kevin moore- combustion behaviors of bimodal aluminum size distributions in thermites

63
COMBUSTION BEHAVIORS OF BIMODAL ALUMINUM SIZE DISTRIBUTIONS IN THERMITES by KEVIN MOORE, B.S.M.E. A THESIS IN MECHANICAL ENGINEERING Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of MASTER OF SCIENCE IN MECHANICAL ENGINEERING Approved Michelle Pantoya Chairperson of the Committee Brandon Weeks Louisa Hope-Weeks Accepted John Borrelli Dean of the Graduate School May, 2005

Upload: tremann

Post on 07-Oct-2014

22 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

COMBUSTION BEHAVIORS OF BIMODAL ALUMINUM

SIZE DISTRIBUTIONS IN THERMITES

by

KEVIN MOORE, B.S.M.E.

A THESIS

IN

MECHANICAL ENGINEERING

Submitted to the Graduate Faculty of Texas Tech University in

Partial Fulfillment of the Requirements for

the Degree of

MASTER OF SCIENCE

IN

MECHANICAL ENGINEERING

Approved

Michelle Pantoya Chairperson of the Committee

Brandon Weeks

Louisa Hope-Weeks

Accepted

John Borrelli Dean of the Graduate School

May, 2005

Page 2: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

ii

ACKNOWLEDGMENTS

First of all, I would like to thank my masters supervisor Dr. Michelle Pantoya for

all of the hard work she has dedicated to me reaching this point in my academic career. I

am so very grateful that she took a chance on me entering her research program, which

has been integral in my development as a student. Her positive attitude and optimistic

approach to life were crucial to my success.

I would also like to thank my supervisor and mentor at Los Alamos National

Laboratory, Dr. Steve Son. I am grateful for the opportunity to work with him and gain

so much experience from his guidance. He was always willing to take time out to talk

with me about whatever is on my mind, for which I am very thankful.

I would like to gratefully acknowledge the support of the Army Research Office

(Contract Number DAAD19-02-1-0214) and our program manager, Dr. David Mann.

I would also like to acknowledge support by the Los Alamos National Laboratory

through the Advanced Energetics Initiative, the Defense Threat Reduction Agency

(DTRA), and the Seaborg Institute.

I would like to thank the Texas Tech University Honors College for their support

through the Undergraduate Research Fellowship Program.

I would like to thank my wife, Lindsay, for the love and support she has given me

through this trying period. I would like to also acknowledge my family for their loving

support and always motivating me to put my best foot forward.

I would also like to thank several co-workers that I have had during my work

towards this point. Bryan Bockmon was an excellent mentor, during the Texas Tech

Page 3: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

iii

Mechanical Engineering Mentoring Program in the spring of 2002, and first introduced

me to the world of energetic materials. I learned so much from Jim Busse and Eric

Sanders, co-workers and lab mentors at Los Alamos National Laboratory, and were an

immense help in performing my summer experiments. Also, I would like to thank John

Granier for all of his help developing experiments here at Texas Tech.

Page 4: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

iv

TABLE OF CONTENTS

ACKNOWLEDGMENTS .................................................................................................. ii

ABSTRACT....................................................................................................................... vi

LIST OF TABLES............................................................................................................ vii

LIST OF FIGURES ......................................................................................................... viii

NOMENCLATURE ........................................................................................................... x

CHAPTER

I. INTRODUCTION............................................................................................. 1

1.1 Thermites ................................................................................................ 1 1.2 Nano-composite Thermites..................................................................... 2 1.3 Aluminum Combustion........................................................................... 4 II. EXPERIMENTAL ............................................................................................ 7

2.1 Sample Preparation ................................................................................ 7 2.2 Laser Ignition of Pellets ....................................................................... 10 2.3 Laser Ignition of Loose Powders ......................................................... 11 2.4 Combustion Velocities of Loose Powders........................................... 12 2.5 Pressure Cell Tests............................................................................... 13 III. THEORY ........................................................................................................ 14

3.1 Flame Structure..................................................................................... 14 3.2 Diffusion vs. Chemical Kinetics .......................................................... 15 IV. RESULTS....................................................................................................... 18

4.1 Ignition Sensitivity............................................................................... 18

Page 5: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

v

4.2 Combustion Velocity ........................................................................... 21 V. DISCUSSION ................................................................................................. 24

5.1 Ignition Sensitivity............................................................................... 24 5.2 Combustion Velocity ........................................................................... 26 VI. CONCLUSIONS ............................................................................................ 36

VII. IMPLICATIONS AND FUTURE RESEARCH............................................ 38

7.1 Nano-scale Aluminum Properties ........................................................ 38 7.2 Mixing Techniques .............................................................................. 39 7.3 Preliminary Molybdenum Trioxide Study........................................... 40 REFERENCES ................................................................................................................. 47

Page 6: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

vi

ABSTRACT

In recent years many studies that incorporated nano-scale or ultrafine aluminum

(Al) as part of an energetic formulation demonstrated significant performance

enhancement. Decreasing the fuel particle size from the micron to nanometer range alters

the material’s chemical and thermal-physical properties. The result is increased particle

reactivity that translates to an increase in the combustion velocity and ignition sensitivity.

Little is known, however, about the critical level of nano-sized fuel particles needed to

enhance the performance of the energetic composite. Ignition sensitivity and combustion

velocity experiments were performed using a thermite composite of Al and molybdenum

trioxide (MoO3) at the theoretical maximum density (TMD) of a loose power (5% TMD)

and a compressed pellet (50% TMD). A bimodal Al particle size distribution was

prepared using 4 or 20 µm Al fuel particles that were replaced in 10% increments by 80

nm Al particles until the fuel was 100% 80 nm Al. These bimodal distributions allow the

unique characteristics of nano-scale materials and their interactions with micron scale Al

particles to be better understood.

Page 7: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

vii

LIST OF TABLES

1. Reactant Particle Description................................................................................................7 2. Calculated thermal and physical parameters for all Al-MoO3 composites ........................28

Page 8: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

viii

LIST OF FIGURES

1. A SEM micrograph of an 80 nm and 4 µm Al mixture ..................................................8 2. Schematic diagram illustrating aluminum distribubimodal mixture ratios.....................9 3. Schematic diagram of the unconfined loose powder ignition sensitivity test

apparatus. .....................................................................................................................12 4. Schematic diagram of the combustion velocity test apparatus for loose powders. .....12 5. Typical thermite flame structure diagram.....................................................................14 6. Still frame images of ignition and flame propagation of an Al-MoO3 pellet with

an Al distribution of 60 wt % 80 nm and 40 wt % 4 µm.. ...........................................18 7. Ignition time for pellets as a function of percent 80 nm Al content mixed with

MoO3 and 4 µm Al at 50 W laser power ....................................................................19 8. Ignition time for pellets as a function of percent 80 nm Al content mixed with

MoO3 and 20 µm Al at 50 W laser power ..................................................................20 9. Ignition time for loose powders as a function of percent 80 nm Al content mixed

with MoO3 and 4 or 20 µm Al at 50 or 100 W laser power.........................................20 10. Combustion velocity as a function of weight percent 80nm Al content for pellets

with 4 µm Al and MoO3 .............................................................................................22 11. Combustion velocity as a function of weight percent 80nm Al content for pellets

with 20 µm Al and MoO3 ...........................................................................................22 12. Combustion velocity as a function of weight percent 80nm Al content for loose

powder with 4 or 20 µm Al and MoO3.......................................................................23 13. Pressure output as a function of percent nano-Al content for loose powder

Al-MoO3 mixtures with 4µm and 20µm Al. ..............................................................29 14. Combustion of 100% 80nm Al and MoO3 loose powder mixture..............................31 15. Combustion of loose powder Al and MoO3 mixture with 10% 80nm Al and

90% 20µm Al .............................................................................................................32 16. Combustion of 100% 4µm Al and MoO3 loose powder mixture................................33

Page 9: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

ix

17. SEM images of products of 50% 80 nm Al and 50% 20 µm Al pellet mixtures

with MoO3....................................................................................................................34

Page 10: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

NOMENCLATURE

Variable

A Arrhenius pre-exponential factor

B Mass transfer number

C Species concentration

D Mass diffusivity

Ea Activation energy

k Surface reaction rate constant

"m& Mass flux

r Radial dimension

R Gas constant

R Spherical fuel particle outer radius

T Temperature

Y Mass fraction

Greek

ρ Pure fuel density

ω Reaction rate

x

Page 11: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

1

CHAPTER I

INTRODUCTION

1.1 Thermites

A thermite reaction is often described as a reaction between a metal and a metallic

or non-metallic oxide, which results in a more stable metallic oxide and the

corresponding metal or non-metal of the reactant oxide [1]. An important product of

these reactions is a high heat of reaction, commonly referred to as energy density [1].

Once ignited, the thermite reaction is self-sustaining due to this highly exothermic heat

release. Thermite reactions are typically thought to be conductively driven reactions.

Most solid energetic materials, including explosives like trinitrotoluene (TNT)

and nitroglycerine, are composed of many molecules of the same engineered material.

