Вы находитесь на странице: 1из 6

Full Paper

DOI: 10.1002/prep.201400238

Enhanced Combustion Characteristics of Bismuth Trioxide-


Aluminum Nanocomposites Prepared through Graphene
Oxide Directed Self-Assembly
Rajagopalan Thiruvengadathan,[a, b] Clay Staley,[a, b] Jordan M. Geeson,[a] Stephen Chung,[a]
Kristofer E. Raymond,[b] Keshab Gangopadhyay,[a, b] and Shubhra Gangopadhyay*[a]

Abstract: We present a facile, spontaneous, and surfactant- from 1.15 to 1.55 km s1, and specific impulse from 41 to
free method to controllably self-assemble aluminum and 71 s. The sensitivity of the self-assembled aluminum and
bismuth trioxide nanoparticles through the introduction of bismuth trioxide to electrostatic discharge was reduced by
graphene oxide as a self-assembly directing agent. The self- four orders of magnitude, without decreasing the combus-
assembled nanocomposites demonstrate significant com- tion performance. Graphene oxide directed self-assembly
bustion performance improvements in comparison to ran- can be used to synthesize nanocomposites with diverse
domly mixed aluminum and bismuth trioxide nanoparticles combustion properties and controlled ignition sensitivity,
with enhanced pressure generation from 60 to 200 MPa, which lays the foundation for preparing multi-functional,
pressurization rate from 3 to 16 MPa ms1, burning rate highly-reactive, combustion systems in the future.
Keywords: Graphene Directed self-assembly Combustion Nanocomposites Reactive materials Thermites

1 Introduction

The synthesis of nanocomposites is a strong and broaden- to its high density, energy content, and gas production by
ing field of research seeking to advance the capabilities of weight, which have encouraged many other recent studies
conventional energetic material systems [110]. Energetic [9, 1216]. The pressure generating capability, linear com-
nanocomposites are generally comprised of discrete fuel bustion rate, thrust behavior, and electrostatic discharge
and oxidizer particles with nanoscale dimensions. Nanopar- (ESD) sensitivity of GO self-assembled Al/Bi2O3 were mea-
ticle fuels and oxidizers enable very rapid reactivity as sured in direct comparison to randomly mixed Al/Bi2O3
a result of reducing the heat and mass diffusion lengths re- nanopowders. Combustion behaviors were measured
quired for combustion. The arrangement and intimacy of across a range of % Theoretical Maximum Densities (TMD)
the fuel and oxidizer nanoparticles in energetic nanocom- to evaluate the effect of GO self-assembly in various com-
posites profoundly impacts reactivity. Maximizing fuel and bustion regimes [1517]. We postulate that the enhanced
oxidizer interfacial contact is paramount for achieving opti- energy content realized through GO directed self-assembly
mum reaction kinetics, and numerous approaches have extends towards predictably improving the combustion
been reported with the goal of intelligently self-assembling performance of Al/Bi2O3 nanocomposites. The formation of
nanocomposites [3, 4, 7, 9, 11]. We have previously demon- macroparticulates from the self-assembly process and
strated that functionalized graphene in the form of gra- introduction of carbon also augers well for reducing the
phene oxide (GO) can be employed as a self-assembly di- ignition sensitivities of the nanocomposites, which in their
recting agent to produce highly-reactive macroparticulates pure state can be extremely sensitive. The fundamental
of nanocomposites [9]. GO was shown to enhance the
energy content of aluminum and bismuth trioxide (Al/ [a] R. Thiruvengadathan, C. Staley, J. M. Geeson, S. Chung,
Bi2O3) nanocomposites up to 92 % as a function of GO K. Gangopadhyay, S. Gangopadhyay
weight percentage. Enhanced energy content was due to Department of Electrical and Computer Engineering
University of Missouri
improved fuel and oxidizer intermixing from self-assembly Columbia, Missouri 65211, USA
as well as the participation of GO as a reactant during com- *e-mail: gangopadhyays@missouri.edu
bustion. [b] R. Thiruvengadathan, C. Staley, K. E. Raymond, K. Gangopadhyay
The motivation of this work was to evaluate the role of NEMS/MEMS Works, LLC
GO directed self-assembly in modifying the combustion be- 8850 W. Westlake Road
haviors and ignition sensitivities of Al/Bi2O3 nanocompo- Columbia, Missouri, 65203, USA
sites. Al/Bi2O3 was the material system of interest here due [] Authors contributed equally to work

