Anda di halaman 1dari 28

This article was downloaded by: [Yale University Library]

On: 26 February 2013, At: 03:47


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Experimental Nanoscience


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tjen20

Nano-metal oxide: potential catalyst on


thermal decomposition of ammonium
perchlorate
a a
Shalini Chaturvedi & Pragnesh N. Dave
a
Department of Chemistry, K.S.K.V. Kachchh University, Mundra
Road, Bhuj-370 001, Gujarat, India
Version of record first published: 25 May 2011.

To cite this article: Shalini Chaturvedi & Pragnesh N. Dave (2012): Nano-metal oxide: potential
catalyst on thermal decomposition of ammonium perchlorate, Journal of Experimental
Nanoscience, 7:2, 205-231

To link to this article: http://dx.doi.org/10.1080/17458080.2010.517571

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-


conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Journal of Experimental Nanoscience
Vol. 7, No. 2, March–April 2012, 205–231

Nano-metal oxide: potential catalyst on thermal decomposition of


ammonium perchlorate
Shalini Chaturvedi and Pragnesh N. Dave*

Department of Chemistry, K.S.K.V. Kachchh University, Mundra Road, Bhuj-370 001,


Gujarat, India
(Received 5 May 2010; final version received 15 August 2010)
Downloaded by [Yale University Library] at 03:47 26 February 2013

In this review, an attempt to collect the summarised data of literature on catalytic


effect of nano-oxides, such as mono oxides, mixed oxide, binary and ternary
ferrites and rare earth metal oxides on the thermal decomposition of ammonium
perchlorate (AP) is made. Influence of size effect of oxides on thermal
decomposition of AP and comparison of bulk and nanosized oxides is also
discussed here. Several experimental results revealed that due to small size and
large surface area nanosized metal oxides are more potential catalysts on thermal
decomposition of AP compared to their bulk size oxides.
Keywords: ammonium perchlorate; thermal decomposition; metal oxide;
nanoparticle; catalytic activity

1. Introduction
Metal oxides are an important class of chemicals having a wide range of applications in
many areas of chemistry, physics and material science [1–5]. In technological applications,
metal oxides are used in the fabrication of microelectronic circuits, sensors and fuel cells
and as catalysts [6]. In the emerging field of nanotechnology, our goal is to make
nanostructures with special properties with respect to those of bulk species [7–12]. Metal
oxide nanocrystals can exhibit unique physico-chemical properties due to their nanosize
and high density of cover or edge surface sites. Nanoscale oxide particles are gaining
increasing technical importance in classic areas of application [13], such as catalysts,
passive electronic components or ceramic materials. Metal oxide nanoparticles are also
widely used in industrial applications as catalysts, ceramic, pigments and so on. In
addition, they play an important role in the selective surface modification of various
substrates in the form of coatings [14,15].
Composite solid propellants (CSPs) are the major source of chemical energy in space
vehicles and missiles. Ammonium perchlorate (AP) is widely used as an oxidiser in CSPs
[16,17]. The ballistics of a composite propellant can be improved by adding a catalyst,
such as ferric oxide (Fe2O3), copper oxide (CuO), copper chromite (CuO  Cr2O3) and

*Corresponding author. Email: pragneshdave@gmail.com

ISSN 1745–8080 print/ISSN 1745–8099 online


ß 2012 Taylor & Francis
http://dx.doi.org/10.1080/17458080.2010.517571
http://www.tandfonline.com
206 S. Chaturvedi and P.N. Dave

nickel oxide (NiO), etc., which accelerates the rate of decomposition of AP [18–21].
Recent investigations have shown that nanoparticles of transition metal oxides
(TMOs), without any agglomeration can increase the burning rate [22] of propellants.
The efficiency of catalytic action increases sharply in nanosized oxide particles than
microscale oxide particles [23,24]. The size distribution, morphology and nanostructure of
particles are very important characteristics and they do affect the kinetics of decompo-
sition of AP.

2. What are nanomaterials?


Nanomaterials are those which have structured components with at least one dimension
less than 100 nm. Materials that have one dimension in the nanoscale (and are extended in
the other two dimensions) are layers, such as a thin films or surface coatings. Some of the
Downloaded by [Yale University Library] at 03:47 26 February 2013

features of computer chips come in this category. Materials that are nanoscale in two
dimensions (and extended in one dimension) include nanowires and nanotubes. Materials
that are nanoscale in three dimensions are particles, for example precipitates, colloids and
quantum dots (tiny particles of semiconductor materials). Nanocrystalline materials, made
up of nanometre-sized grains, also fall into this category. Some of these materials have
been available for some time; others are genuinely new [25–27].
Two principal factors cause the properties of nanomaterials to differ significantly from
other materials: increased relative surface area and quantum effects. These factors can
change or enhance properties such as reactivity, strength and electrical characteristics. As
the particle decreases in size, a greater proportion of atoms are found at the surface
compared to those inside. For example, a particle of size 30 nm has 5% of its atoms on its
surface, at 10 nm 20% of its atoms and at 3 nm 50% of its atoms. Thus nanoparticles have
much greater surface area per unit mass compared with larger particles. As growth and
catalytic chemical reactions occur at surfaces, this means that a given mass of material in
nanoparticulate form will be much more reactive than the same mass of material made up
of larger particles [27].

2.1. Properties of nanomaterials


In tandem with surface–area effects, quantum effects can begin to dominate the properties
of matter as size is reduced to the nanoscale. These can affect the optical, electrical and
magnetic behaviour of materials, particularly as the structure or particle size approaches
the smaller end of the nanoscale. Materials that exploit these effects include quantum dots,
and quantum well lasers for optoelectronics.
For other materials, such as crystalline solids, as the size of their structural components
decreases, there is much greater interface area within the material; this can greatly affect
both mechanical and electrical properties. For example, most metals are made up of small
crystalline grains; the boundaries between the grain slow down or arrest the propagation
of defects when the material is stressed, thus giving it strength. If these grains can be made
very small, or even nanoscale in size, the interface area within the material greatly
increases, which enhances its strength. For example, nanocrystalline nickel is as strong as
hardened steel [27].
Journal of Experimental Nanoscience 207

2.2. Preparation and characterisation of nanomaterials


2.2.1. Preparation
The synthesis of new materials made of particles, rods and wires with dimensions in the
nanometre scale is among the most active areas of research in science due to the unique
properties of these materials compared to conventional materials made from micrometre-
sized particles. Many technologies have been explored to fabricate nanostructures and
nanomaterials. These technical approaches can be grouped in several ways [28].
One way is to group them according to growth media:
(1) Vapour phase growth, including laser reaction pyrolysis for nanoparticle synthesis
and atomic layer deposition for thin film deposition.
(2) Liquid phase growth, including colloidal processing for the formation of
nanoparticles and self-assembly of monolayers.
Downloaded by [Yale University Library] at 03:47 26 February 2013

(3) Solid phase formation, including phase segregation to make metallic particles in
glass matrix.
(4) Hybrid growth, including vapour–liquid–solid growth of nanowires.
Another way is to group the techniques according to the form of the products:
(1) Nanoparticles by means of colloidal processing, flame combustion and phase
segregation.
(2) Nanowries or nanowires by template-based electroplating, solution–liquid–solid
growth and spontaneous anisotropic growth.
(3) Thin film by molecular beam epitaxy and atomic layer deposition.
(4) Nanostructured bulk materials by self-assembly of nanosized particles.
Several methods are reported in the literature for the preparation of nanomaterials,
such as co-precipitation, sol–gel processing, microemulsions, template synthesis, etc.

2.2.2. Characterisation
Nanoparticle characterisation is necessary to establish understanding and control of
nanoparticle synthesis and applications. Characterisation is done by using a variety of
different techniques, mainly drawn from materials science. Common techniques are
transmission electron microscopy, scanning electron microscopy (TEM, SEM), atomic
force microscopy (AFM), dynamic light scattering (DLS), X-ray photoelectron spectros-
copy (XPS), powder X-ray diffractometry (XRD), Fourier transform infrared (FT-IR)
spectroscopy, matrix-assisted laser desorption-time-of-flight (MALDI-TOF) mass spectro-
metry and ultraviolet-visible spectroscopy, etc. [27].

