molecular characterization of energetic materials

133
MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS A Dissertation by SANJEEV R. SARAF Submitted to the Office of Graduate Studies of Texas A&M University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY December 2003 Major Subject: Chemical Engineering

Upload: others

Post on 24-Dec-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

MOLECULAR CHARACTERIZATION

OF

ENERGETIC MATERIALS

A Dissertation

by

SANJEEV R. SARAF

Submitted to the Office of Graduate Studies of Texas A&M University

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

December 2003

Major Subject: Chemical Engineering

Page 2: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

MOLECULAR CHARACTERIZATION

OF

ENERGETIC MATERIALS

A Dissertation

by

SANJEEV R. SARAF

Submitted to Texas A&M University in partial fulfillment of the requirements

for the degree of

DOCTOR OF PHILOSOPHY

Approved as to style and content by: _____________________________ _____________________________ _____________________________ _____________________________ _____________________________

December 2003

Major Subject: Chemical Engineering

M. Sam Mannan (Chair of committee)

David M. Ford (Member)

Michael B. Hall (Member)

Dan F. Shantz (Member)

Kenneth R. Hall (Head of Department)

Page 3: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

iii

ABSTRACT

Molecular Characterization of Energetic Materials. (December 2003)

Sanjeev R. Saraf, B. Chem. Engg., U.D.C.T, Mumbai, India

Chair of Advisory Committee: Dr. M. Sam Mannan

Assessing hazards due to energetic or reactive chemicals is a challenging and

complicated task and has received considerable attention from industry and regulatory

bodies. Thermal analysis techniques, such as Differential Scanning Calorimeter (DSC),

are commonly employed to evaluate reactivity hazards. A simple classification based on

energy of reaction (-∆H), a thermodynamic parameter, and onset temperature (To), a

kinetic parameter, is proposed with the aim of recognizing more hazardous

compositions. The utility of other DSC parameters in predicting explosive properties is

discussed.

Calorimetric measurements to determine reactivity can be resource consuming,

so computational methods to predict reactivity hazards present an attractive option.

Molecular modeling techniques were employed to gain information at the molecular

scale to predict calorimetric data. Molecular descriptors, calculated at density functional

level of theory, were correlated with DSC data for mono nitro compounds applying

Quantitative Structure Property Relationships (QSPR) and yielded reasonable

predictions. Such correlations can be incorporated into a software program for apriori

prediction of potential reactivity hazards. Estimations of potential hazards can greatly

help to focus attention on more hazardous substances, such as hydroxylamine (HA),

which was involved in two major industrial incidents in the past four years. A detailed

discussion of HA investigation is presented.

Page 4: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

iv

To mom, dad, aka, aaji

and

all my family members in Nagpur

Page 5: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

v

ACKNOWLEDGEMENTS

I would like to thank Dr. Sam Mannan for all his guidance and encouragement

during this study and for his invaluable mentorship. I’m grateful to Dr. Rogers for his

contributions to all the projects and for being extremely patient. I would like to thank Dr.

Ford for serving on my committee and allowing me access to his SGI. I would like to

express my appreciation to my committee members: Dr. Dan Shantz and Dr. Michael

Hall.

I learnt a lot from my collaborators and would like to express my gratitude to

individuals I worked with: David Frurip, Sima Chervin, Seshu Dharmavarm, and

Abdulrehman Aldeeb. I would like to thank Dr. Lisa Pérez for her help with theoretical

calculations; she has always been a source of inspiration. I thank the Texas A&M

Supercomputing Facility for computer time and the Laboratory for Molecular Simulation

(LMS) for software and support.

I would like to acknowledge my family members, friends, roommates, former

teachers, and students and staff of the Mary Kay O’ Connor Process Safety Center for

being supportive of my work.

Everyday I promise myself not to commit any faux pas and tell God that I’ve

been good so far; but soon I wake up. Therefore I would like to take this opportunity to

apologize for any misconduct.

Page 6: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

vi

TABLE OF CONTENTS

Page

ABSTRACT………………………………………………………………………… iii

DEDICATION……………………………………………………………………… iv

ACKNOWLEDGEMENTS………………………………………………………… v

TABLE OF CONTENTS…………………………………………………………… vi

LIST OF TABLES………………………………………………………………….. viii

LIST OF FIGURES………………………………………………………………… ix

CHAPTER

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

II EXPERIMENTAL CHARACTERIZATION OF REACTIVE

HAZARDS…………………………………………………………….

