Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal of Energetic Materials

ISSN: 0737-0652 (Print) 1545-8822 (Online) Journal homepage: http://www.tandfonline.com/loi/uegm20

Review on Thermal Decomposition of Ammonium


Nitrate

Shalini Chaturvedi & Pragnesh N. Dave

To cite this article: Shalini Chaturvedi & Pragnesh N. Dave (2013) Review on Thermal
Decomposition of Ammonium Nitrate, Journal of Energetic Materials, 31:1, 1-26, DOI:
10.1080/07370652.2011.573523

To link to this article: https://doi.org/10.1080/07370652.2011.573523

Published online: 09 Jul 2012.

Submit your article to this journal

Article views: 959

View related articles

Citing articles: 45 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=uegm20
Journal of Energetic Materials, 31: 1–26, 2013
Copyright # Taylor & Francis Group, LLC
ISSN: 0737-0652 print=1545-8822 online
DOI: 10.1080/07370652.2011.573523

Review on Thermal Decomposition


of Ammonium Nitrate

SHALINI CHATURVEDI AND PRAGNESH N. DAVE


Department of Chemistry, Kachchh University, Kachchh, Bhuj, India

In this review data from the literature on thermal decomposition of ammonium


nitrate (AN) and the effect of additives to their thermal decomposition are summar-
ized. The effect of additives like oxides, cations, inorganic acids, organic com-
pounds, phase-stablized CuO, etc., is discussed. The effect of an additive mainly
occurs at the exothermic peak of pure AN in a temperature range of 200 C to
140 C.

Keywords additive; ammonium nitrate; differential scanning calorimetry (DSC);


propellants; thermal decomposition

Introduction
Solid propellants are extensively used in space and military missions. They consist of
two major components, namely, a gas-producing solid oxidizer and an organic poly-
meric fuel, which also acts as a binder. Commonly used oxidizers are perchlorates
and nitrates. Most solid rocket motors employ ammonium perchlorate (AP) as the
main oxidizer [1,2]. It is characterized by its high performance and good burning
rate. However, propellants based on ammonium perchlorate have certain disadvan-
tages. One of the preliminary products of combustion of ammonium perchlorate–
based propellants is HCl, which in the presence of water forms hydrochloric acid,
which produces smoke and is highly toxic, in addition to the corrosion of the launch
pad. Another threat from chlorine compounds emitted by AP is to the stratospheric
ozone layer and consequent global warming. Moreover, these propellants do not
meet the requirements of decreased sensitivity, and disposal of AP after the life cycle
of the propellants creates many problems. To overcome these environmental prob-
lems, clean burning propellants, which eliminate chlorine as an ingredient, are put
forward.
Ammonium nitrate is of interest as a potential replacement for ammonium per-
chlorate as a solid propellant oxidizer [3,4]. Ammonium nitrate (AN, NH4NO3) is
one of the most commercially important ammonium compounds in terms of usage
in the chemical industry and in agriculture. It has extensive application in the area
of nitrogen fertilizers and explosives [5–7]. Though it acts as a source of ammonia
and nitrate ion vital to plants in the form of nitrogen fertilizer, in explosives and pro-
pellants the nitrate ion is a source of oxygen and its application is as an oxidizer. AN

Address correspondence to Pragnesh N. Dave, Department of Chemistry, Kachchh


University, Mundra Road, Kachchh, Bhuj 370001, India. E-mail: pragneshdave@gmail.com

1
2 S. Chaturvedi and P. N. Dave

is the principal component of most industrial explosives. Several compositions of


AN such as ammonium nitrate–fuel oil (ANFO), amatol, etc., are well-known explo-
sives. However, its use in the field of propellants=pyrotechnics, unlike potassium
nitrate, which is the principal constituent of black powder or gun powder and was
used in the earliest solid rockets, or AP, which is the main oxidizer of modern solid
propellants, is rather limited. Its principal use in propellants is restricted to
low-burning-rate, low-performance applications, such as gas generators for turbo
pumps of liquid propellant rocket engines or emergency starters for jet aircraft [8].
Ammonium nitrate has attracted the attention of researchers in different disci-
plines for a variety of reasons. Firstly, the intrinsic phase transitions of this solid
[9] have been studied in the field of solid-state physics in order to examine the
details of such transformations [10–12]. Secondly, ammonium nitrate is a common
fertilizer, often showing the unwanted phenomenon of agglomeration. It is believed
that the phase transitions of the solid that take place near ambient temperatures
(20–30 C) contribute to this. Therefore, researchers in fertilizer technology are
interested in methods that may help to prevent such phenomena [13–15]. Thirdly,
ammonium nitrate is an explosive compound and because, as noted above, it is
produced in very large quantities, the conditions under which such disastrous
effects can take place should be carefully scrutinized and scrupulously avoided
[16–22].
Pure AN is considered relatively safe in comparison with other reactants due to
its high stability in the lower temperature range. However, in the process of its manu-
facture, storage, transport, and use, AN is often contaminated by impurities, such as
inorganic acid, organic oil, and others. Auto-ignition and explosion may take place
due to the catalytic effects of the impurity on the decomposition of AN.

Properties of AN
AN is a colorless, crystalline salt and highly soluble in water. Although it is hygro-
scopic, it does not form hydrates. It is also soluble in alcohol, acetic acid, and nitric
acid. AN dissolves in liquid ammonia to form what is known as Divers’ solution and
can be used to strip ammonia from gases. AN has a negative heat of solution in
water and can therefore be used to prepare freezing mixtures. The chemical reactivity
of AN has been well documented by Mellor [23]. The boiling point of pure material
is around 210 C at 11 mmHg and it distills without decomposition. It decomposes
around 230 C at 760 mmHg, and above 325 C it deflagrates. If confined, AN may
explode between 260 and 300 C [24]. Data on solubility, vapor pressure, boiling
point, specific heat of aqueous AN solutions, and many other properties, especially
those relevant to its use as a component of explosive mixtures, are well documented
[6,7,23,25] (Table 1).
Ammonium nitrate is an interesting molecular-ionic crystal that has been stud-
ied using many of the currently available scientific techniques. AN exhibits a series of
temperature-dependent phases at atmospheric pressure that are characterized by suc-

cessively greater motional freedom of the NHþ 4 and NO3 ions. Three of these phases
(commonly referred to as phases VII, V, and IV) occur in the 258 to 27 C tempera-
ture range with the transition point (V to VII) at approximately 178 C [26] and To
(IV to V) at 27 C [27]. The highest temperature phase (phase I) is analogous to the
plastic phase observed in many molecular crystals because large reorientations of

both NHþ 4 and NO3 ions occur.
Thermal Decomposition of Ammonium Nitrate 3

Table 1. Properties of AN

Property Property value


Molecular formula NH4NO3
Molecular weight 80
Heat of combustion 1,447.7 J=g
Heat of formation 4,594 J=g
Heat of explosion 1,447.7 J=g
Heat of fusion 76.7 J=g
Density 1.725 g=cm3
Color Colorless
Melting point 169.6 C
Specific volume 0.580 cm3=g
Solubility in water at 20 C 66 g=100 g
Oxygen content 60%
Available oxygen 20%
Estimated flame temperature 1500 C
Detonation velocity 1,250–4,650 m=s
Coefficient of thermal expansion at 20 C 9.82  104%= C
Specific heat from 0 to 31 C 1.72 J=mol
Vapor pressure at 205 C 7.4 mmHg

Phase I cubic 170 to 125 C


Phase II tetragonal 125 to 84 C
Phase III orthorhombic 84 to 32 C
Phase IV orthorhombic 32 to 16 C
Phase V tetragonal below 16 C
A variety of spectroscopic investigations, including nuclear magnetic resonance
(NMR) imaging [28], X-ray diffraction (XRD) [29], neutron diffraction and scatter-
ing [30,31], infrared absorption [32–34], and Raman scattering [35–37], were conduc-
ted for AN [38].
The structure of the low-temperature phase V was determined by neutron dif-
fraction in two different studies [39–41]. It is an ordered structure with orthorhombic
symmetry, with space group Pccn and Z ¼ 8. On heating, phase V transforms to
phase IV with orthorhombic symmetry with space group Pmmn with two NH4NO3,
formula units per unit cell. In the presence of moisture, phase IV undergoes a room
temperature transition to disordered phase III with orthorhombic unit cell belonging
to the Pnma space group. On heating, phase III undergoes a phase transition to a
tetragonal phase II with space group 4̄21m. Phase II, on heating transforms into a
disordered cubic phase I, belonging to the Pm3 m space group with one molecule
per unit cell. Phase I remains stable up to the melting point of ammonium nitrate.
The transition between different phases of AN has important consequences on prac-
tical applications. It was observed that AN composite propellant grain cracks on
storage at room temperature due to IV ! III transition [42]. Many approaches have
been made to overcome the problem of IV ! III phase transition and make the salt
fit for reliable and practical use. One of the approaches is phase stabilization of
ammonium nitrate [43–45]. Studies have shown that doping of AN with metal oxides
4 S. Chaturvedi and P. N. Dave

significantly influences the phase transitions of AN and hence detected clues of a


possible intermediate formation during the solid-state reaction [46]. Phase stabiliza-
tion of AN with metal oxides is beneficial with respect to burning rate, ignition, and
hygroscopicity [47–49].