3. Ammonium perchlorate
Ammonium perchlorate (NH4ClO4) is the most widely used energetic material. It is an
important oxidiser used in solid rocket propellants known as ammonium perchlorate
composite propellants (APCP) and the Space Shuttle Solid Rocket Boosters, manufac-
tured by Alliant Techsystems (ATK), as well as many other solid rockets including some
fireworks, amateur and hobby high powered rockets, and larger rockets used for space
208 S. Chaturvedi and P.N. Dave

launch and military purposes. Unfortunately, AP is also one of the least understood
energetic materials. Over the past several decades, numerous papers have been devoted to
the decomposition mechanism and structural properties of AP. The decomposition of AP
is rather complicated; mainly because this simple molecule consists of four different
elements N, H, Cl and O. If one considers all the potential oxidation states of these four
elements, over 1000 possible chemical reactions can be written for the decomposition of
AP [29]. In addition, what makes AP completely different from other propellant and
explosive ingredients is the presence of chlorine. Most, if not all, other energetic materials
consist of only carbon, hydrogen, nitrogen and oxygen. (Exceptions to this are the recently
developed organic difluoramino structures.) Furthermore, impurities play a significant
role in the decomposition, combustion and burn rate of AP materials. Because of these
complications, the reaction chemistry of AP is still largely not understood [30]. AP is the
most common oxidiser in CSPs. Thermal decomposition characteristics of AP influence
Downloaded by [Yale University Library] at 03:47 26 February 2013

the combustion behaviour of the propellants [31,12]. AP-based CSPs require combustion
modifiers to achieve higher burning rates and conventionally TMOs are used as the
burning rate modifiers [31–34]. Previous studies suggested that the burning rate modifiers
are active mainly in the condensed phase and hence activity of the catalysts during
condensed phase thermolysis of AP can be a good indicator to the catalytic activity of the
additive during combustion of CSPs [31].

3.1. Properties
AP is a white crystalline substance; it is used as the most common oxidiser in CSPs. The
characteristic of the thermal decomposition of AP are believed to influence the
performance of CSPs and are remarkably sensitive to additives. The TMOs are used as
a burning rate modifier for CSPs [35–37]. Two most important crystal modifications of AP
are: orthorhombic (region of existence: T5240 C; Figure 1) and cubic one, which is stable
at temperatures T4240 C. Heat of phase transition of AP from orthorhombic to cubic
modification is 11.3 kJ mol1 [39]. Investigation with the help of NMR, neutron diffraction

Figure 1. Combined plot of TGA and DSC experiments carried out at a heating rate of 5 C min1
of AP. Source: Ref. [38].
Journal of Experimental Nanoscience 209

[40] and ESR [41] shows that both the ammonium and perchlorate ions can rotate
freely. In the cubic structure, both ions rotate; in the orthorhombic one, only ammonium
does.
It was assumed that the phase transition of the first type at 240 C is preceded by the
transition of the second type, which is the main reason of deviation from the normal
behaviour observed in AP under thermal decomposition [42].
Polymorphous transition in AP crystals was studied in [43]. Several experimental
results show that there are two mechanisms of the transformation of the rhombic
modification into cubic one. At low temperature, the transition follows atom-by-atom
mechanism. When temperature is increased up to 243 C, the transition is of martensite
type. In both cases, the polymorphous transition proceeds through the formation and
growth of the nuclei of the new phase; the growth of martensite nuclei occurs not
continuously but jumpwise. The interface boundary can be coherent during the transition
Downloaded by [Yale University Library] at 03:47 26 February 2013

form rhombic modification into the cubic one [43]. During this process, the (011) plane of
the rhombic modification corresponds to the (001) plane of the cubic one. One more
curious fact was observed during investigation of thermal decomposition of AP: the rate of
thermal decomposition sharply decreases at the moment of phase transition [44,45]
because rearrangement of the lattice from the rhombic into cubic at the moment of
transition results in sharp worsening of the conditions for the accumulation of products
catalysing thermal decomposition in the system. In earlier works [46], this effect was
explained by a decrease in the lattice parameters.
The experimental results show the behaviour of AP under the pressure up to 25 GPa at
temperatures slightly above 240 C (the point of transition from the rhombic modification
to the cubic one); melting is observed [47].

3.2. Mechanism of thermal decomposition


Several studies were carried out to understand the mechanism of decomposition of AP
[48,49]. Reviews by Hall and Pearson [50], Jacobs and Whitehead [51] and Kishore and
Sunitha [52] give good accounts of the mechanism of thermal decomposition of AP.
Several mechanisms are proposed, but still it remains a matter of debate.

3.2.1. Electron transfer mechanism


According to this mechanism [17,46], decomposition occurs due to electron transfer from
anion to cation.
0 0
ClO þ
4 þ NH4 ! ClO4 þ NH4 :

Probability of electron realisation is higher because distance between the ions is


smaller. Because of this, an efficient acceptor of electrons is considered not to be any
ammonium ions but only those located in interstices.
After accepting an electron, ammonium radical decomposed into ammonia and
hydrogen atom.
NH04 ! NH3 þ H:
210 S. Chaturvedi and P.N. Dave

Hydrogen migrates over the lattice. Electron migrates exactly in the same manner over
the anion sublattice:
ClO04 þ ClO 
4 ¼ ClO4 þ ClO4 :
0

As a result of the interaction between ClO4 radical and H, HClO4 is formed. It may
continue interacting with H:
HClO4 þ H ! H2 O þ ClO3 :

The ClO3 radical is a trap for electrons. Having trapped an electron, it is transformed
into ClO þ
3 ion. After that, chlorite ion and ClO4 radical can decompose, interact with NH4
ions, etc. As a result of interaction, secondary products are formed; among them, the
major ones are chlorine, nitrogen hemioxide and water.
Raevsky and Manelis [53] consider electron transition within the energy band theory,
Downloaded by [Yale University Library] at 03:47 26 February 2013

that is, the transition of an electron from the valence band into conduction band.
The AP is a typical dielectric; hence at low temperature of thermal decomposition, the
process cannot be sustained by electron transfer because of its low probability. This rules
out the possibility of electron transfer mechanism. The primary products detected in the
experiments by different researchers were ammonia and perchloric acid. This allowed the
assumption that the primary stage of process of thermal decomposition of AP is proton
transfer.

3.2.2. Proton transfer mechanism


Similar values of activation energies of thermal decomposition and sublimation [54],
identical composition of the products of decomposition [55] and sublimation [56],
inhibition of the reaction in ammonia vapour [55] and acceleration in the vapour of
perchloric acid [57,58], favour proton transfer mechanism. A scheme of thermal
decomposition of AP proposed by Jacobs [54] can be presented as a version of proton
mechanism [17]:

NHþ
4 þ ClO4 ¼ NH3 ðaÞ þ HClO4 ðaÞ ¼ NH3 ð gÞ þ HClO4 ð gÞ:

Step I involves a pair of ions in perchlorate ammonium lattice. Step II involves the
decomposition step that starts with proton transfer from the cation NHþ 4 to the anion
ClO 4 via molecular complex and which then in Step III decomposes into ammonia and
perchloric acid (Figures 1 and 2). The molecules of NH3 and HClO4 either react in the
adsorbed layer on the surface of perchlorate or desorb and sublime interacting in the gas
phase. Many reactions occur rapidly in the gas phase between NH3 and HClO4, forming
the side products, such as O2, N2O, Cl2, NO and H2O at low temperature (5350 C). An
essential feature of the process of thermal decomposition of AP is that the process occurs
not at the very surface but in pores beneath it at a distance of about several micrometres.
This is how the thermal decomposition of AP differs from its sublimation.
During a process occurring in the adsorbed layer, it was assumed and reported in
earlier reports [17] that perchloric acid is desorbed more rapidly than ammonia, which
causes incomplete oxidation of ammonia, creating a saturated atmosphere [54] of NH3. As
a result, high-temperature decomposition decelerates and undergoes incomplete transfor-
mation. During the second exothermic decomposition, NO, O2, Cl2 and H2O products
were formed in the gas phase reactions.
Journal of Experimental Nanoscience 211
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 2. A DSC comparison between AP in an open sample holder and that in a sample holder
with a pierced lid at 12 C min1. Source: Ref. [59].

The process of thermal decomposition both at low and high temperatures has a

common start, which is proton transfer from the cation NHþ 4 to the anion ClO4 . The
difference between the decomposition of low-temperature orthorhombic modification and
the high-temperature cubic one arises when the secondary processes occur. These processes
during the decomposition of orthorhombic and cubic modifications differ both in the
character of chemical processes and in the topography of their course. The decomposition
of the low-temperature orthorhombic modification proceeds in the pores beneath the
surface in the sites where the secondary products of dissociation can be accumulated, and
in the sites where the conditions for regeneration of active centres exist, either due to
pressure of the gases formed in a pore or due to decomposition of perchloric acid and
interaction of the products of its decomposition with ammonia. It is the conditions for
secondary processes, providing the feedback between the accumulation of reaction
products and the formation of new active centres that act as the main reasons of
decomposition of the orthorhombic modification. Because of this, the action of additives,
irradiation and other factors affecting the reactivity of the initial crystal is mainly
restricted to the initial period of reaction, its initiation. The situation changes when higher
temperature is employed, and when we come across decomposition of the cubic
modification. In this case, the major role is played by the processes taking place on the
surface of the crystal: adsorption and desorption of ammonia and chloric acid; so, the
role of the primary processes connected with proton transfer increases. A result is larger
effect of the introduction of dopants, which change the concentration of protons in the
lattice during the decomposition of the cubic modification, and stronger effect of
ammonia. The hypothesis formulated by Kaidymov [60,61] is likely to be worth
mentioning; it says that the catalytic activity of the orthorhombic modification with
respect to the decomposition of perchloric acid, an important stage through which thermal
decomposition proceeds, is much higher than that of the cubic modification. So, in spite of
212 S. Chaturvedi and P.N. Dave

advances in investigating the mechanism of thermal decomposition of AP, the mechanism


is still unclear.

