ArticlePDF Available

Metal Iodate-Based Energetic Composites and Their Combustion and Biocidal Performance

Authors:
  • Naval Nuclear Laboratory

Abstract and Figures

Anthrax and other biological agents, pose a considerable potential public threat. Bacterial spores, in particular, are highly stress resistant and cannot be completely neutralized by common bactericides. This paper reports on synthesis of metal-iodate-based aluminized electrospray assembled nanocomposites which neutralize spores through a combined thermal and chemical mechanism. Here metal iodates (Bi(IO3)3, Cu(IO3)2, and Fe(IO3)3 act as a strong oxidizer to nanoaluminum to yield a very exothermic and violent reaction, and simultaneously generating iodine as a long lived bactericide. These microparticle assembled nanocomposites when characterized in terms of reaction times and temporal pressure release show significantly improved reactivity. Furthermore, sporicidal performance superior to conventional metal oxide-based thermites clearly shows the advantages of combining both a thermal and biocidal mechanism in spore neutralization.
Content may be subject to copyright.
Metal Iodate-Based Energetic Composites and Their Combustion and
Biocidal Performance
H. Wang,
G. Jian,
W. Zhou,
J. B. DeLisio,
V. T. Lee,
,§
and M. R. Zachariah*
,
Department of Chemical and Biomolecular Engineering and Department of Chemistry and Biochemistry, University of Maryland,
College Park, Maryland 20742, United States
Department of Cell Biology and Molecular Genetics, University of Maryland, College Park 20740, Maryland, United States
*
SSupporting Information
ABSTRACT: The biological agents that can be weaponized, such as Bacillus anthracis,
pose a considerable potential public threat. Bacterial spores, in particular, are highly
stress resistant and cannot be completely neutralized by common bactericides. This
paper reports on synthesis of metal iodate-based aluminized electrospray-assembled
nanocomposites which neutralize spores through a combined thermal and chemical
mechanism. Here metal iodates (Bi(IO3)3, Cu(IO3)2, and Fe(IO3)3) act as a strong
oxidizer to nanoaluminum to yield a very exothermic and violent reaction, and
simultaneously generate iodine as a long-lived bactericide. These microparticle-
assembled nanocomposites when characterized in terms of reaction times and temporal
pressure release show signicantly improved reactivity. Furthermore, sporicidal
performance superior to conventional metal-oxide-based thermites clearly shows the
advantages of combining both a thermal and biocidal mechanism in spore
neutralization.
KEYWORDS: energetic materials, metal iodate, nanothermite, biocidal, electrospray
1. INTRODUCTION
Thermites is a class of energetic materials that can undergo fast
redox reaction between a fuel (e.g., Al) and an oxidizer (CuO,
Fe2O3,Bi
2O3, etc.), which ,once initiated, release large amounts
of thermal energy.
+→ ++ΔH
A
l3
2MO 1
2Al O 3
2M
23
(1)
Decreasing the reactant length scales from the micron to the
nanoscale greatly increases the interfacial contact and reduces
the diusion distance between fuel and oxidizer, resulting in as
much as 1000×higher reactivity.
16
The thermochemical
properties of the materials used result in energy densities, for
the most common mixtures, that are a factor of 2 or more
higher on a volumetric basis than conventional organic-based
energetic materials (e.g., TNT or RDX).
722
The potential threats from biological-based weapons, such as
those employing Bacillus anthracis, pose a signicant challenge
to global security. Of particular concern are spores of virulent
bacteria that are highly stress resistant and cannot be
completely killed by high-pressure processing (HPP), heat, or
toxic chemicals such as iodophor.
2325
Conventional energetic
materials produce a thermal event over a relatively short time,
which may not be sucient for total inactivation. Thus, a
strategy has developed in which, in addition to the thermal
event, a remnant biocidal agent delivered simultaneously would
have a much longer exposure time resulting in a more eective
inactivation. As such, both silver- or halogen-containing
thermites, even diuoroiodate compound-based explosives,
have been considered as biological agent defeating ingre-
dients.
2629
The two most common methods of incorporating
biocidal agents into energetic systems are to either directly add
silver or halogen into the metallized energetic materials or to
employ silver- or iodine-containing oxidizers into the thermite
formulation.
3032
The latter option has the potential for Ag and
I to act as part of the energy release landscape rather than as a
passive species. Some extremely strong oxidizers of the latter
kind, such as NaIO4/KIO4, do not release iodine, while others
like I2O5and AgIO3have storage issues because of their
hygroscopicity and light sensitivity.
31,32
The ideal oxidizer
should (1) be easy to handle and (2) decompose to allow the
oxygen to react with the fuel and release molecular iodine.
+→+++Δ
HA
l1
2M(IO ) 1
2Al O 1
2M1
4I
323 2 (2)
In this paper, three metal iodate nanoparticles, Bi(IO3)3,
Cu(IO3)2, and Fe(IO3)3, were synthesized and subsequently
assembled with aluminum nanoparticles using an electrospray
technique to create microsized nanothermites. The conned
burning pressure of the metal iodate-based thermite was found
to be 5×higher than conventional Al/CuO nanothermite.
The rapid reaction mechanism and the iodine release from the
metal-iodate-based thermite was investigated using rapid
heating mass spectrometry. The eectiveness of synthesized
Received: May 26, 2015
Accepted: July 10, 2015
Research Article
www.acsami.org
© XXXX American Chemical Society ADOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
iodate-based thermites as a biocidal agent was tested, and
results show superior performance in the inactivation of spores.
2. EXPERIMENTAL SECTION
Chemicals. Copper oxide (50 nm), bismuth oxide (90210 nm),
and iron oxide (50 nm) were purchased from Sigma-Aldrich, and
aluminum (ALEX, 50 nm) was purchased from Argonide Corp. Al
nanoparticles contain 70% active aluminum (by weight), which is
conrmed by thermogravity analysis.
33
Collodion solution (48%
nitrocellulose, i.e., NC, in ethanol/diethyl ether, by weight) was
purchased from Fluka Corp., and diethyl ether (99.8%)/ethanol
(99.8%) mixture (volume ratio: 1:3) was employed to dissolve the
collodion. The copper iodate, bismuth iodate, and iron iodate
nanoparticles were synthesized by the following procedures. The
formation energies of these oxidizers were calculated on the basis of
the data at https://materialsproject.org/. The formation energies of
Bi(IO3)3, Cu(IO3)2, and Fe(IO3)3are calculated as 1282, 725, and
1293 kJ/mol, respectively.
Synthesis of the Copper Iodate Nanoparticles. Copper iodate
nanoparticles were synthesized by milling copper(II) nitrate trihydrate
(Sigma-Aldrich) and potassium iodate (Sigma-Aldrich) mixture (mass
ratio: 1:3). After the milling process, the sample was washed with 30
mL of deionized water, and centrifuged for 30 min at 13 500 rpm
(Hermle Z300). The whole process was repeated 4 times to enable full
removal of any impurities. The sample was then dried at 100 °C
overnight. The yield of copper iodate nanoparticles using this milling
process was 75%.
Preparation of the Bismuth Iodate and Iron Iodate
Nanoparticles. Bismuth iodate (Bi(IO3)3) and iron iodate (Fe-
(IO3)3) nanoparticles were synthesized by a precipitation method.