These materials are referred to as monomolecular energetic materials. Within each of

these molecules are fuel, typically carbon, and oxidizing molecules which allow the

reaction to occur without the presence of additional oxygen from the environment.

Monomolecular energetics have kinetically controlled reactions which can produce very

high deflagrations or even detonations. Unlike monomolecular energetics, thermites are

mixtures of fuel and oxidizing particles, typically ranging from 1 to 100 microns in size.

Due to the inherent separation of the fuel and oxidizing particles, the reaction is diffusion

controlled, which usually translates to much slower deflagrations. Although thermites

have relatively high energy densities, the slow burning rate causes the energy release rate

to be much slower than that of monomolecular energetics.

Page 12: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

2

This slow energy release rate has limited the utility of thermite materials mostly

to areas outside of explosives. Welding underwater or in hard to reach places is often

solved by the use of thermites; molten metals from the reaction can seep between the

metal pieces and form a bead during the cooling process [1]. Self propagating high

temperature synthesis (SHS) is another important application of thermite-type materials

[3-6]. SHS, which often involves the intermetallic reaction between two or more metals,

allows the development of materials with more mechanically favorable microstructures

for specific applications [3]. One-such application is the formation of a ceramic layer on

the inside of metal piping induced by the reaction of a thermite on the inner surface of the

pipe [1].

1.2 Nano-composite Thermites

Particle size reduction of fuel and oxidizing particles within the micron range in

thermite systems has resulted in significantly improved combustion performance.

Shimizu and Saitou [7] have experimentally shown that the reaction rate of the Fe2O3-

V2O5 system increased as the number of contact points between the fuel and oxidizing

particles increased. Brown et al. [8] have seen similar results in the Sb-KMnO4 system.

As the calculated number of contact points in this system doubled due to smaller particle

sizes, the burning rate of the mixtures tripled [8]. Tomasi and Munir [9] have shown that

as the particle size of the Nb2O5-Al2Zr system decreased from 25 to 2 microns, the

reaction burn rate increased by a factor of five. Drennan and Brown [10] have shown

burn rates double as the molybdenum particle size of MoBaO2 and MoSrO2 systems is

reduced from 35 microns to 15 microns. It is apparent that the particle size of the mixture

Page 13: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

3

plays a significant role in the combustion performance for all of these systems. The

increase in reactivity for these cases is attributed to a better mixing between fuel and

oxidizing particles. Smaller particle sizes allow the fuel and oxidizing particles to

effectively be closer together, creating smaller diffusion distances that must be overcome

during the reaction.

Recent developments in particle formation technology have created the cost

efficient production of nano-scale particles [11]. One potential for nano-particles to

increase combustion performance is due to diffusion distances decreasing for particles of

this size. In addition to new particle formation technologies, new characterization

technologies, like small angle neutron scattering, have been developed to determine

particle size distributions more accurately for particles in the micron range [12-13].

The reduction of aluminum particle size to the nano-scale in the Al-MoO3 system

has shown promising results. An initial study by Son et al. [14] shows increased burning

rates in nano-sized aluminum thermite mixtures; an important finding of their

experiments shows that radiation is of little importance and that convective heat transfer

during the reaction process is a more prominent factor. Bockmon et al. [15] have shown

that burning rates of Al-MoO3 nano-powder mixtures can reach confined burning rates of

up to 1 km/s. These speeds, which are attributed to a prominent convective burning, are

much faster than thermites are traditionally known to burn. Traditional thermites burn on

the order of 1 m/s [1].

Direct comparison of nano and micron aluminum particle sizes in Al-MoO3

mixtures by Granier and Pantoya [16] yields very dramatic results. Granier and Pantoya

[16] examined the ignition and combustion velocity of Al-MoO3 pressed composites as a

Page 14: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

4

function of Al particle size. Their work showed that reducing the Al particle size from

micron to nanometer dimensions decreased the ignition delay time by two orders of

magnitude (from 10 s to 10 ms). The significantly enhanced ignition sensitivity was

attributed to the increased reactivity of the nanometer particles [17]. Pantoya and Granier

[17] showed through thermal analysis that the nano-Al and MoO3 reaction is initiated at

temperatures below the melting of Al (660 °C) and sublimation of MoO3 (770 °C) and

that the reaction takes place in the solid-solid phase. They also showed that for micron Al

reacting with MoO3, the reaction is initiated after Al melting and MoO3 sublimation such

that ignition is controlled by a liquid-gas diffusion mechanism. These results indicate that

the nano-Al particles are more reactive because they react in the solid phase and at lower

temperatures than the micron-Al particles.

1.3 Aluminum Combustion

An important part of the study of Al-MoO3 combustion is the behavior of

aluminum during the ignition process. Much work has been done related to the reactivity

of aluminum, both on the micron and nano scale, over the past 20 years [18-31]. In a

study by Popenko et al. [18], a mixture of ultrafine Al powder was combined with micron

Al powder for an examination of the bimodal Al distribution combustion behavior in air.

They analyzed the presence of bound nitrogen in the products of bimodal Al and air

combustion and found that for mixtures consisting of less than 70 % micron Al powder

the percent of bound nitrogen remained constant. The interesting finding was that the

bound nitrogen content in the combustion products of these mixtures decreases

considerably if the ultrafine Al concentration in the mixture is less than 20 % and this

Page 15: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

5

behavior is attributed to the concurrent processes of sintering and incomplete combustion

[18].

In a study by Trunov et al. [23], differential scanning calorimetry (DSC) and

thermogravimetric analysis (TGA) were used to analyze micron sized aluminum behavior

at increased temperatures. A focus of this study was on the phase changes of the alumina

oxide layer that covers the outside of each aluminum particle. Trunov determined the

initial oxide layer to be amorphous and between two to three nanometers thick for his

samples. During the TGA analysis, a weight gain was observed between 550-650 °C.

During this temperature range, the alumina layer reportedly underwent a phase change

from amorphous to γ-phase. The γ-phase has a density that is 20% higher than the

amorphous phase and after this transition the outer surface of the aluminum particle may

not be completely surrounded by an oxide layer. Exposed aluminum would oxidize as

oxygen from the atmosphere diffuses to the particle surface, which could account for the

weight gain observed.

Other factors may also contribute to the enhanced ignition sensitivity, such as

altered absorption properties which enable nano-Al particles to more readily absorbed

energy than their micron scale counterparts. In fact, Yang et al. [32] showed that the

absorption coefficient of 30-nm Al particles is significantly greater than micron scale

particles and a strong function of particle size (see Fig. 5 of Ref. 32).

Using nanometer combined with micron scale Al particles in rocket propellant

applications has strong advantages. For example, all Al particles are pyrophoric and

therefore passivated with an unreactive oxide shell (e.g., Al2O3). As the particle surface

area to volume ratio increases the presence of Al2O3 increases and becomes a significant

Page 16: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

6

portion of the overall mixture. Because propellant payloads can be restrictive, the

unwanted levels of an unreactive oxide that may add weight and reduce energy density

are undesirable [11, 35]. For this reason, adding small amounts of nanometer to micron-

scale Al particles may facilitate increased reactivity without the unwanted burdens of

excessive amounts of Al2O3. In a study by Dokhan et al. [36] the burning behavior of

ammonium perchlorate (AP) solid propellant with bimodal aluminum particle size

distributions was examined. They showed a significant increase in burn rate with only a

20% addition of nanometer Al. At this level, Dokhan et al. [36] showed Al combustion

takes place closer to the propellant burning surface allowing increased radiative and

conductive heat feed back that increases the temperature at the burning surface and

correspondingly increases the burn rate.

Due to the increased ignition sensitivity and burning rates of the nano-composite

thermites, a recent surge of interest is focused on developing thermites that may replace

traditional lead-based compounds in gun primers [37, 38]. Reducing the presence of

toxins such as lead in firearms will not only reduce health risks to personnel but will also

improve the environment. In particular, nanometer Al mixed with MoO3 and acetylene

black (a form of carbon) is being studied as a replacement for lead compounds [37].

This study will examine the ignition sensitivity and combustion velocity of Al and

MoO3 composites as a function of the Al particle size distribution. Mixtures are prepared

using 4 or 20 micron combined with 80 nm diameter Al particles in discrete mixture

ratios. Both powder and compressed pellet combustion was studied. The goal is to

investigate the influence of nanometer Al on the ignition sensitivity and combustion

velocity of thermites.

Page 17: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

7

CHAPTER II

EXPERIMENTAL

2.1 Sample Preparation

Table 1 shows physical data for the Al particles and the material supplier for the

Al and MoO3. The Al particles are encapsulated within a protective Al2O3 shell. The

active Al content is the percent of Al powder that is not in the form of Al2O3. As can be

seen with the 80 nm diameter Al powder, the Al2O3 shell becomes an appreciable portion

of the total powder, causing the active Al content to reduce to only 73 % (Table 1). The

active Al content is reported by the manufacturer. The particle diameters listed in Table

1 are an average value reported by the manufacturer. Micrographs, one of which is shown

in Figure 1, reveal that the reported average particle diameter is consistent with the

particles observed using scanning electron microscopy (SEM). The average oxide shell

thickness is calculated assuming that the oxide layer is uniformly surrounding spherical

particles and all particles are of the average particle size. Based on these assumptions and

the reported active Al content and the average particle diameter, the thickness of the shell

is calculated and tabulated in Table 1.