4. Effect of nanosized oxides as compared to bulk sized oxides


A specific feature of thermal decomposition of AP is extremely high sensitivity to the
action of various additives. These additives can accelerate or decelerate the process of low-
temperature thermal decomposition, which is important for the storage of AP and mix
compositions based on it. The additives may affect the deflagration delay time. It was
assumed that there existed a correlation between the effect of the additives on the rate of
thermal decomposition and their influence on the combustion rate of mix compositions
incorporating AP as the major component. This stimulated the attention of many research
groups, especially those dealing with the development of solid propellants, to the problem.
Downloaded by [Yale University Library] at 03:47 26 February 2013

The catalytic activity of TMOs in the thermal decomposition of AP has been reported by
several researchers [31–35]. It is generally observed that since catalytic activity is primarily
a surface phenomenon, size reduction of the catalyst increases the surface area and hence
the catalytic activity is also increased. Activities of the solid catalysts increase manyfold by
size reduction into the nanometre scale. These researchers discovered that the most
efficient additive accelerating thermal decomposition of AP is manganese dioxide. A less
efficient catalyst of thermal decomposition is ferric oxide [17]. Aluminium oxide did not
affect the rate of thermal decomposition. Here, the summarised reported data of effect of
nanosized TMOs on thermal decomposition of AP are given.
Both mono-and divalent oxides of copper were studied by the researcher as catalysts of
low-temperature thermal decomposition of AP. Jacobs and Kureishy [62] showed that
addition of Cu2O decreases deflagration delay time of perchlorate both due to the catalytic
action itself and due to heat evolution during oxidation to copper oxide. The use of Cu2O
as efficient catalyst of thermal decomposition of perchlorate is also described in [63,64].
In the presence of copper(II) oxide, thermal decomposition of AP without additive starts
at 180 C, while usually it starts at a temperature about 200 C; catalytic action of copper
oxide manifests itself even in the case when its amount is only 0.5% mass. Now several
researchers [37,39] studied the effect of nanosized CuO on the AP thermal decomposition
and results show that catalytic activity increased by many times as compared to their bulk
size. Xu et al. [65] demonstrated that CuO nanocrystals exhibit a particular chemical
reactivity due to their high concentration of dislocations and large surface areas. In
addition, the different morphologies of nanocrystals are expected to have different effects
on the thermal decomposition of AP [66]. A clear comparison is studied by Singh et al.
[12]; according to their investigations, catalytic activity of CuO is enhanced approximately
by five times (Table 1). They have also studied the effect by varying the per cent content of
additive in AP; the results show that 0.5% of nanosized CuO is a better catalyst than 1%
of their bulk size (Figure 3) and with the increase percentage of additive, rate of
decomposition of AP increased. Position of second exothermic peak of AP shifted
approximately to 120 C lower as compared to pure AP.
The catalytic action of nickel oxide on thermal decomposition of AP is similar to the
action of copper oxides. Catalytic activity of nickel oxide can be changed by doping or by
using thermal treatment. With temperature increase of above 500 C, the catalytic activity
of the additive decreases substantially [37]. As compared to bulk size, the effect of
Journal of Experimental Nanoscience 213

Table 1. Decomposition rate and catalytic activities of AP, AP þ TMONC and AP þ TMO (old).

Decomposition Decomposition
rate (for 25% rate (for 25%
decomposition decomposition CA (catalytic CA (catalytic
Temperature min1) of min1) of activity) of activity) of
Sample ( C) TMONC TMO (old) TMONC TMO (old)

AP 260 0.142 0.142 – –


AP þ CuO (1 wt%) 260 0.819 0.167 5.77 1.174
AP þ NiO (1 wt%) 260 0.745 0.156 5.25 1.100
AP þ Co2O3 (1 wt%) 260 1.600 0.180 11.27 1.268
AP þ MnO2 (1 wt%) 260 0.595 0.151 4.19 1.062

Note: TMONC, transition metal oxide nanocrystal. Source: Ref. [12].


Downloaded by [Yale University Library] at 03:47 26 February 2013

nanoparticles (4 nm) of nickel oxide on the thermal decomposition of NH4ClO4 was
studied by Singh et al. [12]. According to their investigation, catalytic activity enhanced
approximately four times (Table 1) and with the increased percentage of additive, rate of
decomposition of AP increased (Figure 3). Similar investigation by Wang et al. [37]
shows that complete decomposition of AP with nanosize NiO (10 nm) done approxi-
mately at 350 C in the presence of bulk-sized NiO occurs at 400 C and it was also shown
that addition of about 2 wt% of NiO to perchlorate brought about decrease in
decomposition temperature and increase of heat of the reaction from 580 to 1490 J g1
(Figure 4).
Iron (III) oxide is a catalyst of thermal decomposition of AP mainly at high
temperature (320–380 C). The effect of the catalyst is observed at the later stages of the
process [67,68]. It was shown that the efficiency of catalytic action of ferric oxide increases
sharply when passing from microscale of the size of catalyst particles to the nanosize [37].
Many properties of nanosized -Fe2O3, such as high surface area, large number of surface
atoms and oxygen vacancies on its surface enable it to act as an excellent catalyst [69].
Above all, it is structurally simple, highly stable, easy to synthesise and inert to side
reactions. With reduced particle size of Fe2O3, the surface-to-volume ratio increases.
Hence, nano-ferric oxide becomes more efficient and active in thermal decomposition of
AP. The authors studied the thermal decomposition of AP/hydroxyl-terminated polybu-
tadiene (HTPB)-based composite propellant both in the presence and absence of nano-
ferric oxide and achieved high heat release, up to 40% with the retention of well-separated
necklace-like structure of -Fe2O3 in HTPB [70]. Joshi et al. [24] studied the effect of
various percentages of nanosized Fe2O3 as additive in AP. Results show enhanced thermal
decomposition with increased percentage of additive. The 3.5 nm -Fe2O3 particles showed
pronounced effect in lowering the high temperature decomposition to 40 C (0.5 wt%),
35 C (1 wt%), 59 C (2 wt%) and 77 C (5 wt%) compared to pure AP. Such a marked
reduction in high temperature exotherm is attributed to the presence of large number of
active sites, higher surface area and smaller particle size of synthesised -Fe2O3. The
lowering in decomposition temperature was observed with respect to decrease in particle
size from 30 to 3.5 nm at 0.5, 1 and 2 wt% compositions (Table 2).
214 S. Chaturvedi and P.N. Dave

CuO
80
AP

AP + 1% CuO (old)

% Mass loss AP + 1% CuO


40
AP+ 2% CuO
AP + 0.25% CuO

AP + 0.5% CuO
0
0 50 100 150 200
Time (min)
NiO
60
Downloaded by [Yale University Library] at 03:47 26 February 2013

AP
AP +1% NiO
% Mass loss

40
AP + 2% NiO
AP + 0.5% NiO
20
AP +0.25% NiO
AP + 1% NiO (Old)
0
0 50 100 150 200
Time (min)

MnO2
60
AP
AP + 1% MO
% Mass loss

40
AP + 2% MO
AP + 0.5% MO
20
AP + 0.25% MO
AP + 1% MO (Old)
0
0 50 100 150 200
Time (min)
Co2O3
60
AP
AP+1% CO
% Mass loss

40
AP+2% CO
AP + 0.5% CO
20
AP +0.25% CO
AP + 1% CO (Old)
0
0 50 100 150 200
Time (min)

Figure 3. Isothermal TG of various percentages of metal oxide nanoparticles in AP at 260 C.


Source: Ref. [12].
Journal of Experimental Nanoscience 215
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 4. TG curves for: (a) AP þ NiO nanoparticles; (b) AP þ bulk NiO and (c) pure AP.
Source: Ref. [37].

Table 2. Kinetic parameters of Fe2O3 of different sizes in various compositions of AP.

HTD Ea A k DH
( C) (kJ mol1) (min1) (s1) (kJ g1)

AP þ 30 nm
Fe2O3
0.5 wt% 445 108.08 2.27  107 1.37  103 1.105
1.0 wt% 441 80.47 1.58  105 2.58  103 3.814
2.0 wt% 410 144.04 5.55  1010 5.49  103 2.987
5.0 wt% 373 121.50 2.13  109 4.97  103 1.447
AP þ 3.5 nm
Fe2O3
0.5 wt% 425 122.04 1.54  106 4.95  103 1.115
1.0 wt% 432 105.30 1.00  109 3.69  103 3.494
2.0 wt% 408 102.53 1.84  107 4.98  103 4.574
5.0 wt% 390 104.75 4.35  107 4.02  103 1.244
Compositions
Pure AP 467 324.24 4.31  1027 5.61  104 0.834
AP þ 2% commercial Fe2O3 460 174.59 1.99  1012 2.13  103 1.782

Source: Ref. [24].