Bi(IO3)3nanoparticles were synthesized by the following process: 485
mg bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) was dissolved in 8
mL nitric acid solution (2 M), and 528 mg iodic acid (HIO3) was
dissolved in 8 mL deionized (DI) water. These were then mixed by
dropwise addition of Bi(NO3)3solution. The yield of bismuth iodate
nanoparticles using this milling process was 90%. Fe(IO3)3
nanoparticles were synthesized by mixing the Fe(NO3)3solution
and HIO3solution (1:3, by mole) at concentrations of 25 and 66 mg/
mL, respectively. The obtained brown solution was kept at 100 °C
overnight, and the subsequent yellow-green precipitate was collected
and further washed and dried for characterization, with a yield of
75%. Both of the Bi(IO3)3and Fe(IO3)3powders were handled with
the same washing, centrifuging, drying, and breaking process as the
Cu(IO3)2.
Preparation of the Physically Mixed Metal Iodate and Metal-
Oxide-Based Composites. The traditional approach to create a
nanocomposite thermite is by physical mixing. For example, to make
the bismuth-iodate-based thermite, 160 mg Bi(IO3)3powder was
dispersed in 1.8 mL of ethanol and sonicated for 60 min. Then, 51 mg
of aluminum nanoparticles (and 0.20 mL collodion for the case of
added NC, from Fluka Corp.) is added and sonicated for another 60
min. The suspension was stirred for 24 h, and then allowed to air-dry
in a hood. The dry powder was then gently broken to a ne powder. A
similar procedure was employed for the other iodates, but the mass
was adjusted to maintain a stoichiometry.
Electrospray Procedure. The electrospray assembly process is
essentially as previously described in ref 22. For example to make the
bismuth-iodate-based thermite 160 mg Bi(IO3)3powder was dispersed
in 1.8 mL of ethanol and sonicated for 60 min. To this was added 51
mg of aluminum nanoparticles (and 0.20 mL collodion for the case
with NC), and the mixture was sonicated for another 60 min. Then,
after stirring for 24 h, the suspension was ready to be electrosprayed.
The electrospray system consisted of a syringe pump to feed the
precursor at a constant speed of 4.5 mL/h, through a stainless steel
0.43 mm inner diameter needle. The distance between the needle and
substrate was 10 cm, with an applied voltage of 19 kV.
SEM/EDS, TEM, and TGA/DSC. Scanning electron microscope
(SEM) characterization was conducted with a Hitachi SU-70
instrument coupled to an energy dispersive spectrometer (EDX).
Transmission electron microscope (TEM) analysis employed a JEOL
2100F eld-emission instrument. Thermogravimetry/dierential scan-
ning calorimetry (TG/DSC) results were obtained with a TA
Instruments Q600 at a rate of 10 °C/min up to a maximum
temperature of 1000 °C in a nitrogen atmosphere.
Combustion Cell and Burning Products. The details of
combustion cell experiment can be found in refs 5and 34. A conned
combustion cell with a constant volume (13 cm3) was used to
measure the pressure and burn time of the samples. In this study, 25.0
mg of the loose thermite sample was placed inside a cell and was
Figure 1. (a) Schematic of electrospray approach. (bd) Low- and high-magnication SEM images of (b) Al/Bi(IO3)3composites, (c) Al/Cu(IO3)2
composites, and (d) Al/Fe(IO3)3composites. Note: all the above composites contain 5% NC (by weight).
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
B
ignited by a nichrome coil on top of the loose powder. An attached
piezoelectric pressure sensor (PCB) together with an in-line charge
amplier and signal conditioner were used to record the pressure
history. The optical emission history was simultaneously collected by a
lens tube assembly, containing a planoconvex lens ( f= 50 mm) and a
photodetector to collect the broadband emission. The burn time was
dened as the peak width at half height of the optical emission.
T-Jump Ignition and Time-Resolved Mass Spectrometry
Measurement. The details of the time-of-ight mass spec system
used can be found in refs 5and 3335. In a typical experiment, a ne
Pt wire (76 μm in diameter, 10 mm in length) was coated with a thin
layer of sample (35 mm in length). An applied voltage of 24 V was
put across the wire to rapidly heat to 1800 K at a heating rate of 5 ×
105Ks
1. The wire temperature can be calculated by determining the
resistance changing with time. Time-resolved mass spectra were used
for characterization of the species produced during the rapid heating. A
Vision Research Phantom v12.0 digital camera was employed to
capture the burning of the thermite in air. The resolution used was 256
×256 pixels, and the frame rate used was 67 065 fps (14.9 μs per
frame).
Spore Inactivation. The above combustion cell was employed as a
container to evaluate the spore inactivation eectiveness of the
samples. First, metal-iodate-based thermite was placed in the cell of the
internal chamber in the combustion cell (Figure 2a). Aluminum foil
was adopted as the substrate for Bacillus thuringiensis spores, a known
surrogate for Bacillus anthracis.A10μL portion of spore solution (7
×105/μL, 1.4 μg) was dripped onto the foil surface followed by
spreading and drying. Aluminum foils deposited with or without
spores were placed close to the wall of the chamber, 15 mm from
thermite sample. After ignition, the combustion cell was kept closed
for 30 min, following which the foils were extracted and sent for
bacterial spore counting assay. Before the test, explosion-product-
loaded control foils were deposited with the same amount of spores.
Each foil was then immersed in 1 mL of fresh Lysogeny Broth (LB)
medium and incubated at 37 °C for 3 h. Our preliminary test reveals
that this time period is necessary for the initially immobile living
spores to detach from the foil and germinate in the media, enabling us
to measure the spore counts quantitatively later. Since the exact range
of spore viability was unknown, a serial dilution was used. A 100 μL
portion of the medium was extracted and diluted subsequently, with
each dilution of a factor of 10. Then 10 μL of these diluted media with
dierent bacteria concentrations were spread in parallel lines on an
agar plate. Agar plates were kept overnight at room temperature prior
to numeration of the colony forming units (CFUs) in each line. The
nal counts of spores in each sample were obtained statistically in
terms of these serial CFUs. Each experiment was repeated in triplicate.
3. RESULTS AND DISCUSSION
3.1. Metal-Iodate-Based Thermite Synthesis and
Assembly. In this work, micron-scale metal-iodate-based
energetic composites were produced by electrospraying a
precursor solution containing nanoscale components, compris-
ing fuel (nano-Aluminum), the biocidal oxidizer (metal
iodates), and a small quantity of energetic binder nitrocellulose.
As shown in Figure 1a, because of the combined action of
electrostatic forces and surface tension, the precursor
suspension shatters into nearly monodiserse microsized
droplets. Subsequent rapid evaporation of the solvent drives
gelling process within the droplet until the aggregate particles
jam, creating a porous structure. More details of the
electrospray assembly technique can be found in our prior
work.
22,36
The metal iodate nanoparticles used in this study include
bismuth iodate (90 nm), iron iodate (70 nm), and copper
iodate (65 nm), which were prepared by either precipitation
or milling (see Experimental Section and Supporting
Information). For comparison purposes, physically mixed
metal-iodate-based nanothermites were also prepared, and the
SEM images can be found in Figure S9 in the Supporting
Information.
Figure 1bd shows typical SEM images of the three metal-
iodate-based thermites. The obtained Al/Bi(IO3)3(Figure 1b)
composites have a size distribution of 35μm, have a porous
structure, and contain a well-dispersed mixture of the fuel and
oxidizer nanoparticles. Al/Cu(IO3)2composites have similar
structure features and a relatively narrow size of 24μm
(Figure 1c), respectively, while the Al/Fe(IO3)3composites
Figure 2. (a) Schematic of combustion cell. (b) Pressure and optical emission proles of the Al/Bi(IO3)3nanothermite composites. (c) Pressure
proles of the three metal-iodate-based thermites. (d) Burning times of metal-iodate-based nanothermite composites. Note: The pressure traces of
Cu(IO3)2- and Fe(IO3)3-based thermites have been oset to the right, for better readability.