Table 1. Reactant Particle Description

Particle % Active Al Content

Al2O3 Layer Thickness Supplier

80 nm Al 73 4.3 nm Nanotechnologies 4 um Al 91 70 nm Alfa Aesar

20 um Al 99 30 nm Sigma Aldrich MoO3 - - Technanogy/Climax

Page 18: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Aluminum was mixed with MoO3 in a 40/60 wt % ratio which corresponds to a

fuel rich equivalence ratio of 1.3, based on active Al content. This mixture ratio was

shown to be an optimal composition for achieving the highest combustion velocity and

shortest ignition delay time [16]. The mixtures used to make pellets were dispersed in a

hexane solution and sonicated to break up agglomerates and ensure a homogeneous

mixture. The wet solution was poured into a tray and slightly heated to allow hexane

evaporation.

Figure 1. A SEM micrograph of an 80 nm and 4 µm Al mixture.

For pellet tests, a well-mixed, dried powder was separated into 230-270 mg

quantities and cold pressed with a hydraulic press (0.5 – 1.5 MPa) and a uniaxial die. All

final pellets were 6.51 mm in diameter and 3.9 mm in length. Higher pressures (1.5

8

Page 19: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

MPa) were needed to form the less dense nano-powder formulations to the same

dimensions of the micron-powder formulations. Theoretical maximum density (TMD)

calculations are based on the weighted average of Al, MoO3, and Al2O3 present in the

mixture. For pellets, the powders were pressed to a TMD of 50 % ( 2 g/cc) while the

pour density of powders was 5 % TMD (

≈ 0.2 g/cc).

Eleven mixtures of Al/MoO3 were prepared, each with a varying distribution of

Al particle size ranging from 100 % 80 nm to 100 % 4 or 20 µm diameter. The

distribution of active aluminum by particle size for each of the samples is illustrated in

Fig. 2.

Distribution of Active Aluminum

0

10

20

30

40

50

60

70

80

90

100

1 2 3 4 5 6 7 8 9 10 11

Mixture

Perc

ent A

ctiv

e A

lum

inum

Nano AlMicron Al

Figure 2. Schematic diagram illustrating aluminum distribubimodal mixture ratios.

9

Page 20: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

10

In all experiments, the thermite mixtures were burned in an ambient environment.

A study by Asay et al. [39] has experimentally shown that burning rates of Al/MoO3

powders were the same when in ambient and vacuum environments.

2.2 Laser Ignition of Pellets

The pellets were ignited using a 50-W CO2 laser (Universal Laser Systems Inc,

Scottsdale, AZ), power meter and associated optics. Ignition and flame propagation were

recorded using a Phantom IV (Vision Research, Wayne, NJ) high-speed camera which

captured images at 32,000 frames per second (fps). Details of this experimental apparatus

are discussed elsewhere [16].

Ignition is defined in the context of thermite combustion as the onset of a fully

sustained self-propagating reaction. There are several techniques for measuring an

ignition delay time [16]. The technique applied here is based on the “first-light” approach

in which ignition time is determined as the time lapse between sample exposure to laser

beam and detection of the first light. This may not guarantee ignition but is a commonly

used technique for experimentally determining ignition times [40]. The high-speed

camera is synchronized with the CO2 laser and detects light intensity. In this way, the

reaction light is used as the illuminating source to visualize the ignition process.

Burn rate is a measure of the burning solid surface of an energetic material and

often is used in reference to a single particle. Propellant combustion studies typically

refer to a mass burning rate which characterizes the regression rate of the combusting

solid propellant. In this thermite combustion, the flame consumes and spreads through

discrete particles packed in a highly porous matrix. The physics of flame propagation in

Page 21: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

11

this arrangement entails flame spreading and is a strong function of the packing

arrangement of particles and the porous structure of the material. The term combustion

velocity will be used to characterize the speed of the leading edge of the reaction zone

identified by visible light emission recorded from the high-speed photographic data.

2.3 Laser Ignition of Loose Powders

A schematic of the setup used in the laser ignition of powders is shown in Figure

3. A Coherent 250-W CO2 laser (Santa Clara, CA) was used to ignite the samples. A

NaCl beam splitter (Vigo Photovoltaic, Warsaw, Poland) diverted 8% of the laser beam

to a tri-metal detector used to determine the time of the start of the laser pulse. A

Thorlabs DET210 photo-diode (Newton, NJ) with a response time of 1 nanosecond,

measured light intensity of the reaction and determined the ignition time, based on a

Tektronix (Richardson, TX) oscilloscope with a response time of 1 nanosecond. The

ignition time was defined as the time at which the photo-diode trace reached 3% of its

peak voltage.

Page 22: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Figure 3. Schematic diagram of the unconfined loose powder ignition sensitivity test apparatus.

2.4 Combustion Velocities of Loose Powders

The unconfined loose powder tests were completed using an apparatus

schematically shown in Figure 4. A piezo-electric starter ignited the powder mixture

(120 mg) poured into the tray. A profiling glide leveled a consistent cross section

Figure 4. Schematic diagram of the combustion velocity test apparatus for loose powders.

12

Page 23: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

13

throughout the sample. The pour density of the powder was about 0.02 g/cc

corresponding to a TMD of 5 %. As the reaction progressed inside the tray, two holes, 1

mm in diameter, in the bottom of the tray, which were 2 cm apart, allowed light from the

reaction to be emitted into fiber-optic cables. Photo-diodes were used to convert this

light pulse to a voltage that was displayed on an oscilloscope.

The average velocity of the unconfined burn was determined from the two voltage

pulses on the oscilloscope. The time of reaction propagation between the holes was

determined by the time delay between the two voltage pulses. Due to the small size of

the hole, the voltage pulses did not occur until the reaction passed directly over the

pinholes. The average velocity was calculated by dividing the distance between the holes

by the time of reaction propagation between the holes. Previous tests have shown

through high speed photography that the reaction reaches its steady state velocity after a

very short distance, which supports the accuracy of this average velocity measurement.

2.5 Pressure Cell Tests

For the pressure cell experiments, a YAG Continuum MiniLite 100mJ laser

(Santa Clara, CA), by way of a Thorlabs fiber-optic cable, was used to ignite the samples

inside a parr-bomb type chamber. An aluminum cup contained the mixture at a constant

volume of 0.151 cm3. The aluminum cup did not react with the samples and was

weighed after each test to ensure this. A PCB Piezotronics (Depew, NY) pressure

transducer, with a response time of 1 microsecond, was used to determine maximum

pressure output for each test.

Page 24: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

CHAPTER III

THEORY

3.1 Flame Structure

Ignition of the thermite reaction is traditionally thought to be due to a phase

change in one or more reactants. Due to the diffusive nature of the reaction, melting and

sublimation are common phase changes that allow elevated temperature fuel and

oxidizing particles to come in better contact, which then facilitates ignition of the

reaction. The Al/MoO3 reaction occurs according to the following chemical equation:

2Al + MoO3 Al2O3 + Mo. ∆Hcomb=4279cal/cm3

The phase changes of this reaction could be the melt of aluminum at 660 °C

and/or the sublimation of MoO3 at 700 °C. A simplified diagram of the flame structure is

shown in Figure 5.

Figure 5. Typical thermite flame structure diagram

14

Page 25: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

As the reaction occurs, molten products spew away from the pellet or powder

specimen. Some of the energy from the reaction convectively heats a region of the

reactants before ignition of those particles. The thickness of the pre-heat zone is

determined by the porosity and thermal characteristics of the sample material.

3.2 Diffusion vs. Chemical Kinetics

The combustion behavior of these reactants can be explained by describing the

rate at which reactants (Al and MoO3) convert to products (Al2O3 and Mo) [41]. The

reaction can be reduced to considering a single small sphere of fuel (Al) and surrounding

oxidizer. First assume that the reaction occurs at the surface of the solid Al sphere and

the reaction is either limited by the chemical reaction rate or the oxidizer diffusion rate.

The reaction rate (ω) can be expressed as shown in Eq. (3.1) [41] assuming the following

global reaction.

32

2OAlAlCkC=ω (3.1)

2Al + MoO3 Al2O3 + Mo

In Eq. (3.1) and are the concentrations of the reacting chemical species

Al and Al

AlC32OAlC

2O3 respectively. Each concentration is raised to the power equal to the

corresponding stoichiometric coefficient. The specific reaction-rate constant, k, can be

expressed as an Arrhenius law according to Eq. (3.2) [41], where A is the pre-exponential

factor and is the activation energy of the reaction. aE

)exp(RTEAk a−

= (3.2)

15

Page 26: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Assuming that the spherical nano-scale particles burn on the surface as oxidizer

diffuses to the surface, the mass-burning rate ( ) of the sphere is given by Eq. (3.3) [41],

where is the mass flux at the surface, D, is the mass diffusivity and B is the mass

transfer number.

m&

"Rm&

)1ln(4"4 2 BDRmRm R +== ρππ && (3.3)

The mass flux from the particle surface would equal the fuel reaction rate. The

reaction rate is first order in oxygen concentration and second order in fuel concentration

as in Eq. (3.1).