Xu et al. [71] investigation shows that different shapes of additive also affect the
thermal decomposition of AP. For nanorods, irregular particles and micro-octahedrons of
-Fe2O3, the temperature shift for high temperature exothermic peaks of AP was 48.9 C,
16.8 C and 10 C, respectively (Figure 5). Therefore, -Fe2O3 nanorods exhibit better
216 S. Chaturvedi and P.N. Dave
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 5. DSC curves of AP decomposition in the absence and presence of different -Fe2O3
samples. (a) pure AP; (b) AP þ sample B (micro-octahedrons); (c) APþ sample C1 (irregular
particles) and (d) AP þ sample A (nanorods). Source: Ref. [71].

catalytic activity than the other two samples, while the catalytic performance of irregular
particles is slightly better than that of micro-octahedron and the particle with smaller size
possess better performance and more active sites, which can further enhance the rate of
heterogeneous decomposition of deprotonised HClO4 gas on the surfaces of catalyst
particles.
Copper chromite is also used as a catalyst in the thermal decomposition of AP;
investigation was carried by authors [72–75]. According to their investigation, at low
temperatures up to 270 C, copper chromite has a slight effect on the rate, increasing it by a
factor of two. However, at higher temperature (280–340 C), its effect becomes more
noticeable. Copper chromite affects the deflagration delay time by decreasing it. Since
industrial catalysts of combustion based on copper chromite contain not only copper
chromite but also other compounds of copper and chromium. Nanosized copper chromite
(CuCr2O4) showed best catalytic effects as compared to their bulk size [76] in lowering
the high temperature decomposition and heat releases of 5.430 kJ g1. The decrease in the
activation energy and the increase in the rate constant for copper chromite confirmed the
enhancement in catalytic activity of AP. A comparative investigation was also done here
between copper chromite and copper oxide; result shows that the nano-copper chromite
(CuCr2O4) showed best catalytic effects as compared to nano-cupric oxide (CuO) in
lowering the high temperature decomposition by 118 C at 2 wt%. Heat releases of 5.430
and 3.921 kJ g1 were observed in the presence of nano-CuO and CuCr2O4, respectively.
Singh et al. [12] investigation also confirms the size effect of TMOs on their catalytic
activity; its result shows that the catalysts affect both LTD and HTD of AP. Higher
percentage mass loss is observed in the thermal curves for AP þ TMO mixtures than that
for pure AP. Onset temperatures for HTD for the mixtures are considerably lower than
Journal of Experimental Nanoscience 217
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 6. Non-isothermal TG thermal curves of AP and its mixtures with nanosize TMO (1% by
mass) at 10 C min1 under air atmosphere.
Source: Ref. [12].

that for pure AP. Gasification of AP in the presence of catalysts during HTD not only
begins early, but is also complete at lower temperatures. Thus, endset temperatures of
about 505 C, 490 C, 460 C, 453 C and 445 C were observed for AP, AP þ MnO2,
AP þ NiO, AP þ CuO and AP þ Co2O3, respectively (Figure 4). Differential thermal
analysis (DTA) thermal curves for decomposition of AP in the presence of metal oxide
nanocrystals show noticeable changes in the decomposition pattern (Figure 6) and confirm
the TG result. The comparative data for the catalytic activities of different TMOs of nano
and bulk size are reported. Nanosized oxides were found to enhance the thermal
decomposition of AP comparatively to a greater extent. The order of catalytic activity for
AP thermal decomposition was found to be:

Co2 O3 4 CuO 4 NiO 4 MnO2 :


Thermolysis of AP is also measured by the ignition delay of AP in the presence of these
catalysts. AP in the presence of TMO nanocrystal (1% by mass) as catalyst ignites with
noise and fumes under rapid heating. These metal oxide nanocrystals act as good catalyst
during deflagration of AP with decreasing activation energy for ignition.
Spinel oxides (ferrospinels or ferrites) are of considerable interest due to their diverse
applications in optical, electronic, catalytic and magnetic materials, etc [77]. Ferrite spinels
may also contain a mixture of two divalent metal ions, in which the ratio of these divalent
metal ions may vary, are called mixed ferrite. The cation distribution of mixed ferrite
significantly affects the surface properties of ferrospinels making them catalytically active.
Because of their small size and large number of cations, for co-ordination sites, nanosized
ferrites are capable of enhancing the rate of chemical reactions. Singh et al. [78]
218 S. Chaturvedi and P.N. Dave

investigation results show that ternary ferrites nanoparticles can be used as catalysts in AP
thermal decomposition. The DTA and non-isothermal thermogravimetry (TG) curves in
static air (Figures 7 and 8, respectively) and in flowing N2 atmosphere (Figure 9) for pure
AP, clearly indicate that thermal decomposition of AP takes place in two steps [46,73]. TG
thermal curves (Figures 4 and 5) for AP with ferrites do confirm that the catalysts affect
both LTD and HTD processes and further gasification of AP in the presence of catalyst
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 7. DTA thermal curves of AP and its mixtures with nanosize TMO (1% by mass) at
10 C min1 under air atmosphere. Source: Ref. [12].

Figure 8. Non-isothermal TG thermal curves of AP and the ferrite nanoparticles (1% by mass) at
10 C min1 under static air.
Note: NiZnFe2O4 (NZF), CuCoFe2O4 (CuCoF), NiCuFe2O4 (NCuF), CuZnFe2O4 (CuZF),
CoNiFe2O4 (CoNF) and CoZnFe2O4 (CoZF). Source: Ref. [78].
Journal of Experimental Nanoscience 219

during HTD process not only begins early but is also complete at lower temperature
(30–60 C). It is inferred that the temperature difference observed under static air and inert
atmosphere is due to experimental conditions. DTG peaks also confirm the lowering of the
temperature of HTD (Table 3). Differential scanning calorimetry (DSC) and DTA results
also show the same result. IR spectra results of pure AP and AP with nanoferrites results
are shown in Figure 4, which exhibit enhanced intensity of Cl–O (1087 cm1) and the
mode of ClO 1
4 (625 cm ) with temperature. Up to 523 K, as (N–H) 3179 cm
1
and (N–H)
1
bending 1407 cm , intensity decreases gradually; but above 523 K, a marked decrease
occurs. The ClO 4 modes are the most temperature sensitive ones because of the increased
rotational motion at higher temperatures. These changes in both the cation and anion
internal modes could be related to increased ion liberation as well as LTD decomposition.
Addition of a catalyst to AP enhances the rate, of N–H bond heterolysis in NHþ 4 and, O–H
bond making in HClO4. Decomposition reactions of energetic materials often involve both
Downloaded by [Yale University Library] at 03:47 26 February 2013

bond-breaking and bond-forming steps. Approximately, 1000 reactions may be involved in


the decomposition and combustion of AP [73,79], because of the four elements and the full
range of oxidation states utilised by nitrogen and chlorine and supports to proton transfer
mechanism.
Catalytic properties of quaternary ferrite nanoparticle on AP, composite solid rocket
propellants (CSPs) and HTPB were studied [80], which enhance in thermal decomposition
of AP, CSPs and HTPB observed, while the burning rate of CSPs in the presence of ferrites
increases manyfold. Singh et al. [81] studied the catalytic properties of binary ferrites on
thermolysis of AP; their investigation shows HTD peaks observed at approximately
100 C, a lower temperature as compared to pure AP. TG, ignition delay data confirms the
catalytic action of binary ferrites. The order of catalytic action of ferrites observed was
CoFe2 O4 4 CuFe2 O4 4 NiFe2 O4 :

Figure 9. Non-isothermal TG thermal curves of AP and the ferrite nanoparticles (1% by mass) at
10 C min1 under N2 atmosphere.
Note: NiZnFe2O4 (NZF), CuCoFe2O4 (CuCoF), NiCuFe2O4 (NCuF), CuZnFe2O4 (CuZF),
CoNiFe2O4 (CoNF) and CoZnFe2O4 (CoZF). Source: Ref. [78].
220 S. Chaturvedi and P.N. Dave

Table 3. DTG, DSC and DTA phenomenological data of AP and AP with ferrite nanocrystallites.

DTG DSC DTA

Peak Peak Peak


Samples temperature ( C) Nature temperature ( C) Nature temperature ( C) Nature

AP 300 Exo 300 Exo 305 Exo


370 Exo 358 Exo 443 Exo
NZF 295 Exo 285 Exo 300 Exo
342 Exo 338 Exo 370 Exo
CoNF 295 Exo 283 Exo 295 Exo
340 Exo 331 Exo 370 Exo
CoZF 295 Exo 283 Exo 295 Exo
Downloaded by [Yale University Library] at 03:47 26 February 2013

340 Exo 336 Exo 365 Exo


CuCoF 295 Exo 285 Exo 300 Exo
344 Exo 338 Exo 365 Exo
CuZF 295 Exo 287 Exo 295 Exo
341 Exo 330 Exo 365 Exo
NCuF 295 Exo 282 Exo 295 Exo
342 Exo 322 Exo 370 Exo

Source: Ref. [78].