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
C
(Figure 1d) were found to have a more open structure, with a
larger size ranging from 5 to 7 μm.
3.2. Reactivity of Metal-Iodate-Based Nanothermite.
The reactivity of nanothermite was assessed in a constant
volume combustion cell (13 cm3), from which the pressure
history and optical emission during combustion can be
obtained simultaneously. Figure 2a shows the schematic of
the combustion cell, which is composed of a pressure sensor
and an optical detector. Pressurization rates obtained from the
pressure history data, which mimic the burning rate, are used to
evaluate the reactivity. Burning times were also obtained by
measuring the peak width at half-maximum (fwhm) of the
optical emission trace.
5,37
Supporting Information Table S1
summarizes the combustion cell results for all the nano-
thermites in this study.
Figure 2b shows the pressure and optical-emission trace for
Al/Bi(IO3)3composites. The peak pressure and burning time
are 4.7 MPa and 170 μs, respectively. The burning time is
considerably longer than the corresponding pressure rise time,
implying that decomposition of iodates occurs prior to the
primary reaction of the thermite, suggesting that the thermite
burning rate is limited by the aluminum fuel release, rather than
oxidizer availability. The temporal pressure of the three metal-
iodate-based nanothermites are shown in Figure 2c. All three
pressure traces show a rapid rise, of 1μs, achieving a peak
pressure of 45 MPa with the Cu > Bi > Fe.
For comparison, the reactivity of metal iodates thermite
prepared by conventional physical mixing method was also
evaluated. The pressurization rate results of metal-iodate-based
thermites produced by two approaches were listed in Table 1.
The results show that the pressurization rate of electrosprayed
Al/Bi(IO3)3composites is approximately 2 orders of magnitude
higher than the physical mixed thermites, while the Al/
Cu(IO3)2and Al/Fe(IO3)3systems are more than 4 orders of
magnitude higher, relative to the physically mixed thermites.
The increase in reactivity of the electrosprayed sample can be
attributed to the unique porous inner structure combined with
energetic gas generator nitrocellulose which we have previously
shown minimizes sintering among nanoparticles.
22
Similar to what we have observed for Al/CuO composites,
22
the physically mixed Bi(IO3)3,Cu(IO
3)2,andFe(IO
3)3
nanothermites have a much longer burning time than that of
the corresponding electrosprayed nanothermites, as Figure 2d
shows.
The reactivity dierence between metal-iodate-based and
metal oxide-based thermites was compared. The peak pressure
and pressurization rate of the two thermites were simulta-
neously obtained, as listed in Table 2. The reactivity of the
metal-iodate-based thermite is much higher than the reactivity
of the corresponding metal-oxide-based thermite even though
the size of metal iodate nanoparticles is larger. Specically, the
so-called weak thermite, Al/Fe2O3, has a pressure and
pressurization rate of only 0.06 MPa and 0.02 GPa/s,
respectively. While the corresponding iron-based iodate, Al/
Fe(IO3)3, achieves 1.9 MPa and 590 GPa/s, as shown in Figure
3a. As discussed in our previous work,
38
Al/Fe2O3nano-
thermite combustion pressure and optical emission signals
appear concurrently (Figure 3b), and it has been suggested that
oxidizer decomposition is the rate limited step. In contrast, Al/
Fe(IO3)3features a very rapid pressure increase followed by a
prolonged optical emission prole, indicating the burning is
rate limited by the aluminum fuel. These results show the
advantages of using strong oxidizers, with rapid oxygen release
kinetics in nanothermite formulations.
The adiabatic ame temperatures (AFT) of the three metal-
iodate- and metal-oxide-based thermites have been calculated
and are listed in Table 2. Since there are no experimental data
for thermochemical properties of these metal iodates, the
theoretical estimate results were adopted.
39
These results are
then employed in the Cheetah code, assuming the sample is in
the maximum density and the volume of the burning product is
constant.
40,41
Generally, the ame temperatures of the three
metal-iodate-based thermite reactions are roughly the same,
4050 K, which is >1000 K higher than that of the
corresponding metal-oxide-based thermites. The high reaction
temperature can largely improve the potential biocidal
performance for its additional thermal activities which can
synergistically function with the released iodine.
To investigate the ignition mechanism, the nanothermite
samples were coated on a Pt wire with a diameter of 76 μm and
joule heated in air at a heating rate of 5 ×105Ks
1,as
schematically shown in Figure 4a. The Al/Bi(IO3)3composites
reveal a violent reaction, with an ignition temperature of 590
°C. The ignition temperatures of Cu(IO3)2-based and Fe-
(IO3)3-based nanothermites is 560 and 550 °C, respectively,
Table 1. Peak Pressure, Pressurization Rate, and Burning
Time of Metal-Iodate-Based Nanothermites Made by
Electrospraying (ES) and Physical Mixing (PM)
a
thermite pressure
(MPa) pressurization
(GPa/s) burning
time (μs)
ES Al/Bi(IO3)34.5 3816 235
PM Al/Bi(IO3)30.73 53 298
ES Al/Cu(IO3)24.9 3966 238
PM Al/Cu(IO3)20.14 0.07 2162
ES Al/Fe(IO3)34.0 3186 161
PM Al/Fe(IO3)30.17 0.10 3667
a
All the composites contain 5% NC (by weight). 25.0 mg sample in
each test.
Table 2. Pressure Cell Results of Physically Mixed Nanothermites
a
oxidizer pressure (MPa) pressurization (GPa/s) burning time (μs) AFT (K) stoichiometric ratio
Bi2O3,90210 nm 1.0 54 240 2333 1.3
Bi(IO3)3,90 nm 2.3 770 150 4062 1.0
CuO, 50 nm 1.1 100 220 3054 1.0
Cu(IO3)2,200 nm 1.8 225 170 4061 1.0
Fe2O3,50 nm 0.06 0.02 3330 3027 1.0
Fe(IO3)3,70 nm 1.9 590 170 4043 1.0
a
Sample mass: 25.0 mg. All the composites contain no NC. The AFT of Al/Bi2O3thermite is only 3031 K. (Because this was a fuel lean mixture,
stoichiometric ratio = 1, to optimize pressurization rate.)
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
D
which is much lower than that of corresponding CuO-based
(770 °C) and Fe2O3-based nanothermites (1140 °C). The low
ignition temperature is indicative that the metal iodate salts
release oxygen at a relatively low temperature.
5
To further
conrm this, the wire within the extraction region of a mass
spectrometer, i.e., T-jump/time-of-ight mass spectrum (T-
Jump/TOFMS), was employed to investigate the oxygen
release properties of the three metal iodates.
37
Reactivity of a thermite can be evaluated by the explosion
characterization parameters such as peak pressure, pressuriza-
tion rate, and burn time in a conned container (Figure 2a).
Especially, the pressurization rate is directly proportional to the
burn rate of the thermite. In the previous work, it is found that
the reactivity is closely related to its oxygen release behavior at
high heating rates.
