You are on page 1of 16

Journal of Energetic Materials, 32: 146–161, 2014

Copyright # Taylor & Francis Group, LLC


ISSN: 0737-0652 print=1545-8822 online
DOI: 10.1080/07370652.2013.796428

Effect of Copper Oxide, Titanium Dioxide, and


Lithium Fluoride on the Thermal Behavior and
Decomposition Kinetics of Ammonium Nitrate

ANUJ A. VARGEESE,1 S. J. MIJA,2 AND


KRISHNAMURTHI MURALIDHARAN1
1
School of Chemistry and Advanced Centre of Research in High Energy
Materials, University of Hyderabad, Hyderabad, India
2
Electrical Engineering Department, National Institute of Technology
Calicut, Calicut, Kerala, India

Ammonium nitrate (AN) is crystallized along with copper oxide, titanium dioxide,
and lithium fluoride. Thermal kinetic constants for the decomposition reaction of the
samples were calculated by model-free (Friedman’s differential and Vyzovkins
nonlinear integral) and model-fitting (Coats-Redfern) methods. To determine the
decomposition mechanisms, 12 solid-state mechanisms were tested using the Coats-
Redfern method. The results of the Coats-Redfern method show that the decompo-
sition mechanism for all samples is the contracting cylinder mechanism. The phase
behavior of the obtained samples was evaluated by differential scanning calorimetry
(DSC), and structural properties were determined by X-ray powder diffraction
(XRPD). The results indicate that copper oxide modifies the phase transition beha-
vior and can catalyze AN decomposition, whereas LiF inhibits AN decomposition,
and TiO2 shows no influence on the rate of decomposition. Possible explanations
for these results are discussed.
Supplementary materials are available for this article. Go to the publisher’s online
edition of the Journal of Energetic Materials to view the free supplemental file.

Keywords ammonium nitrate; copper oxide; decomposition kinetics; lithium


fluoride; thermal properties; titanium dioxide

Introduction
As an inorganic oxidizer, ammonium nitrate (AN) finds application in composite
propellant formulations, especially in gas generator systems and automotive air
bags. The potential environmental effects of the exhaust products of conventional
ammonium perchlorate (AP) containing rocket propellants have been assessed by
different investigators [1]. There is increasing interest in developing ecofriendly solid
propellant systems [2]. AN is receiving attention as a possible substitute for the
oxidizer AP. However, applications of AN in highly reliable systems are limited
because of the near–room temperature crystallographic phase transition, low
energetics, hygroscopicity, and inability for sustained burning.

Address correspondence to Anuj A. Vargeese, Department of Chemistry, National


Institute of Technology Calicut, Chathamangalam, Calicut 673601, Kerala, India. E-mail: dr_
anuj@ymail.com

146
Ammonium Nitrate 147

Interest in understanding the physical and thermal properties of AN has a long


history [2–5]. Despite many investigations, the actual phase transition behavior of
AN and its causes still remain unanswered. According to the literature, the addition
of a small quantity (5%) of foreign salts stabilizes AN [2,6]. Incorporation of these
foreign salts into the crystal lattice of AN can be achieved by different methods.
In melt stabilization, metal oxides are reacted with molten AN to form stabilized
diamine complexes [6–8]. Phase-stabilized AN in the form of poly(vinyl pyrolidone)-
AN glass has also been reported [9].
AN melts at around 170 C and dissociative decomposition begins above this
temperature [10,11]. In addition, in the melt crystallization method a larger quantity
of stabilizing agent is required to achieve phase stabilization [2,6]. In the present
study, we used a significantly smaller amount (1% by weight) of micrometer-size
particles and a low temperature (100 C), considerably below the decomposition
temperature of AN to obtain the modified AN samples. The influence of
micrometer-size CuO, TiO2, and LiF particles on the phase behavior and decompo-
sition of AN was investigated.
The influence of transition metal oxides in industrial process and thermal
decomposition reactions are well known [12], and additives significantly influence
the thermal stability of AN [13,14]. Isoconversional kinetic methods provide
information on the multistep kinetics and mechanisms of the thermal decomposition
reactions [15] and help determine the thermal stability of the material under
investigation [11]. Hence, to understand the thermal stability and decomposition
behavior, a decomposition kinetics analysis of the AN and modified AN samples
was carried out using two model-free (isoconversional) and one model-fitting method.

Experimental Materials and Methods


The chemicals used for the experiments were of commercial (Merck India Ltd.,
Hyderabad, India) analytical reagent grade and used without further purification.