2,," RFRoxR YkYm == ω& (3.4)

Assuming that the reaction rate occurs at the particle surface, the fuel mass

fraction ( ) in the gas phase is small compared to oxygen such that: RFY ,

0, ≈∞FY

1, =RFY

1, <<RoxY .

The mass transfer number can be expressed as

Roxox YYB ,, −= ∞ (3.5)

and

1,, << ∞oxRox YY .

The small number approximation can replace the natural log function in the mass

flux term.

)()1ln(" ,, RoxoxR YYRDB

RDm −≈+= ∞

ρρ& (3.6)

16

Page 27: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Solving for in Eq. (3.4) and substituting into Eq. (3.6), the mass flux at the

particle surface is given in Eq. (3.7).

RoxY ,

)(" ,

RDk

RDk

Ym oxR ρ

ρ

+= ∞& (3.7)

In diffusion controlled reactions, the reaction rates are fast compared to the

oxygen diffusion rates. Here k >> ρD/R and the mass flux is a function of the particle

radius according to Eq. (3.8).

RDYm oxR

ρ∞= ,"& (3.8)

If the reaction rates are comparable to or slower than the oxygen diffusion rate,

then the chemistry controls the reaction. In this case, k << ρD/R and the mass flux is

independent of particle radius (Eq. (3.9)).

∞= ," oxR kYm& (3.9)

For diffusion controlled reactions, the mass flux (Eq. (3.8)) and subsequent

reaction rates are inversely proportional to particle radius and directly proportional to

oxygen concentration. Eq. (3.8) suggests that smaller particles (nano-regime) will

produce higher mass flux rates than larger particles (micron regime) and thus produce

increased burn rates.

17

Page 28: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

CHAPTER IV

RESULTS

4.1 Ignition Sensitivity

Figure 6 is a sequence of still frame images illustrating ignition and flame

propagation of a pellet composed of 60 wt % 80 nm and 40 wt % 4 µm Al particles

combined with MoO3. The sequence of images was captured with the high-speed camera

at 32,000 fps (corresponding to 31 ms intervals). Ignition occurs at the pellet center front

face and the reaction propagates radially then axially through the pellet. The flame front

is stable and planar and fully self-sustained.

s

Figure 6. Still frame images of ignian Al distribution of 60 wt % 80 nm

m1 11.762

s

111.792 m

s

111.823 m

s

111.855 m

s

m1 11.886

s

111.916 m

s

111.947 m

tion and flame propagation of an Al-MoO3 pellet with and 40 wt % 4 µm. The time from trigger is given.

18

Page 29: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Figures 7 and 8 show the ignition delay time as a function of percent nano-Al

within the pellet mixture. The standard deviation bars represent the range of

measurements for 4-6 pellets from each of the 11 mixtures, the data symbol corresponds

to the average ignition delay time. The ignition times reported here for all mixtures

containing nano-Al are consistent with previous ignition time measurements made by

Granier and Pantoya [16] for 100 % nano-Al – MoO3 mixtures.

Figure 9 shows the ignition delay time as a function of percent nano-Al for a

loose powder mixture. Both mixtures containing either 4 or 20 µm Al are shown for laser

powers set at 50 or 100 W.

0

50

100

150

200

250

0 20 40 60 80 1Percent 80nm Al

Igni

tion

Del

ay T

ime,

ms

00

Figure 7. Ignition time for pellets as a function of percent 80 nm Al content mixed with MoO3 and 4 µm Al at 50 W laser power

19

Page 30: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

0

500

1000

1500

2000

2500

3000

0 20 40 60 80

Percent 80nm Al

Igni

tion

Del

ay T

ime,

ms

100

Figure 8. Ignition time for pellets as a function of percent 80 nm Al content mixed with MoO3 and 20 µm Al at 50 W laser power

1

10

100

1000

10000

0 20 40 60 80 1Percent 80nm Al

Ign

itio

n D

ela

y T

ime

, m

s

00

20µm and 80nm at 100W

4µm and 80nm Al at 100W

20µm and 80nm Al at 50W

4µm and 80nm Al at 50W

20

Figure 9. Ignition time for loose powders as a function of percent 80 nm Al content mixed with MoO3 and 4 or 20 µm Al at 50 or 100 W laser power.

Page 31: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

21

4.2 Combustion Velocity

The Phantom imaging software uses a designated light intensity transition and the

elapsed time between frames to calculate the combustion velocity. The combustion

velocity as a function of 80 nm Al content is plotted in Fig. 10 for the 4 µm mixture

pellet, in Fig. 11 for the 20 µm mixture pellet and Fig. 12 for the 4 and 20 µm mixture

powders.

In Figure 11, the data for 0 and 10 % 80-nm Al content are not shown. In the case

of 0% nm Al, the pellet was exposed to the laser for a relatively long period of time

before ignition. The exposure to this heat flux volumetrically heated the pellet, igniting

the pellet under different initial thermal conditions than for the pellet mixtures containing

nano-Al. For this reason, this data was not included. For the 10 % 80 nm Al sample, the

pellet could not sustain a self propagating wave and velocity measurements could not be

made.

Page 32: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

0

5

10

15

20

25

30

35

40

0 20 40 60 80

Percent 80nm Al

Com

bust

ion

Vel

ocity

, m/s

100

Figure 10. Combustion velocity as a function of weight percent 80nm Al content for pellets with 4 µm Al and MoO3

0

5

10

15

20

25

30

35

40

45

0 20 40 60 80

Percent 80nm Al

Com

bust

ion

Velo

city

, m/s

100

Figure 11. Combustion velocity as a function of weight percent 80nm Al content for pellets with 20 µm Al and MoO3

22

Page 33: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

The unconfined loose powder combustion velocities for both micron-scale

mixtures are plotted versus the percent nano Al content in Fig. 12. The piezo-electric

starter was not able to ignite mixtures with 0% or 10% nano Al content, therefore no data

are shown for these mixtures.

0

50

100

150

200

250

300

350

400

450

0 20 40 60 80 100

Percent 80nm Al

Co

mb

ust

ion

Ve

loci

ty,

m/

s

20µm and 80nm Al4µm and 80nm Al

Figure 12. Combustion velocity as a function of weight percent 80nm Al content for loose powder with 4 or 20 µm Al and MoO3.

23

Page 34: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

24

CHAPTER V

DISCUSSION

5.1 Ignition Sensitivity

The increased ignition sensitivity of mixtures with nanometer scale Al may be

explained by the increased reactivity of nano-sized metal particles. Pantoya and Granier

[16] showed through thermal analysis on nano and micron Al combined with MoO3

mixtures that the initiation mechanisms for the two particle length scales are very

different. They showed that for micron composites, Al oxidation occurs after Al melting

(660 °C) and MoO3 sublimation (770 °C) and in the solid-solid phase [16]. Ignition

temperatures corresponding to micron composites were measured at roughly 1000 °C and

the diffusion mechanism is in the liquid-gas state [16]. Figures 7-9 show that only 10 wt

% nano-Al content is required to reduce ignition delay times by two orders of magnitude.

Enough energy is generated from the localized reaction between nano-Al and

MoO3 to ignite neighboring particles. However, if the micron scale particles are too large

(20 µm) and only 10 % nano Al is within the matrix, enough heat is lost to the larger

scale particles to prevent a self-sustained reaction. In this case, shortly after a localized

ignition spot is formed, the reaction is extinguished. It is noted that Popenko et al. [18]

similarly observed that if the nano-Al concentration in a mixture with micron scale Al

powder is less than 10 %, combustion is difficult to initiate.

The effective thermal conductivity of the composite may also influence

ignition sensitivity. The effective thermal conductivity is a function of the thermal

conductivity values of each component within the matrix, the volume fraction, and the

Page 35: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

25

distribution of the matrix and dispersed phases(s). Hasselman and Donaldson [42]

theoretically investigated the role of material distribution in the effective thermal

conductivity of composites and found that the presence of an interfacial thermal barrier

could have a significant effect on composite thermal conductivity. In this system, the

Al2O3 shell encapsulating nano-Al particles could represent an interfacial thermal barrier.

As localized Al oxidation reactions occur, the Al2O3 may provide enough thermal

insulation to facilitate localized energy build up. When energy is constrained, heat losses

associated with thermal diffusion away from the localized reaction are reduced. This

reduction in heat losses effectively causes localized hot spots thermally insulated from

micron-scale Al particles.

Another factor that could contribute to the increased ignition sensitivity

associated with nano-Al is the physical change of the alumina layer during heating. The

oxide layer of the nano-Al particles is less than 5 nm thick and amorphous (Table 1).