42
Figure 4b shows the temporal prole for
oxygen from metal iodates in T-Jump/TOFMS, from which the
oxygen release temperature can be obtained, and the oxygen
release behavior of dierent oxidizers at high heating rates can
be compared. It is found that the onset temperature of oxygen
release for Bi(IO3)3, Cu(IO3)2, and Fe(IO3)3nanoparticles is
490, 450, and 475 °C, respectively, and thus insensitive to
structure. Interestingly, it is worth noting that while Bi(IO3)3
releases oxygen at 490 °C, and Bi2O3releases gas phase oxygen
at 1350 °C, they both show a similar ignition temperature of
590 °C.
43
The results imply that Bi(IO3)3and Bi2O3have
dierent initiation mechanisms. For Bi(IO3)3, it is obvious that
oxygen release from oxidizer contributes to the ignition of
nanothermite, while its condensed phase appears to be the
initiation mechanism for the Al/Bi2O3nanothermite.
As Figure 4b shows, all of the oxygen release traces show a
rapid rise. The oxygen release from Cu(IO3)2initiates at 450
°C, and reaches it maximum in 0.20 ms, signicantly
outperforming the commercial CuO nanoparticles, whose rise
time is 1.0 ms,
33
revealing its fast oxygen release kinetics at high
heating rates. Oxygen release proles of Fe(IO3)3and Bi(IO3)3
nanoparticles are very similar to that of Cu(IO3)3, implying, not
surprisingly, that all metal iodate nanoparticles release oxygen
in a similar manner. It is worth noting that the decomposition
of iodates occurs prior to the thermite ignition, suggesting that
gas phase oxygen release is critical to ignition.
The metal-iodate-based nanothermites were also rapidly
heated in the T-Jump/TOFMS to determine temporal
Figure 3. Pressure and optical emission proles of (a) Al/Fe(IO3)3and (b) Al/Fe2O3nanothermite reactions.
Figure 4. (a) T-Jump wire ignition method, and burning snapshots of Al/Bi(IO3)3composites. (b) Temporal prole of oxygen release upon heating
the three dierent metal iodates. (c) T-Jump TOFMS results for the Bi(IO3)3-based, Cu(IO3)2-based, and Fe(IO3)3-based thermite made by
electrospray (5% NC, by weight). Note: The heating pulse time was 3.0 ms.
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
E
speciation. As shown in Figure 4c, iodine and oxygen species
are detected in all three metal-iodate-based thermites reactions.
The intensities of oxygen and iodine species, in all of the
samples, reach their maximum values between 500 and 600 °C,
revealing the major weight loss event happening during that
period, and is consistent with TGA results (Supporting
Information).
3.3. Sporicidal Performance. The ability of metal-iodate-
based thermites to produce high heat, pressure, and iodine
makes them ideal biocidal energetic materials with potential
synergistic killing mechanisms. To test the sporicidal perform-
ance of metal iodate-based thermites, a spore inactivation
experiment was conducted in the combustion cell (V=13cm
3,
see Figure 2a). Aluminum foil coupons deposited with or
without 7×106CFU Bacillus thuringiensis spores were placed
within the combustion chamber around the wire coil that
ignites the thermite reaction. The control experiment of Al/
Bi(IO3)3thermite reaction without spores shows that iodine
was released after reaction and uniformly dispersed within the
layer of postcombustion products (Figure 5a,b). By loading
dierent quantities of Al/Bi(IO3)3thermite in the combustion
cell, the survival ratios of Bacillus thurigiensis spores on coupons
were measured after the thermite reactions (Figure 5c). Spores
exposed to 5.0 mg and 25.0 mg of thermite reaction products
had a 2 log or a >4 log reduction in their viability (i.e., the
ability to form bacterial colonies), respectively. Alternatively,
similar tests by employing the reaction of 5.0 mg Al/Bi2O3
thermites show that the viability of spores decreased by 28%.
The major dierence of these two thermite systems is that
iodate-based composites can generate free iodine through a
thermite reaction, which has been widely conrmed as a
biocidal agent. Therefore, except for the pressure and thermal
stresses which were also generated from the combustion of Al/
Bi2O3thermites, the combustion of Al/Bi(IO3)3thermites
exposes spores to the additional biocidal stress of iodine. We
estimated that the released iodine eciently contributed to 27%
loss of spore viability (28% 1%), while the other eects from
pressure, heat, and other reaction products such as Al2O3and
Bi contributed to the majority of the loss as conrmed by the
result after the Al/Bi2O3reaction (Figure 5c). Figure 5c shows
that the postreaction products of Al2O3and Bi induced viability
reduction of spores by 30%, implying that the other heat and
pressure stresses contributed to a 42% loss of spore viability
(72% 30%). Given that other reactions of metal iodate
thermites (Al/Bi(IO3)3and Al/Cu(IO3)2) also led to uniform
iodine release (Supporting Information Figures S13 and S14)
and demonstrate higher sporicidal capabilities compared with
the corresponding oxide-based thermite reactions (Supporting
Information Figure S15), it is proposed that the metal iodate
thermite system can be recognized as an eective biocidal
agent.
The SEM images of spores (Figure 5d) further demonstrate
that ne particles were deposited on the spore surfaces after the
5.0 mg Al/Bi(IO3)3thermite reaction. There is some precedent
for cell death in the absence of morphological damage.
43
However, interestingly, there is no gross morphological change
in the spores after the thermite reaction, suggesting that the
thermite reaction has not compromised the spore structure,
although the viability has been reduced by 2 orders of
magnitude (Figure 5c). Previously, we have shown that, during
rapid heating, spores maintained most of their viability until the
spore coat was melted.
44
Herein, it is further found that, by
applying iodine, the spores were killed before changing
morphologies, indicating the superiority in ecacy of the
Figure 5. (a, b) SEM and energy dispersive X-ray spectroscopy (EDS) results of Al/Bi(IO3)3thermite reaction products. (c) CFU counts from the
spores treated by 5.0 mg Al/Bi(IO3)3postcombustion products, 5.0 mg Al/Bi2O3thermite reaction, and 5.0 mg and 25.0 mg Al/Bi(IO3)3thermite
reactions. (d) SEM images of spores before (top) and after (bottom) undergoing the Al/Bi(IO3)3thermite reaction (5.0 mg). Note: all the
composites are produced by electrospraying with 5% NC (by weight). The labeled numbers in part c are survival ratios of spores.
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
F
Bi(IO3)3-based thermite complex in killing spores compared
with simple rapid heating.
4. CONCLUSION
This work reports the formation of metal-iodate-based
thermites which have superior performance as sporicidal
agents. Three dierent metal iodate nanoparticles were
synthesized and incorporated into energetic composites with
nanoaluminum by electrospray or physical mixing, forming
highly reactive nanothermites. We characterized the reactivity
using combustion cell and fast-heating wire experiments. The
pressure and pressurization rate of metal-iodate-based thermites
are several times higher than the corresponding oxide-based
thermites, and the reactivity could be further promoted by
using the electrospray technique to assemble the metal-iodate-
based thermites. The thermal decomposition properties of
metal iodates, as well as the reaction mechanisms of the related
thermites, were separately investigated by TGA/DSC, and
TOF-MS. The results show that the metal iodates decompose
into their corresponding metal oxide, oxygen, and iodine before
the aluminothermic reaction takes place. The sporicidal
performance of the metal-iodate-based thermites was also
assessed, and the results showed that they outperform the
corresponding metal-oxide-based thermites. The working
mechanism was proposed as a synergistic eect of heat/
pressure and iodine production from the highly reactive metal-
iodate-based thermite reaction. This combination has a much
higher sporicidal rate than the individual eect of either heat/
pressure (from the metal-oxide-based thermite reaction) or
iodine (from the burning residue of metal-iodate-based
thermite reaction).