The effect of spinels NiFe2O4, NiFeCrO4 and NiCr2O4, obtained by solid-phase


interaction between the corresponding oxides, on thermal decomposition of AP and its
mixtures with polystyrene was also studied [82]. Relative efficiency of the catalysts for pure
perchlorate decreases in the series NiCr2O44NiFeCrO44NiFe2O4 and NiFeCrO44
NiCr2O44NiFe2O4 for the mixture. The catalytic action of these catalysts on polystyrene
decomposition is characterised by the series
NiCrO4 4 NiCr2 O4 4 NiFe2 O4 :
Yu et al. [83] studied the catalytic effect of orthorhombic structural perovskite NdCrO3
nanocrystals with a size of 60 nm for thermal decomposition of AP. This was investigated
by DSC and TG–MS. The results revealed that the NdCrO3 nanoparticles had effective
catalysis on the thermal decomposition of AP. Adding 2% of NdCrO3 nanoparticles to AP
decreased the temperature of thermal decomposition by 87 C and increased the heat of
decomposition from 590 to 1073 J g1. Gaseous products of thermal decomposition of AP
were NH3, H2O, O2, HCl, N2O, NO, NO2 and Cl2. The mechanism of catalytic action was
based on the presence of superoxide ion O 2 on the surface of NdCrO3, and the difference
of thermal decomposition of AP with 2% of NdCrO3 and pure AP was mainly caused by
the different extent of oxidation of ammonium.
Oxides of cobalt are also studied as catalyst in thermal decomposition of AP as
reported [84]; when AP was mixed with CoO nanocrystals, the HTD process disappeared
completely. CoO nanocrystals were prepared by a solvothermal reaction. The average
particle sizes of CoO nanocrystals were about 22 nm prepared in the pure ethanol solution
or 44 nm in the presence of a small amount of hydrazine. As-prepared samples were
Journal of Experimental Nanoscience 221
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 10. TG, DTG and DSC curves of pure AP (a) and AP þ 2% CoC2O4 (b).
Source: Ref. [85].

explored as the additives to AP. It is found that the decomposition heat of AP was
significantly improved as followed by an apparent decrease in the maximum decompo-
sition temperature in comparison with those of the additives of 18 nm Co3O4 of pure
phase. The decomposition temperature and heat of AP systematically varied with the
content of CoO additives and the use of the lids, which are ascribed to the dual effects of
intermediate product and the release of O2 ions from the surface Co3O4 thin layers.
Another investigation [85] result shows CoC2O4 have intensive catalytic effect on the
thermal decomposition of AP. It can raise the decomposition speed of AP, increase the
exothermic heat and lower the decomposition temperature. Adding 2% of CoC2O4 in AP
decreases the high temperature peak of AP decomposition by 104 C and increases the
exothermic heat of decomposition from 655 to 1469 J g1 (Figure 10). It can be foreseen
that the transition metal oxalate will be an attractive catalyst used in the AP-based
propellant. In situ catalytic thermal decomposition of AP has been investigated over
222 S. Chaturvedi and P.N. Dave

CoC2O4 and the results show that the new ecological nano-cobalt oxides exhibit better
catalytic performance in the thermal decomposition of AP. The oxidation of adsorbed
ammonia by cobalt oxides via the superoxide active centres takes place on the surface of
cobalt oxide. The presence of O2 can accelerate the oxidation thermal decomposition
process of AP, with a clear increase of DSC heat release.
The catalytic effect of copper cobaltite was investigated; CuxCo3xO4 spinel in the
range (05x53) on the thermal decomposition of AP has been studied [36]. The results
revealed that the addition of CuO (x ¼ 0.1 or 0.3) led to an observable increase in the
catalytic activity of Co3O4 towards the decomposition of AP. The creation of more holes
within the p-type semiconducting catalyst is responsible for enhancing the decomposition
of AP.
A comparative investigation of catalytic effect of synthesised nanoparticle and
commercial MgO [86] done on the thermal decomposition of AP result shows that MgO
Downloaded by [Yale University Library] at 03:47 26 February 2013

nanocrystalline has no effects on the crystallographic transition temperatures, but


decreases the area of endothermic peak (Figure 11). The number of exothermic peak
decreases from two peaks to one peak; namely the second exothermic peak is superposed
to the first one. The shapes of exothermic peaks all become narrow and sharp. The areas of
exothermic peaks, which indicate the decomposition heat of AP, increase on a large scale.
Comparing with commercial product, nanosized MgO clearly showed a good catalytic
activity on the decomposition of AP following their decomposition heat value. The area of
exothermic peak of nanosized MgO is about 20.1 units; but that of commercial is about
7.5 units, as shown in Table 4. These results indicate that the catalytic decomposition effect
on AP of nanosized MgO is better than that of commercial product.
Catalytic effect of TMOs (CuO, Co2O3)/AP composite nanoparticle which are
prepared by solvent–non-solvent method are also studied [87]. The result revealed that
the TMOs nanoparticles in composite particles can easily facilitate the thermal
decomposition of AP at a higher temperature as compared to the same nanoparticles
which are added in AP by simple mixing (Figures 12 and 13). According to proton transfer
mechanism during the higher temperature decomposition step, nanoparticles absorb the
gaseous reactive molecules on their surface and catalyse the reaction. TMO nanoparticles

Figure 11. Catalytic decomposition curves of AP with different blend ratios of MgO powders.
Source: Ref. [86].
Journal of Experimental Nanoscience 223

added by simple mixing can easily aggregate and cannot be distributed homogeneously
with AP which will affect their catalytic activity. While in TMOs/AP composite
nanoparticles, TMOs are homogeneously covered by AP having large interface area and
they enhance the catalytic effect of TMOs. Investigation shows that Co2O3 has better
catalytic effect at higher temperatures as compared to CuO nanoparticles.
The Ni/TiO2 nanoparticles were prepared by a liquid-phase chemical reduction
method. The results show that Ni particles in Ni/TiO2 nanoparticles exhibit better
dispersion and the size of most Ni particles is 10 nm or so. The catalytic activity of Ni/TiO2
nanoparticles on the thermal decomposition of AP was investigated [88]. Results show that
composite process of Ni and TiO2 can improve the catalytic activity of Ni nanoparticles on
the thermal decomposition of AP, which is mainly attributed to the improvement of Ni
dispersion in Ni/TiO2 nanoparticles. The catalytic activity of Ni/TiO2 nanoparticles
increases with increasing weight ratio of Ni to AP, while TiO2 nanoparticles have reverse
Downloaded by [Yale University Library] at 03:47 26 February 2013

Table 4. Catalytic decomposition data of AP with different MgO powders.

Onset End
Sample temperature ( C) temperature ( C) Intensity Peak area

AP þ 2%A 350 375 2910 20.1


AP þ 2%B 311 365 2410 7.5

Note: A, nanosized MgO; and B, commercial MgO. Source: Ref. [86].

Figure 12. DSC curve 1 – pure AP, 2 – mixture of pure CuO with AP, 3-CuO/AP nanocomposite.
Source: Ref. [87].
224 S. Chaturvedi and P.N. Dave
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 13. DSC curve (a) pure AP, (b) mixture of pure Co2O3 with AP, (c) Co2O3/AP
nanocomposite. Source: Ref. [87].

Table 5. Values of temperature peak, mass loss, exothermic quantity and phase transition enthalpy
of AP and the mixtures.

Phase transition First First mass Second Second mass Total H


Sample H (J g1) Tp ( C) loss (%) Tp ( C) loss (%) (exo) (kJ g1)

Near AP 87 321 14 441 86 0.6


Ap þ LaFeO3 88 317 35 374 63 1.03
AP þ LaCoO3 86 307 98 – – 1.42
AP þ LaNiO3 88 312 37 377 61 1.09

Source: Ref. [90].

effect on thermal decomposition of AP as shown in the investigation [89]. The thermal


decomposition temperature of AP decreased when the specific surface area of TiO2
increased. Al2O3, TiO2 and SnO2 catalyse the decomposition of perchloric acid; these
oxides have no effect on the decomposition of AP. In contrast, cobalt, nickel and copper
oxides which are known as the most active additives have rather moderate activity when
used as catalysts of decomposition of perchloric acid.
Catalytic effect of nanoparticles of perovskite-type oxides LaMO3 (M ¼ Fe, Co
and Ni) are also studied [90]. Result showed that decomposition heat of pure AP is
0.6 kJ g1, while in the presence of LaFeO3 and LaNiO3 heat of decomposition of AP is 1.0
and 1.1 kJ g1, respectively. In the presence of LiCoO3, the observed heat of decompo-
sition of AP is 1.4 kJ g1 (Table 5). This shows that LaCoO3 is obviously a more effective
catalyst than LaFeO3 and LaNiO3. The addition of LaCoO3 lowers the decomposition
Journal of Experimental Nanoscience 225
Downloaded by [Yale University Library] at 03:47 26 February 2013

Figure 14. TG curves for: (a) pure AP; (b) AP þ LaFeO3; (c) AP þ LaCoO3; and (d) AP þ LaNiO3.
Source: Ref. [90].