Upon heating, Trunov and co-workers [23] showed that the amorphous alumina shell

undergoes a phase transition just below 500°C to form γ-Al2O3. This phase change

corresponds to an increase in density by about 20 % (from 3050 to 3660 kg/m3) [23].

This density increase will cause the aluminum core to become exposed to air within void

pockets and surrounding MoO3 particles. Because this phase transition occurs at

temperatures corresponding to the ignition temperature of the nanocomposite, exposure

of the solid-Al core may be a rate-determining step in the solid-solid diffusion

mechanism of the nano-Al and MoO3 reaction. This step may not be as critical in the

ignition of micron-composites because the aluminum surface particles of the aluminum

core of a micron-Al particle make up a much smaller portion of the total aluminum

Page 36: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

26

content. When considering the same amount of active aluminum for the nano-Al and

micron-Al particles, nano-Al will have much more aluminum core surface area that can

be exposed during the phase transition of the Al2O3 shell since there are so many more

particles for the same amount of mass. The energy released from the nano-Al during this

step would be much larger than the energy release of the micron-Al, which could greatly

increase ignition sensitivity of these nanoparticles.

5.2 Combustion Velocity

The combustion wave speeds shown in Figs. 10 and 11 increase from roughly 1 to

40 m/s as nanometer Al content increases. The most interesting behavior is observed in

Fig. 11 in which a sharp transition from relatively slow to fast flame propagation occurs

between 50 and 70 % nanometer Al content. A similar trend was previously observed in

porous explosive charges and attributed to a transition from normal to convective burning

[43]. When convective burning takes precedence over thermal conduction and radiation,

energy and mass transfer in the burning zone are driven by gas jets that penetrate into the

pores of the energetic material. Bobolev et al. [43] explain that in some cases penetration

of the combustion into the pores is followed by the establishment of a regime of

stationary convective burning whose rate substantially exceeds the normal burning rate.

In an effort to identify this burning regime, an analysis of porous energetic material

combustion under constant pressure conditions has lead to a stability criterion known as

the Andreev number, An (Eq. (5.1)) [43]. This non-dimensional parameter is similar to

Page 37: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

the Peclet number except tailored for reacting flow through porous media.

An=g

phb

kcUdρ

=const. (5.1)

In this equation, ρb is the bulk density of the composite, U is the measured combustion

wave speed, dh is the hydraulic pore diameter, cp is the heat capacity of the composite and

kg is the thermal conductivity of the gas. If this value exceeds a certain constant,

combustion penetrates into the porous structure and convective mechanisms dominate

flame propagation.

In making this calculation it is necessary to estimate the characteristic pore size

which is termed the hydraulic pore diameter, dh, given by Eq. (5.2) [44].

)1(4

εε−

=o

h Ad (5.2)

In this equation ε is the bulk void volume, Ao is the specific surface area based on the

solid volume and is calculated as the solid surface area divided by the solid volume

(As/Vs), and (1-ε) is the solid volume fraction [44].

Table 2 shows the calculated values of the physical and thermal parameter for

both the 4 and 20 micron containing bimodal mixtures. These calculations are based on

the assumption of kg = 0.137 W/mK corresponding to air at an average temperature of

2000 K. For each estimate, the An number is significantly less than 1.0. Bobolev et al.

[43] calculated critical An numbers between 3 and 10 and generalized that if the An > 6.0

combustion will penetrate into the pores and convective mechanisms will play the

primary role in accelerating the flame front.

27

Page 38: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Table 2. Calculated thermal and physical parameters for all Al-MoO3 composites

0 2272 885.4 0.165 1.2678E-03 1.98E-08 5.27E-14 3.24E-0710 2270 884.4 0.982 7.5389E-03 1.97E-08 8.83E-15 1.78E-0720 2268 883.4 1.798 1.3810E-02 1.96E-08 4.80E-15 2.83E-0730 2266 882.5 2.615 2.0081E-02 1.95E-08 3.29E-15 1.92E-0740 2264 881.5 3.431 2.6352E-02 1.94E-08 2.50E-15 1.33E-0750 2262 880.5 4.248 3.2623E-02 1.93E-08 2.01E-15 2.85E-0760 2260 879.5 5.064 3.8894E-02 1.92E-08 1.68E-15 3.19E-0770 2258 878.5 5.881 4.5165E-02 1.91E-08 1.44E-15 3.70E-0780 2256 877.5 6.698 5.1437E-02 1.90E-08 1.26E-15 3.93E-0790 2253 876.5 7.514 5.7708E-02 1.89E-08 1.11E-15 3.56E-07100 2251 875.5 8.331 6.3979E-02 1.87E-08 9.99E-16 5.44E-07

0 2278 889.8 0.033 2.5289E-04 2.01E-08 2.67E-1310 2276 888.4 0.863 6.6255E-03 2.00E-08 1.02E-1420 2274 887.0 1.692 1.2998E-02 1.99E-08 5.16E-15 7.60E-0830 2272 885.5 2.522 1.9371E-02 1.98E-08 3.45E-15 6.89E-0840 2269 884.1 3.352 2.5743E-02 1.96E-08 2.58E-15 7.68E-0850 2267 882.7 4.182 3.2116E-02 1.95E-08 2.06E-15 1.53E-0760 2264 881.2 5.012 3.8488E-02 1.94E-08 1.71E-15 3.43E-0770 2261 879.8 5.841 4.4861E-02 1.92E-08 1.46E-15 6.26E-0780 2258 878.4 6.671 5.1234E-02 1.91E-08 1.27E-15 6.33E-0790 2254 876.9 7.501 5.7606E-02 1.89E-08 1.12E-15 5.90E-07100 2251 875.5 8.331 6.3979E-02 1.87E-08 9.99E-16 4.37E-07

Hydraulic Pore Diameter (m)

Andreev Numberε/(1−ε)Αο (1/m)% nano Al

contentTotal Surface

Area (m2)ρ (kg/m3) Cp (J/kgK)

For 80 nm Al mixed with 20 micron Al particles and MoO3

For 80 nm Al mixed with 4 micron Al particles and MoO3

Αο (1/m) ε/(1−ε)Hydraulic Pore Diameter (m)

Andreev Number

% nano Al content ρ (kg/m3) Cp (J/kgK)

Total Surface Area (m2)

The above analysis suggests that the transition from low to high combustion wave speeds

may not result from a shift in the flame propagation mechanism. A better explanation for

this behavior may be related to the ignition insensitivity of the micron-scale Al particles.

Because the micron-scale particles require more time and higher temperatures to ignite,

the nano-Al reactions may proceed too quickly to allow the micron-Al particles to

participate in the reaction. In this way, reactants containing large amounts of micron-Al

particles may experience significant incomplete combustion which would result in

reduced combustion velocities.

For mixtures with mostly nano-Al particles (70-90% nano Al), the micron-Al

particles make up a significantly smaller portion of the volume and impede the nano-Al/ 28

Page 39: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

MoO3 reaction much less. Channels of nano-Al and MoO3 exist throughout the powder

or pellet, allowing the reaction to propagate without micron Al obstructions hindering the

velocity. These channels allow the measured velocity to be the same as a mixture of pure

nano-Al and MoO3 but may still result in incomplete reactions between micron Al and

MoO3.

To test the theory that incomplete combustion is responsible for the reduced

velocities in Figures 10-12, a series of pressure measurements were performed using the

pressure cell described previously. Figure 13 shows the peak pressure of both the 4 and

20 µm Al loose powder mixtures as a function of percent nano-Al content. The pressure

is displayed on a per unit mass basis because a slight pour density change occurred as the

size distribution of the aluminum particles changed. The constant volume sample cup

0

5

10

15

20

25

30

35

0 20 40 60 80 1

Percent 80nm Al Content

Pres

sure

Out

put,

kPa/

mg

00

20µm and 80nm Al4µm and 80nm AlCalculated Pressure

Figure 13. Pressure output as a function of percent nano-Al content for loose powder Al-MoO3 mixtures with 4µm and 20µm Al.

29

Page 40: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

holds more material if composed of the more-dense micron particles. The sample amount

slightly increased as a result, but, more importantly, the volume of the pressure vessel

remained constant. No data points exist for the 0-20% nano-Al content due to the

difficulty of igniting the mixture with the YAG laser.

The maximum peak pressure of the powder samples increases as the amount

of nano-Al in each mixture increases (Figure 13). The pressure increase corresponds

linearly to the increase in nano-Al content. This trend suggests that the micron-Al

particles do not contribute to raising the peak pressure. For example, assuming a

complete reaction between all Al particles and MoO3, the maximum peak pressure for the

mixture can be calculated assuming the peak pressure results from the rapid expansion of

high temperature gasses present in the parr-bomb chamber under roughly adiabatic flame

temperature conditions. Equation 4.3 is the ideal gas law, which assumes that the gas

behavior at the flame temperature is ideal.