ASSOCIATED CONTENT
*
SSupporting Information
Characterization of the three dierent metal iodates nano-
particles was shown, including SEM and TEM images, as well
as thermal decomposition results. The SEM images of
electrosprayed metal-oxide-based thermites and physically
mixed metal-iodate-based thermites. The combustion cell
results of all three kinds of metal-iodate-based thermites and
corresponding metal-oxide-based thermites. The burning snap-
shots of Al/Cu(IO3)2and Al/Fe(IO3)3composites on the wire.
The SEM images and EDS results of the postcombustion
products from Al/Cu(IO3)2and Al/Fe(IO3)3composite
thermite reaction. The Supporting Information is available
free of charge on the ACS Publications website at DOI:
10.1021/acsami.5b04589.
AUTHOR INFORMATION
Corresponding Author
*Phone: 301-405-4311. E-mail: mrz@umd.edu.
Present Address
§
Department of Safety Engineering, Nanjing University of
Science and Technology, Nanjing, Jiangsu 210094, China.
Author Contributions
The manuscript was written through contributions of all
authors. All authors have given approval to the nal version of
the manuscript.
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
This work was supported by the Defense Threat Reduction
Agency and the Army Research Oce. We acknowledge the
support of the Maryland Nanocenter and its NispLab. The
NispLab is supported in part by the NSF as a MRSEC Shared
Experimental Facility. H.-Y.W. is grateful for the nancial
support from China Scholarship Council.
REFERENCES
(1) Rossi, C.; Esté
ve, A.; Vashishta, P. Nanoscale Energetic Materials.
J. Phys. Chem. Solids 2010,71,5758.
(2) Weismiller, M. R.; Malchi, J. Y.; Lee, J. G.; Yetter, R. A.; Foley, T.
J. Effects of Fuel and Oxidizer Particle Dimensions on the Propagation
of Aluminum Containing Thermites. Proc. Combust. Inst. 2011,33,
19891996.
(3) Moore, K.; Pantoya, M. L.; Son, S. F. Combustion Behaviors
Resulting from Bimodal Aluminum Size Distributions in Thermites. J.
Propul. Power 2007,23, 181185.
(4) Pantoya, M. L.; Granier, J. J. Combustion Behavior of Highly
Energetic Thermites: Nano versus Micron Composites. Propellants,
Explos., Pyrotech. 2005,30,5362.
(5) Jian, G. Q.; Feng, J. Y.; Jacob, R. J.; Egan, G. C.; Zachariah, M. R.
Super-reactive Nanoenergetic Gas Generators Based on Periodate
Salts. Angew. Chem., Int. Ed. 2013,52, 97439746.
(6) Granier, J. J.; Pantoya, M. L. Laser Ignition of Nanocomposite
Thermites. Combust. Flame 2004,138, 373383.
(7) Blobaum, K. J.; Reiss, M. E.; Plitzko, J. M.; Weihs, T. P.
Deposition and Characterization of a Self-propagating CuOx/Al
Thermite Reaction in a Multilayer Foil Geometry. J. Appl. Phys.
2003,94, 29152922.
(8) Schoenitz, M.; Ward, T. S.; Dreizin, E. L. Fully dense Nano-
Composite Energetic Powders Prepared by Arrested Reactive Milling.
Proc. Combust. Inst. 2005,30, 20712078.
(9) Blobaum, K. J.; Wagner, A. J.; Plitzko, J. M.; Van Heerden, D.;
Fairbrother, D. H.; Weihs, T. P. Investigating the Reaction Path and
Growth Kinetics in CuOx/Al Multilayer Foils. J. Appl. Phys. 2003,94,
29232929.
(10) Petrantoni, M.; Rossi, C.; Salvagnac, L.; Coné
dé
ra, V.; Estè
ve,
A.; Tenailleau, C.; Alphonse, P.; Chabal, Y. J. Multilayered Al/CuO
Thermite Formation by Reactive Magnetron Sputtering: Nano versus
Micro. J. Appl. Phys. 2010,108, 084323.
(11) Zhang, K.; Rossi, C.; Ardila Rodriguez, G. A. Development of a
Nano-Al/CuO Based Energetic Material on Silicon Substrate. Appl.
Phys. Lett. 2007,91, 113117.
(12) Umbrajkar, S. M.; Seshadri, S.; Schoenitz, M.; Hoffmann, V. K.;
Dreizin, E. L. Aluminum-Rich Al-MoO3Nanocomposite Powders
Prepared by Arrested Reactive Milling. J. Propul. Power 2008,24, 192
198.
(13) Ward, T. S.; Chen, W.; Schoenitz, M.; Dave, R. N.; Dreizin, E. L.
A Study of Mechanical Alloying Processes Using Reactive Milling and
Discrete Element Modeling. Acta Mater. 2005,53, 29092918.
(14) Kim, S. H.; Zachariah, M. R. Enhancing the Rate of Energy
Release from Nanoenergetic Materials by Electrostatically Enhanced
Assembly. Adv. Mater. 2004,16, 18211825.
(15) Tillotson, T. M.; Gash, A. E.; Simpson, R. L.; Hrubesh, L. W.;
Satcher, J. H., Jr.; Poco, J. F. Nanostructured Energetic Materials Using
SolGel Methodologies. J. Non-Cryst. Solids 2001,285, 338345.
(16) Seo, H. S.; Kim, J. K.; Kim, J. W.; Kim, H. S.; Koo, K. K.
Thermal Behavior of Al/MoO3Xerogel Nanocomposites. J. Ind. Eng.
Chem. 2014,20, 189193.
(17) Rossi, C.; Zhang, K.; Esteve, D.; Alphonse, P.; Tailhades, P.;
Vahlas, C. Nanoenergetic Materials for MEMS: a Review. J.
Microelectromech. Syst. 2007,16, 919931.
(18) Subramanium, S.; Hasan, S.; Bhattacharya, S.; Gao, Y.;
Apperson, S.; Hossain, M.; Shede, R. V.; Gangopadhyay, S.; Render,
R.; Kapper, P.; Nicolich, S. Self-assembled Ordered Energetic
Composites of CuO Nanorods and Nanowells and Al Nanoparticles
with High Burn Rates. Mater. Res. Soc. Symp. Proc. 2005,896,9.
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
G
(19) Shende, R.; Subramanian, S.; Hasan, S.; Apperson, S.;
Thiruvengadathan, R.; Gangopadhyay, K.; Gangopadhyay, S.;
Redner, P.; Kapoor, D.; Nicolich, S.; Balas, W. Nanoenergetic
Composites of CuO Nanorods, Nanowires, and Al-Nanoparticles.
Propellants, Explos., Pyrotech. 2008,33, 122130.
(20) Malchi, J. Y.; Foley, T. J.; Yetter, R. A. Electrostatically Self-
assembled Nanocomposite Reactive Microspheres. ACS Appl. Mater.
Interfaces 2009,1, 24202423.
(21) Sé
verac, F.; Alphonse, P.; Estè
ve, A.; Bancaud, A.; Rossi, C.
High-Energy Al/CuO Nanocomposites Obtained by DNA-Directed
Assembly. Adv. Funct. Mater. 2012,22, 323329.