Propellants Explos. Pyrotech. 2015, 40, 729 734 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 729
Full Paper R. Thiruvengadathan, C. Staley, J. Geeson, S. Chung, K. Raymond, K. Gangopadhyay, S. Gangopadhyay

goal of this study was to synthesize highly-reactive nano- ported in our previous works [15, 16]. However, to remain
composites with reduced sensitivity to ESD and enhanced consistent in the preparation of samples with and without
combustion performance through GO directed self-assem- the inclusion of GO, we have always used in 1 : 1 volume
bly. ratio of DMF and IPA as the solvent system. Randomly
mixed Al/Bi2O3 samples were collected by drying well-dis-
persed suspensions of Bi2O3 mixed with Al at 65 8C for 16 h.
The pressurization rates of all nanocomposites prepared
2 Experimental
in this work were determined by measuring the pressure
Bi2O3 spherical nanoparticles (90210 nm avg. diameter as generated during the propagation of reaction front as
per the specification by Accumet Materials, NY) and Al a function of time. For this purpose, a 60 mm3 closed com-
spherical nanoparticles (80 nm avg. diameter, 80 % active bustion cell (6.35 mm diameter, 2 mm depth) equipped
Al, 2.2 nm Al2O3 passivation as per the specification given with a pressure transducer (PCB 119B12, 862 MPa maximum
by Novacentrix, TX) were purchased and used as received. rating, 400 kHz resonance frequency, 2 ms rise time) was
GO was synthesized using the modified Hummers method used. All of the samples were tested at approximately 14 %
[18, 19]. Nanocomposites were prepared by mixing Bi2O3 TMD. The well was completely filled for each test with
and Al in accordance with an equivalence mixing ratio of dried nanocomposites without applying any external pres-
1.0 calculated from the theoretical Al/Bi2O3 reaction. The sure. A Ni-alloy fuse wire with a diameter of 0.13 mm was
Al2O3 dead mass was accounted in calculating equivalence used to ignite each sample. The slope of the rising pressure
mixing ratios. Self-assembled nanocomposites were synthe- signal from 10 % to 90 % peak pressure was used to quanti-
sized by adding various amounts of GO to the Bi2O3 and Al fy pressurization rate [8, 20]. At least three identical tests
(0 % to 5 % GO by weight relative to weight of Al and were performed for each nanocomposite to account for ex-
Bi2O3). perimental error. The pressure sensor was calibrated using
The exact synthesis protocols and fundamental science a hydraulic press with oil gauge before testing to quantify
driving the GO directed self-assembly of Bi2O3 and Al are sensitivity. Further details about the closed cell combustion
reported in our prior publication [9]. Briefly, Al, GO, and test measurement configuration are provided in our previ-
Bi2O3 suspensions were separately prepared by dispersion ous publications [8, 20].
in a 1 : 1 volume ratio of 2-propanol (IPA) and N,N-dimethyl- A theoretical peak pressure was also calculated for each
formamide (DMF) using ultrasonic agitation (output power nanocomposite for comparison to the experimental peak
of 250 W at 44 kHz frequency) for varied durations of 4 to pressures measured using closed cell combustion tests.
8 h. While Al and Bi2O3 dispersions were prepared by sub- First the volume of gas evolved from the nanocomposite
jecting it to ultrasonic agitation for 4 h, and GO dispersions was calculated at standard temperature and pressure using
in 1 : 1 volume mixtures of IPA and DMF were ultrasonically the ideal gas law:
agitated for 8 h. Precursor dispersion quality was evaluated
through zeta potential analysis (Delsa Nano C instrument,
n1 RT 1
Beckman Coulter), which showed + 70, + 40, and 59 mV V1 1
P1
potentials for Al, Bi2O3, and GO suspensions respectively to
qualify precursor suspension stability [9]. The GO and Al
where V1 is the volume of gas generated in an unconfined
suspensions were mixed together, ultrasonically agitated
environment, n1 is the amount of gas in moles, R is the uni-
for 1 h, and then Bi2O3 suspensions were added and the
versal gas constant, T1 is standard temperature, and P1 is
dispersions were sonicated for 1 h further. Following the
standard pressure. The moles of gas produced were de-
last mixing step, all suspensions were removed from the ul-
rived using 0.0047 mol of gas per g for the theoretical Al/
trasonic bath and left undisturbed until complete solids
Bi2O3 reaction, and assuming GO reacts purely as carbon
precipitation upon self-assembly, which occurred in mi-
with ambient oxygen to produce 0.083 mol of gas per g
nutes to hours depending on GO content. Higher GO con-
[21]. C-Bi2O3 and C-Al reaction pathways were neglected
tent (3.5 to 5.0 %) nanocomposites self-assembled within
from the calculation. The dead mass of the Al2O3 shell was
2 min, while lower concentration nanocomposites (< 2.0 %)
included in the calculation by reducing the mass of active
self-assembled in 2436 h. Electron microscope images of
Al accordingly. Assuming an isochoric combustion process,
the precursor materials and self-assembled GO nanocom-
the pressure generated in the closed combustion chamber
posites are presented in our prior work [9].
of volume V2 = 60 mm3 is expressed as:
Combustion performance tests were performed on dried
nanocomposites. For this purpose, GO(x%)/Al/Bi2O3 nano-
composites were obtained by separating the precipitates P1 V 1 T 2
P2 2
from the supernatant (suspension) phase through decanta- V 2T 1
tion. Subsequently, the precipitates were dried at 65 8C
under rough vacuum for 16 h. Al/Bi2O3 control samples where P2 is the closed cell pressure and T2 is the adiabatic
(without GO) were made following standard procedures re- flame temperature of the nancomposite. [22] The adiabatic