Thermal Decomposition of AN
Though AN is stable at room temperature, on storage a small amount of ammonia is
evolved, leaving the salt slightly acidic [1,50]. Berthelot [51] formulated the equations
of decomposition and provided values for heat of explosion, heat of formation, tem-
perature of explosion, etc. Due to its application as an explosive component, the
decomposition chemistry of AN has been quite extensively studied [52–63]. As for
most energetic materials, the thermolysis of AN depends upon various factors such
as pressure and temperature and experimental conditions such as sample size, rate of
heating, sample purity, monitoring techniques, and the presence of foreign sub-
stances. Obviously, no single mechanism can explain all of the aspects of its
decomposition characteristics. Some of the possible modes of decomposition path-
ways are provided in Table 2 [64]. It is generally accepted that thermal decompo-
sition is initiated by an endothermic proton transfer reaction as shown in reaction
1 (Table 2). When the salt is heated from 200 to 230 C, exothermic decomposition
occurs (reaction 6). The reaction is rapid but can be controlled and is the basis for
commercial preparation of nitrous oxide (reaction 2). Above 230 C, the decompo-
sition follows reaction 3. The reaction pathway is reported to follow reaction 4 dur-
ing detonation, and reaction 5 has been suggested when AN undergoes explosion.
Most mechanisms proposed for AN decomposition assume the production of
ammonia and nitric acid and subsequent oxidation of NH3 by the decomposition
products of HNO3. Roser et al. [58] assumed the following known equilibrium reac-
tion for HNO3 that leads to the oxidizing species NO 2 and explained the formation
of N2O and water as the main products as in reaction 2 (Table 2).

2HNO3 ¼ NOþ 
2 þ NO3 þ H2 O

NH3 þ NOþ
2 ! products ðN2 O; H2 OÞ

Table 2. Modes of thermal decomposition of AN

Heat
evolved Gas volume Temperature
Reaction (cal=g) (ml=g) at NTP ( C)
(1) NH4NO3 ! NH3(g) þ HNO3(g) 521
(2) NH4NO3 ! N2O þ 2H2O 108 840 320
(3) NH4NO3 ! 3=4N2 þ 1=2NO2 þ 2H2O 316 910 860
(4) NH4NO3 ! N2 þ 2H2O þ l=2O2 354 980 950
(5) 8NH4NO3 ! 5N2 þ 4NO þ 2NO2 201 945 560
þ 16H2O
(6) NH4NO3 ! 1=2N2 þ NO þ 2H2O 86 980 260
Thermal Decomposition of Ammonium Nitrate 5

However, no details of the mechanism were given. Recent in situ decomposition


under rapid heating and immediate analysis of the products, using rapid scan Four-
ier transform infrared (FTIR) spectroscopy, however, clearly revealed the formation
of HNO3 [62] as an intermediate. In addition to N2O and H2O, NO2 was also formed
during the decomposition [62,63]. Furthermore, NO2 is known to be a product in
other modes of decomposition [65]. Because HNO3 is formed only at high tempera-
ture during the decomposition of AN, the following mode of decomposition of
HNO3 was thought to be more realistic [63]:

2HNO3 ¼ NO2 þ H2 O þ 1=2O2

NO2 formed can subsequently oxidize NH3 as follows in the temperature range
342–387 C.

NH3 þ NOþ
2 ! NH2 þ HNO2

NH2 þ NO2 ! NH þ HNO2

NH þ NO2 ! HNO þ NO

NH2 þ NO ! N2 þ H2 O

2HNO ! NO2 þ H2 O

2HNO2 ! NO2 þ H2 O þ NO

leading to an overall stoichiometry:

4NH3 þ 5NO2 ! N2 O þ 2N2 þ 6H2 O þ 3NO

On the other hand, oxygen from the decomposition of nitric acid reacts rather
slowly with NH3, whereas the reaction between oxygen and NO is instantaneous.
It is therefore likely that the involvement of oxygen may be limited to the further
oxidation of NO formed in the above reaction. The subsequent reactions can be
represented by the following equations [63]:

3NO þ 3=2O2 ! NO2

2NO2 þ H2 O ! HNO2 þ HNO3

NH3 þ HNO3 ! NH4 NO3

NH3 þ HNO2 ! NH4 NO2

NH4 NO2 ! N2 þ H2 O

The overall stoichiometry becomes

6NH3 þ 6NO2 þ 3=2O2 ! N2 O þ 7H2 O þ NH4 NO3 þ 3N2 þ 2NO2


6 S. Chaturvedi and P. N. Dave

The above reaction scheme indeed accounts for the observed decomposition pro-
ducts of AN, namely, AN aerosol, NO, N2O, NH3, and H2O, when heated at high
heating rate (80 C=s) and examined in situ by rapid scan FTIR spectroscopy within
the first second [63]. The AN decomposition can thus be represented as follows:

6NH4 NO3 ! N2 O þ 10H2 O þ NH4 NO3 þ 3N2 þ 2NO2

Contrary to the mechanism discussed so far, Davies and Abrahams [66] assumed
that the decomposition of AN proceeds through the formation of a nitramide inter-
mediate.

NH4 NO3 ! NH2 NO2 þ H2 O

NH2 NO2 ! N2 O þ H2 O

Brower et al. [61] decomposed AN in a sealed capillary and examined the pro-
ducts using IR spectroscopy. No NO2 was observed under these conditions. They
suggested that at high temperature a homolytic mechanism for the decomposition
of HNO3 is more likely. The observed activation energy changes continuously from
118 kJ=mol at low temperature to 193 kJ=mol at high temperature, which is nearly
equal to the N–O bond energy in nitric acid. Water and ammonia strongly inhibit
the ionic reaction at low temperature, but the effect fades away at high temperature.
In this mode, the decomposition proceeds through the formation of nitramide,
although the initial step involves a proton transfer. The following mechanism has
been proposed:

NH4 þ N
3 ¼ NH3 þ HONO2

HONO2 ! HO: þ NO2

HO þ NH3 ! HOH þ NH:2

NH2 þ NO:2 ! NH2 NO2

NH2 NO2 ! N2 O þ H2 O

The enthalpy changes and temperatures involved during the various modes of
decomposition are provided in Table 2. The activation energy reported for the ther-
mal decomposition of AN varies from 125.5 to 171.5 kJ=mol [67,68] and is suggested
to be of first order. Because AN is susceptible to considerable evaporation, the val-
idity of these values is in question. It is also worth mentioning that completely dry
AN evaporates after melting and does not decompose [57].
Many of the decomposition studies reported earlier were carried out at low heat-
ing rates. The importance of heating rate was pointed out by Anderson et al. [69,70].
Though surface decomposition occurs at high heating rates, slower heating rate
experiments result in bulk decomposition. The high heating rate surface decompo-
sition is endothermic and produces NH3 and HNO3, whereas a slow heating rate
bulk decomposition is exothermic and produces N2O and H2O. Fast thermolysis stu-
dies of AN are important because they more or less simulate combustion conditions
[62,63,71]. The reaction scheme evolved out of these studies can indeed throw more
Thermal Decomposition of Ammonium Nitrate 7

light on the complex processes during propellant ignition and combustion. In


addition to the decomposition chemistry of pure AN, a number of authors have
examined the effect of addition of various compounds on thermolysis [52,60,
70–90]. The additives studied include ammonium chloride, fluoride, bromide and
iodide, sodium chloride, chromium salts such as potassium chromate and dichro-
mate, chromium oxide, chromium nitrate, pure chromium, cobalt salts, carbon,
metals such as magnesium and zinc, first row transition metal oxides, nitric acid,
organic species, and various other metal cations. The studies are significant in
relation to the hazards associated with the extensive use of AN in bulk as a fertilizer
or oxidizer. In addition to the safety aspects associated with handling and storage,
catalytic decomposition studies are important in the field of AN propellants. Because
AN does not burn easily and requires a catalyst to sustain combustion, some of the
above compounds have been incorporated in AN-based propellant formulations to
catalyze the combustion reaction. Catalysts such as ammonium dichromate, copper
chromate, and chromium oxide are reported to increase the burning rate of AN pro-
pellants [74]. Ammonium dichromate not only facilitates the decomposition of AN
but decomposes itself into water, nitrogen, and chromium oxide and thus is not
solely catalytic in action. Chromium compounds in particular show a greater cata-
lytic action than other cations. It has been suggested that the NO3 in the acidic melt
is deoxygenated to NO 2 by the dichromate ion. NO 
2 further undergoes a number of
reactions with NH3 or NHþ 2 , forming intermediate nitramides or nitrosamines,
which are then decomposed to nitrogen or nitrous oxide and water. The mechanism
of catalysis is, however, debatable [74,76,91–93]. The influence of fuels on the
decomposition of AN was also explored with a view to produce easily ignitable
metallized fuel-rich compositions. The thermal ignition behavior of various mixtures
of organic fuels, magnesium, and AN was studied by Jain and Oommen [82]. It was
observed that the thermal decomposition=ignition of organic fuel–AN mixtures is
modified significantly in the presence of magnesium metal. Many compositions
consisting of organic fuel–AN have been studied [84]. Some of these include
ammonium oxalate and picrates, 2,4-dinitro-resorcinol, glucose, guanidine nitrate,
pitch, bitumen, and styrene polymers. With a suitable choice of fuels and proper
composition it has been possible to make cast compositions that yield product gases
in a wide range of temperature, from about 360 to 2000 C [94]. The surface tempera-
ture of burning AN and gas-phase AN flame have been a subject of many investiga-
tions [95]. The surface temperature was reported to be around 300 C. Because
the burning rate of AN and AN propellants are of similar order it is assumed
that the decomposition of AN in a composite propellant is similar to that of a
monopropellant.