VmRTP = (4.3)

In this equation, P is the peak pressure, T is evaluated at the adiabatic flame

temperature of the mixture 3200 °C [1], which is sufficiently high enough to consider an

ideal gas assumption. The gas constant (R) for air at 3200 °C is used in conjunction with

the volume (V) of the combustion chamber, 9.26cm3. The mass (m) of gas present in the

chamber is estimated as 8.75 mg based on [1] which states that 25% of the products of

this reaction are gaseous. With these values in Eq. (4.1), the maximum peak pressure for

a complete reaction is 25.2 kPa/mg, and is shown in Fig. 13 as the dashed line. The

measured peak pressures approach 25.2 kPa/mg only when a significant percent of nano-

30

Page 41: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Al is included in the mixture, indicating that the micron-Al particles may not be

contributing to the overall reaction enough to raise the peak pressure to the theoretical

value.

Increased levels of nano-Al may also enhance the radiant heat transport

through the mixture. Yang et al. [32] showed nano-scale Al particles absorb more energy

compared to their micron scale counterparts. The increased levels of absorbed energy

may generate a more intense radiation field and elevate the temperature of the preheat

zone. Higher preheat temperatures will induce increased combustion velocities. If too few

nano-particles are present in the composite, the thickness and intensity of this preheat

zone will diminish and result in comparably slower combustion velocities.

Still frame images and light traces captured during powder ignition experiments

also complement the theory that micron scale particles are either not participating in the

reaction or react much slower than the reaction between nano-Al and MoO3. The six

frames of Figures 14-16 show the combustion of three different powder mixtures. The

frames of Figure 14 show the reaction of 100% 80nm Al and MoO3.

Figure 14. Combustion of 100% 80nm Al and MoO3 loose powder mixture

31

Page 42: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

The reaction appears to consume all particles, which rise into the air and convectively

cool as smoke. More importantly, all of the particles seem to react at the same time,

although a few get forced out onto the sample platform as seen in the third frame. This

whole reaction is outside the range of the photodiode in as little as 9 ms..

A different looking reaction appears in Figure 15, which is a mixture of 10%

80nm Al and 90% 20 µm Al with MoO3. The reaction of the nano Al and MoO3 can be

seen in the first two frames with a convective plume similar to Figure 14. Many other

particles can be seen radiating in these frames. The micron Al particles may only begin

reacting during the nano-Al reaction with MoO3 and forced into the air while still

radiating and reacting. These radiating particles appear to burn much slower and are

Figure 15. Combustion of loose powder Al and MoO3 mixture with 10% 80nm Al and 90% 20µm Al

32

Page 43: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

Figure 16. Combustion of 100% 4µm Al and MoO3 loose powder mixture.

apparent even after the nano-Al plume has disappeared. The time duration of these six

frames is about 13 times the time duration of the frames of Figure 14; therefore, it can be

seen that the micron particles react more slowly than the nano Al with MoO3. This

finding is consistent with the peak pressure measurements which indicate that the micron

particles do not contribute to increasing the measured peak pressure.

The frames of Figure 16 show a reaction with 4 µm Al and MoO3. This reaction

also occurs slowly relative to the reaction in Figure 14, about 2.5 times as slow. The

reaction spreads from the left to the right across the sample cup in the first three frames

and sends a bright plume of radiating particles into the air. Radiating particles can still be

seen floating in the air in the final two frames. It is apparent that the reaction of micron

Al particles, even as small as 4 µm, with MoO3 occurs much more slowly than the nano

Al reaction.

Products of a pellet reaction with 50% 80 nm Al and 50% 20 µm Al mixed with

MoO3 are shown in Figure 17. These micron particles were removed from the test

apparatus after combustion. The surfaces of these particles are of both smooth and rough

nature. The smooth sections are typically indicative of unreacted Al, as seen in Figure 1.

33

Page 44: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

The rough sections, however, indicate aluminum that has reacted with oxygen. These

SEMs show that micron Al particles in this mixture are not completely reacting, which

supports the still frame images of Figures 15 and 16.

A. B.

C.

Figure 17. SEM images of products of 50% 80 nm Al and 50% 20 µm Al pellet mixtures with MoO3

34

Page 45: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

35

The results obtained for bimodal mixtures in a thermite composite are similar to

results of bimodal Al mixtures reacting with air made by Popenko [18]. They suggest

incomplete combustion for nano-Al concentrations less than 20 % [18]. This finding is

consistent with the slower reacting (or incomplete combustion) micron Al particles

observed in this study. Also, Dokhan [36] showed a plateau in burning performance for

70 % nano-Al concentration in bimodal Al mixtures added to AP. Our work suggests 70

% nano-Al concentration will optimize the combustion wave speed in thermites.

The results presented here may have implications towards the use and

handling of thermites. Nano-scale Al particles can be costly, significantly more so than

micron scale particles. It may be advantageous to use mixtures of nano and micron scale

material for large scale formulations. This work shows that ignition sensitivity is

heightened by small additions of nano-Al to the mixture. However, what is gained in

ignition sensitivity is sacrificed in performance through reduced velocities that result

from incomplete (or significantly slower) micron-Al reactions.

Page 46: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

36

CHAPTER VI

CONCLUSIONS

The ignition delay time and combustion wave speed of thermite composites

composed of Al and MoO3 were examined as a function of Al particle size distribution.

Bimodal Al size distributions consisting of 80 nm combined with either 4 or 20 µm

particles showed increased sensitivity to ignition with only a 10 % concentration of nano-

scale particles. The increased ignition sensitivity is explained by the solid-solid phase

ignition mechanism associated with nano-Al particles compared with the liquid-gas phase

mechanism associated with micron-Al particles. The nano-Al particles are more reactive

and stimulate ignition two orders of magnitude faster than with purely micron-Al

containing mixtures. Although the energy generated from the localized reaction between

nano-scale Al and oxidizer is enough to initiate the reaction, in some cases the reaction is

not self-sustaining. For example, with 10 % nano-Al particles combined with 20 micron

particles the initiated reaction is quickly quenched as heat is lost to the surrounding media

too quickly to permit a sustained reaction. This study indicates that at least 20 % nano-Al

particles are required to ensure reduced ignition delay times and are on the same order as

with mixtures containing 100 % nano-Al.

The combustion velocity increases linearly with nano-Al content from roughly 1

to 40 m/s. This is consistent with prior investigations on bimodal Al size distributions

concerning Al with air and Al as a propellant additive. Slow (and incomplete)

combustion of micron Al particles leads to the decrease of the combustion wave speed for

mixtures in which the micron Al particles make up a large part of the volume. For

Page 47: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

mixtures with small amounts of micron Al, the reaction speed is significantly increased

since the nano Al can react with MoO3 in channels that exist around the micron particles.

37

Page 48: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

CHAPTER VII

IMPLICATIONS AND FUTURE RESEARCH

Research in nanotechnology has advanced in the past decade. Within the field of

nano-energetic materials, research is still needed to help characterize the materials for

many combustion and thermal properties. Safety, longevity, and reliability are all factors

that are important in the use and predictability of these materials, and further research in

these areas will expediate the use of nano-energetics in many applications.

7.1 Nano-scale Aluminum Properties

Nano-aluminum obviously has striking combustion property differences from

micron-aluminum, as seen in this work, but a potential for significant differences in its

physical properties may also exist. One of these physical properties that has not been

extensively studied is radiation absorbance. Nano-aluminum appears a very dark gray

color, while micron-aluminum is light gray. This color difference shows optical

properties of aluminum change in the visible light spectrum with decreased particle size,

possibly due to dependent scattering. Groundwork has been laid on the potential for this

property difference [32, 33, 34], but the range of wavelengths researched is not broad.

Integrating spheres have been used for gaining reflectance information of

aluminum surfaces [45] and could be very useful for the evaluation of powder aluminum

samples. Lindseth et al. [45] experimentally determined that the rolling of aluminum

surfaces can reduce the reflectivity of the surface by over 10%. Thermites and other

38

Page 49: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

nano-aluminum mixtures are often ignited using laser irradiation, and the absorbance,

transmittance and reflectance properties of nano-aluminum could explain which

wavelengths are most effective in this ignition process.

An important safety consideration of many materials, especially energetic

materials, is the risk of dust explosions, to which even common grains are susceptible.

The risk of micron-sized aluminum in regard to random dust explosions has been heavily

documented, but due to its recent development, nano-aluminum risks in this area have not

been studied. The enhanced ignition capabilities of nano-aluminum compared to micron-

aluminum could pose a much higher threat to explosion and must be assessed to sustain

proper usage procedures. Nano-sized aluminum is much less dense than micron

aluminum in its powder form and can more easily be dispersed in air due to its increased

surface area. The addition of nano-Al to explosives, for increased energy output, poses

another significant risk that should be assessed. Kwok et al. [30] have shown that the

addition of nano-scale aluminum to ammonium perchlorate decreases the minimum

electro-static discharge ignition energy from greater than 156 mJ to 6 mJ. In order to

safely handle nano-Al, more research is being done to quantify the risk of nano-Al dust

explosions and the proper precautionary measures for prevention.