temperature of AP and increases the speed of decomposition and heat of decomposition


reaction. AP was completely decomposed in lower temperature and short time (Figure 14).
Catalytic activity of LiCoO3 on the burning rate of AP-based propellants was also studied.
By adding 1% LaCoO3 nanocrystals to basic propellant, the burning rate of the modified
propellant was increased around 8%.
Manganese dioxide as the catalyst of thermal decomposition of AP was also the subject
of investigation. In the presence of MnO2 catalyst, transformation degree during thermal
decomposition of AP reaches 100%, while non-catalysed decomposition is usually limited
by a level of 30%. Deflagration delay time decreases, too. Several research works are done
on it [63] and result shows that MnO2 catalyses mainly HTD of perchlorate. The catalytic
activity of mixed valent manganese oxide octahedral molecular sieve (OMS) on the
thermal decomposition of AP also has been studied [91]. Results show that OMS plays a
catalytic role on the thermal decomposition of AP. The decomposition temperature of AP
has been decreased to 145 C and the heat of decomposition has been increased to
0.66 kJ g1 with the addition of 2 wt% OMS. Such effects are mainly attributed to the
mixed valence of Mn3þ and Mn4þ existing in OMS material and its high surface area.
Singh et al. [12] investigated the catalytic activity of nanoparticle of MnO2 which enhanced
the thermal decomposition of AP higher as compared to their bulk size.
Catalytic effect of ZnO on thermal decomposition of AP is also studied. The
mechanism of the accelerating action of ZnO is as follows: (1) the reaction accelerates
because zinc perchlorate originating from the interaction between the oxides and AP form
easily melting eutectics. Due to the transition from the solid state into liquid, the rate of
the process increases [92,93]; (2) the rate of the process increases due to the formation of
intermediate amine compounds [94]; (3) the O 2 ions formed during decomposition
of perchlorates and the surface of the oxides itself are proton traps and thus they simplify
thermal decomposition [95]. Study of the nanocrystal of ZnO on thermal decomposition of
AP [96] has also been done. Investigation revealed that N-doped ZnO sample with peculiar
226 S. Chaturvedi and P.N. Dave

morphology drives the thermal decomposition peak of AP decrease about 163 C with a
strong decomposition heat of about 1.325 J g1, and the activation energy also decreases
from 178.22 to 93.51 kJ mol1. The enhanced catalytic activity of N-doped ZnO sample
can be attributed to oxygen vacancies and other defects induced by the doping of nitrogen.
The effect of ‘p’-type rare earth oxides with partially filled ‘f’ orbitals, such as La2O3,
Pr2O3 and Nd2O3 on the thermal decomposition of AP has been studied [97]. DTA and
TG results reveal that these oxides influence the thermal decomposition pattern of AP
significantly and bring down the decomposition temperature substantially. Isothermal
studies showed substantial acceleration of the decomposition of AP with the incorporation
of these oxides as little as 0.05%. Isothermal data were analysed using various kinetic
models. The activation energy for the catalysed decomposition of AP was significantly
lower. The catalytic effect is explained on the basis wherein the ‘p’-type oxides act as
conduits through metal cation in the electron-transfer mechanism for AP decomposition.
Downloaded by [Yale University Library] at 03:47 26 February 2013

5. Conclusions
Solid-fuelled rockets are used for space, ballistic, tactical purposes and to assist pro-
pulsion. Solid composite propellants in use today are mixtures of prepolymer (binder),
aluminium fuel and oxidiser salts, such as AP. Burning rate catalysts are always of great
interest in order to improve the ballistics of the rocket design. Though thermal decom-
position of AP is one of the most extensively studied thermal decomposition reactions,
it still remains a mystery to a large extent.
Metal oxides, mainly TMOs, are considered as potential burning-rate catalysts for
CSPs. There is continuous interest in the synthesis and application of nanometre particles
of metal or metal oxides, due to their large specific surface area and high activity in most
catalytic processes. It is noted that these nanoparticles are not only effective in the gas–
solid, gas–liquid–solid or liquid–solid catalytic processes, where the nanoparticles as the
catalyst are in the solid phase; but they also are effective in the solid–solid catalytic
processes. For instance, the nanosized metal oxide particles have been proved effectively in
improving the decomposition or explosion of the AP, a key energetic material and the
most common oxidiser in the rocket propellant. However, considering the limited loading
of AP or their composites in the rocket, it is crucial to further improve its decomposition
efficiency to produce a large amount of energy as far as possible and to decrease its
burning temperature for easy operation and control. In contrast to the nanoparticles of
metal oxide, the pure metal nanoparticles are much sensitive to the oxygen and may be
effective to improve the decomposition efficiency of AP. It was found that the catalytic
performance of nanometre-sized oxide particles was superior to that of micrometre-sized
oxide particles on the thermal decomposition of AP.

References

[1] C. Nogeura, Physics and Chemistry at Oxide Surface, Cambridge University Press, Cambridge,
UK, 1996.
[2] H.H. Kung, Transition Metal Oxides: Surface Chemistry and Catalysis, Elsevier, Amsterdam,
1999.
Journal of Experimental Nanoscience 227

[3] V.E. Henrich and P.A. Cox, The Surface Chemistry of Metal Oxides, Cambridge University
Press, Cambridge, UK, 1994.
[4] A.F. Wells, Structural Inorganic Chemistry, 6th ed., Oxford University Press, New York, 1987.
[5] W.A. Harrison, Electronic Structure and the Properties of Solids, Dover, New York, 1989.
[6] G. Ertl, H. Knozinger, and J. Weitkamp (eds.), Handbook of Heterogeneous Catalysis,
Wiley-VHC, Weinheim, 1997.
[7] H. Gleiter, Nanostructured materials: State of the art and perspectives, Nanostruct. Mater.
6 (1995), pp. 3–14.
[8] J.A. Rodriguez, G. Liu, T. Jirsak, J. Herbek, Z. Chang, J. Dvorak, and A. Maiti, Activation of
gold on titania: Adsorption and reaction of SO2 on Au/TiO2(110), Am. Chem. Soc. 124 (2002),
pp. 5242–5250.
[9] M. Baumer and H.J. Freund, Metal deposits on well-ordered oxide films, Prog. Surf. Sci.
61 (1999), pp. 127–198.
[10] M.L. Trudeau and J.Y. Ying, Nanocrystalline materials in catalysis and electrocatalysis:
Downloaded by [Yale University Library] at 03:47 26 February 2013

Structure tailoring and surface reactivity, Nanostrut. Mater. 7 (1996), pp. 245–258.
[11] M. Valden, X. Lai, and D.W. Goodman, Onset of catalytic activity of gold clusters on titania with
the appearance of nonmetallic properties, Science 281 (1998), pp. 1647–1650.
[12] G. Singh, I.P.S. Kapoor, S. Dubey, and P.F. Siril, Preparation, characterization and catalytic
activity of transition metal oxide nanocrystals, J. Sci. Conf. Proc. 1 (2009), pp. 11–17.
[13] C. Feldmann and H.-O. Jungk, Polyol-mediated preparation of nanoscale oxide particles, Angew.
Chem. Int. Ed. 40 (2001), pp. 359–362.
[14] B. O’Regan and M. Gratzel, A low-cost, high-efficiency solar cell based on dye sensitized colloidal
TiO2 films, Nature 353 (1991), pp. 737–740.
[15] A. Hagfeldt and M. Gratzel, Light-induced redox reactions in nanocrystalline systems, Chem.
Rev. 95 (1995), pp. 49–68.
[16] P.W.M. Jacob and H.M. Whitehead, Decomposition and combustion of ammonium perchlorate,
Chem. Rev. 69 (1969), pp. 551–590.
[17] V.V. Boldyrev, Thermal decomposition of ammonium perchlorate, Thermochim. Acta 443 (2006),
pp. 1–36.
[18] S.M. Shen, S.I. Chen, and B.B. Wu, The thermal decomposition of ammonium perchlorate (AP).
Containing a burning rate modifier, Thermochim. Acta 223 (1993), pp. 135–143.
[19] A.A. Said, Thermal decomposition of ammonium metavanadate doped with iron, cobalt or nickel
hydroxides, Therm. Anal. 37 (1991), pp. 959–962.
[20] A.K. Nema, S. Jain, S.K. Sharma, S.K. Nema, and S.K. Verma, Mechanistic aspect of thermal
decomposition and burn rate of binder and oxidiser of AP/HTPB composite propellants comprising
HYASIS-CAT, Int. J. Plastics Tech. 8 (2004), pp. 344–354.
[21] J. Gao, F. Guan, and Y. Zhao, Preparation of ultrafine nickel powder and its catalytic
dehydrogenation activity, Mat. Chem. Phys. 71 (2001), pp. 215–219.
[22] V.S. Sedoi and V.V. Valevich, Characterization of ultra-fine powders produced by the exploding
wire method, 32nd International Annual Conference of ICT, Fraunhofer ICT, FRG, Karlsruhe,
July 3–6 2001, pp. 80-1–80-10.
[23] M.M. Telkar, C.V. Rode, R.V. Choudhary, S.S. Joshi, and A.M. Nalawade, Shape controlled
preparation and catalytic activity of metal nanoparticles for hydrogenation of 2-butyne-1,4-diol and
styrene oxide, J. Appl. Catal. A 273 (2004), pp. 11–19.
[24] S.S. Joshi, P.R. Patil, and V.N. Krishnamurthy, Thermal decomposition of ammonium
perchlorate in the presence of nanosized ferric oxide, Def. Sci. J. 58(6) (2008), pp. 721–727.
[25] N. Taniguchi, On the basic concept of nanotechnology, Proc. Int. Conf. Prod. Eng. Tokyo, Part
2, JSPE, Tokyo 1974, pp. 18–23.
[26] N. Lubick and K. Betts, Silver socks have cloudy lining court bans widely used flame retardant,
Environ. Sci. Technol. 42(11) (2008), p. 3910.
228 S. Chaturvedi and P.N. Dave