AN as a Propellant Oxidizer
The composite solid propellant essentially consists of an oxidizer powder, loaded up
to 85–90%, held together in a matrix of polymeric binder, which acts as a fuel. Most
solid rocket motors currently in use, including the space shuttle’s solid rocket boos-
ters (SRBs), use AP as the main solid oxidizer. One of the primary products of com-
bustion of AP-based propellants is HCl, which in the presence of water forms
hydrochloric acid, which produces smoke and is highly toxic. When the space shuttle
is launched, each of its SRBs produces over 100 tons of HCl [95,96]. It is estimated
that in the initial 120 s of flight the SRBs burn 502 tons of propellant. The resulting
8 S. Chaturvedi and P. N. Dave

residue is 21% HCl and 30.2% Al2O3. Al2O3 is a chemically inert powder but HCl
alters its surface properties so that cleanup and corrosion problems are more severe
than expected. Similarly, Ariane 5, with its two solid fuel boosters each holding
237 tons of propellant, during the 2 metric tons of burning produces 156 tons of
Al2O3 and 96 tons of HCl, all of which falls on the ocean and forest around the
launch pad [97]. The impact of HCl pollution has been established on tests conduc-
ted on animals and vegetation around the launch pad. Though some of these effects
are short lived, the cumulative effect of these acid shocks due to frequent launching
of rockets is a serious concern. Yet another anticipated threat from the chlorine com-
pounds emitted by AP propellants is to the stratospheric ozone layer and global
warming. It is likely that hydroxyl radicals could react with HCl to release atomic
chlorine, which destroys ozone by reducing it to oxygen. To overcome these environ-
mental problems it is necessary to investigate the use of so-called clean burning pro-
pellants that eliminate chlorine as an ingredient. Less than 1.0% by weight HCl in the
SRB exhaust is a common goal in the solid rocket industry. An additional motiv-
ation for developing chlorine-free propellants is the need for less harmful and
eco-friendly (minimum smoke) propellants. Minimum smoke propellants are those
that eliminate both primary smoke (e.g., Al2O3) and secondary smoke (i.e., water
aerosol condensed from the atmosphere) from the exhaust plume [95]. They are
important in missile systems because detection of military launch locations is an area
of primary concern. Because HCl serves as a nucleation site for water vapor, chlorine
must either be eliminated from the propellant or scavenged from the exhaust
products in order to eliminate secondary smoke. Yet another use of chlorine-free
propellants is in gas generators. Solid-propellant gas generators are used to start
and drive turbines for power generation. For such a purpose, the propellants have
to burn with a low flame temperature and the combustion gases should not contain
corrosive constituents such as chlorine or chlorine compounds. In addition, the
amount of solid particles in the combustion gases should be kept as low as possible
[98]. AN is one of the most appropriate oxidizer for propellant compositions that
meet the above requirements. Though the experiments on AN as a high-performance
oxidizer for SRB in place of AP is are in their infancy, gas generator compositions
have established its use as an efficient oxidizer. Both double-base and composite
(without metal) propellants are used extensively with AN or as a mixture of AN
and guanidine nitrate [99] in gas generators. AN composite propellants provide
desired exhaust properties such as (a) high nitrogen content (55 wt%), (b) low water
content (10 wt%), (c) modest amounts of solid particles (8 to 12 wt%), and (d) a rela-
tively nonhazardous exhaust. Further, AN propellant formulations are relatively
insensitive to temperature and impact and have good strength properties over a wide
range of temperature. AN-based systems have several positive features including
clean burning and smokeless exhaust. They can be used as a substitution for AP.
Though AN-based systems have several positive features such as clean burning
and smokeless exhaust they are not free from drawbacks, and substitution of AP
with AN for high-performance systems is not straightforward. The main problems
associated with AN in its use as an oxidizer in propellants include (a) phase trans-
formation around room temperature (32 C), which is accompanied by significant
volume expansion that results in crack formation in the propellant grain; (b) hygro-
scopicity; (c) low burning rate; and (d) low energy. Many attempts have been made
to overcome these problems and to realize a better phase transition and combustion
behavior for AN-based propellants.
Thermal Decomposition of Ammonium Nitrate 9

AN-based composite propellants are attractive primarily due to the clean burn-
ing nature of AN as an oxidizer. However, this propellant has some disadvantages
such as poor ignition and low burning rate. AP-based composite propellants have
excellent ignition and burning characteristics, although the combustion gases include
HCl. It is expected that an AN=AP-based propellant would have acceptable perfor-
mance for practical applications because the disadvantages of the AN-based propel-
lant would be compensated for by the advantages of the AP-based propellant. The
burning rate of the AP-based propellant depends on the AP particle size and the
burning characteristics of the AN=AP-based propellant would be influenced by
the AP particle size. Kohga et al. [100] investigated the burning characteristics of
AN=AP-based propellants with fine AP (3 mm) and coarse AP (180 mm). The thermal
decomposition behaviors of AN=AP propellants were independent of the AP particle
size. The burning rate characteristics of AN=AP propellants were remarkably affec-
ted by the particle size and content of AP. Some of the AN=AP propellants prepared
with fine AP self-quenched or burned improperly: partial flashing flame and flame
propagation along the side of the propellant. Furthermore, the cause of these
unstable combustions was revealed.
The mechanical strength of AN prills is dependent on the phase transition beha-
vior. The technical significance hence attached to the thermal transitions is a prime
reason for the extensive literature available in this area. Many of the studies on the
phase transformation support the generally accepted idea that the mechanical strength
of the prill should be related to the AN(IV)–AN(III) polymorphic transition at 32 C,
which is accompanied by a sensible volume change (3.8%). AN prill is a polymorphic
aggregate. The volume change of the elementary cell causes a change in the crystal
packing in the polycrystalline aggregate and, consequently, the adherence between
crystallites decreases. The product’s mechanical strength thus decreases after several
thermal cycles through the transition temperature and the granular product breaks
down into dust or fine particles and the caking tendency is increased. Attempts to
break up the caked salt by explosives have resulted in serious accidents. Implications
of the volume change associated with the transition on the efficient use of AN as a pro-
pellant oxidizer are even more serious. Because the composite propellant grain, which
contains oxidizer (AN) of usually about 85% by weight, can be subjected to tempera-
tures that cause the IV–III transition during storage or transportation, the volumetric
changes associated with the phase change may lead to defects (cracks) in the propellant
grain. Cracks in a propellant grain are undesirable because they can lead to cata-
strophic burning behavior. It is therefore imperative to understand the phase tran-
sition fully and modify it suitably to make the salt fit for reliable and practical use [1].
Several additive such as oxides, inorganic acids, etc., have been reported to have
a phase stabilization effect on the effectiveness of AN as an oxidizer. It is known that
transition metals=metal compounds are capable of increasing the burning rate and
ignitability of AN propellants. Another welcome side effect of phase-stabilized
ammonium nitrate (PSAN) is the reduced hygroscopicity. Therefore, phase stabiliza-
tion of AN with metal oxide is also beneficial with respect to burning rate, ignition,
and hygroscopicity.

Effect of Additive on the Thermal Decomposition of AN


The need for chlorine-free and minimum smoke propellants has already been dis-
cussed. Realizing the significance of chlorine-free propellants, many attempts have
10 S. Chaturvedi and P. N. Dave

been made to develop clean burning propellants [101–125]. Many of these attempts
are based on AN as the oxidizer. Several studied have been conducted with diff-
erent modifiers to enhance the thermal decomposition of pure AN and AN-based
propellants.
Sun et al. [126] conducted two types of experiments. In the first type, an AN iso-
thermal experiment was executed to confirm the reaction order of the decomposition
reaction of AN and its mixtures with inorganic acid. In the second type, constant
temperature rate increasing experiments were performed to study the heat behavior
of the whole reaction or decomposition process, and the data were used to calculate
the activation energy and preexponential factor of the reaction. Figure 1 shows the
heat flux curves of pure AN and its mixtures with different masses of sulfuric acid.
The curves in Fig. 1 demonstrate that there are three endothermic peaks on the curve
of sample 1 (the heat flow curve of pure ammonium nitrate). The endothermic peaks
at 86 and 127 C are the first and second crystalloid change of ammonium nitrate,
respectively, and the endothermic peak at 167 C is the melting of ammonium nitrate.
In the exothermic peak, the heat flow increased very slowly and gradually with the
temperature increase in the range of 190–232 C and increased sharply above 232 C.
For the heat flow curve of sample 2, the first two endothermic peaks are the same as
those in sample 1. However, the decomposition reaction started before the melting
peak of pure AN. The heat flow increased gradually with the temperature increase
in the range of 145–207 C and increased sharply above 207 C. For the heat flow
curve of sample 3, only one endothermic peak exists on the curve at 86 C. At the
exothermic peak of decomposition, heat flow increased gradually with the tempera-
ture increase in the range of 135–188 C and increased sharply above 188 C. It is
obvious that the decomposition of ammonium nitrate starts at lower temperature
when sulfuric acid is present. Figure 2 shows the measured heat flux curves of pure
AN and its mixtures with different masses of hydrochloric acid. It shows that the
heat flux curve of sample 4 is similar to that of sample 5 but much different from
that of pure AN. According to the results, thermal decomposition of AN starts at

Figure 1. Heat flux curve of pure AN and its mixture with sulfuric acid, heating rate
0.2 C=min; sample mass 2.0 g.
Thermal Decomposition of Ammonium Nitrate 11

Figure 2. Heat flux curve of pure AN and its mixture with hydrochloric acid, heating rate
0.2 C=min; sample mass 1.0 g (color figure availble online).