(22) Wang, H. Y.; Jian, G. Q.; Egan, G. C.; Zachariah, M. R.
Assembly and Reactive Properties of Al/CuO Based Nanothermite
Microparticles. Combust. Flame 2014,161, 22032208.
(23) Nicholson, W. L.; Munakata, N.; Horneck, G.; Melosh, H. J.;
Setlow, P. Resistance of Bacillus Endospores to Extreme Terrestrial
and Extraterrestrial Environments. Microbiol. Mol. Biol. Rev. 2000,64,
548572.
(24) Setlow, P. Spores of Bacillus Subtilis: their Resistance to and
Killing by Radiation, Heat and Chemicals. J. Appl. Microbiol. 2006,101,
514525.
(25) Leggett, M. J.; McDonnell, G.; Denyer, S. P.; Setlow, P.;
Maillard, J. Y. Bacterial Spore Structures and their Protective Role in
Biocide Resistance. J. Appl. Microbiol. 2012,113, 485498.
(26) Johnson, C. E.; Higa, K. T. Presented at MRS Meeting Iodine-
Rich Biocidal Reactive Materials, Boston, 2530 (11, 2012).
(27) Mulamba, O.; Hunt, E. M.; Pantoya, M. L. Neutralizing Bacterial
Spores Using Halogenated Energetic Reactions. Biotechnol. Bioprocess
Eng. 2013,18, 918925.
(28) He, C.; Zhang, J.; Shreeve, J. M. Dense Iodine-Rich Compounds
with Low Detonation Pressures as Biocidal Agents. Chem. - Eur. J.
2013,19, 75037509.
(29) Fischer, D.; Klapotke, T.; Stierstorfer, J. R. Synthesis and
Characterization of Guanidinium Difluoroiodate, [C(NH2)-
3]+[IF2O2]and its Evaluation as an Ingredient in Agent Defeat
Weapons. Z. Anorg. Allg. Chem. 2011,637, 660665.
(30) Aly, Y.; Zhang, S.; Schoenitz, M.; Hoffmann, V. K.; Dreizin, E.
L.; Yermakov, M.; Indugula, R.; Grinshpun, S. A. Iodine-containing
Aluminum-based Fuels for Inactivation of Bioaerosols. Combust. Flame
2014,161, 303310.
(31) Feng, J. Y.; Jian, G. Q.; Liu, Q.; Zachariah, M. R. Passivated
Iodine Pentoxide Oxidizer for Potential Biocidal Nanoenergetic
Applications. ACS Appl. Mater. Interfaces 2013,5, 88758880.
(32) Sullivan, K. T.; Piekiel, N. W.; Chowdhury, S.; Wu, C.;
Zachariah, M. R.; Johnson, C. E. Ignition and Combustion
Characteristics of Nanoscale Al/AgIO3: A Potential Energetic Biocidal
System. Combust. Sci. Technol. 2010,183, 285302.
(33) Jian, G. Q.; Chowdhury, S.; Sullivan, K.; Zachariah, M. R.
Nanothermite Reactions: Is Gas Phase Oxygen Generation from the
Oxygen Carrier an Essential Prerequisite to Ignition? Combust. Flame
2013,160, 432437.
(34) Jian, G. Q.; Piekiel, N. W.; Zachariah, M. R. Time-resolved Mass
Spectrometry of Nano-Al and Nano-Al/CuO Thermite under Rapid
Heating: a Mechanistic Study. J. Phys. Chem. C 2012,116, 26881
26887.
(35) Zhou, L.; Piekiel, N.; Chowdhury, S.; Zachariah, M. R. T-Jump/
time-of-flight Mass Spectrometry for Time-Resolved Analysis of
Energetic Materials. Rapid Commun. Mass Spectrom. 2009,23, 194
202.
(36) Wang, H. Y.; Jian, G. Q.; Yan, S.; DeLisio, J. B.; Huang, C.;
Zachariah, M. R. Electrospray Formation of Gelled Nano-Aluminum
Microspheres with Superior Reactivity. ACS Appl. Mater. Interfaces
2013,5, 67976801.
(37) Sullivan, K. T.; Piekiel, N. W.; Wu, C.; Chowdhury, S.; Kelly, S.
T.; Hufnagel, T. C.; Fezzaa, K.; Zachariah, M. R. Reactive Sintering: an
Important Component in the Combustion of Nanocomposite
Thermites. Combust. Flame 2012,159,215.
(38) Sullivan, K.; Zachariah, M. R. Simultaneous Pressure and
Optical Measurements of Nanoaluminum Thermites: Investigating the
Reaction Mechanism. J. Propul. Power 2010,26, 467472.
(39) Jain, A.; Ong, S. P.; Hautier, G.; Chen, W.; Richards, W. D.;
Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; Persson, K.
A. The Materials Project: A materials genome approach to accelerating
materials innovation. APL Mater. 2013,1, 011002.
(40) Howard, W. M.; Souers, P. C.; Vitello, P. A. Cheetah 6.0 User
Manual; LLNL-SM-416166; Lawrence Livermore National Labora-
tory: Livermore, CA, 2010.
(41) Hobbs, M. L.; Baer, M. R.; McGee, B. C. JCZS: An
Intermolecular Potential Database for Performing Accurate Deto-
nation and Expansion Calculations. Propellants, Explos., Pyrotech. 1999,
24, 269279.
(42) Zhou, L.; Piekiel, N. W.; Chowdhury, S.; Zachariah, M. R. Time-
resolved Mass Spectrometry of the Exothermic Reaction between
Nanoaluminum and Metal Oxides: the Role of Oxygen Release. J. Phys.
Chem. C 2010,114, 1426914275.
(43) Davis, C. Enumeration of Probiotic Strains: Review of Culture
Dependent and Alternative Techniques to Quantify Viable Bacteria. J.
Microbiol. Methods 2014,103,917.
(44) Zhou, W. B.; Wu, M. O.; Watt, S. K.; Jian, G. Q.; Lee, V. T.;
Zachariah, M. R. Inactivation of Bacterial Spores Subjected to Sub-
second Thermal Stress. Chem. Eng. J. 2014,279, 578588.
ACS Applied Materials & Interfaces Research Article
DOI: 10.1021/acsami.5b04589
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX
H
... In recent years, there has been extensive domestic and international research focused on iodine-containing compound-based MICs, encompassing iodine oxide [34,35], iodates [36][37][38][39], and periodates [40][41][42]. These studies have revealed that these MICs possess notable characteristics such as high energy density, gas generation properties (with I 2 as a product), and sterilization capabilities. ...
Article
Full-text available
To address the challenges related to significant pressure loss and weakened combustion performance of energetic materials when ignited on Micro-Electro-Mechanical Systems (MEMS), periodate-based MICs (Metastable Intermolecular Composites) have emerged as promising candidates due to their high energy density and combustion pressure. However, the integration of periodate-based MICs onto MEMS has not been effectively achieved thus far. Herein, we propose a simple, environmentally friendly, and cost-effective approach to fabricate a MEMS-compatible Al/Cu2HIO6/PVDF (polyvinylidene difluoride) energetic composite film. Our method involves utilizing a Cu(OH)2 array as a template and employing an in-situ and spin-coating process. The resulting Al/Cu2HIO6/PVDF film was characterized using various techniques including SEM, TEM, XRD, and XPS. These analyses reveal a tightly packed nano array morphology with a porous structure. DSC-TG analysis was conducted to investigate the thermal reaction of the Al/Cu2HIO6/PVDF composite film. The results demonstrate that the thermal reaction involved a complex multi-step process, with higher heat release (1121 J g-1), a faster reaction rate, and a lower initial reaction temperature (294.5 °C) compared to the Al/CuO/PVDF composite film. the reaction process of Al/Cu2HIO6/PVDF film was also been proposed by analyzing the products at different temperatures. combustion diagnostic tests were carried out to evaluate the combustion performance of the Al/Cu2HIO6/PVDF composite film. The results indicate that as CuO transformed into Cu2HIO6, the film exhibits a reduced ignition delay (8 ms) and combustion duration (50 ms), higher combustion temperature (3710 K), and stronger flame intensity (8630 a.u.).