[27] A.S. Edelstein and R.C. Cammarata (eds.), Nanomaterials: Synthesis, Properties, and
Applications, 2nd ed., CRC Press, Londres, 1998, pp. 89–95.
[28] B.L. Cushing, V.L. Kolensnichenko, and C.J. O’ Connor, Recent advances in the liquid-phase
syntheses of inorganic nanoparticles, Chem. Rev. 104 (2004), pp. 3893–3946.
[29] T.B. Brill and B.T. Budenz, Flash pyrolysis of ammonium percholrate-hydroxyl-terminated-
polybutadiene mixtures including selected additives, in Solid Propellant Chemistry, Combustion,
and Motor Interior Ballistics, Vol. 185, Progress in Astronautics and Aeronautics, V. Yang, T.
Brill, and W.-Z. Ren, Eds, AIAA, Reston VA, pp. 3–32, 2000, pp. 3–32.
[30] S.M. Peiris, G.I. Pangilinan, and T.P. Russell, Structural properties of ammonium perchlorate
compressed to 5.6 GPa, J. Phys. Chem. A 104 (2000), pp. 11188–11193.
[31] G. Singh and S.P. Felix, Evaluation of transition metal salts of NTO as burning rate modifier for
HRPB-AN composite solid propellants, Combust. Flame 132(3) (2003), pp. 422–432.
[32] C.L. Carnes and K.J. Klabunde, The catalytic methanol synthesis over nanoparticle metal oxide
catalysts, J. Mol. Catal. Chem. 194 (2003), pp. 227–236.
Downloaded by [Yale University Library] at 03:47 26 February 2013

[33] K. Kishore and M.R. Sunitha, Effect of transition metal oxides on decomposition and deflagration
on composite solid propellant systems: A survey, AIAA J. 17 (1979), pp. 1118–1125.
[34] G. Singh, I.P.S. Kapoor, and D.K. Pandey, Studies on energetic compound: Part 24 – Hexamine
metal perchlorate as high energetic burning rate catalysts, J. Energetic Mater. 20 (2002),
pp. 223–244.
[35] N.B. Singh and A.K. Ojha, Formation of copper oxide through NaNO3-KNO3 eutectic melt and
its catalytic activity on thermal decomposition of ammonium perchlorate, Thermochim. Acta 390
(2002), pp. 67–72.
[36] A.A. Said and R. Al-Qusami, The role of copper cobaltite spinel, CuxCo3xO4 during the thermal
decomposition of ammonium perchlorate, Thermochim. Acta 275 (1996), pp. 83–91.
[37] Y. Wang, J. Zhu, X. Yang, L. Lu, and X. Wang, Preparation of NiO nanoparticles and
their catalytic activity in the thermal decomposition of ammonium perchlorate, Thermochim. Acta
437 (2005), pp. 106–109.
[38] S. Vyazovkin and C.A. Wight, Kinetics of thermal decomposition of cubic ammonium perchlorate,
Chem. Mater. 11 (1999), pp. 3386–3393.
[39] M.W. Evans, R.B. Beyer, and L. McCulley, Initiation of deflagration waves at surfaces of
ammonium perchlorate–copper chromite–carbon pellets, J. Chem. Phys. 40(9) (1964),
pp. 2431–2438.
[40] J.A. Ibers, Nuclear magnetic resonance study of polycrystalline NH4ClO4, J. Chem. Phys. 32
(1960), pp. 1448–1449.
[41] A.V. Dubovitsky, N.Ya. Buben, and G.B. Manelis, Investigation of the movement of NHþ 4 ion in
ammonium perchlorate by means of electron paramagnetic resonance, Russ. J. Struct. Chem. 5
(1964), pp. 40–43.
[42] A.G. Keenan and R.F. Siegmund, Thermal decomposition of ammonium perchlorate, Quart. Rev.
Chem. Soc. London 23(3) (1969), pp. 435–452.
[43] E.Yu. Ivanov and V.V. Boldyrev, On the mechanism of polymorphous transition in ammonium
perchlorate crystals, Dokl. AN SSSR. 248(4) (1979), pp. 862–863.
[44] P.W. Jacobs and W.L. Ng, Thermal decomposition of ammonium perchlorate single crystals,
J. Solid State Chem. 9(4) (1974), pp. 305–322.
[45] M.M. Pavlyuchenko, E.A. Prodan, and L.I. Peshrovskaya, On the effect of temperature on the
rate of topochemical transformations in ammonium perchlorate, Russ. J. Phys. Chem. 51(3)
(1977), pp. 577–580.
[46] L. Bircomshaw and B. Newman, Thermal decomposition of ammonium perchlorate, Proc. R. Soc.
A227 (1955), pp. 228–237.
[47] M. Foltz and J. Meinschein, Ammonium perchlorate phase transitions to 26 GPa and 700K in
diamond anvil cell, Mater. Lett. 24(6) (1995), pp. 407–414.
Journal of Experimental Nanoscience 229

[48] G.B. Manelis (ed.), Mechanism of Thermal Decomposition of Ammonium Perchlorate: A


Collection of Papers, Institute of Chemical Physics AN SSSR, Chernogolovka, 1981, pp. 1–129
(in Russian).
[49] F. Solymosi, Structure and Stability of Salts of Halogen Oxyacids in the Solid Phase, Akademiai
Kiado, Budapest, 1977, pp. 195–261.
[50] A.R. Hall and G.S. Pearson, Ammonium perchlorate. A review of its role in composite propellant
combustion, Rocket Propulsion Establishment, Westcoff Tech. Rep. 67/1, 1967, p. 164.
[51] P.W. Jacobs and H.M. Whithead, Thermal decomposition and combustion of ammonium
perchlorate, Chem. Rev. 4 (1969), pp. 551–590.
[52] K. Kishore and M.R. Sunitha, Comprehensive view of the combustion models of composite solid
propellants, AIAA J. 17(11) (1979), pp. 1216–1224.
[53] A.V. Raevsky and G.B. Manelis, On the mechanism of decomposition of ammonium perchlorate,
Dokl. Akad. Nauk SSSR. 151(4) (1963), pp. 886–889.
[54] R.V. Jacobs and A.A. Russel-Jones, On the mechanism of decomposition of ammonium
Downloaded by [Yale University Library] at 03:47 26 February 2013

perchlorate, Raketnaya Tekhnika I Kosmotavtika 5(4) (1967), pp. 275–278, (in Russian).
[55] B.S. Svetlov and V.A. Koroban, On the inhibition of thermal decomposition of ammonium
perchlorate by the products of decomposition, Kinetika I Kataliz. 8 (1967), pp. 456–459.
[56] B.S. Svetlov and V.A. Koroban, On the mechanism of thermal decomposition of ammonium
perchlorate, Fizika Goreniya I Vzryva. 6(1) (1970), pp. 12–18, (in Russian).
[57] V.V. Boldyrev, Yu.P. Savintsev, T.V. Mulina, and G.V. Shchetinina, On the physicochemical
reasons of the formation and growth of reaction nuclei during thermal decomposition of ammonium
perchlorate, Kinetika I Kataliz. 11 2(B5) (1970), pp. 1131–1139.
[58] Yu.P. Savintsev, T.V. Mulina, and V.V. Boldyrev, Thermal Decomposition of NH4ClO4,
Combustion and Explosion, Nauka, Moscow, 1972, p. 756, (in Russian).
[59] A.J. Lang and S. Vyazovkin, Effect of pressure and sample type on decomposition of ammonium
perchlorate, Combust. Flame 145 (2006), pp. 779–790.
[60] B.I. Kaidymov and V.S. Gavazova, The influence of polymorphous transition in ammonium
perchlorate on the catalytic effect of homogeneous and heterogeneous additives in its thermal
decomposition, Fizika Goreniya I Vzryva. 10(6) (1974), pp. 801–810 (in Russian).
[61] B.I. Kaidimov and V.S. Gavazova, A sharp change in the inhibiting effect of ammonia on thermal
decomposition of ammonium perchlorate as a result of its phase transition, J. Inorg. Nucl. Chem.
36(12) (1974), pp. 3848–3849.
[62] P. Jacobs and A. Kureishy, The effect of additives on thermal decomposition of ammonium
perchlorate, Eighth Symposium in Combustion, The Williams and Wilkins, Baltimore, MD,
1962, p. 672.
[63] K. Kurotani, Some studies on solid propellants. Kinetics and thermal decomposition of ammonium
perchlorate/Science, Sci. Rep. Aeronautical Research Institute, University of Tokyo, Rep. 372,
Tokyo, 1962, p. 23.
[64] A.A. Shidlovsky, L.F. Shmagin, and V.V. Bulatova, The effect of some additives on thermal
decomposition of ammonium perchlorate, Proc. High School USSR Ser. Chem. Technol. 8 (1965),
pp. 533–538.
[65] Y. Xu, D. Chen, X. Jiao, and K. Xue, CuO microflowers composed of nanosheets: Synthesis,
characterization, and formation mechanism, Mater. Res. Bull. 42 (2007), pp. 1723–1731.
[66] J. Wang, S. Hi, Z. Li, X. Jing, M. Zhang, and Z. Jiang, Synthesis of chrysalis-like CuO
nanocrystals and their catalytic activity in the thermal decomposition of ammonium perchlorate,
J. Chem. Sci. 121(6) (2009), pp. 1077–1081.
[67] A. Hermony (Makovky), The catalytic decomposition of ammonium perchlorate, Eighth
Symposium in Combustion, The Williams and Wilkins, Baltimore, MD, 1962, p. 656.
[68] F. Solymosi and L. Revesz, Catalysis of reactions in the solid phase. Thermal decomposition of
ammonium perchlorate in the presence of ferric oxide, Kinetika I Kataliz. 4 (1963), pp. 88–96.
230 S. Chaturvedi and P.N. Dave