730 www.pep.wiley-vch.de 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Propellants Explos. Pyrotech. 2015, 40, 729 734
Bismuth Trioxide-Aluminum Nanocomposites

flame temperature of the Al/Bi2O3 reaction of 3253 K was 3 Results and Discussion
used for all nanocomposites regardless of GO content.
Combustion wave speed was measured using an optical Directly after synthesis, nanocomposites with higher GO
method, wherein a combination of photodiodes and fiber content by weight (3.5 % and 5 %) exhibited complete solid
optics, and a National Instruments data acquisition (NI phase precipitation within 2 min after being left undistur-
DAQ) board connected to a computer were used. A Lexan bed, while lower GO content nanocomposites ( 2 %) fully
tube of 3.2 mm inner diameter and of 95 mm length was precipitated from the solvent phase within 2436 h. The
loaded with 250 mg of nanocomposites. The % TMD esti- complementary surface potentials of GO (59 mV), Al (+
mated in this test configuration ranged from 4 to 5.5 % 70 mV), and Bi2O3 (+ 40 mV) in suspension (as obtained
TMD for various compositions. The loaded Lexan tube was from measurements using Zeta potential analyzer) pro-
inserted into an aluminum block fitted with four fiber optic motes electrostatic attraction in suspension [9]. Subsequent
photodetectors spaced 10 mm apart. The tube was open- chemical and electrostatic bonding in the short range (as
ended in all of the experiments. A spark gap igniter was verified by XPS measurements) yield macroscale particu-
positioned at one end of the tube, and triggered to initiate lates of self-assembled GO/Al/Bi2O3 that precipitate from
the reactions. Combustion wave speed was measured by suspension [9]. It is important to note that the precursor
monitoring the detection of light from the flame front at GO, Al, and Bi2O3, and Al/Bi2O3 suspensions remained stable
each photodiode as a function of time. The photodiodes for several hours to days. Moreover, without GO directed
were positioned far away from the ignition end so that the self-assembly, Bi2O3 and Al phase separate over time. There-
combustion wave speed was measured in the steady state. fore, to limit phase separation, randomly mixed Al and
Similar experimental arrangements for the combustion Bi2O3 nanocomposites (0 % GO content) were prepared
wave speed measurements were reported earlier by several using standard procedures, which are described in our pre-
research groups [8, 14, 23]. vious publications [9, 15, 16].
A thrust measurement configuration was used to investi- The pressure-time profiles, theoretical peak pressure, ex-
gate the effects of GO on nanocomposite combustion per- perimental peak pressure, pressurization rates, and linear
formance at approximately 55 % TMD. Dried nanocompo- combustion rates of the GO/Al/Bi2O3 nancomposites are
site powders were loaded in 1.6 mm diameter, 6 mm deep shown in Figure 1 and Figure 2 respectively. Peak pressure
combustion chambers in an iterative filling and cold press- generation increased with greater GO content from ca.
ing procedure (315 MPa applied pressure) until completely 85 MPa for randomly mixed Al/Bi2O3 (0.0 % GO) to ca.
full. The filled engines were mounted on a Kister 9222 200 MPa for 5.0 % GO composites. Pressurization rate also