... To that end sandwiched structures of solid propellants with laminates of AP and HTPB have been studied. Previous studies have emphasized the importance of the leading-edge flames (LEF) between oxidizer rich layer (AP) and fuel layer (HTPB) in non-aluminized propellants, which is affected by the packing of different size AP particles [32][33][34][35][36][37][38]. On the other hand, morphological restructuring of the metallic fuel has not been a focus [39]. ...
... This is because a conventional energetic materials can result in low neutralization efficiency since they rely solely on a thermal neutralization pulse which is spatially distributed, may not provide enough thermal energy for long enough to kill the spores [10]. Therefore, it has been proposed that simultaneously delivering a rapid thermal pulse with a remnant biocidal agent would prolong the exposure time and improve the inactivation process [11]. Iodine-containing energetic materials have shown the most promise because of their excellent biocidal properties [12] compared to other biological energetic materials [13][14][15]. ...
Article
Full-text available
It was found that all iodine-containing biocidal energetic materials has a relatively long ignition delay upon combustion. A shorter ignition time (from the thermal trigger to the peak pressure) for Al/iodine oxides might further boost their combustion performance due to their high reactivity and high gas release rate. To achieve this goal, a secondary oxidizer, CuO, is incorporated into Al/I2O5 at different mass content keeping the overall thermite stoichiometry constant. The ternary Al/I2O5/CuO thermites were characterized in ignition using a T-jump ignition temperature set up, and in combustion in a constant volume combustion cell. Consequently, all ternary thermites outperform traditional Al/I2O5 counterpart with an optimum for 80/20 wt% of I2O5/CuO. This later composition ignites in 0.01 ms (30 times shorter than Al/I2O5) and produces peak pressure and pressurization rate of ∼4 and 26 times greater than those produced by Al/I2O5. A series of additional characterizations using Fourier-transform infrared spectroscopy, Differential Scanning Calorimetry, Electrical/Thermal conductivity measurement, etc., permitted to unravel the cause of such improvement and to propose a reaction mechanism for this ternary Al/I2O5/CuO system. From an applications point of view, this study proposes a facile, inexpensive and efficient way to enhance the combustion performance of Al/I2O5 biocidal nanoenergetic materials.
Article
This work is aimed to develop a stable Mg‐based thermite with high iodine concentration and preliminarily set a method to test inactivation factor (IF) values of biocidal materials with proximity to the combat situation. First, Mg−I 2 composites were mechanochemically prepared and up to 10 wt% of iodine could be retained until the material heated to 265 °C. Then biocidal thermites were mechanochemically prepared using Mg−I 2 as a fuel and Ca(IO 3 ) 2 as an oxidizer. For Mg−I 2 −Ca(IO 3 ) 2 thermites, apparent aging was observed in the material with a stoichiometric fuel‐to‐oxidizer ratio. In heated filament ignition tests, thermites had ignition temperatures of ~760 °C and ~820 °C at heating rates of 1800 K/s and 13000 K/s, respectively, and were slightly lower than that of pure Mg. Particle combustion test showed that the thermite with more fuel fraction burned faster. A setup based on the constant volume explosion experiment was established to test the inactivation effect of biocidal materials. Results showed that pure Mg had the best inactivation effect, and IF value doesn't have a clear correlation with max pressure achieved by explosions of biocidal materials. For results within the same max pressure range, IF value is proportional to the iodine concentration within the materials. Interestingly, for Mg, IF value was exponentially changed with max pressurization rate, this leads to the unexpected high IF value of Mg. This study shows excessive iodine or oxidizer could deactivate the biocidal material, and the heat release rate may play a crucial role in biocidal performance.
Article
Full-text available
Accelerating the discovery of advanced materials is essential for human welfare and sustainable, clean energy. In this paper, we introduce the Materials Project (www.materialsproject.org), a core program of the Materials Genome Initiative that uses high-throughput computing to uncover the properties of all known inorganic materials. This open dataset can be accessed through multiple channels for both interactive exploration and data mining. The Materials Project also seeks to create open-source platforms for developing robust, sophisticated materials analyses. Future efforts will enable users to perform ‘‘rapid-prototyping’’ of new materials in silico, and provide researchers with new avenues for cost-effective, data-driven materials design.
Article
Full-text available
The fight against biological warfare has prompted investigation of the chemistry and exothermic energy from energetic material reactions as a means for the neutralization of bacterial spores. The interaction between energetic reactions containing biocides and spore forming bacteria is not well understood. The goal of this work is to fundamentally examine the mechanisms of neutralization for Bacillus thuringiensis utilizing a halogenated energetic material reaction. Spore neutralization is attributed to a thermal effect from the reaction heat and the associated chemical influence of the halogen gas (i.e., produced from combustion). Results show heat transfer in the spore enhances the effectiveness of the halogen gas in the neutralization process and that elevated temperatures increase spore permeability, facilitating gas penetration and accelerating spore neutralization. Based on experimental results, a mathematical model was developed to predict spore behavior during reaction exposure over varying time scales. In the millisecond range, the model showed that the coupled thermal-biocidal gas mechanism will require elevated temperatures of 360A degrees C to produce 80% neutralization in tens of milliseconds while thermal conditions alone would require nearly 1,000A degrees C for the same neutralization. These results provide molecular-level insights into the components underpinning biological processes leading to spore neutralization.
Article
Full-text available
Probiotics are live microorganisms that confer a benefit on the host. Standard culture techniques are commonly used to quantify probiotic strains, but cell culture only measures replicating cells. In response to the stresses of processing and formulation, some fraction of the live probiotic microbes may enter a viable but non-culturable state (VBNC) in which they are dormant but metabolically active. These microbes are capable of replicating once acclimated to a more hospitable host environment. An operating definition of live probiotic bacteria that includes this range of metabolic states is needed for reliable enumeration. Alternative methods, such as fluorescent in situ hybridization (FISH), nucleic acid amplification techniques such as real-time quantitative PCR (RT-qPCR or qPCR), reverse transcriptase (RT-PCR), propidium monoazide-PCR, and cell sorting techniques such as flow cytometry (FC)/fluorescent activated cell sorting (FACS) offer the potential to enumerate both culturable and VBNC bacteria. Modern cell sorting techniques have the power to determine probiotic strain abundance and metabolic activity with rapid throughput. Techniques such as visual imaging, cell culture, and cell sorting, could be used in combination to quantify the proportion of viable microbes in various metabolic states. Consensus on an operational definition of viability and systematic efforts to validate these alternative techniques ultimately will strengthen the accuracy and reliability of probiotic strain enumeration.