a much lower temperature (128 C) as compared to pure AN and AN with sulfuric


acid (Fig. 1). The heat fluxes of the samples 4 and 5 increased very sharply with
an increase in temperature, and the temperature range of their decomposition reac-
tion was much narrower than that of pure AN, which means that when AN is mixed
with hydrochloric acid, it will decompose more violently.
Figure 3 shows the measured pressure curves of samples 1, 2, and 4 during the
decomposition processes at a constant temperature rate of 0.1 C=min from room
temperature to 300 C. It is clear from the experimental results that the pressure
increased when the temperature increased and the rates of pressure increase for each

Figure 3. Measured pressure curves of sample during their decomposition process, heating
rate 0.1 C=min.
12 S. Chaturvedi and P. N. Dave

sample were much different, but the measured pressures for each sample were almost
the same when the temperature was below 75 C or higher than 250 C. Although the
sulfuric and hydrochloric acid can catalyze the decomposition reaction of AN and
decrease the onset temperature of AN decomposition, they cannot change the
decomposition products of AN.
The above-mentioned results show that both sulfuric and hydrochloric acid can
catalyze the AN decomposition reaction. However, the catalysis mechanisms of vari-
ous substances are different. A conceivable explanation of the catalysis mechanism
of sulfuric acid given in Figure 4 [127] shows that sulfuric acid can replace the
NO 3 ion in AN, thus promoting the buildup of nitric acid, which will increase the
decomposition of AN. Because this process is the first step of AN decomposition,
which occurs at relatively low temperature, the heat release curve starts at a lower
temperature than that of pure AN. This mechanism is also applicable to other strong
acids such as nitric acid, etc. [127].
An explanation of the catalysis mechanism of the Cl ion to the AN decompo-
sition in acidic conditions [128] shows that during the process of pure AN decompo-
sition, Eq. (1) is the controlling step because of its slow reaction rate.

NH4 NO3 ¼ N2 O þ 2H2 O

When mixed with hydrochloric acid, this step is substituted by Eq. (2), the ion
Cl and Hþ activate NOþ 2 and NH3, thus increasing the generation rate of the inter-
mediate product [NH3 NOþ 2 ] at lower temperature. The activation energy of the reac-
tion ( ) is decreased, giving priority to the catalysis mechanism of Cl for the AN
decomposition in acidic conditions [128].
Keskar et al. [129] showed that the thermogravimetric curve (TG) (Fig. 6) of a
reaction mixture of UO2 þ 0.17 NH4NO3 in air shows a weight loss of 3.60% in the
temperature range of 160–215 C followed by a weight gain of 2.1% in the tempera-
ture range of 225–400 C. In the UO2 þ 0.17 NH4NO3 reaction, the decomposition of
only NH4NO3 (0.17 mol) was expected to show a weight loss of 4.79% with an endo-
thermic differential thermal analysis (DTA) peak, but the observed weight loss was
3.60%. This could be due to the partial oxidation of UO2 to UO2.2 (the expected total
weight change for simultaneous decomposition of NH4NO3 and oxidation of UO2 to
UO2.2 is 3.63%). The weight loss was associated with an exothermic DTA peak at
200 C. The exothermic nature of the DTA peak also substantiates the oxidation
of UO2 during decomposition of NH4NO3. The weight loss was followed by a weight

Figure 4. (a) Sample 1, 210 C.


Thermal Decomposition of Ammonium Nitrate 13

gain in the temperature range of 225–400 C associated with an exothermic DTA


peak at 350 C. This exothermic DTA peak was due to the oxidation of UO2.2 to
U3O8. In addition, the DTA curve showed endothermic peaks at 125 and 170 C
due to the phase transition and melting of NH4NO3, respectively. It is well known
that when heated in air, UO2 undergoes oxidation to form U3O7 as an intermediate
oxide at 350 C and on further heating up to 600 C gives U3O8 as the end product.
To understand the role of oxygen from air in the absence of ammonium nitrate, the
oxidation of UO2 was also carried out in a thermoanalyzer in a stream of dry air,
heating UO2 up to 600 C, and the resultant TG curve is shown in Fig. 4 for compari-
son. From the TG curves of UO2 þ 0.17 NH4NO3 and oxidation of UO2 in air, it was
confirmed that the second step of oxidation (225–400 C) in the former reaction is
due to the oxidation of UO2.2 formed during the decomposition of NH4NO3. It
was confirmed that when heated with NH4NO3, the oxidation of UO2 toU3O8
was complete at much lower temperature (400 C) than in air oxidation (600 C).
TG curves of reaction mixtures of UO2 with 0.33 and 0.85 mol of NH4NO3 also
showed weight loss followed by a weight gain similar to the TG curve of
UO2 þ 0.17 NH4NO3.
To determine the effect of various anions on the thermal stability of AN, it was
necessary to introduce the additives with inert counterions. When sodium nitrate was
added it was found to have no effect on the rate of ammonium nitrate decompo-
sition; therefore, the sodium ion was considered an inert counterion. The ammonium
ion, however, was not inert. When AN formulations were spiked with a variety of
sodium or ammonium salts (SOþ  2 
4 ; BF4 ; B4 O7 ; HCO3 ), sodium salt–containing
AN formulations were slightly more stable than ammonium salt–containing AN for-
mulations. It was concluded that the added ammonium cations accelerate the
decomposition of AN formulations, although this destabilization may not always
be evident if the anion is stabilizing. Although this conclusion is based on only a
slight difference in the effect of the ammonium and sodium salts, it is concurrent
with the fact that the ammonium ion is a weak base, and the salts of weak bases acid-
ify aqueous solutions:

NHþ
4 þ H2 O ! NH4 OH þ H
þ

In acidifying the solution, added NHþ 4 destabilizes AN formulations. Con-


versely, salts of weak acids should give basic aqueous solutions and stabilize AN.
Sulfate, the anion of the weak acid bisulfate, tended to retard AN decomposition.
It was found that a number of sodium salts of weak acids (carbonic, acetic, formic,
oxalic, and hydrofluoric), when added to AN or ANFO, retarded decomposition.
The salts of strong acids, which leave aqueous solutions neutral, did not affect the
rate of AN decomposition. Salts of weak acids and urea stabilized AN formulations,
even in the presence of destabilizing species such as iron salts. A study of the effect of
various cations on AN thermal stability began with the iron salts that might possibly
be present in mining operations. The Bureau of Mines reported that pyrite and its
weathering product FeSO4 accelerate the decomposition of AN and ANFO
[130,131]. Oxley et al. [132] examined the effect of a number of iron-containing addi-
tives. AN, ANFO, and AN emulsion were mixed with 5 wt% of various
iron-containing salts: Fe(NO3)3, Fe2(SO4)3, FeSO4, FeS, and FeS2 (in ore). All accel-
erated the decomposition of AN, ANFO, and AN emulsion at 270 C. To systemati-
cally study the effect of cations on AN stability, it was necessary to select
14 S. Chaturvedi and P. N. Dave

appropriate counterions. Nitrate, the anion of a strong acid, was chosen as an inert
anion; sulfate, as the anion of a weak acid, was expected to have a stabilizing effect;
oxides were examined because they are ubiquitous. The thermal stability of AN with
most common metals was assessed by differential scanning calorimetry (DSC). As
expected, metal sulfates and oxides tended to be more thermally stable than the
nitrates. Most oxides were water insoluble, and those that were soluble were basic.
The exception was Cr2O3, which was the only oxide that had a detrimental effect
on the thermal stability of AN.
If an oxidation=reduction cycle is important in promoting AN decomposition,
then metals with multiple oxidation states should be most destabilizing. This was
not the case. Cobalt and copper nitrate exhibited no strong destabilizing effect on
ammonium nitrate, even though both commonly use two oxidation states
(Co2þ=Co3þ and Cuþ=Cu2þ). However, aluminum, which has only one oxidation
state, A1þ3 , had a strong destabilizing effect on AN. Because both Cr

and A13þ
had such a strong destabilizing effect, this suggested that they have some common
property that promotes decomposition. Chromium nitrate, aluminum nitrate, and
iron nitrate, all of which strongly promote AN thermal decomposition, have two
properties in common: they were the only 3þ oxidation state metal salts examined,
and they all have a rather small ionic radius.
Gunawan and Zhang [133] demonstrated that the mass loss and heat flow of
pure ammonium nitrate and a mixture of ammonium nitrate and alumina are very
similar, indicating that the type of crucible (platinum or alumina) has no significant
effect on ammonium nitrate decomposition. Moreover, the activation energy values
calculated using a set of heating rates of 0.5, 1, 2.5, and 10 C min1 for the decompo-
sition of 15 mg of pure ammonium nitrate using the platinum crucible with and
without added alumina powder were 93.7 and 94.1 kJ mol1, respectively. Therefore,
the difference between the activation energy values obtained with the two types of
crucibles was also insignificant.
A mixture of 5 g ammonium nitrate was mixed with 15 g pyrite; the mixture was
placed in the reactor and argon was flowed to create an inert atmosphere. The mixture
was then heated to 105 C and the temperature and concentration of gases were mon-
itored. A set temperature of 105 C was selected based on the prediction of critical
temperature calculated for the reactor diameter of 21 mm for which the critical tem-
perature was predicted to be 105.9 C. Figure 5 shows the results of this verification
experiment using the gas-sealed isothermal reactor. It is interesting to see that,
although the onset temperature of thermal runaway was around 110 C, it was evident
at 105 C. NO gas was emitted, indicating that the ammonium nitrate decomposition
in the presence of the pyrite had started. Additional experiments were conducted at
lower temperatures, 90, 95, and 100 C. It was found that the mixture did not react
when held at those temperatures for up to 24 h. A further experiment was then con-
ducted by heating the ammonium nitrate–pyrite mixture to 100 C and holding for
about 400 min. No reaction was observed. The temperature was then increased again
to 103 C, and at this temperature thermal runaway occurred after 350 min, as
shown in Fig. 8. A close inspection of the temperature and gas emission profiles
(Fig. 6) revealed that NO gas was formed and emitted at around 105.5 C. It is inter-
esting to note that both NO and SO2 were detected from the isothermal reactor when
ammonium nitrate was mixed with the pyrite, though it is known that endothermic
decomposition of ammonium nitrate does not emit SO2 [134–139]. It is clear that both
NO and SO2 are produced from the oxidation of pyrite by ammonium nitrate [140].
Thermal Decomposition of Ammonium Nitrate 15

Figure 5.