Synthesis of CuO, TiO2, and Lif


A hydrothermal method using copper acetate [Cu(CH3COO)2] and urea [CO(NH2)2]
was employed for the synthesis of micrometer-size CuO particles [16]. TiO2 with
uniform particle size was synthesized by the hydrolysis of a titanium isopropoxide
solution [17]. LiF was prepared by the reaction of lithium hydroxide monohydrate
and hydrofluoric acid [18]. The formation and morphology of the synthesized
CuO, TiO2, and LiF were confirmed by scanning electron microscopy (SEM) and
X-ray powder diffraction (XRPD).

Crystallization of AN with CuO, TiO2, and Lif


The evaporative crystallization of AN with CuO, TiO2, and LiF was carried out as
follows: 1 g of AN was dissolved in 2 mL deionized water with continuous stirring.
After complete dissolution, 1% by weight of the additive was transferred to the
reaction vessel. After 10 min of stirring, the reaction vessel was transferred to
a constant-temperature paraffin oil bath kept at 100 C with heating and stirring
continuously until dry crystals formed. The crystals thus obtained were collected,
dried under vacuum, and used for further analysis. Commercially available AN
148 A. A. Vargeese et al.

obtained from the manufacturer (Merck) was used for the thermogravimetric and
kinetic analyses.

X-ray Powder Diffraction


The XRPD patterns were recorded on a Bruker D8 Advance XRD at room
temperature (25 C). AN samples were scanned over a 2h range of 10–70 and the
AN mixtures were scanned over 20–70 with a sampling interval of 0.01 . From
the obtained XRPD patterns, the interplanar distances (d) were calculated
using Bragg’s equation and were compared with the Joint Committee on Powder
Diffraction Standards (JCPDS) data.

Morphological (SEM) Analysis


SEM images were obtained at different magnifications on an FEI XL30 ESEM and
used for particle size and morphology analysis of CuO, TiO2, and LiF. For the SEM
analysis, the powder sample was dispersed in ethanol, stirred continuously for
30 min, drop-casted on a glass slide, and then coated with gold using a sputter coater.

Thermal Analysis
The AN mixtures were thermally cycled on a Mettler-Toledo DSC 1 from 20 to
100 C at a heating rate of 5 C=min. The cycling sequence was 20–100 C ! 100–
20 C ! 20–100 C ! 100–20 C ! 20–100 C. The samples were held for 2 min at
isothermal conditions at the final temperature before reversing the temperature.
For each experiment,  2 mg of sample was loaded into an aluminum pan, covered
with a pierced lid, and the purge gas (nitrogen) flow rate was maintained at
100 mL=min.
Thermogravimetric analysis (TGA) measurements were carried out on a Met-
tler-Toledo TGA=DSC 1. In all experiments, samples of 1–2 mg were heated in open
90-mL alumina pans in a flowing nitrogen atmosphere (30 mL=min). The protective
gas used for the balancing chamber was also nitrogen, at a flow rate of 30 mL=min.

Decomposition Kinetic Analysis


Two model-free isoconversional methods and one model-fitting method were used to
compute the Arrhenius parameters of the decomposition reaction.

Model-Free (Isoconversional) Kinetic Analysis


TGA measurements were used to determine the decomposition kinetics of the
samples. Nonisothermal TGA runs were conducted at heating rates 2.5, 5, 10, and
20 C=min and the extent of conversion (a) was computed from the weight loss data
using the standard method as described elsewhere [19]. As per the International Con-
federation for Thermal Analysis and Calorimetry (ICTAC) committee recommenda-
tions [11], the dependence of apparent activation energy (Ea) on a was computed
using isoconversional methods employing kinetic data obtained from four noni-
sothermal runs. Das of 0.01 and 0.025 were respectively used to compute the appar-
ent activation energy from the differential and the integral isoconversional methods.
For comparison and plotting, constant a values were shown at an interval of 0.05.
Ammonium Nitrate 149

Two isoconversional methods, namely Friedman’s differential method [20] and


Vyazovkin’s nonlinear integral method [21] were employed for the kinetic analysis
and are explained below.

Friedman’s Differential Method. The differential isoconversional method of


Friedman [20] is based on the Arrhenius equation (Eq. (1)):
 
da E
¼ A exp  f ðaÞ; ð1Þ
dt RT

where A is the preexponential factor, E is the activation energy, T is the temperature,


and f(a) is the reaction model. Friedman analysis applies the logarithm of the
conversion rate da=dt as a function of the reciprocal temperature at different
degrees of conversion a. This modification results in Eq. (2):
   
da Ea
ln ¼ ln½Aa f ðaÞ  ð2Þ
dt a;i RTa;i

Hence, by plotting ln(da=dt) versus 1=T at constant a values gives a family of


straight lines with slope Ea=R. In a similar way, the apparent activation energy
of AN and AN mixtures was computed.