[69] P.R. Patil, S.K. Hait, V.N. Krishnamurthy, and S.S. Joshi, Characterisation of nano iron oxide
and its effect on thermal decomposition of ammonium perchlorate, International High Energy
Material Conference and Exhibition; DRDL, Hyderabad, 2005.
[70] P.R Patil, V.N Krishnamurthy, and S.S. Joshi, Differential scanning calorimetric study of HTPB
based composite propellants in presence of nano ferric oxide, Propell. Explo. Pyrotech. 31(6)
(2006), pp. 442–446.
[71] H. Xu, X. Wang, and L. Zhang, Selective preparation of nanorods and micro-octahedrons of
Fe2O3 and their catalytic performances for thermal decomposition of ammonium perchlorate,
Powder Technol. 185 (2008), pp. 176–180.
[72] F. Solymosi, Initiation of ammonium perchlorate—ignition by chromic oxide titanium dioxide
catalysts, Combust. Flame 9(2) (1965), pp. 141–148.
[73] W.A. Rosser, S.H. Inami, and H. Wise, Thermal decomposition of ammonium perchlorate,
Combust. Flame 12(5) (1968), pp. 427–435.
[74] F. Solymosi, S. Borcsok, J. Rasko, and J. Kiss, The effect of composite propellant catalysts on the
Downloaded by [Yale University Library] at 03:47 26 February 2013

stability of HClO4 and HClO4–NH3 system, Abstract of 14th Symposium (International)


Combustion, The Combustion Institute Pittsburgh, University Park, PA, 1972, p. 239.
[75] F. Shadman-Yazdi and E. Petersen, The effect of catalysts on the deflagration Limits of
ammonium perchlorate, Combust. Sci. Technol. 5 (1972), pp. 61–63.
[76] P.R. Patil, V.N. Krishnamurthy, and S.S. Joshi, Effect of nano-copper oxide and copper chromite
on the thermal decomposition of ammonium perchlorate, Propell. Explo. Pyrotech. 33(4) (2008),
pp. 266–270.
[77] D.H. Chen and X.R. He, Synthesis of nickel ferrite nanoparticles by sol-gel method, Mater. Res.
Bull. 36 (2001), pp. 1369–1377.
[78] G. Singh, I.P.S. Kapoor, S. Dubey, P.F. Siril, J.H. Yi, F.Q. Zhao, and R.-Z. Hu, Effect of mixed
ternary transition metal ferrites nanocrystallites on the thermal decomposition of ammonium
perchlorates, Thermochem. Acta 477 (2008), pp. 42–47.
[79] N.E. Ermolin, O.P. Korobeinichev, A.G. Tereshenk, and V.M. Foomin, Kinetic calculations and
mechanism definition for reactions in an ammonium perchlorate flame, Combust. Explos. Shock
Waves 18 (1982), pp. 180–189.
[80] P. Srivastava, I.P.S. Kapoor, and G. Singh, Nano ferrites: Preparation characterization and
catalytic properties, J. Alloys Compounds 485(1–2) (2009), pp. 88–92.
[81] G. Singh, I.P.S. Kapoor, S. Dubey, and P.F. Siril, Kinetics of thermal decomposition of
ammonium perchlorate with nanocrystals of binary transition metal ferrites, Propell. Explo.
Pyrotech. 34 (2009), pp. 72–77.
[82] B. Dubey, N. Singh, J. Srivastara, and A. Ojha, The catalytic behavior of NiFe2xCrxO4
(05  42) during thermal decomposition of ammonium perchlorate, polystyrene and their
composite propellants, Indian J. Chem. A 40 (2001), pp. 841–847.
[83] Z. Yu, Y. Sun, W. Wei, L. Lu, and X. Wang, Preparation of NdCrO3 nanoparticles and their
catalytic activity in the thermal decomposition of ammonium perchlorate by DSC/TG-MS,
J. Therm. Anal. Calorim. 97(3) (2009), pp. 903–909.
[84] L. Li, X. Sun, X. Qiu, J. Xu, and G. Li, Nature of catalytic activities of CoO nanocrystals in
thermal decomposition of ammonium perchlorate, Inorg. Chem. 74(19) (2008), pp. 8839–8846.
[85] Y. Zongxue, C. Lifen, L.U. Lude, Y. Xujie, and W. Xin, DSC/TG-MS study on in situ catalytic
thermal decomposition of ammonium perchlorate over CoC2O4, Chin. J. Catal. 30(1) (2009),
pp. 19–23.
[86] G. Duan, X. Yang, J. Chen, G. Huang, L. Lu, and X. Wang, The catalytic effect of nanosized
MgO on the decomposition of ammonium perchlorate, Powder Tech. 172 (2007), pp. 27–29.
[87] M.A. Zhenye, L.I. Fengsheng, and C. Aisi, Preparation and thermal decomposition behavior of
TMOS/AP composite nanoparticles, Nanoscience 11(2) (2006), pp. 142–145.
Journal of Experimental Nanoscience 231

[88] L. Cheng, M. Zhenye, Z. Lixiong, and Q. Renyuan, Preparation of Ni/TiO2 nanoparticles and
their catalytic performance on the thermal decomposition of ammonium perchlorate, Chin. J.
Chem. 27(10) (2009), pp. 1863–1867.
[89] K. Fujimura and A. Miyake, The effect of specific surface area of TiO2 on the thermal
decomposition of ammonium perchlorate, J. Therm. Anal. Calorim. 99(1) (2010), pp. 27–31.
[90] Y. Wang, X. Yang, L. Lu, and X. Wang, Experimental study on preparation of LaMO3 (M ¼ Fe,
Co, Ni) nanocrystals and their catalytic activity, Them. Chem. Acta 433 (2006), pp. 225–230.
[91] T. Fu, F. Liu, L. Liu, L. Guo, and F. Li, Catalytic thermal decomposition of ammonium
perchlorate using manganese oxide octahedral molecular sieve (OMS), Catal. Commun. 10(1)
(2008), pp. 108–112.
[92] F. Solymosi and L. Revesz, Thermal decomposition of ammonium perchlorate in presence of zinc
oxide, Nature 192 (1961), pp. 64–65.
[93] F. Solymosi and K. Fonagy, The effect of cadmium oxide and cadmium perchlorate on
decomposition and ignition of ammonium perchlorate, In Proceedings of the 11th International
Downloaded by [Yale University Library] at 03:47 26 February 2013

Symposium on Combustion, The Combustion Institute, Pittsburgh, PA, 1966, p. 268.


[94] K.C. Patil, V.R. Pai Verneker, and S.R. Jain, The role of lithium and magnesium perchlorate
amines in ammonium perchlorate decomposition, Combust. Flame 27(3) (1976), pp. 295–298.
[95] R.J. Acheson and P. Jacobs, Thermal decomposition of magnesium perchlorate and of ammonium
perchlorate mixtures, J. Phys. Chem. 74 (1970), pp. 281–288.
[96] M. Zheng, Z.S. Wang, J.Q. Wu, and Q. Wang, Synthesis of nitrogen-doped ZnO nanocrystallites
with one-dimensional structure and their catalytic activity for ammonium perchlorate decomposi-
tion, J. Nanopart. Res. 12(6) (2009), pp. 2211–2219.
[97] D.V. Survase, D.B. Sarwade, and E.M. Kurian, Effect of La2O3, Pr2O3 and Nd2O3 on the
thermal decomposition of ammonium perchlorate, J. Energetic Mater. 19(1) (2001), pp. 23–40.

Anda mungkin juga menyukai