The activation energy and preexponential factor of ammonium nitrate decompo-


sition were determined to be 102.6 kJ mol1 and 4.55  107 s1 in the absence of
pyrite and 101.8 kJ mol1 and 2.57  109 s1 in the presence of pyrite. The critical
temperatures for ammonium nitrate decomposition, with and without pyrite, were
shown to decrease with increasing diameter of the blast holes charged with the explo-
sives. The presence of pyrite reduces the temperature and accelerates the rate of
decomposition of ammonium nitrate. It has been further shown that pyrite can sig-
nificantly decrease the critical temperature of ammonium nitrate decomposition,
potentially causing undesired premature detonation of the explosives in rock blasting
practices using ANFO in mining operations, especially in reactive mining grounds
containing pyrite minerals. For a typical blast hole diameter of 200 mm, the presence
of pyritic reactive shale has been shown to drastically decrease the critical tempera-
ture from 98.6 to 57.7 C. A concept of using the critical temperature as an indication
of the thermal stability of the explosives which was reduced by decreasing critical
temperature. The results of this study have significant implications for rock blasting

Figure 6. (A) TG; (B) DTA curves for mixture of UO2 and NH4NO3 in a 1:0.17 molar ratio in
air; and (C) TG curve of UO2 in air.
16 S. Chaturvedi and P. N. Dave

practices using ANFO in mining operations, especially in reactive mining grounds,


because pyrite minerals present in some of the reactive ground have been shown
to significantly reduce the thermal stability of ammonium nitrate.
Skaribas et al. [141] investigated AN samples containing varying amounts of
Cr3þ cations and their mode of thermal decomposition. These were analyzed under
nonisothermal conditions in a thermogravimetric balance. Pure AN decomposes
endothermically, but the addition of Cr3þ in amounts greater than one mole of
Cr3þ per 3,000 mol of AN acts catalytically, changing the decomposition to an exo-
thermic mode. At lower concentrations, the diffusion limitations prevent the conden-
sation of Cr3þ species into Cr2 O 7 , which are the catalytically active species. The
rates of decomposition, described by a typical Coats-Redfern kinetic equation
[142], show two stages. Satisfactory results were obtained using a modified
Coats-Redfern equation and by taking into account the continuous increment of
chromium catalyst in the melt. The second stage of decomposition showed an
increased activation energy compared with the first. This was probably caused by
a change in the rate-determining step of the reaction path, which was most likely
the decomposition of di-chromates.
Skordilis and Pomonis [143] thermogravimetrically studied the decomposition of
ammonium nitrate in mixtures with varying amounts of metal cation M (where M is
Mn, Co, Cu). The ratios studied were NH4NO3: M ¼ 30,000:1, 15,000:1, 5,000:1,
1,500:1, 150:1, 25:1, and 10:1 in order to identify the threshold limits for the inver-
sion from an endothermic to an exothermic decomposition. This transition takes
place at NH4NO3:M ¼ 150:1. Similar mixtures containing Al in ratios
NH4NO3:Al:M ¼ 15,000:5:1, 500:165:1, 300:100:1, 150:50:1, 15:5:1, and 8:2:1 were
also studied by differential thermal analysis (DTA)=and differential thermal gravi-
metric (DTG) with respect to their thermochemical behavior and the influence of
the mode of decomposition on the final specific surface area (Sw, m2g1) of the
remaining product.
The addition of cations of transition metals Mn, Co, or Cu to NH4NO3 resulted
in an inversion of the endothermic decomposition (samples M-n with n > 4) to an
exothermic decomposition. This inversion of the thermal effect was gradual with
increasing amounts of cations and the turning point was to be around sample M-5
where NH3NO4:M ¼ 150:1. Some small differences were apparent for the three
cations. For example, copper produced a sharper exothermic decomposition; this
was also true in the case of manganese. However, cobalt produced a less violent
decomposition, although it occurred at lower temperatures than with Mn and Cu.
The transition point from the endo- to the exothermic decomposition is depicted
in Fig. 2. From sample (NH4NO3:M ¼ 150:1), the exo-effect starts to become appar-
ent. In sample (NH3NO4: M ¼ 25:1), the endo-effects cease to be apparent, whereas
they are stronger and are the only ones present in sample (NH3NO4:M ¼ 10:1).
Copper phase-stabilized ammonium nitrate (Cu-PSAN) was prepared by the
incorporation of copper (II) oxide (CuO) in the ammonium nitrate (AN) crystal lat-
tice by a melt process. Remya Sudhakar and Mathew [39] studied the phase transi-
tions of ammonium nitrate and dimensional changes occurring during this phase
stabilization process. Crystallographic aspects of the solid-state reaction between
AN and CuO and the thermal behavior of phase-stabilized ammonium nitrate
(PSAN) were discussed. It was found that the solid-state reaction occurred by an
intermediate solid solution formation. Lattice parameters and unit cell volume were
calculated as a function of temperature from XRD patterns, which showed a
Thermal Decomposition of Ammonium Nitrate 17

Figure 7. Temperature and gas profiles of the mixture of 5 g ammonium nitrate and 15 g pyrite
heated up to 105 C in the isothermal gas-sealed reactor with argon as a carrier gas

nonlinear and anisotropic behavior with increasing temperature. Lattice expansion


along the b-axis and contraction along the a- and c-axes were observed. Unit cell vol-
ume increased with temperature. The DSC curve of Cu-PSAN is shown in Fig. 8.
The first two DSC endothermic peaks near 86.7 and 125.3 C were due to phase tran-
sitions III ! II and II ! I, respectively. From the curves it is clear that the room tem-
perature transition, viz. IV ! III was completely absent in Cu-PSAN. This shows
that Cu-PSAN is phase stabilized. The endothermic peak corresponding to the melt-
ing occurred near 175.5 C, after which an endothermic decomposition of PSAN

Figure 8. Temperature and gas profiles of the mixture of 5 g ammonium nitrate and 15 g pyrite
heated up to 100 C and then to 103 C in the isothermal gas-sealed reactor with argon as a
carrier gas.
18 S. Chaturvedi and P. N. Dave

Figure 9. DSC curve of pure AN.

occurred with Tmax ¼ 262.9 C. Further decomposition was an in situ oxidation by


the nitrate groups of PSAN as seen from the exotherm with Tmax ¼ 280 C.
On comparison of the DSC pattern of pure AN (Fig. 9) with that of Cu-PSAN
(Fig. 10) it is clear that Cu-PSAN is stabilized. As can be seen from the DSC trace of
Cu-PSAN, there was no endothermic peak below 50 C. The first endothermic peak
occurred only at 86.7 C. This drastic shift of the phase transition temperature to the
higher side, from room temperature regime to 86.7 C, clearly indicates that CuO
doping is effective for the phase stabilization of AN phase IV. Copper oxide signifi-
cantly influences the transition temperature of AN by the formation of a solid sol-
ution of copper (II) diammine dinitrate in the AN crystal lattice. The reaction
proceeded as the equation [144]:

2NH4 NO3 þ CuO ! ½CuðNH3 Þ2 ðNO3 Þ2 þ H2 O

Figure 10. DSC curve of Cu-PSAN.


