TR-71-41, Aerojet Liquid Rocket, Sacramento, CA, Dec. 1971.
Webb, K. Cavitating Venturi Characterization Test - Special Test Data Report. WSTF-92-26736, NASA
Johnson Space Center White Sands Test Facility, Las Cruces, NM, Dec. 1992
Webb, K. Simulated Space Station Propulsion Module (SSPM) Rapid Compression - Special Test Data
Report. WSTF-93-27076, NASA Johnson Space Center White Sands Test Facility, Las Cruces,
NM, April 1993
Webb, K. Space Station Propulsion Module (SSPM) Material and Instrumentation Comparison Testing -
Special Test Data Report. WSTF-93-27012, NASA Johnson Space Center White Sands Test
Facility, Las Cruces, NM, April 1993
Webb, K. D., D. D. Davis, and R. L. Bunker. Evaluation of the Thermal and Material Degradation Hazards
of Gold-Nickel Braze in Hydrazine Systems. TR-599-001, Lyndon B. Johnson Space Center
White Sands Test Facility, Las Cruces, NM, January 17, 1992.
Webb, K., Poe, R., Peterson, J. Gamma Ray Observatory Rapid Compression - Special Test Data
Report. WSTF-92-26101/02, NASA Johnson Space Center White Sands Test Facility, Las
Cruces, NM, Sep. 1992
Weber, L. Prevention of Hydrogen Embrittlement of High-Strength Steel. U. S. Patent 3919014, Nov. 11,
1975.
Wedlich, R. C. and D. D. Davis. "Non-isothermal Kinetics of Hydrazine Decomposition". Thermochimica
Acta, 171 (1990)
Wedlich, R., D. D. Davis, and T. Peters. "Evaluation of Thermal Hazards of Hydrazine Decomposition by
Accelerating Rate Calorimetry." In 1988 JANNAF Safety and Environmental Protection
Subcommittee Meeting. pp. 33 - 42. Johns Hopkins University, Laurel, Md.: Chemical Propulsion
Information Agency, May 1988.
Westbrook, C. K. and P. A. Urtiew. Chemical Kinetics of Critical Parameters in Gaseous Detonations.
In Nineteenth Symp. (Intl.) Combust., pages 615 - 625. Pittsburgh, PA: Combustion Institute,
1982.
Whitwell, J. C. and R. K. Toner. Conservation of Mass and Energy. Waltham, Mass.: Blaisdell
Publishing, 1969.
Worthington, N. S. Investigation of Intergranular Corrosion and Stress Corrosion Cracking of Selected
APU Alloys in a Hydrazine Environment. TR-387-001, NASA Johnson Space Center White
Sands Test Facility, Las Cruces, NM, May 1986.
Yaws, Carl L. Physical Properties, A Guide to the Physical, Thermodynamic and Transport Property Data
of Industrially Important Chemical Compounds. New York, New York: McGraw-Hill Publishing,
1977.
Young, R. C., and J. C. Travis. An Analysis of Commercial Vapor Detectors as Continuous Monomethyl
Hydrazine Monitoring Devices. JANNAF Safety & Environmental Protection Subcommittee
Meeting, Johns Hopkins University Applied Physics Lab, Chemical Propulsion Information
Agency, 1987.
Young, R. C., C. F. McBrearty, D. L. Curran, N. A. Leavitt, and D. M. Blaies. Follow-on Development and
Testing of the interim Active Vanillin Sampler (IAVS). In 1992 JANNAF Safety and
Environmental Protection Subcommittee Meeting, Naval Postgraduate School, Monterey, CA:
Chemical Propulsion Information Agency, August 10-14, 1992.
Young, R. C., W. R. Helms, and J. C. Travis. NASA/KSC's Development of a Portable Hydrazines Vapor
Detector with 10-PPB Detection Capability. In 1990 JANNAF Safety and Environmental
Protection Subcommittee Meeting, Publication 543. Lawrence Livermore National Laboratory,
Livermore, CA: Chemical Propulsion Information Agency (June 1990) pp. 104-114.
Zabetakis, M. B. Fire and Explosion Hazards at Temperature and Pressure Extremes. Paper presented
at the Joint Meeting of the American Institute of Chemical Engineers-I. Chem. E., London, June
1965.
Zakin, M. R., L. S. Bernstein, and D. L. Ellis. Development of a Conductive Polymer Sensor for Real-
Time Monitoring of Hydrazine and hydrazine at the Part-Per-Billion Level. In 1992 JANNAF
AIAA SP-084-1999
150
Safety and Environmental Protection Subcommittee Meeting, Naval Postgraduate School,
Monterey, CA: Chemical Propulsion Information Agency, August 10-14, 1992.
Zakin, M. R., L. S. Bernstein, and R. A. Moody. Dosimetry of Hydrazines Using Conductive Polymer-
Based Sensing Devices. In 1990 JANNAF Safety and Environmental Protection Subcommittee
Meeting, Publication 543, Lawrence Livermore National Laboratory, Livermore, CA: Chemical
Propulsion Information Agency, June 1990.
Zeldovich, Y. B., G. I. Barenblatt, V. B. Librovich, and G. M. Makhviladze. The Mathematical Theory of
Combustion and Explosions. New York: Consultants Bureau, 1985.
Zung, L. B.. B. P. Breen, and R. Kushida. A Basic Study on the Ignition of Hypergolic Liquid Propellants.
Paper No. 68-43 presented at Western States Section/The Combustion Institute Stanford
Research Institute, Menlo Park, October 29, 1968.
151
American Institute of
Aeronautics and Astronautics
1801 Alexander Bell Drive, Suite 500
Reston, VA 20191-4344
ISBN 1-56347-346-1

Anda mungkin juga menyukai