Vyazovkin’s Nonlinear Integral Method. Vyazovkin’s method is a nonlinear


integral isoconversional method that can be used for nonisothermal experiments.
Vyazovkin’s method provides more precise apparent activation energy values by
performing numerical integration [21]. For a set of n experiments carried out at
different heating rates, the apparent activation energy can be determined at any
particular value of a by finding the value of Ea for which the given function, Eq.
(3), is a minimum. The minimization procedure is repeated for each value of a to find
the dependence of the apparent activation energy on the extent of conversion.
Xn Xn
i j6¼i
½IðEa ; Ta;i Þbj =½IðEa ; Ta;i Þbi  ¼ min; ð3Þ

where

Z T  
E
IðE; TÞ ¼ exp  dT: ð4Þ
0 RT

b in Eq. (3) represents the heating rates, the indexes i and j denote the set of
experiments performed under different heating rates, and n is the total number of
experiments performed.
The third-degree approximation (Eq. (5)) proposed by Senum and Yang [22] was
used in the present study to evaluate the integral Eq. (4).

expðxxÞ x2 þ 10x þ 18
f ðxÞ ¼  3 ; ð5Þ
x x þ 12x2 þ 36x þ 24

where
150 A. A. Vargeese et al.

Table 1 Kinetic parameters estimated by different methods

E (kJ mol1) E (kJ mol1) E (kJ mol1) A (s1)


(Friedman) (Vyazovkin) (Coats-Redfern) (Coats-Redfern)
AN 119 110 117 1.82  1011
AN þ CuO 97 98 116 2.5  1011
AN þ TiO2 115 117 120 1.67  1011
AN þ LiF 141 137 121 5.02  1010

E E
x¼ and IðE; TÞ ¼ f ðxÞ: ð6Þ
RT R
MatLab 7.0.1 was used to perform the kinetic computations.

Model-Fitting (Coats-Redfern) Method


The kinetics of thermal decomposition of AN was analyzed by applying the
Coats-Redfern approximation [23]. During the integration of the single-step kinetic
equation (Eq. (1)), applying one of the approximations to the temperature integral
and incorporating the constant heating rate (b) resulted in the following
Coats-Redfern equation:
  
gðaÞ AR 2RT E
ln 2 ¼ ln 1  : ð7Þ
T bE E RT

If the correct g(a) is used, the plot of ln[g(a)=T2] against 1=T should give a
straight line with a high correlation coefficient in a linear regression analysis from
which the values of E and A can be derived. Table I (see online supplement) shows
different expressions of g(a) for 12 different solid-state mechanisms used in this study
for the estimation of reaction mechanisms using the Coats-Redfern method. The
TGA curve obtained at a heating rate of 5 C=min was used for the computation
of E and A using this method.

Results and Discussion


Characterization of CuO, TiO2, and Lif
Micrometer-size particles of CuO, TiO2, and LiF were synthesized as described in the
‘‘Experimental Materials and Methods’’ section. The synthesized particles were char-
acterized by SEM and XRPD. Figures 1–3 shows the SEM images of CuO, TiO2,
and LiF, respectively, and Fig. 4 shows their XRPD patterns. The CuO particles
(Fig. 1) were spherical in shape with a size of 2–3 mm, as evident from SEM images
obtained at 1,200  magnification. The XRPD pattern was compared with JCPDS
05-0661 (CuO), which showed that the strong peaks of CuO (corresponding to hkl
values (11 1), (200=111), and (202)) were intact. As seen in the SEM image (Fig. 3),
the well-separated TiO2 particles had a size of 0.5 mm. The XRPD pattern indicated
the formation of an anatase phase with a small amount of brookite phase and the
reflections matched either JCPDS 21-1272 (anatase) or 29-1360 (brookite) of TiO2.
The LiF particles observed by SEM were rectangular prisms with a groove
Ammonium Nitrate 151

Figure 1. SEM image of CuO.

Figure 2. SEM image of TiO2.

Figure 3. SEM image of LiF.


152 A. A. Vargeese et al.

Figure 4. XRPD pattern of CuO, TiO2, and LiF.

morphology (Fig. 3) with a size of  10 mm, and the XRPD reflections were found to
be in good agreement with JCPDS 89-3610 (LiF). In general, the catalytic activity of
the catalysts for thermal decomposition reactions was dependent on the material
property of the catalyst: surface area, shape, spatial distribution, surface compo-
sition, electronic structure, thermal and chemical stability, etc.