Thermal Decomposition of Ammonium Nitrate 19

Farhat et al. [145] studied the decomposition of aqueous NH4 NO3 (50–55 wt%)
in the presence of various mono- and bimetallic catalysts. Investigations were per-
formed using thermal analysis (DTA-TGA), a batch reactor, and a dynamic flow
reactor with on-line mass spectrometry (MS) analysis. The first results were as fol-
lows: heating in the absence of catalyst endothermic peak observed due to vaporiza-
tion of water and evidence showed that AN dissociated into ammonia and nitric acid
even at high temperature. In the presence of catalysts, a dramatic change was
observed. All monometallic catalysts (supported Pt, Fe, Cu, or Zn) presented a true
catalytic decomposition reaction linked to exothermic peaks. However, only
Pt-based catalyst was able to trigger decomposition at lower temperature (210 C).
Nevertheless, tests in the batch reactor revealed mediocre results associated with very
slow AN decomposition. Bimetallic catalysts M-M0 =Al2O3-Si (M, M0 ¼ Fe, Cu, Zn,
Pt) were also evaluated. The catalytic decomposition depended on the active phase
and on the preparation method of the bimetallic catalysts. The addition of zinc or
copper to non-reduced platinum catalyst (PtCuAl-NR and PtZnAl-NR) increased
the catalytic effect of platinum and the results displayed a beneficial effect by violent
one-step decomposition. On the contrary, the addition of zinc or copper to reduced
platinum metal led to less active catalysts and the decomposition occurred in two
steps. This difference in activity could be mainly related to the formation zinc–plati-
num and copper–platinum alloys when adding the second metallic precursor to
non-reduced platinum, followed by a final reduction. Regardless of the catalyst
(PtAl, PtCuAl-NR, or PtZnAl-NR) used, the results obtained using a dynamic reac-
tor revealed the presence of the same gaseous and condensed products—major nitro-
gen and nitric acid and no oxygen—and the formation of nitrogen oxides NO and
N2O depended on the catalyst’s nature: minor for PtAl and medium for PtCuAl-NR
and PtZnAl-NR.
Kaljuvee et al. [146] clarified the influence of different lime-containing sub-
stances—mainly Estonian limestone and dolomite—as internal additives on the ther-
mal behavior of AN. Commercial fertilizer-grade AN was investigated. The amount
of additives used was 5, 10, or 20 mass%, or calculated as a mole ratio of AN=(CaO,
MgO) ¼ 2:1 in the blends. Experiments were carried out under dynamic heating con-
ditions up to 900 C (10 C min1) in a stream of dry air or N2 using Setaram Labsys
2000 equipment coupled to an FTIR. The results of analyses of the gaseous com-
pounds evolved at thermal decomposition of neat AN are different decomposition
occurs in air or in N2 atmosphere. In the thermal treatment of blends of AN with
CaCO3, MgCO3, limestone, and dolomite samples the decomposition of AN pro-
ceeded through a completely different mechanism, depending on the origin and
the content of additives, partially or completely, through the formation of Mg(NO3)2
and Ca(NO3)2.
Li and Hui [147] studied the crystal transformation and thermal decomposition
of AN with additives using DSC. The method and mechanism of enhancing the
stability of AN under normal temperature and stimulating its thermal activity
under high temperature were investigated. The experimental results proved that
some organic additives had a stabilizing effect on AN crystal phase, which was indi-
cated by the decreasing thermal effect of crystal transformation. The additives,
including sawdust, cotton fiber, phosphoric acid, potassium chloride, and potass-
ium bichromate, accelerated the thermal decomposition of AN under high tempera-
ture and exothermic temperature observed at low temperature and energy release
more abruptly.
20 S. Chaturvedi and P. N. Dave

Oxley et al. [148] evaluated the thermal stability of ammonium nitrate water-in-
oil emulsions. The thermal decomposition kinetics and resultant products were
examined. As a baseline, the thermal stability of ammonium nitrate with individual
components of the emulsion was determined. Only mineral oil had any effect on the
decomposition. Generally, ammonium nitrate mixed with hydrocarbons has
enhanced thermal stability. However, ammonium nitrate mixed, rather than emulsi-
fied, with mineral oil can decompose along a lower energy pathway than pure
ammonium nitrate. The extent of decomposition along that pathway is not great
before the decomposition process results in a buildup of ammonia and, thus, in
termination of that pathway. Mineral oil appears to be unique among the hydro-
carbons in its ability to destabilize ammonium nitrate. The experiments utilized
DSC in combination with conventional isothermal techniques.
Rubtsov et al. [149] investigated the kinetics of heat release in the interaction of
AN with different organic compounds using an automatic differential calorimeter.
The factors controlling the decomposition rate of AN=organic compound composi-
tions included an acidic oxidizing medium, which is characteristic for AN due to
reversible dissociation into initial ammonia and nitric acid; the greater volatility of
NH3 in contrast to HNO3; positive oxygen balance; the intrinsic thermal stability
of an organic component; the capability to develop liquid eutectics; and the possibility
of reactions of one component with decomposition products of other components.
Zhang and Zou [150] suggested that ignition and combustion difficulties with
AN-based propellant can be resolved by lowering the AN decomposition tempera-
ture and accelerating the thermal decomposition of AN. The effects of metal oxide,
organic metal salt (OME), energetic binder, and hexanitrohexaazaisowurtzitane
(HNIW) on the thermal decomposition of AN were studied by DSC and TG-DTG.
The results showed that catalysts lowered the endothermic decomposition tempera-
ture of AN to a certain extent and energetic components accelerated the exothermic
decomposition process of AN.
Simoes et al. [151] studied the simultaneous thermal analysis (DSC-TG) used in
the thermal characterization of a specific type of PSAN containing 1.0% of NiO (sta-
bilizing agent) and 0.5% of Petro (anti-caking agent) as additives. Repeated runs
covering the nominal heating rate range from 2.5 to 20 C=min1 showed the good
reproducibility of the main features of the thermal behavior of PSAN; that is, two
phase transitions, melting, and decomposition. Nonisothermal kinetic analysis was
used to estimate the Arrhenius parameters of the decomposition process by applying
both a single curve method and two isoconversion methods to the TG data. The
results were discussed considering the range of applicability of the methods as well
as the influence of the experimental conditions and=or techniques in the kinetic
analysis results in a broader sense. A systematic approach based on the results of iso-
conversion methods and statistical tools was adopted to obtain reliable estimates of
the Arrhenius parameters for the thermal decomposition of PSAN. Under the con-
dition of study, an activation energy of 81.4  3.5 kJ mol1 and preexponential factor
of (5.63  0.01)  105 s1 were found to describe the process over a wide range of the
kinetic analysis.

Conclusion
Both DSC and isothermal results showed a general and strong correlation between
the basicity of an additive and its ability to stabilize AN to thermal decomposition.
Thermal Decomposition of Ammonium Nitrate 21

Interestingly, DSC scans of AN more closely reflect isothermal studies at 260 C than
those at 320 C. In fact, additives had a greater effect on the low-temperature
decomposition of AN than they did on high-temperature decomposition of AN. A
number of additives were identified that dramatically increased the temperature of
the AN DSC exotherm. In general, an increase in the DSC exothermic peak
maximum signaled a decrease in the rate of AN decomposition, but an increase in
the exotherm had to be greater than 20 C to have a notable effect on the decompo-
sition rate.
Though this study has successfully identified several additives that stabilize AN
and might be acceptable in fertilizer formulations, these formulations have not been
shown to be non-detonatable. The small-scale explosivity device (SSED) clearly dif-
ferentiated between fueled and unfueled AN formulations in terms of explosivity,
and the device detected the decrease in detonability accompanying dilution of AN
from 100 to 90, 67, 50, 30, and 10%. However, within a given dilution of AN there
was little range to distinguish whether a stabilizing additive can also reduce deton-
ability. At this point, large-scale testing of some of the best formulations is necessary
to determine the failure diameter.
AN-based composite propellants are attractive primarily due to the clean burn-
ing nature of AN as an oxidizer. However, this propellant has some disadvantages
such as poor ignition and low burning rate. AP-based composite propellants have
excellent ignition and burning characteristics, although the combustion gases include
HCl. It is expected that an AN=AP-based propellant would have an acceptable per-
formance for practical applications because the disadvantages of the AN-based pro-
pellant would be compensated for by the advantages of the AP-based propellant.

Acknowledgments
The authors are grateful to the Chemistry Department of KSKV University, Bhuj,
for laboratory facilities. One of the authors (S.C.) is also thankful to CSIR for a
Research Associate (RA) fellowship.

References
[1] Oommen, C. and S. R. Jain. 1999. Journal of Hazardous Materials, 67: 253–281.
[2] Kohga, M. and Y. Hagihara. 1998. Propellants, Explosives, Pyrotechnics, 23: 182–187.
[3] Sinditiskii, V. P., V. Y. Egorshev, A. I. Levshenkob, and V. V. Serushkin. 2005. Propel-
lants, Explosives, Pyrotechnics, 30: 269–280.
[4] Singh, G. and S. F. Prem. 2003. Combustion and Flame, 135: 145–150.
[5] Othmer, K. 1978. Encyclopedia of Chemical Technology, 3rd edition. New York: Wiley.
[6] Othmer, K. 1992. Encyclopedia of Chemical Technology. Vol. 2, 4th ed. New York:
Wiley.
[7] Ullmann’s Encyclopedia of Industrial Chemistry, Vol. A2, 5th ed. Weinheim, Germany:
Wiley VCH.
[8] Sutton, G. P. 1986. Rocket Propulsion Elements. 5th ed. New York: Wiley.
[9] Rao, C. N. R. 1978. Phase Transitions in Solids. New York: McGraw-Hill.
[10] Davey, R. J., P. D. Guy, and A. J. Ruddick. 1985. Journal of Colloid and Interface
Science, 108: 189.
[11] Carrie, M. T. and D. W. James. 1986. Australian Journal of Chemistry, 39: 771.
[12] Rilo, I. P., I. P. Klyus, and I. G. Skrypnik. 1975. Zh. Prikl. Khim., 48(6): 1192.
[13] Sjolin, C. 1971. Journal of Agricultural and Food Chemistry, 19: 83.
22 S. Chaturvedi and P. N. Dave