7.2 Mixing Techniques

The enhanced performance of nano-composites over their micron counterparts is

largely due to the superior mixing distribution of the fuel and oxidizing particles. In

order to harness the full potential of these materials, techniques to produce optimum

39

Page 50: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

mixing of these particles are necessary. Also, inconsistent performance of these materials

that could hinder their reliability could be reduced through research in this area. Current

mixing techniques utilize sonication of alcohols and other non-reactive solvents to break

up particle agglomerations and create a homogeneous mixture of fuel and oxidizing

particles. Research regarding mixing behaviors has begun [46], and sol-gel chemistry

solutions through nano-structuring are also being investigated [47-49]. Mixing

techniques will continue to be an important factor in the development and production of

nano-energetic applications in the future.

7.3 Preliminary Molybdenum Trioxide Study

After completing the bimodal aluminum distribution study, a preliminary study

was performed on the combustion effects related to environmental alterations of

molybdenum trioxide. During lab operations, a color change was noticed in molybdenum

trioxide nanoparticles that were exposed to light and the effects of the color change on

combustion speeds were unknown. This drew the attention to what environmental

conditions caused this change and what other environmental conditions could cause the

material to react differently.

Metal oxides like molybdenum oxide have been known for over 40 years to

exhibit photochromic properties [50-57]. Ultraviolet (UV) light has been shown to cause

several chemical, mechanical, and electrical properties of metal oxides to change.

Indium oxide thin films have been shown to become much more electrically conductive

through exposure to UV radiation [55]. When in an oxygen-present environment,

40

Page 51: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

amorphous zinc oxide thin films will crystallize due to UV irradiation [56].

Photodarkening in these same zinc oxide films has been attributed to photoreduction

effects. Li has shown similar UV-induced color formation and photoreduction

observations in tungsten oxide nanoparticles, which are claimed to describe a change of

the oxide from WO3 to W2O5 [58].

Molybdenum oxide also exhibits photochromic properties. S. K. Deb [57] reports

the formation of “color centers” which result in a significant increase in optical density

due to UV irradiation. He describes this color change as being due to an increase in

oxygen vacancies in the lattice structure. He also describes a thermal bleaching process,

which entails heating the sample to 300 ºC in the presence of oxygen. The bleached

molybdenum oxide particles are rid of these color centers and do not form new color

centers in the presence of UV irradiation. In the past ten years, new technologies have

allowed the formation of nanoparticle molybdenum oxide (MoO3), which seem to have a

stronger photochromic properties. Absorbance measurements by Li have shown

nanoparticle MoO3 to be 15 times more absorbent of UV irradiation than their bulk MoO3

counterparts.

Raman spectroscopy has also been used to characterize MoO3 [59-60]. According

to S. H. Lee [59], heat treatment of amorphous thin films, in a manner similar to the

thermal bleaching described previously, causes alpha MoO3 crystal formation to occur.

Raman spectroscopy experiments from Lee shown sharp peak formation previously not

seen with the amorphous samples.

41

Page 52: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

More information in this area of research is needed to describe variability and

inconsistency seen in performance parameters, such as combustion rate and ignition

sensitivity. Plantier has shown standard deviations of Al/Fe2O3 formulations to be as

high as 15% of the mean combustion velocity [49]. Granier has shown ignition time and

combustion speed standard deviations in Al/MoO3 formulations of greater than 20% of

the mean values [61].

Although the ability of MoO3 optical properties to change through exposure to

UV irradiation is known, little is known about how the combustion properties are affected

by similar environmental factors. A change in combustion characteristics could lead to a

greater knowledge of causes of variations in the performance of nano-composites. This

information would lead to the development of more consistent nano-composites and

expediate the use of environmentally responsible primer formulations. Therefore, the

scope of this preliminary study was to evaluate how the combustion behavior of nano-

composite Al/MoO3 formulations is affected by MoO3 particles that have been altered

through exposure to fluorescent light, UV irradiation and high relative humidity levels,

common environmental factors which may change the combustion performance of these

materials.

The aluminum used in this study was 120 nm, and the same molybdenum oxide

product from Climax was used as in the bimodal distribution study. This molybdenum

oxide was tested in its original form as received from Climax and after a heat treatment of

400°C. SEMs of the untreated and heat treated MoO3 are shown in Figure 18.

42

Page 53: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

a. b.

Figure 18. SEM micrographs of a. Untreated MoO3 and b. Heat treated MoO3

Powders were prepared in a process similar to the preparation process mentioned

previously and were tested in an open channel setup with spark ignition. The

molybdenum oxide, both untreated and heat treated, was exposed to three different

environmental conditions for up to 4 days before being mixed with the aluminum: UV

light exposure, fluorescent light exposure, and a 99% relative humidity environment.

Figures 19-21 show the combustion velocities as a function of exposure to

fluorescent light, UV light, and high humidity levels.

43

Page 54: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

0

50

100

150

200

250

300

350

400

450

500

0 20 40 60 80 10Exposure to Fluorescent Light (hrs)

Bur

n Ve

loci

ty (m

/s)

0

Untreated Heat Treated

Figure 19. Burn rate of Al/MoO3 as a function of fluorescent light exposure

0

50

100

150

200

250

300

350

400

450

0 20 40 60 80 10

Exposure to UV (hrs)

Bur

n Ve

loci

ty (m

/s)

0

UntreatedHeat Treated

Figure 20. Burn rate of Al/MoO3 as a function of UV light exposure

44

Page 55: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

0

50

100

150

200

250

300

350

400

0 20 40 60 80 10

Exposure to 99% Relative Humidity (hrs)

Bur

n Ve

loci

ty (m

/s)

0

UntreatedHeat Treated

Figure 21. Burn rate of Al/MoO3 as a function of 99% relative humidity

The untreated MoO3 powder changed color from yellow to blue after both

fluorescent and UV light exposure, which has been attributed to a desporption of oxygen

from the molecules [57]. However, these types of exposure had little effect on the burn

rate of the mixture. The white heat treated MoO3 did not change color after both light

exposures, due to the differing crystal phase structure shown in Figure 18. Like the

untreated MoO3, light exposure had little effect on the burn rate.

Although light had little effect on the combustion performance for both MoO3

samples, exposure to high humidity levels drew a clear difference between the samples.

The untreated MoO3 sample burned at speeds less than a meter per second after just one

45

Page 56: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

day of exposure, which was probably due to water molecules that had absorbed onto the

MoO3 particles. The heat treated MoO3 sample showed no change after two days

exposure and a moderate decrease thereafter. The heat treated MoO3 has not been

thoroughly studied and could provide an alternative for untreated MoO3 in harsh

environmental applications.

Based on this preliminary study, it seems environmental factors can play a

significant role in the performance of thermites. Increased humidity levels can reduce the

combustion performance of thermites very quickly, an light exposure has the potential to

decrease the performance over a longer period of time. In order to utilize these materials

in ordinance applications, environmental factors must be addressed to accurately predict

how these materials will perform. Future research could provide a more definite

correlation of the effects of these and other environmental factors on thermite materials.

46

Page 57: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

REFERENCES

1. Wang, L.L., Munir, Z.A., and Maximov, Y.M. Journal of Materials Science 28 (1993) 3693-3708.

2. Fischer, S.H. and Grubelich, M.C. “Theoretical Energy Release of Thermites, Intermetallics, and Combustible Metals.” Proceedings of the 24th International Pyrotechnics Seminar, Monterey, CA, July 1998.

3. Varma, A. “Form from Fire.” Scientific American. August (2000) 58-61.

4. Mukasyan, A.S., Rogachev, A.S., Varma, A. Chemical Engineering Science 54 (1999) 3357-3367.

5. Varma, A., Rogachev, A.S., Mukasyan, A.S., and Hwang, S. Proc. Natl. Acad. Sci. USA 95 (1998) 11053-11058.

6. Hunt, E.M. and Pantoya, M.L. “Nanocomposite Reactions in the Self-Propagating High-Temperature Synthesis of Nickel Alumnide Systems” Proceedings of the Third Joint Meeting of the U.S. Sections of The Combustion Institute (2003).

7. Shimizu, A. and Saitou, J. Solid State Ionics 38 (1990) 261-269.

8. Brown, M.E., Taylor, S.J., and Tribelhorn, M.J. Propellants, Explosives, Pyrotechnics 23 (1998) 320-327.

9. Tomasi, R., and Munir, Z.A. J. Am. Ceram. Soc. 82 [8] (1999) 1985-1992.

10. Drennan, R.L. and Brown, M.E., Thermochimica Acta 108 (1992) 223-246.

11. Pesiri, D., Aumann, C.E., Bilger, L., Booth, D., Carpenter, R.D., Dye, R., O’Neill, E., Shelton, D. and Walter, K.C. “Industrial Scale Nano-Aluminum Powder Manufacturing” To appear in Journal of Pyrotechnics (2004)

47

Page 58: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

12. Peterson, R.P., Mang., J.T., Son, S.F., Asay, B.W., Fletcher, M.F., Hjelm, R.P. and Roemer, E.L. “Microstructural Characterization of Energetic Materials” Proceedings of the 29th International Pyrotechnics Seminar (2002).