Phase Behavior of Modified AN Mixtures


Differential scanning calorimetry (DSC) thermal cycling of the AN mixtures was car-
ried out. The study indicated that only the AN with CuO showed a modified phase
behavior. The DSC thermal cycling data of the AN-CuO mixture is shown in
Fig. 5. Because AN-TiO2 and AN-LiF did not show promising results, they are pro-
vided in the supplementary files for this article. From Fig. 5 it can be seen that the
AN-CuO mixture showed only one transition at 55 C in the temperature range of
20–100 C. Generally, pure AN undergoes a III-II transition at 84 C, and this tran-
sition was suppressed. On cooling the sample, instead of the expected two reverse
transitions, the sample showed only one reverse transition at 44 C. Hence, it was
concluded that the AN-CuO mixture behaves as a phase-modified AN (PMAN). In
addition, a consistent transition at  60 C and no other transition in the range
100 to 60 C has been reported in the literature for AN melt doped with CuO [24].
Figure 6 shows the XRPD patterns of AN-CuO, AN-TiO2, and AN-LiF
mixtures. The XRPD results were compared with JCPDS data for AN and a series
of different possible compounds, especially those of amine complexes. These
comparisons revealed that the AN-CuO sample exhibited peaks corresponding to
AN (JCPDS 83-520) and copper tetramine nitrate (JCPDS 70-195). Few low-intensity
peaks of copper tetramine nitrate (CTAN) at 2h values (counts per second values
Ammonium Nitrate 153

Figure 5. DSC thermal cycling data for the AN-CuO mixture.

Figure 6. XRPD pattern of AN mixtures.


154 A. A. Vargeese et al.

in bracket) 13.3 (43), 30.6 (48), 39.3 (79), 50.3 (38), and 50.9 (40) were observed, indi-
cating its presence. To avoid any ambiguity, peaks within  0.2 of the AN 2h values
were assigned to AN reflections. The peak at 2h ¼ 13.3 (011) can be assigned to
CTAN, because AN has only one reflection at 17.9 (001) in the 2h range of 10 to
20. These results showed the presence of crystallized AN along with the in situ formed
CTAN. Because the quantity of CuO added was 1% by weight, it was expected that
the amine complex concentration in the system will also be of the same order, and this
might have resulted in the low-intensity peaks. The XRD peak positions of the
AN-CTAN mixture showed small variations from the JCDPS values, probably due
to the elongation of the cell lengths originating from CTAN as an impurity
accompanied by internal stresses or stacking faults. The AN-TiO2 and AN-LiF
showed only reflections that correspond to AN.
The formation of CTAN can be explained as follows. The addition of black CuO
powder to the hot acidic AN solution (pH ¼  4.5) could result in hydrolyzation,
leading to the formation of Cu(OH)2 [25]. This formation was indicated by the
change in the solution color to light blue, because the Cu(OH)2 has a pale blue color.
In the presence of NH3, produced due to the dissolution of AN, the Cu(OH)2 may be
converted to CTAN [26,27]. In the initial stages, most of the Cu will be converted to
Cu(OH)2 and remain in hydroxide form. Concentrating the solution by evaporation
shifts the equilibrium toward the formation of CuðNH3 Þ2þ 4 , and this was visually
observed by the more deep blue color compared to the hydroxide. At this stage,
the uncreated AN crystals began precipitating out, and this complex was homo-
geneously mixed with the AN crystals. The final AN-CTAN mixture collected was
of deep blue color.
Reportedly, the CuO and AN reaction initially results in the formation of
[Cu(NH3)2](NO3)2, the copper diamine complex, and at room temperature the
diamine complex slowly converts to the tetramine complex and stabilizes AN [24].
The possibility of incorporation of Cu2þ, interstitially or substitutionally, in the
crystal lattice and the subsequent modification of the phase transition of AN was
also suggested [24]. However, the substitution mechanism is favored only if the host
cations have a cationic radii within  15% of the cation being replaced. NH4þ has
a reported ionic radius of 1.48 Å, and the replacement of this NH4þ by Cu2þ
(0.73 Å) might not be favored.

Decomposition Kinetics of AN Mixtures


Though the literature reports a variety of decomposition pathways that AN can
undergo [2], it is widely accepted that the thermal decomposition is initiated by
the vaporization of melted AN accompanied by an endothermic proton transfer
reaction R1 [28]. Molten AN at atmospheric pressure is incapable of burning, and
the addition of small amount of minerals, catalysts, fuels, or explosives is required
for sustained burning [29]. The combustion behavior of AN mixtures with different
additives, fuels, and energetic materials was intensely studied [29]. The thermal
decomposition mechanism of AN is highly dependent on temperature, and only
the dissociative sublimation that occurs around 200 C is endothermic in nature
[2]. Because all other reported decomposition pathways are exothermic in nature,
and the differential thermal analysis curve indicated an endothermic decomposition,
the possibility of the reaction shown in R1 seems apparent in the present context.
However, many secondary reactions occur rapidly in the gas phase between the
Ammonium Nitrate 155

dissociation products ammonia and nitric acid. In the decomposition of noncata-


lyzed AN, NH3 and HNO3 are formed in the first step (R1) and then the HNO3
further dissociates to form NOþ 
2 ; NO3 , and H2O (R2). The NH3 formed according
to R1 is subsequently oxidized by the NO2þ through the reaction shown in R3,
leading to an overall decomposition reaction given by R4 [28, 29]. It was suggested
that the H2O and NH3 formed during the decomposition reaction presumably
prevents the sustained burning of pure AN [29].