thrust transducer, top ignited, and thrust force and burning increased from 5.1 MPa ms1 for Al/Bi2O3 to 16.4 MPa ms1
duration were measured to acquire total impulse. Specific for 3.5 % GO composites, but reduced to 10.5 MPa ms1 for
impulse, a metric of thrust efficiency, was calculated using 5.0 % GO composites. Theoretical peak pressure was shown
the thrust-time data and loaded nanocomposite mass [24]. to increase with GO content. Linear combustion rates ex-
Further details of the thrust measurement configuration are hibited a similar trend vs. GO content as pressurization rate
provided in our prior work [15, 16]. measurements. Average linear combustion rates of
ESD sensitivity measurements were performed using 1.15 km s1, 1.55 km s1, and 1.26 km s1 were observed for
a test system fabricated by Electro-Tech Systems (ETS) Inc randomly mixed Al/Bi2O3, 3.5 % GO composites, and 5.0 %
(Model 931) in accordance with standard US military proto- GO composites, respectively.
cols as provided in MIL-STD-1751A. This instrument in- The relationship between GO content and increasing
cludes a capacitor bank (30020,000 pF), which can be peak pressure is attributed to greater gaseous product evo-
charged to a voltage ranging from 100 V to 26 kV. Typically lution and the larger energy content available for nano-
5 mg of nanocomposite powder was placed in a stainless composites with greater amounts of GO [9]. Differential
steel sample holder. The % TMD used in ESD measurements scanning calorimetry and thermogravimetric analysis (DSC/
for different samples in this work is about 4 %. The high- TGA) of GO/Al/Bi2O3 showed GO functional group decom-
voltage pin electrode (a stainless steel needle) was placed position and subsequent participation of GO as a fuel
1 mm above the surface of the powder prior to subjecting (carbon) for Bi2O3 that occurs prior to the primary Al/Bi2O3
the powder to 24 consecutive discharge cycles at a known reaction [9, 25]. The evolution of gaseous CO species and
energy level typically beginning with the lowest energy GO functional group decomposition promotes greater peak
level of 0.16 mJ. If ignition was not observed, the voltage pressures for higher GO content nanocomposites. The en-
and/or capacitance were increased, and the same sample hancements in pressurization and linear combustion rates
was again subjected to 24 consecutive discharges at from randomly mixed to 3.5 % GO composites are attribut-
a slightly higher energy level until ignition occurred. A ed to improved fuel and oxidizer intermixing from self-as-
sample was considered ESD safe for an energy level if it did sembly and the energetic decomposition of GO. Decreased
not ignite upon 24 consecutive discharges. pressurization rates and linear combustion rates from 3.5 %
to 5.0 % GO content composites are likely due to imbal-

Propellants Explos. Pyrotech. 2015, 40, 729 734 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.pep.wiley-vch.de 731
Full Paper R. Thiruvengadathan, C. Staley, J. Geeson, S. Chung, K. Raymond, K. Gangopadhyay, S. Gangopadhyay

Figure 2. Linear combustion rates of GO/Al/Bi2O3 against GO con-


tent (data points are average values from three independent tests
one standard deviation).