[14] Novikova, O. S., I. V. Tsekanskaya, V. I. Titova, and T. I. Gantarenko. 1977. Zh. Fiz.
Khim., 51: 257.
[15] Filipescu, L., D. Fatu, T. Coseac, M. Mocioi, and E. Segal. 1986. Thermochimica Acta,
97: 229.
[16] Kolaczkowski, A. and A. Biskupski. 1981. Journal of Chemical Technology and Biotech-
nology, 31: 424.
[17] Kolaczkowski, A., A. Biskupski, and J. Schrveder. 1981. Journal of Chemical Tech-
nology and Biotechnology, 31: 327.
[18] Bennet, D. 1972. Journal of Applied Chemistry and Biotechnology, 22: 973.
[19] Rubstov, Y. I., A. Rubstov, A. I. Kazakov, S. Y. Moroskin, and L. P. Andrienko. 1984.
Zh. Prikl. Khim., 57(9): 1926.
[20] Bespalov, G. N., L. B. Filatova, and A. A. Shidlovkii. 1968. Russian Journal of Physical
Chemistry, 42(10): 1386.
[21] Rubtsov, Y. I., A. I. Kazakov, N. G. Vais, A. P. Alekseev, I. I. Strizhevskii, and E. B.
Moshkovish. 1988. Zhurnal Prikladnoi Khimi, 61(1): 131.
[22] Szabo, Z. G., E. Hollos, and J. Trompler. 1985. Zeitschrift für Physikalische Chemie,
144: 187.
[23] Mellor, J. W. 1992. A Comprehensive Treatise on Inorganic and Theoretical Chemistry.
Vol. 2. London: Longmans Green.
[24] Fedroff, B. T. 1960. Encyclopedia of Explosives and Related Items. Vol. 1. Picatinny
Arsenal: Dover, NJ.
[25] Jensen, A. V. 1972. Hazards of Chemical Rockets and Propellants Handbook: Vol. 2.
Solid Rocket Propellant Processing, Handling, Storage and Transportation. Springfield,
VA: National Technical Information Services, U S. Department of Commerce.
[26] Nagatani, M., T. Seiyama, M. Sakiyama, H. Suga, and S. Seki. 1967. Bulletin of the
Chemical Society of Japan, 40: 1833.
[27] Brown, R. N. and A. C. McLaren. 1962. Proceedings of the Royal Society A, 266: 329.
[28] Riggin, M.-T., R.-R. Knispel, and M. M. Pintar. 1972. Journal of Chemical Physics, 56:
2911.
[29] Amoros, J. L., F. Arrese, and M. Canut. 1962. Zeitschrift fur Kristallographie, 117: 92.
[30] Choi, G. S., J. E. Mapes, and E. Prince. 1972. Acta Crystallographica, B28: 1357.
[31] Baiorek, A., T. A. Machekhina, and K. Parlinski. Inelastic Scattering of Neutrons
in Solids and Liquids. Paper presented at IAEA Conference on Neutron Scattering,
Bombay, India, December 15–19, 1964.
[32] Theoret, A. and C. Sandorfy. 1964. Canadian Journal of Chemistry, 42: 57.
[33] Fernandes, J. R., S. Ganguly, and C. N. R. Rao. 1979. Spectrochimica Acta, 35A: 1013.
[34] Adam, D. M. and S. K. Sharma. 1976. Journal of the Chemical Society Faraday Transac-
tions, 72: 2069–2074.
[35] James, D. W., M. T. Carrick, and W.-H. Leong. 1974. Chemical Physics Letters, 28:
117.
[36] Tang, H. C. and B. H. Torrie. 1978. Journal of Physics and Chemistry of Solids, 39:
845–850.
[37] Madina, J. A. 1982. PhD Thesis, King’s College London.
[38] Iqbal, Z. 1974. Chemical Physics Letters, 28: 117.
[39] Remya Sudhakar, A. O. and S. Mathew. 2006. Thermochimica Acta, 451: 5–9.
[40] Choi, C. S. and H. Prask. 1983. Acta Crystallographica B, 39: 414–420.
[41] Ahtee, M., K. J. Smolander, B. W. Lucas, and A. W. Hewat. 1983. Acta Crystallogra-
phica C, 39: 651–655.
[42] Carvalheira, P., G. M. H. L. Gadiot, and W. P. C. de Klerk. 1995. Thermochimica Acta,
269–270: 273–293.
[43] Mathew, S., K. Krishnan, and K. N. Ninan. 1998. Propellants, Explosives, Pyrotechnics,
23: 150–154.
[44] Engel, W. 1973. Explosive Stoffe, 1: 9–13.
Thermal Decomposition of Ammonium Nitrate 23

[45] Szarmes, K. R. V. and J. M. Ramaradhya. 1971. Stabilisation of ammonium nitrate.


Canadian Patent, #879:586.
[46] Oberth, A. E. 1991. Phase-stabilization of ammonium nitrate by zinc diammine com-
plexes. U.S. Patent #5,071:630.
[47] Guth, E. D. Potassium fluoride stabilized ammonium nitrate. U.S. Patent #4,552,736.
[48] Simoes, P. N., L. M. Pedroso, A. A. Portugal, and J. L. Campos. 2000. Thermochimica
Acta, 364: 71–85.
[49] Mathew, S., K. Krishnan, and K. N. Ninan. 1999. Defence Science Journal, 49: 71–78.
[50] Emmet, J. P. 1930. American Journal of Science, 18: 255.
[51] Berthelot, M. 1883. Sur la Force des Matierres Explosiva. Paris: Gauthier-Villars.
[52] Saunders, H. L. 1922. Journal of the Chemical Society, 121: 698.
[53] Shah, M. S. and T. M. Oza. 1932. Journal of the Chemical Society, 725: 735–736.
[54] Robertson, A. J. B. 1948. Journal of the Society of Chemical Industry, 67: 221.
[55] Cook, M. A. and A. Taylor. 1951. Industrial and Engineering Chemistry, 43: 1098.
[56] Wood, B. J. and H. Wise. 1955. Journal of Chemical Physics, 23: 693.
[57] Friedman, L. and J. Bigeleisen. 1950. Journal of Chemical Physics, 18: 1325.
[58] Roser, W. A., S. H. Inami, and H. Wise. 1963. Journal of Physical Chemistry, 67: 1753.
[59] Roser, W. A., S. H. Inami, and H. Wise. 1964. Transactions of the Faraday Society, 60:
1618.
[60] Kapor, J. H., O. G. Jansen, and P. J. van den Berg. 1970. Explosive Stoffe, 8: 181.
[61] Brower, K. R., J. C. Oxley, and M. Tiwari. 1989. Journal of Physical Chemistry, 93:
4029.
[62] Russell, T. P. and T. B. Brill. 1989. Combustion and Flame, 76: 393.
[63] Patil, D. G., S. R. Jain, and T. B. Brill. 1992. Propellants, Explosives, Pyrotechnics,
17: 99.
[64] Taylor, J. and G. P. Sillitto. 1949. Flame and Explosion Phenomena. Third International
Symposium on Combustion. Baltimore: Williams & Wilkins.
[65] Fedroff, B. T. 1960. Encyclopedia of Explosives and Related Items. Vol. 1. Picatinny
Arsenal: Dover, NJ.
[66] Davies, T. L. and A. J. J. Abrahams. 1925. Journal of the American Chemical Society, 47:
1043.
[67] Urbanski, T. 1983. Chemistry and Technology of Explosives. Oxford: Pergamon.
[68] Sarner, S. F. 1966. Propellant Chemistry. New York: Reinhold Publishing.
[69] Anderson, W. H., K. W. Bills, A. O. Dekker, E. Mishuck, G. Moe, and R. D. Schultz.
1958. Journal of Propulsion, 28: 831.
[70] Anderson, W. H., K. W. Bills, E. Mishuck, G. Moe, and R. D. Schultz. 1959. Combus-
tion and Flame, 3: 301.
[71] Brill, T. B., P. J. Brush, and D. G. Patil. 1993. Combustion and Flame, 92: 178.
[72] Mellor, J. W. 1992. A Comprehensive Treatise on Inorganic and Theoretical Chemistry.
Vol. 2. London: Longmans Green.
[73] Keenan, A. G. 1955. Journal of the American Chemical Society, 77: 1379.
[74] Skaribas, S., T. C. Vaimakis, and P. J. Pomonis. 1990. Thermochimica Acta, 158: 235.
[75] Colvin, C. I., A. G. Keenan, and J. B. Hunt. 1963. Journal of Chemical Physics, 38: 3033.
[76] Keenan, A. G., K. Notz, and N. B. Franco. 1969. Journal of the American Chemical
Society, 91: 3168.
[77] Hardesty, J. O. and R. O. E. Davis. 1946. Industrial and Engineering Chemistry, 38: 1298.
[78] Skordilis, S. S. and P. J. Pononis. 1993. Thermochimica Acta, 216: 137.
[79] Glaskova, A. P., Yu. A. Kazarova, and A. V. Savelev. 1983. Fizika Goreniya i Vzryva,
19: 65.
[80] Glaskova, A. P. and V. A. Kazarova. 1986. Archivum. Combustionis., 6: 62.
[81] Hussain, G. and G. J. Rees. 1993. Fuel, 72: 1475.
[82] Jain, S. R. and C. Oommen. 1989. Journal of Thermal Analysis, 35: 1119.
[83] Petrakis, D. E., A. T. Sdovkos, and P. J. Pomonis. 1992. Thermochimica Acta, 196: 447.
24 S. Chaturvedi and P. N. Dave

[84] Guiochon, G. and L. Jacque. 1957. Comptes Rendus, 244: 1927.