NH4 NO3 ! NH3 þHNO3 ðR1Þ

2HNO3 Ð NOþ 
2 þ NO3 þ H2 O ðR2Þ

NH3 þ NOþ
2 ! N2 O þ H3 O
þ
ðR3Þ

NH4 NO3 ! N2 O þ 2H2 O: ðR4Þ

Activation energy may be correlated to bond energy data and is generally considered
as a measurement of the energy barrier to a controlling (rate limiting) bond rupture
or bond redistribution step [30]. The activation energy provides reasonable
information about critical energy needed to start the decomposition reaction of
the compound. Hence, a reduction in the activation energy of the decomposition
reaction can be directly correlated to the catalytic activity of the added compounds.
Though the reported activation energy for AN decomposition varies from 86.2 to
206.9 kJ=mol, previous researchers in this area agree that the overall decomposition
reaction of AN is described by the first-order reaction kinetics [31]. A recent work
reported that the thermal gasification of AN has a similar kinetics in the solid and
liquid phases and requires an activation energy of  90 kJ=mol [10].

Isoconversional Kinetic Analysis


Decomposition of AN-CuO Mixture. The dependence of apparent activation
energy on the extent of conversion for AN and its mixtures calculated using
Vyazovkin’s method is shown in Fig. 7. The isoconversional Arrhenius plots for
the thermal decomposition of AN and its mixture obtained using Friedman’s
method are shown in Fig. 8. The ln(da=dt) values corresponding to a values of 0.4
and 0.6 were plotted against the inverse of the absolute temperature. The figure
shows that the linear fits have a constant slope for a given sample, indicating con-
stant activation energy throughout the process. The Ea vs. a dependence obtained
by Friedman’s method is shown in Fig. 9, and the activation energies obtained by
different methods are shown in Table 1.
The activation energies obtained for the AN decomposition reaction were
119 kJ=mol (Friedman) and 110 kJ=mol (Vyazovkin). For PMAN, the activation
energy calculations showed lower values, 97 kJ=mol (Friedman) and 98 kJ=mol
(Vyazovkin), indicating the catalytic influence of CTAN on the AN decomposition
reaction. However, the catalytic activity may be due to either CTAN or its decompo-
sition products or both. In the present case, many factors can play an important role
in the AN-CTAN mixture decomposition. First, because the CTAN contains both
a strong base and an acid, either the ammonia or nitric acid or both molecules
can act as an autocatalyst for the PMAN decomposition reaction. Second, on
heating to higher temperatures, the CTAN can undergo a general decomposition
156 A. A. Vargeese et al.

Figure 7. Activation energy (Ea) dependencies for AN mixtures, computed using Vyazovkin’s
method [21].

reaction and release Cu [32]. However, it is quite likely that CuO is quickly formed in
reaction with the nitric acid. An earlier investigation on the dissociation products of
CTAN, which accounts for CuO as the major residue [33], supports this presump-
tion. Because CuO is often used as a decomposition catalyst for AP and AP-based
propellants [34], it does not seem unreasonable to assume that CuO can act as a cata-
lyst for the AN thermal decomposition reaction. Thus, the Cu could enhance the
decomposition reaction by reacting early with the nitric acid, leading to the forma-
tion of CuO and the subsequent reaction of CuO with the remaining AN. Hence, due
to the interaction of Cu and CuO, decomposition products are removed at a faster

Figure 8. Isoconversional plots of samples based on Freidman’s method [20] (color figure
available online).
Ammonium Nitrate 157

Figure 9. Activation energy (Ea) dependencies for AN mixtures, computed using Friedman’s
method [20].

rate, and as per Le Chatelier’s principle, the equilibrium (R1) is displaced to the
right, which in turn will encourage the dissociative decomposition of AN. Possibly
this catalytic activity would have resulted in the lowering the activation energy
required for the AN-CTAN decomposition reactions.