age than the predicted value. This is likely due to not ac-
counting for the increased heat of reaction of GO in the
calculation and gaseous product evolution from GO func-
tional group decomposition.
The thrust behaviors of the GO/Al/Bi2O3 nanocomposites
are presented in Figure 3. With greater GO content, the
average thrust forces reduced while the burning durations
increased. Although it is intuitive to expect faster burning
durations for composites with larger amounts of GO, as ob-
served in the low % TMD tests (pressurization and linear
combustion rate), in high % TMD conditions, combustion is
dominated by a conductive thermal transfer mechanism as
opposed to the convective thermal transfer mechanism
that governs low % TMD combustion [5, 15]. Consequential-
Figure 1. Closed pressure cell measurements of GO/Al/Bi2O3 ly, the increased thrust durations as a function of GO con-
against GO content: (A) Pressure-time data. (B) Peak pressures and tent are due to the participation of GO during combustion
pressurization rates (data points are average values from three in- as a slow burning reactant, which impedes the primary Al
dependent tests one standard deviation). and Bi2O3 reaction under conductive burning conditions.
Specific impulse (thrust efficiency) improved continuously
for composites with greater GO content from 41 s for ran-
anced stoichiometry offsetting the benefits of self-assembly domly mixed Al/Bi2O3 to 71 s for 5.0 % GO composites. Im-
as GO participates as a fuel during combustion [9]. provements in specific impulse were due to complete com-
The theoretical peak pressure increased for nanocompo- bustion resulting from self-assembly and enhanced gas
sites with greater GO content due to the extra gaseous production for higher GO content composites.
product evolution associated with the GO additive. For the The ESD sensitivity of the GO/Al/Bi2O3 nanocomposites is
low GO content nanocomposites (0.0 % to 1.0 %) the theo- shown in Figure 4. Randomly mixed Al/Bi2O3 failed at
retical peak pressure was substantially higher than the ex- 0.125 mJ (measured for 40 nm Bi2O3 and 41 nm Al in [13]),
perimental peak pressure, likely due to partial phase sepa- but self-assembled GO/Al/Bi2O3 composites showed tre-
ration leading to incomplete combustion, and thermal mendous reductions in ESD sensitivities. A maximum pass-
energy loss to the surroundings associated with the finite ing energy of 1.2 mJ was observed for 5.0 % GO compo-
time of the reaction. The difference between experimental sites. The possible reasons for the reduced ESD sensitivity
peak pressure and theoretical peak pressure reduced with of self-assembled GO/Al/Bi2O3 composites to ESD are dis-
increasing GO content until 5.0 % nanocomposites, where cussed as follows: Although the electrical conductivity of
experimental pressure was measured to be higher on aver- GO is lower than that of pristine graphene, a recent work

732 www.pep.wiley-vch.de 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Propellants Explos. Pyrotech. 2015, 40, 729 734
Bismuth Trioxide-Aluminum Nanocomposites

Figure 3. Thrust behaviors of GO/Al/Bi2O3 against GO content: (A) thrust-time data, (B) average thrust force, thrust duration, and specific
impulse (data points are average values from three independent tests one standard deviation).

has shown the electrical conductivity of GO (with similar ESD sensitivity by increasing the electrical conductivity of
carbon to oxygen ratio of 2.3 as obtained in our work the composite [27].
versus 2.6 in their work) is 3.2 102 S cm1, which is quite Furthermore, the lower ESD sensitivity measured for GO/
high [26]. Besides, the XPS data analysis reported in our Al/Bi2O3 composites in comparison to randomly mixed Al/
earlier publication[9] confirms the removal of oxygen Bi2O3 is attributed to surface charge reduction due to the
during the self-assembly with Al nanoparticles through formation of macroparticulates from the nanoparticle pre-
short range covalent interactions, which suggests the GO cursors due to self-assembly [9]. The surface charge density
obtains even greater electrical conductivity upon self-as- of the nanocomposite powders can be inferred from zeta
sembly. In a study on the development of energetic nano- potential analysis of the suspensions by interchanging zeta
composites with variable ESD ignition thresholds reported potential with the surface field as has been illustrated in
by Foley et al., it has been shown that one can reduce the the literature [28]. Our prior work has demonstrated a re-
duction in surface charge density from 23658 e mm2 for Al/
Bi2O3 nanocomposites to 300 e mm2 for GO/Al/Bi2O3 nano-
composites due to self-assembly. [9] Thus, the drastic re-
duction of surface charge density by nearly two orders as
a result of self-assembly on GO, and enhanced electrical
conductivity of the resulting nanocomposites are the signif-
icant factors in decreasing the ESD sensitivity by four
orders. Therefore in addition to facilitating self-assembly
and supporting enhanced reaction kinetics, GO proved to
be a viable mechanism for engineering the ESD sensitivity
of the self-assembled nanocomposites, which is a very val-
uable tool in the design of combustion systems.