13. Mang, J.T., Hjelm, R.P., Peterson, P.D., Jorgensen, B.S. and Son, S.F. “Characterization of Components of Nano-Energetics by Small-Angle Scattering Techniques.” Proceedings of the 31st International Pyrotechnics Seminar (2004).

14. Son, S.F., Asay, B.W., Busse, J.R., Jorgensen, B.S., Bockmon, B. and Pantoya, M. “Reaction Propagation Physics of Al/MoO3 Nanocomposite Thermites” Proceedings of the 28th International Pyrotechnics Symposium (2001).

15. Bockmon, B.S., Pantoya, M.L., Son, S.F. and Asay, B.W., “Burn Rate Measurements in Nanocomposite Thermites,” Proceedings of the Americal Institute of Aeronautics and Astronautics Aerospace Sciences Meeting, Paper No. AIAA-2003-0241 (2003).

16. Granier, J. J. and Pantoya, M. L., Combustion and Flame 138 (2004) 373-383.

17. Pantoya, M. L. and Granier, J. J., to appear in Propellants Explosives Pyrotechnics (2004).

18. Popenko, E.M., Il’in, A. P., Gromov, A. M., Kondratyuk, S. K., Surgin, V. A., and Gromov, A. A., Combustion, Explosion, and Shock Waves 38 [2] (2002) 157-162 .

19. Il’in, A.P., Gromov, A.A. and Yablunovskii, G.V., Combustion, Explosion, and Shock Waves, 37 [4] (2001) 418-422.

20. Il’in, A.P., Yablunovskii, G.V., Gromov, A.A., Popenko, E.M. and Bychin, N.V. Combustion, Explosion, and Shock Waves 35 [6] (1999) 656-659.

21. Kwon, Y., Gromov, A.A., Ilyin, A.P., Popenko, E.M., Rim, G. Combustion and Flame 133 (2003) 385-391.

48

Page 59: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

22. Il’in, A.P., Popenko, E.M., Gromov, A.A., Shamina, Y.Y. and Tikhonov, D.V., Combustion, Explosion, and Shock Waves 38 [6] (2002) 665-669.

23. Trunov M.A., Schoenitz M., Zhu X. and Dreizin E.L., “Effect of Polymorphic Phase Transformations in Al2O3 Film on Oxidation Kinetics of Aluminum Powders,” submitted to Combustion and Flame, June 2004.

24. Dreizin, E.L., Progress in Energy and Combustion Science 26 (2000) 57-78.

25. Dreizin, E.L. Combustion, Explosion, and Shock Waves 39 [6] (2003) 681-693.

26. Dreizin, E.L. Combustion and Flame 105 (1996) 541-556.

27. Driezin, E.L., Combustion and Flame 117 (1999) 841-850.

28. Jones, D.E.G., Brousseau, P., Fouchard, R.C., Turcotte, A.M. and Kwok, Q.S.M. Journal of Thermal Analysis and Calorimetry 61 (2000) 805-818.

29. Ilyin, A., Gromov, A., et al. Propellants, Explosives, Pyrotechnics 27 (2002) 361-364.

30. Kwok, Q.S.M., Fouchard, R.C., Turcotte, A., Lightfoot, P.D., Bowes, R., and Jones, D.E.G., Propellants, Explosives, Pyrotechnics 27 (2002) 229-240.

31. Sandstrom, M.M., Oschwald, D.M. and Son, S.F. “Laser Ignition of Aluminum Nanoparticles in Air.” Proceedings of the 31st International Pyrotechnics Seminar (2004).

32. Yang, Y., Wang, S., Sun, Z., and Dlott, D. D, J. Appl. Phys. 95 [7] (2004).

33. Yang, Y., Sun, Z., Wang., S., Dlott, D.D. J. Phys. Chem. B 107 (2003) 4485-4493.

34. Wang, S., Yang, Y., Sun, Z. and Dlott, D.D. Chemical Physics Letters 368 (2003) 189-194.

49

Page 60: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

35. Kramer, M. “Nanotechnology Aids the Advanced Energetics Program” AFRL Technology Horizons, March 2003.

36. Dokhan, A., Price, E. W., Seitzman, J. M., Sigman, R. K., “The Effects of Bimodal Aluminum with Ultrafine Aluminum on the Burning Rates of Solid Propellants,” Proceedings of the Combustion Institute, 29 [2] (2002) 2939-2945.

37. Giles, J., Nature 427 (2004) 580-581.

38. Naud, D.L., Hiskey, M.A., Son, S.F., Busse, J.R. and Kosanke, K. Journal of Pyrotechnics 17 (2003)

39. Asay, B. Son, S., Busse, J., and Oschwald, D. “Observations on the Mechanism of Reaction Propagation in Metastable Intermolecular Composites.” Proceedings of the APS Schock Physics Meeting, Portland OR, July 2003

40. Ostmark, H. and Roman, N., J. Appl. Phys. 73 [4] (1993) 1993-2003.

41. Turns, S.R., An Introduction to Combustion: Concepts and Applications, McGraw-Hill Companies, Inc. New York, 2000

42. Hasselman, D. P. H and Donaldson, K.Y., “The Role of Particle Size in the Effective Thermal Conductivity of Composites with an Interfacial Thermal Barrier,” Journal of Wide Bandgap Materials, 7 [4] (2000).

43. Bobolev, V.K., Margolin, A.D., and Chuiko, S.V., “ Stability of Normal Burning of Porous Systems at Constant Pressure,” Combustion, Explosion and Shock Waves, 2 [4] (1966) 24-32.

44. Kaviany, M., Principles of Heat Transfer In Porous Media, Springer-Verlag, New York (1991)

45. Lindseth, I., Bardal, A., and Spooren, R. Optics and Lasers in Engineering 32 (2000) 419-435.

46. Lee, C.S., Lee, J.S. and Oh, S.T., Materials Letters 57 (2003) 2643-2646.

50

Page 61: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

47. Tillotson, T.M., Gash, A.E., Simpson, R.L., Hrubesh, L.W., Satcher, Jr., J.H. and Poco, J.F. Journal of Non-Crystalline Solids, 285 (2001).

48. Gash, A.E., Tillotson, T.M., Satcher, J.H., Poco, J.F., Hrubesh, L.W., Simpson, R.L. Chemistry of Materials, 13 [3] (2001) 999 -1007.

49. Plantier, K., Pantoya, M.L., and Gash. “Combustion Behavior of Sol-Gel Synthesized Nanocomposite Thermites,” Proceedings of the Third Joint Meeting of the U.S. Sections of The Combustion Institute, March 2003

50. Fujita, Y. and Kwan T. Bull. Chem. Soc. Jpn. 31 [380] (1958)

51. Zelaya-Angel, O., Menezes, C. and Sanchez-Sinencio, F. J. Appl. Phys. 51 [11] (1980)

52. Yao, J. N., Hashimoto, J. N. and Fujishima, A. Nature 355 [624] (1992)

53. Reyes-Betanzo, C., Herrera-Perez, J. L., Cocoletzi, G. H., Zelaya-Angel, O. J. Appl. Phys. 88 [1] (2000)

54. Itoh, M., Hayakawa, K., Oishi, S. J. Phys.: Condens. Matter 13 [6853] (2001)

55. Bender, M., Natsarakis, N., Gagaoudakis, E., Hourdakis, E., Douloufakis, E., Cimalla, V., and Kiriakidis, G. J. Appl. Phys. 90 [10] (2001)

56. Asakuma, N., Hirashima, H., Imai, H., J. Appl. Physics. 92 [10] (2002)

57. Deb, S. K. and Chopoorian, J. A. J. Appl. Physics. 37 [13] (1966)

58. Li, S., Germanenko, I., and El-Shall, M. S. J. Cluster Science, 10 [4] (1999)

59. Lee, S., Seong, M. J., Tracy, C. E., Mascarenhas, A., Pitts, J. R., and Deb, S. Solid State Ionics. 147 [129] (2002)

51

Page 62: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

60. Zhao, Y., Jingguo, L., Zhou, Y., Zhang, Z., Xu, Y., Naramoto, H., and Yamamoto, S. J. Phys.: Condens. Matter. 15 [L547] (2003)

61. Granier, J., and Pantoya, M., Laser ignition study of medium density nanocomposite thermites. Submitted to Combustion and Flame (May 2003)

52

Page 63: Kevin Moore- Combustion Behaviors of Bimodal Aluminum Size Distributions in Thermites

PERMISSION TO COPY

In presenting this thesis in partial fulfillment of the requirements for a master’s

degree at Texas Tech University or Texas Tech University Health Sciences Center, I

agree that the Library and my major department shall make it freely available for research

purposes. Permission to copy this thesis for scholarly purposes may be granted by the

Director of the Library or my major professor. It is understood that any copying or

publication of this thesis for financial gain shall not be allowed without my further

written permission and that any user may be liable for copyright infringement.

Agree (Permission is granted.)

Kevin Moore__________________________________ ________________ Student Signature Date Disagree (Permission is not granted.) _______________________________________________ _________________ Student Signature Date

53