Decomposition of AN-TiO2 Mixture. The activation energies for AN-TiO2 were


115 kJ=mol (Friedman) and 117 kJ=mol (Vyazovkin). Due to the known interaction
of TiO2 and NH3 [35] (AN dissociation product in the present case), a significant
reduction in activation energy might be anticipated. However, no significant vari-
ation in the activation energy was actually observed. When the TiO2 powder was
thermally treated above 100 C, dehyroxylation occurred, leaving incompletely coor-
dinated titanium atoms at the surface, which behaved as Lewis acid centers. These
active sites=centers can fix either a water molecule or an ammonia molecule and sup-
port surface reactions to enhance the decomposition rate of AN [36]. In the present
study, though the TiO2 was dried at 100 C, the active sites likely became hydroxy-
lated during the solution crystallization, resulting in reduced catalytic efficiency.

Decomposition of AN-LiF Mixture. In the case of AN-LiF, the activation ener-


gies obtained were 141 kJ=mol (Friedman) and 137 kJ=mol (Vyazovkin), indicating
higher stability. Because LiF acts as a burn rate retardant for AP-based propellants
[37], it is reasonable to assume that it may act as a negative catalyst for the AN
decomposition reaction. The endothermic decomposition of lithium fluoride or alter-
ation of the decomposition pathway of AN can increase the required activation
energy. The dependence of Ea on a is an indication that there is a change in the
decomposition pathway of AN. If AN undergoes the normal first-order decompo-
sition kinetics, Ea must be independent of a. Hence, the Ea vs. a dependence indicates
a possible alteration of the decomposition pathway.
Though the apparent relationship between activation energy and the extent of
conversion showed similar trends, the numerical values of apparent activation
158 A. A. Vargeese et al.

energy obtained using Friedman’s method and nonlinear (Vyazovkin) isoconver-


sional method showed slight differences. The confidence interval calculated for each
value showed that the differences were statistically negligible; moreover, a 10% error
in Ea values is rather common. However, these small differences could be due to the
approximation of the temperature integral used in the derivations of the kinetic
equations [38]. Because the differential isoconversional method of Friedman employs
instantaneous rate values, it is very sensitive to experimental noise and tends to be
numerically unstable, especially when the rate is estimated by numerical differen-
tiation of the experimental data. Indeed, investigators have compared these two
methods and shown evidence of approximately 20% difference in the absolute value
of Ea [39,40]. If the isoconversional method gives actual information regarding the
decomposition reaction, for multistep kinetics, where the overall reaction is governed
by different mechanisms, the Ea varies with the extent of conversion. This reflects the
variation in relative contributions of single steps to the overall reaction rate.

Kinetic Analysis by Model-Fitting Method


The model-fitting approach proposed by Coats-Redfern [23] was used to evaluate
the Arrhenius parameters. The reliability of activation energy values obtained by this
method has been disputed [10]; however, in the present study the trend exhibited by
the activation energy values for different AN samples was similar to that obtained
by isoconversional methods. From the 12 different reaction models tested for the
decomposition, the standard statistical procedure of linear regression analysis
yielded the highest correlation coefficient for the contracting cylinder model. This
was in good agreement with the reported work on the decomposition of solid and
liquid AN [10]. The regression coefficients obtained for different samples using
different reaction models were tabulated and are provided in in Table II (see online
supplement). Evidently, the studies showed that a small quantity of CuO can modify
the phase transition in AN and can catalyze the decomposition process as well.
Hence, during the catalyzed decomposition of AN the decomposition products
possibly undergo further reactions on the additive=catalyst surface. These surface
reactions consequently alter the reaction pathway and activation energy required
for the process. CuO helps in the complete oxidation of ammonia or removal of
H2O, known inhibitors of the AN decomposition reaction, thereby lowering the
activation energy.

Conclusions
AN crystallized along with CuO, TiO2, and LiF exhibited thermal behavior and
decomposition kinetics different from that of pure AN. The aqueous media reaction
of AN and CuO resulted in the formation of a mixture of AN and CTAN. The DSC
thermal cycling of the AN-CTAN mixture exhibited only one transition at around
55 C in the temperature range 20–100 C, indicating a modification of the phase tran-
sition behavior. The model-free and model-fitting kinetic analyses showed signifi-
cantly lower activation energy for the decomposition reaction of AN crystallized
with CuO. On the other hand, no significant variation in activation energy was
observed for AN crystallized with TiO2, and a higher activation energy was observed
for AN crystallized with LiF. From the 12 possible reaction models analyzed using
the Coats-Redfern method, a contracting cylinder mechanism seems to be the mech-
anism responsible for the decomposition process. In the case of CuO-modified AN,
Ammonium Nitrate 159

autocatalysis and the interaction of Cu species with the dissociative decomposition


products of AN brought down the activation energy barrier. The unavailability of
active sites on the TiO2 surfaces and alteration of the decomposition pathway by
LiF had an influenced in the case of the AN mixtures with TiO2 and the LiF.