4 Conclusions
We have shown that GO directed self-assembly can be har-
nessed to tune the combustion performance and ignition
sensitivity of Al/Bi2O3 nanocomposite through both the
Figure 4. Electrostatic discharge sensitivities of GO/Al/Bi2O3 against self-assembly process and introduction of carbon as a terti-
GO content measured using ETS model 931 with a minimum ary reactant in the nanocomposite. Drastic improvements
0.16 mJ spark. The randomly mixed Al/Bi2O3 (0 % GO) datum of
in peak pressure generation, pressurization rate, combus-
0.125 mJ lowest energy failed is provided by reference [17]. The
fractions above the lowest energy failed data points indicate the tion rate, and specific impulse were observed when using
discharge number when ignition occurred out of 24 consecutive GO to direct the self-assembly of Al/Bi2O3 as opposed to
charge-discharge cycles. randomly mixing Al and Bi2O3 nanopowders. Additionally,

Propellants Explos. Pyrotech. 2015, 40, 729 734 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.pep.wiley-vch.de 733
Full Paper R. Thiruvengadathan, C. Staley, J. Geeson, S. Chung, K. Raymond, K. Gangopadhyay, S. Gangopadhyay

the ESD sensitivity of Al/Bi2O3 was reduced by four orders Composites (Al/MoO3, Al/WO3, Al/CuO, and Al/Bi2O3), J. Propul.
of magnitude through GO directed self-assembly, without Power 2007, 23, 707.
impeding combustion performance. GO directed self-as- [15] C. S. Staley, K. E. Raymond, R. Thiruvengadathan, S. J. Apper-
son, K. Gangopadhyay, S. M. Swaszek, R. J. Taylor, S. Gango-
sembly is a viable process for synthesizing highly-reactive padhyay, Fast-Impulse Nanothermite Solid-Propellant Minia-
nanocomposites with application-tailored combustion be- turized Thrusters, J. Propul. Power 2013, 29, 1400.
haviors and ignition sensitivities. [16] C. S. Staley, K. E. Raymond, R. Thiruvengadathan, J. J. Herbst,
S. M. Swaszek, R. J. Taylor, K. Gangopadhyay, S. Gangopadhyay,
Effect of Nitrocellulose Gasifying Binder on Thrust Per-
References formance and High-g Launch Tolerance of Miniaturized Nano-
thermite Thrusters, Propellants Explos. Pyrotech. 2014, 39, 374.
[1] C. R. Becker, S. Apperson, C. J. Morris, S. Gangopadhyay, L. J. [17] S. J. Apperson, A. V. Bezmelnitsyn, R. Thiruvengadathan, K.
Currano, W. A. Churaman, C. R. Stoldt, Galvanic Porous Silicon Gangopadhyay, S. Gangopadhyay, W. A. Balas, P. E. Anderson,
Composites for High-Velocity Nanoenergetics, Nano Lett. S. M. Nicolich, Characterization of Nanothermite Material for
2010, 11, 803. Solid-Fuel Microthruster Applications, J. Propul. Power 2009,
[2] L. J. Currano, W. A. Churaman, Energetic Nanoporous Silicon 25, 1086.
Devices, J. Microelectromech. Syst. 2009, 18, 799. [18] L. J. Cote, F. Kim, J. Huang, LangmuirBlodgett Assembly of
[3] S. H. Kim, M. R. Zachariah, Enhancing the Rate of Energy Re- Graphite Oxide Single Layers, J. Am. Chem. Soc. 2008, 130,
lease from Nanoenergetic Materials by Electrostatically En- 1043.
hanced Assembly, Adv. Mater. 2004, 16, 1821. [19] W. S. Hummers Jr., R. E. Offeman, Preparation of Graphitic
[4] J. Y. Malchi, T. J. Foley, R. A. Yetter, Electrostatically Self-assem- Oxide, J. Am. Chem. Soc. 1958, 80, 1339.
bled Nanocomposite Reactive Microspheres, ACS Appl. Mater. [20] R. Thiruvengadathan, G. M. Belarde, A. Bezmelnitsyn, M. Shub,
Interfaces 2009, 1, 2420. W. Balas-Hummers, K. Gangopadhyay, S. Gangopadhyay, Com-
[5] M. L. Pantoya, J. J. Granier, Combustion Behavior of Highly En- bustion Characteristics of Silicon-Based Nanoenergetic Formu-