[85] Colvin, C. I., P. W. Fearnow, and A. G. Keenan. 1965. Inorganic Chemistry, 4: 173.
[86] Kummel, R. and F. J. Picshel. 1974. Inorganic and Nuclear Chemistry, 6: 513.
[87] Guiochon, G. 1960. Annali di Chimica, 5: 295.
[88] Guiochon, G. 1960. Memorial des Poudres, 42: 47.
[89] Barelay, K. S. and J. M. Crewe. 1967. Journal of Applied Chemistry, 17: 21.
[90] Anderson, W. H., K. W. Bills, E. Mishuck, G. Moe, and R. D. Schultz. 1959. Combus-
tion and Flame, 3: 301.
[91] Beckstead, M. W. 1989. A model for AN composite propellant combustion. 26th
JANNAF Combustion Meeting, Vol. 4. Columbia, CPIA Publication 529.
[92] Moser, M. D. 1988. Workshop report, AN combustion. 25th JANNAF Combustion
Meeting, Vol. 3. Columbia, CPIA Publication 529.
[93] Mihlfeith, C. M., J. R. Goleniewski, J. H. Thacher, and A. G. Butcher. 1990. Burn rate
mechanism, Vol. 1. AL-TR-89-016. Magna, UT: Herculas, Inc.
[94] Ellern, H. 1968. Military and Civilian Pyrotechnics. New York: Chemical Publishing.
[95] Brewster, M. Q. and T. A. Sheridan. 1990. Final Report on Combustion Studies of Clean
Burning Propellants. Ogden, UT: Thiocol.
[96] Vaugham, W. W. and B. Anderson. 1993. Aerospace America, 5.
[97] Ward, M. 1996. New Sci., June 8:14.
[98] Korting, P. A. O. G., F. W. M. Zee, and J. J. Meulenbrugge. 1990. Journal of Propulsion
and Power, 6: 250.
[99] Sutton, G. P. 1986. Rocket Propulsion Elements. 5th ed. New York: Wiley.
[100] Kohga, M., S. Yoshida, and K. Hasue. 2009. Influence of AP Particle Size on the
Bunring Characteristics of AN=AP-Based Composite Propellants. Advancements in Ener-
getic Materials and Chemical Propulsion: New York.
[101] Borman, S. 1994. Chemical & Engineering News, 72: 18.
[102] Quinn Brewster, M., T. A. Scheridan, and A. Ishihara. 1992. Journal of Propulsion and
Power, 8: 760.
[103] Fink, B. and C. Perut. CA 118:105884.
[104] Doll, D. W. and G. K. Laud. 1992. Journal of Propulsion and Power, 8: 1185.
[105] Xuegang, Z. and J. G. Huojian. 1996. CA 125:118906.
[106] Longevialle, Y., G. Berteleau, M. Golfier, and H. Mace. 1995. Proceedings of the Inter-
national Symposium on Energetic Materials and Technology, Phoenix, AZ.
[107] Akira, I., S. Takeo, Y. Toshio, S. Masataka, and H. Keiichi. 1992. Int. Annual. Conf.
ICT.
[108] Lund, G. K., M. J. Spinti, and D. W. Doll. Solid propellant formualtions producing acid
neutralizing exhaust. U. S. Patent #5180452.
[109] Takuo, K., S. Noboru, and K. Kogyo. 1992. CA 117:153981.
[110] James, B. D. and R. G. John. CA 120:34032.
[111] Spinti, M., D. Doll, and R. Carpenter. 1990. Clean burning propellant for large solid
rocket motors. Edwards Air Force Base, CA: Astronautics Lab. Report No. AL-TR-
89-033.
[112] Smedley, J. E. 1990. Clean propellant demonstration. Edwards Air Force Base, CA:
Astronautics Lab. Report No. AL-TR-90-029.
[113] Jacox, J. 1990. Clean propellant development and demonstration. Edwards Air Force
Base, CA: Astronautics Lab. Report No. F04611-89-C-0082.
[114] Bennet, R. R. NASA Conference Publ. 1995. Aerospace Environmental Technology
Conference.
[115] Jacox, J. L. and D. L. Bradford. 1994. Aerospace Environmental Technology Conference.
[116] Klaus, M., B.-M. Jutta, S. Helmut, B. Klaus Martin, and E. Walter. European Patent
Application #EP 705,809, CA 124:347574.
[117] Weyland, H. H. and R. R. Miller. 1990. Clean propellant for large solid rocket motors.
Edwards Air Force Base, CA: Astronautics Lab. Report No. AL-TR-90-016.
Thermal Decomposition of Ammonium Nitrate 25

[118] Hideo, H., O. Tpsjop, and G. Kakaku. 1996. CA 125: 225920.


[119] Hinshaw, C., R. B. Wardle, and T. K. Hingshmith. Propellant Formulations Based on
Dinitramide Salts and Energetic Binders. Thiokol.
[120] Russell, T. P., G. J. Piermarini, S. Block, and P. J. Miller. 1996. Journal of Physical
Chemistry, 100: 3248.
[121] Mebel, A. M., M. C. Lin, and M. Morokuma. 1995. Journal of Physical Chemistry, 99:
6842.
[122] Muthiah, R., T. L. Varghese, S. S. Rao, K. N. Ninan, and V. N. Krishnamurthy. 1998.
Propellants, Explosives, Pyrotechnics, 23: 90.
[123] Schoyer, H. F. R. and A. J. Schnorhk. 1995. Journal of Propulsion and Power, 11(4): 856.
[124] Fleming, W. C. U. S. Patent #5583315.
[125] Perut, C., V. Bodart, and B. Cristofoli. 1993. Int. Annu. Conf. ICT.
[126] Sun, J., Z. Sun, Q. Wang, H. Ding, T. Wang, and C. Jiang. 2005. Journal of Hazardous
Materials B, 127: 204–210.
[127] Zhu, Z. H. and Z. L. Guo. 2000. Discussion on the prior evaluation of the explosion of
ammonium nitrate. China Safety Science Journal, 10: 70–74.
[128] MacNeil, J. H., H. T. Zhang, and P. Berserh. 1997. Catalytic decomposition of
ammonium nitrate in superheated aqueous solution. Journal of the American Chemical
Society, 119: 9738–9744.
[129] Keskar, M., T. V. Rao, and S. K. Sali. 2010. Thermochimica Acta, 510: 68–74.
[130] Miron, Y., T. C. Ruhe, and R. W. Watson. 1979. Reactivity of ANFO with Pyrite
Weathering Products. Bureau of Mines.
[131] Forshey, D. R., T. C. Ruhe, and C. M. Mason. 1968. The Reactivity of ANFO with
Pyrite Bearing Ores. Bureau of Mines.
[132] Oxley, J. C., S. M. Kaushik, and N. S. Gilson. 1992. Thermal stability and compatibility
of ammonium nitrate explosive on a small and large scale. Thermochimica Acta, 212:
77–85.
[133] Gunawan, R. and D. Zhang. 2009. Journal of Hazardous Materials, 165: 751–758.
[134] Rosser, W. A., S. H. Inami, and H. Wise. 1963. The kinetics of decomposition of liquid
ammonium nitrate. Transactions of the Faraday Society, 67: 1753–1757.
[135] Brower, K. R., J. C. Oxley, and M. Tewari. 1989. Evidence for homolytic decompo-
sition of ammonium nitrate at high temperature. Journal of Physical Chemistry, 93:
4029–4033.
[136] Patil, D. G., S. R. Jain, and T. B. Brill. 1992. Thermal decomposition of energetic mate-
rials 56. On the fast thermolysis mechanism of ammonium nitrate and its mixtures with
magnesium and carbon. Propellants, Explosives, Pyrotechnics, 17: 99–105.
[137] Medard, L. A. 1989. Accidental Explosions: Vol. 2. Types of Explosives Substances.
Chichester: Ellis Horwood.
[138] Russel, T. P. and T. B. Brill. 1989. Thermal decomposition of energetic materials 31—
Fast thermolysis of ammonium nitrate, ethylenediammonium dinitrate and hydrazinium
nitrate and the relationship to the burning rate. Combustion and Flame, 76: 393–401.
[139] Davis, T. L. and A. J. J. Abrams. 1925. The dehydration of ammonium nitrate. Journal
of the American Chemical Society, 47: 1043–1045.
[140] Gunawan, R., S. Freij, D. Zhang, F. Beach, and M. Littlefair. 2006. A mechanistic study
into the reactions of ammonium nitrate with pyrite. Chemical Engineering Science, 61:
5781–5790.
[141] Skaribas, S., T. C. Vaimakis, and P. J. Pomonis. 1990. Thermochimica Acta, 158:
235–246.
[142] Coats, A. W. and J. P. Redfern. 1964. Nature, 201: 68.
[143] Skordilis, C. S. and P. J. Pomonis. 1993. Thermochimica Acta, 216: 137–146.
[144] Mathew, S., N. Eisenreich, and W. Engel. 1995. Thermochimica Acta, 269–270: 475–489.
[145] Farhat, K., W. Cong, Y. Batonneau, and C. Kappenstein. 2009. 5th AIAA=ASME=
SAE=ASEE Joint Propulsion Conference and Exhibit. Denver, CO. Paper number 4963.
26 S. Chaturvedi and P. N. Dave

[146] Kaljuvee, T., E. Edro, and R. Kuusik. 2008. Journal of Thermal Analysis and Calor-
imetry, 92(1): 215–221.
[147] Li, Y. and J.-M. Hui. 2005. Chinese Journal of Explosives and Propellants, 28(1): 76–80.
[148] Oxley, J. C., S. M. Kaushik, and N. S. Gilson. 1989. Thermochimica Acta C, 153(1):
269–286.
[149] Rubtsov, Y. I., A. I. Kazakov, D. B. Lempert, and G. B. Manelis. 2006. Propellants,
Explosives, Pyrotechnics, 31(6): 421–434.
[150] Zhang, J. and Y.-W. Zou. 2005. Energetic Materials, 13(4): 229–231.
[151] Simoes, P. N., L. M. Pedroso, A. A. Portugal, and J. L. Campos. 1998. Thermochimica
Acta, 319: 65.

You might also like