Acknowledgments
The authors thank Dr. V. N. Krishnamurthy and Professor S. P. Tewari for fruitful
discussions, DRDO for financial support to ACRHEM, and A. Rajesh, CIL,
University of Hyderabad, for collecting the thermal analysis data.

References
[1] Bennett, R. R., J. R. Whimpey, R. Smith-Kent, and A. J. McDonald. 1997. Effects of
rocket exhaust on the launch site environment and stratospheric ozone. International
Journal of Energetic Materials and Chemical Propulsion, 4: 92–105.
[2] Oommen, C. and S. R. Jain. 1999. Ammonium nitrate: A promising rocket propellant
oxidizer. Journal of Hazardous Materials, 67: 253–281.
[3] Hendricks, S. B., E. Posnjak, and F. C. Kracek. 1932. Molecular rotation in the solid
state. The variation of the crystal structure of ammonium nitrate with temperature.
Journal of the American Chemical Society, 54: 2766–2786.
[4] Marino, R. A. and B. Suryanarayana. 1985. Investigation of phase transitions in
ammonium nitrate by nitrogen-15 nuclear magnetic resonance. Journal of Energetic
Materials, 3: 57–74.
[5] Vargeese, A. A., S. S. Joshi, and V. N. Krishnamurthy. 2010. Use of potassium
ferrocyanide as habit modifier in the size reduction and phase modification of ammonium
nitrate crystals in slurries. Journal of Hazardous Material, 180: 583–589.
[6] Engel, W. 1973. Phase stabilization of ammonium nitrate. Explosivstoffe, 21: 9–13.
[7] Morozov, I. V., Y. M. Korenev, and S. I. Troyanov. 1996. Synthesis and crystal structure
of new amminecopper (II) nitrates: [Cu(NH3)2](NO3)2 and [Cu(NH3)](NO3)2. Zeitschrift
für Anorganische und Allgemeine Chemie, 622: 2003–2007.
[8] Morosin, B. 1976. The crystal structure of copper (II) tetraamine nitrate. Acta
Crystallographica Section B: Structural Science, 32: 1237–1240.
[9] Lang, A. J. and S. Vyazovkin. 2008. Ammonium nitrate–polymer glasses: A new concept
for phase and thermal stabilization of ammonium nitrate. The Journal of Physical
Chemistry B, 112: 11236–11243.
[10] Vyazovkin, S., J. S. Clawson, and C. A. Wight. 2001. Thermal dissociation kinetics of
solid and liquid ammonium nitrate. Chemistry of Materials, 13: 960–966.
[11] Vargeese, A. A., K. Muralidharan, and V. N. Krishnamurthy. 2011. Thermal stability of
habit modified ammonium nitrate: Insights from isoconversional kinetic analysis.
Thermochimica Acta, 524: 165–169.
[12] Rudloff, W. K. and E. S. Freeman. 1970. Catalytic effect of metal oxides on thermal
decomposition reactions. II. Catalytic effect of metal oxides on the thermal decompo-
sition of potassium chlorate and potassium perchlorate as detected by thermal analysis
methods. The Journal of Physical Chemistry, 74: 3317–3324.
[13] Manelis, G. B., G. M. Nazin, Y. I. Rubtsov, and V. A. Strunin. 2003. Thermal
Decomposition and Combustion of Explosives and Powders. Boca Raton, FL: CRC Press.
[14] Kwoka, Q. S.M., P. Kruusb, and D. E.G. Jonesc. 2004. Wettability of ammonium nitrate
prills. Journal of Energetic Materials, 22: 127–150.
[15] Vyazovkin, S., A. K. Burnham, J. M. Criado, L. A. Pérez-Maqueda, C. Popescu, and
N. Sbirrazzuoli. 2011. ICTAC kinetics committee recommendations for performing
kinetic computations on thermal analysis data. Thermochimica Acta, 520: 1–19.
160 A. A. Vargeese et al.