Article
Full-text available
It is generally agreed that a key parameter to high reactivity in nanothermites is intimate interfacial contact between fuel and oxidizer. Various approaches have been employed to combine fuel and oxidizer together in close proximity, including sputter deposition [1], and arrested milling methods [2]. In this paper, we demonstrate an electrospray route to assemble Al and CuO nanoparticles into micron composites with a small percentage of energetic binder, which shows higher reactivity than nanothermite made by conventional physical mixing. The electrospray approach offers the ability to generate microscale particles with a narrow size distribution, which incorporates an internal surface area roughly equivalent to the specific surface area of a nanoparticle. The size of the micron scale composites could be easily tuned by changing the nitrocellulose content which is used as the binder. The composites were burned in a confined pressure cell, and on a thin rapidly heated wire to observe burning behavior. The sample of 5 wt.% nitrocellulose showed the best response relative to the physical mixing case, with a 3× higher pressure and pressurization rate. The ignition characteristics for these micron particles are essentially equivalent to the nanothermite despite their significantly larger physical size. It appears that electrospray assembly process offers to potential advantages. 1. Enhanced mixing between fuel and oxidizer; 2. Internal gas release from nitrocellulose that separates the particles rapidly to prevent sintering. The later point was shown by comparing the product particle size distribution after combustion.
Article
Ammonium and guanidinium difluoroiodate(V), [NH4](+)[IF2O2] (1a) and [C(NH2)(3)](+)[IF2O2] (1b), and diazidoglyoxime, [N3C=N-OH](2) (2) were synthesized and the molecular structures in the solid state of 1b and 2 were elucidated by single-crystal X-ray diffraction. 1b: P (1) over bar, a = 6.6890(5), b = 10.2880(6), c = 10.30.92(8) angstrom, a = 105.447(6), beta = 108.568(7), gamma = 91.051(5)degrees, V = 644.08(8) angstrom(3), rho = 2.650 g.cm (3); 2: P2(1)/n, a = 4.4211(3), b = 13.7797(9), c = 4.9750(3) angstrom, beta = 98.735(6), V = 299.57(3) angstrom(3), rho = 1.886 g.cm (3). The suitability of compounds 1a and 1b as active ingredients for agent defeat weapons (ADW) with biocidal activity has been shown in detonation tests using geobacillus stearothermophilus spores. In addition, a complete energetic characterization of the promising primary explosive 2 is given.
Article
Rapid heat pulse is the primary method for neutralizing large quantities of spores. Characterizing heat inactivation on a millisecond time scale has been limited by the ability to apply ultrafast, uniform heating to spores. Using our system for immobilization of spores on metal surfaces, bacterial spores were subjected to high temperatures (200 ˚C to 800 ˚C) and heating rates (∼103 ˚C/s to ∼105 ˚C/s). Spore inactivation increased with temperature and fit a sigmoid response. We observed the critical peak temperature (Tc) which caused a 2-fold reduction in spore viability was 382 ˚C and 199 ˚C for heating rates of ∼104 ˚C/s and ∼105 ˚C/s, respectively. Repetitive heating to the same peak temperature had little effect on viability. In contrast, stepwise heating to elevated peak temperatures inactivated spores in a manner similar to a single pulse heating to the same peak temperature. These results indicate that the maximum temperature rather than the overall heating time is primarily responsible for spore neutralization at ∼104 ˚C/s heating rate. The mechanism of spore inactivation was further investigated at two heating rates (∼104 ˚C/s and ∼105 ˚C/s). Viability reduction was mainly due to DNA damage at the heating rate of ∼104 ˚C/s as mutant strains defective for sspA sspB and recA were more sensitive to heat than the wide-type strains. At the higher heating rate (∼105 ˚C/s), spore inactivation was correlated with physical damage from ultrafast vapor pressurization inside spores. This new approach of pulse heating generates a temperature, time, and kill relationship for Bacillus spores at sub-second timescales.
Article
Exponential-13,6 (EXP-13,6) potential parameters for 750 gases composed of 48 elements were determined and assembled in a database, referred to as the JCZS database, for use with the Jacobs Cowperthwaite Zwisler equation of stare (JCZ3-EOS)((1)). The EXP-13,6 force constants were obtained by using literature Values of Lennard-Jones (LJ) potential functions, by using corresponding states (CS) theory, by matching pure liquid shock Hugoniot data, and by using molecular volume to determine the approach radii with the well depth estimated from high-pressure isentropes. The JCZS database was used to accurately predict detonation velocity, pressure, and temperature for 50 different explosives with initial densities ranging from 0.25 g/cm(3) to 1.97 g/cm(3). Accurate predictions were also obtained for pure liquid shock Hugoniots, static properties of nitrogen, and gas detonations at high initial pressures.
Article
Sol–gel process using molybdenum alkoxides was employed to prepare Al/MoO3 xerogel nanocomposites as a thermite with better performance by improvement of interfacial contact area between the oxidizer and fuel. Micromorphology and thermite reaction characteristics of Al/MoO3 xerogel nanocomposites were analyzed by scanning electron microscopy (SEM) and thermogravimetry/differential scanning calorimetry (TG/DSC), respectively. In the present Al/MoO3 xerogel system, it was found that exothermic enthalpy increases as the Al/Mo mole ratio increases and then decreases when Al/Mo mole ratio is larger than 6 indicating that optimum mole ratio of Al/Mo is 6 with reaction enthalpy of 420.58 J/g.
Article
Inactivation of aerosolized biologically viable Bacillus atrophaeus endospores (stimulant of the B. anthracis bio-agent) in combustion products of air-acetylene flames seeded with different aluminum-based powders was investigated. A flow of bioaerosol was introduced into the environment above the flame. The mixing of the combustion products with bioaerosol particles occurred when the combustion temperatures were cooled off to approximately 170–260 °C (the cross-sectional weighted average temperature). The flame was seeded with pure Al powder as well as with composite Al·I2 and Al·B·I2 powders prepared by mechanical milling. The iodine content was close to 20 and 15 wt.% for Al·I2 and Al·B·I2 powders, respectively. The burn rate was highest for particles of pure Al and lowest for particles of Al·B·I2. It was also observed that in the flame, particles of Al·B·I2 had the lowest temperature compared to other materials. Despite a lower iodine concentration, the combustion products from the flame seeded with Al·B·I2 exhibited the highest levels of inactivation of the aerosolized spores. The flame products of Al·I2 have also shown an effective inactivation. The inactivation levels observed for the unseeded flame and flame containing pure Al, were much lower and similar to each other; these inactivation levels were consistent with relatively weak thermal stress experienced by the bioaerosol at the relatively low temperatures of the exhaust gas. The highest level of inactivation observed for the combustion products of Al·B·I2 composite powder is attributed to its lower burn rate and respectively more homogeneous mixing of the iodine-containing products with the exhaust gases.
Article
Iodine pentoxide (I2O5), also known as diiodine pentoxide, is a strong oxidizer which has been recently proposed as an iodine-rich candidate oxidizer in nanoenergetic formulations, whose combustion products lead to molecular iodine as a biocidal agent. for potential biocidal applications. However, its highly hygroscopic nature hinders its performance as a strong oxidizer and an iodine releasing agent, and prevents its implementation. In this work, we developed a gas phase assisted aerosol spray-pyrolysis which enables creation of iron oxide passivated I2O5. Transmission electron microscopy elemental imaging as well as Temperature-Jump mass spectrometry confirmed the core shell nature of the material and the fact that I2O5 could be encapsulated in pure unhydrated form. Combustion performance finds an optimal coating thickness that enables combustion performance similar to a high performing CuO based thermite.