ergetic Thermites: Nano vs. Micron Composites, Propellants lations with Reduced Electrostatic Discharge Sensitivity, Pro-
Explos. Pyrotech. 2005, 30, 53. pellants Explos. Pyrotech. 2012, 37, 359.
[6] C. Rossi, K. Zhang, D. Esteve, P. Alphonse, P. Tailhades, C. [21] S. H. Fischer, M. C. Grubelich, Theoretical Energy Release of
Vahlas, Nanoenergetic Materials for MEMS: a Review, J. Micro- Thermites, Intermetallics, and Combustible Metals, 24th Inter-
electromech. Syst. 2007, 16, 919. national Pyrotechnics Seminar, pp. 23186: IIT Research Insti-
[7] F. Sverac, P. Alphonse, A. Estve, A. Bancaud, C. Rossi, High- tute, Kanpur, India, 1998.
Energy Al/CuO Nanocomposites Obtained by DNA-Directed [22] P. W. Cooper, S. R. Kurowski, Introduction to the Technology of
Assembly, Adv. Funct. Mater. 2012, 22, 323. Explosives, VCH, Weinheim, 1996.
[8] R. Thiruvengadathan, A. Bezmelnit syn, S. Apperson, C. Staley, [23] B. S. Bockmon, M. L. Pantoya, S. F. Son, B. W. Asay, J. T. Mang,
P. Redner, W. Balas, S. Nicolich, D. Kapoor, K. Gangopadhyay, S. Combustion Velocities and Propagation Mechanisms of Meta-
Gangopadhyay, Combustion Characteristics of Novel Hybrid stable Interstitial Composites, J. Appl. Phys. 2005, 98.
Nanoenergetic Formulations, Combust. Flame 2011, 158, 964. [24] G. P. Sutton, O. Biblarz, Rocket Propulsion Elements, John Wiley
[9] R. Thiruvengadathan, S. Chung, S. Basuray, B. Balasubramani- & Sons, New Jersey, 2010, p. 768.
an, C. Staley, K. Gangopadhyay, S. Gangopadhyay, A Versatile [25] N. Piekiel, K. Sullivan, S. Chowdhury, M. Zachariah, The Role of
Self-Assembly Approach Toward High Performance Nanoener- Metal Oxides in Nanothermite Reactions: Evidence of Condensed
getic Composite using Functionalized Graphene, Langmuir Phase Initiation, OMB NO. 0704-0188 DTIC Document, Fort Bel-
2014, 30, 6556. voir, VA, USA, 2010.
[10] R. A. Yetter, G. A. Risha, S. F. Son, Metal Particle Combustion [26] Z. J. Li, B. C. Yang, S. R. Zhang, C. M. Zhao, Graphene Oxide
and Nanotechnology, Proc. Combust. Inst. 2009, 32, 1819. with Improved Electrical Conductivity for Supercapacitor Elec-
[11] R. Shende, S. Subramanian, S. Hasan, S. Apperson, R. Thiruven- trodes, Appl. Surf. Sci. 2012, 258, 3726.
gadathan, K. Gangopadhyay, S. Gangopadhyay, P. Redner, D. [27] T. Foley, A. Pacheco, J. Malchi, R. Yetter, K. Higa, Development
Kapoor, S. Nicolich, Nanoenergetic Composites of CuO Nano- of Nanothermite Composites with Variable Electrostatic Dis-
rods, Nanowires, and Al-Nanoparticles, Propellants Explos. Pyro- charge Ignition Thresholds, Propellants Explos. Pyrotech. 2007,
tech. 2008, 33, 122. 32, 431.
[12] K. Martirosyan, L. Wang, A. Vicent, D. Luss, Synthesis and Per- [28] S. Basuray, H. C. Chang, Designing a Sensitive and Quantifiable
formance of Bismuth Trioxide Nanoparticles for High Energy Nanocolloid Assay with Dielectrophoretic Crossover Frequen-
Gas Generator Use, Nanotechnology 2009, 20, 405609. cies, Biomicrofluidics 2010, 4, 013205.
[13] J. A. Puszynski, C. J. Bulian, J. J. Swiatkiewicz, Processing and
Ignition Characteristics of Aluminum-bismuth Trioxide Nano- Received: October 7, 2014
thermite System, J. Propul. Power 2007, 23, 698. Revised: March 7, 2015
[14] V. E. Sanders, B. W. Asay, T. J. Foley, B. C. Tappan, A. N. Pacheco, Published online: April 16, 2015
S. F. Son, Reaction Propagation of Four Nanoscale Energetic

734 www.pep.wiley-vch.de 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Propellants Explos. Pyrotech. 2015, 40, 729 734

Вам также может понравиться