[16] Jia, W., E. Reitz, P. Shimpi, E. G. Rodriguez, P. Gao, and Y. Lei. 2009. Spherical CuO
synthesized by a simple hydrothermal reaction: concentration dependent size and its
electrocatalytic application. Material Research Bulletin, 44: 1681–1686.
[17] Mahshid, S., M. Askari, and M. S. Ghamsari. 2007. Synthesis of TiO2 nanoparticles by
hydrolysis and peptization of titanium isopropoxide solution. Journal of Material
Processing Technology, 189: 296–300.
[18] Sarraf-Mamoory, R., S. Nadery, and N. Riahi-Noori. 2007. The effect of precipitation
parameters on preparation of lithium fluoride (LiF) nano-powder. Chemical Engineering
Communications, 194: 1022–1028.
[19] Brown, M. E. 2001. Introduction to Thermal Analysis: Techniques and Application.
New York: Kluwer Academic Publishers.
[20] Friedman, H. L. 1964. Kinetics of thermal degradation of char-forming plastics from
thermogravimetry. Application to a phenolic plastic. Journal of Polymer Science, Part
C: Polymer Symposium, 6: 183–195.
[21] Vyazovkin, S. 1997. Evaluation of activation energy of thermally stimulated solid-state
reactions under arbitrary variation of temperature. Journal of Computational Chemistry,
18: 393–402.
[22] Senum, G. I. and R. T. Yang. 1977. Rational approximations of the integral of the
Arrhenius function. Journal of Thermal Analysis, 119: 445–447.
[23] Coats, A. W. and J. P. Redfern. 1964. Kinetic parameters from thermogravimetric data.
Nature, 201: 68–69.
[24] Owens, F. J. 1982. Electron paramagnetic resonance study of the role of CuO additives
in altering phase transition behaviour of ammonium nitrate. Journal of Applied Physics,
53: 368–371.
[25] Palmer, D. A. and P. Benezeth. 2004. Solubility of copper oxides in water and steam.
In Proceedings of the 14th International Conference on the Properties of Water and
Steam, August 29–September 3, Kyoto, Japan.
[26] Lane, R. W. and H. J. McDonald. 1946. Kinetics of the reaction between copper
and aqueous ammonia. Journal of the American Chemical Society, 68: 1699–1704.
[27] Ziemniak, S. E., M. E. Jones, and K. E. S. Combs. 1992. Copper (II) oxide solubility
behaviour in aqueous sodium phosphate solutions at elevated temperatures. Journal of
Solution Chemistry, 21: 179–200.
[28] Rosser, W. A., S. H. Inami, and H. Wise. 1964. Decomposition of liquid ammonium
nitrate catalyzed by chromium compounds. Transactions of the Faraday Society, 60:
1618–1625.
[29] Sinditskii, V. P., V. Y. Egorshev, A. I. Levshenkov, and V. V. Serushkin. 2005.
Ammonium nitrate: Combustion mechanism and the role of additives. Propellants,
Explosives, Pyrotechnics, 30: 269–280.
[30] Schmid, R. and V. N. Sapunov. 1982. Non-formal Kinetics: In Search of Chemical
Reaction Pathways. Weinheim, Germany: Verlag Chemie GmbH.
[31] Gunawan, R. and D. Zhang. 2009. Thermal stability and kinetics of decomposition
of ammonium nitrate in the presence of pyrite. Journal of Hazardous Materials,
165: 751–758.
[32] Gorbunov, V. V. and L. F. Shmagin. 1972. Burning of copper (II) tetramine salts.
Combustion, Explosion and Shock Waves, 8: 429–431.
[33] Southern, T. M. and W. W. Wendlandt. 1970. The thermal decomposition of metal
complexes—XX: Some amine copper (II) nitrate complexes. Journal of Inorganic and
Nuclear Chemistry, 32: 3783–3792.
[34] Solymosi, F. and E. Krix. 1962. Catalysis of solid phase reactions effect of doping of
cupric oxide catalyst on the thermal decomposition and explosion of ammonium
perchlorate. Catalysis, 1: 468–480.
[35] Il’chenko, N. I. 1976. Catalytic oxidation of ammonia. Russian Chemical Reviews,
45: 1119–1134.
Ammonium Nitrate 161

[36] Vargeese, A. A. and K. Muralidharan. 2011. Anatase–brookite mixed phase nano TiO2
catalyzed homolytic decomposition of ammonium nitrate. Journal of Hazardous
Materials, 192: 1314–1320.
[37] Kubota, N. and N. Hirata. 1985. Inhibition reaction of LiF on the combustion of
ammonium perchlorate propellants. Symposium (International) on Combustion, 20:
2051–2056.
[38] Jankovic, B., B. Adnadevica, and J. Jovanovica. 2007. Application of model-fitting and
model-free kinetics to the study of non-isothermal dehydration of equilibrium swollen
poly (acrylic acid) hydrogel: Thermogravimetric analysis. Thermochimica Acta, 452:
106–115.
[39] Vyazovkin, S. 2001. Modification of the integral isoconversional method to account for
variation in the activation energy. Journal of Computational Chemistry, 22: 178–183.
[40] Budrugeac, P., D. Homentcovschi, and E. Segal. 2001. Critical analysis of the isoconver-
sional methods for evaluating the activation energy II. The activation energy obtained
from isothermal data corresponding to two successive reactions. Journal of Thermal
Analysis and Calorimetry, 63: 465–469.

You might also like