4

1. Experimental Techniques…………………………………………... 4 1.1 Thermal Analysis…………………………………………… 4 1.2 Sensitivity Tests……………………………………………. 11

2. Classifying Reactive Hazards………………………………………. 16 2.1 Classification Based on Sensitivity Tests…………………... 17 2.2 Classification Based on Calorimetric Data…………………. 17 2.3 Proposed Classification……………………………………. 21

3. Further Investigation of DSC Parameters to Quantify Reactivity…. 27 3.1 Experimental Data…………………………………………... 29 3.2 Results and Discussion……………………………………… 31

4. Conclusions………………………………………………………… 35

III STRUCTURE BASED PREDICTION OF REACTIVITY

HAZARDS ……………………………………………………………

38 1. Review of the Available Methods…………………………………. 39

1.1 Rules of Thumb……………………………………………. 39 1.2 Oxygen Balance Method………………………………….. 42 1.3 CHETAH…………………………………………………… 43

Page 7: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

vii

CHAPTER Page

1.4 CART……………………………………………………… 46 2. Advanced Prediction Techniques…………………………………… 48 3. Application of Transition State Theory for Thermal Stability

Prediction…………………………………………………………… 49

3.1 Model Development……………………………………….. 50 3.2 Experimental Details……………………………………….. 58 3.3 Results and Discussion…………………………………….. 59

4. Correlating Calorimetric Data with Molecular Descriptors……….. 64 4.1 Data Set Selection………………………………………….. 65 4.2 Discussion of a Few Descriptors…………………………… 66 4.3 Correlations………………………………………………… 69 4.4 Correlations Using the Semi-empirical Method, AM1……. 73

5. Conclusions and Future Work……………………………………….. 76

IV DETAILED INVESTIGATION OF A REACTIVE SYSTEM……… 81

1. Background………………………………………………………… 81 2. Ab initio Heat of Formation for HA………………………………. 82

2.1 Computational Methods…………………………………… 84 2.2 Results and Discussion…………………………………….. 87 2.3 Choice of Best Values……………………………………… 91

3. Investigation of Hydroxylamine Runaway Behavior ……………... 95 3.1 Experimental Observations……………………………….. 95 3.2 Theoretical Calculations…………………………………… 96

4. Integrating Reactivity Data and Risk Analysis for Improved Process Design……………………………………………………... 100

5. Conclusions………………………………………………………… 104

V CONCLUSIONS..…………………………………………………….. 105

LITERATURE CITED……………………………………………………………... 107

APPENDIX A……………………………………………………………………... 117

APPENDIX B……………………………………………………………………... 119

APPENDIX C……………………………………………………………………... 120

APPENDIX D……………………………………………………………………... 121

VITA………………………………………………………………………………... 123

Page 8: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

viii

LIST OF TABLES

TABLE

Page

1.1 Energies Associated with Typical Reactions…………………………… 2

2.1 Comparison of Available Calorimeters………………………………… 5

2.2 Explosivity Rank ……………………………………………………….. 18

2.3 Instability Rating Based on IPD………………………………………... 20

2.4 Summary of RSST Data for 30 wt % DTBP in Various Solvents……… 25

2.5 Summary of DSC and Sensitivity Data………………………………… 30

2.6 Explosion Propagation vs. Actual Rank………………………………... 36

2.7 Summary of Sensitivity Tests Predictions……………………………… 37

3.1 Functional Groups Indicative of Reactive Hazards…………………… 40

3.2 Oxygen Balance and Hazard Rank……………………………………... 42

3.3 Summary of Heat of Formation and Maximum Heat of Decomposition. 53

3.4 Summary of Experimental and Predicted Values for To (Tonset)………... 60

3.5 Comparison of Onset Temperatures……………………………………. 63

3.6 Onset Temperatures (observed and predicted)………………………... 71

3.7 Experimental Energy of Reaction .…………………………………….. 74

3.8 Summary of Energy of Reaction Values……………………………… 75

3.9 Decomposition Energies for Typical Functional Groups………………. 77

3.10 Typical Values for Energetic Polymerization…………………………... 78

3.11 -∆H, To Values Based on Functional Groups…………………………... 79

4.1 Experimental Heats of Formation (1atm and 298.17 K)……………….. 86

4.2 Summary of Calculated Heats of Formation (∆Hf)…………………….. 88

4.3 Accurate Values for HA Heat of Formation……………………………. 92

Page 9: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

ix

LIST OF FIGURES

FIGURES

Page

2.1 A typical DSC run………………………………………………………. 7

2.2 RSST…………………………………………………………………….. 8

2.3 RSST run on 50-wt% hydrogen peroxide………………………………. 8

2.4 APTAC………………………………………………………………….. 9

2.5 APTAC run on 50-wt% hydroxylamine/water system………………… 10

2.6 Gap test………………………………………………………………….. 12

2.7 BAM impact sensitivity apparatus……………………………………… 13

2.8 Koenen test……………………………………………………………… 15

2.9 Time/pressure test assembly…………………………………………….. 16

2.10 Proposed classification for reactive chemicals………………………….. 23

2.11 Heat of reaction and onset temperature for DTBP in various solvents… 26

2.12 Schematic of DSC curves for two energetic materials………………….. 28

2.13 Correlation between UN Gap test and DSC data……………………….. 32

2.14 Relationship between the Koenen test and DSC data…………………... 33

2.15 Relationship between the Time/Pressure test and DSC data…………… 34

3.1 Hazard evaluation output from CHETAH 7.2…………………………... 45

3.2 Hypothesized potential energy surface (PES) for a runaway reaction….. 54

3.3 Comparison of onset temperatures …………………………………… 61

3.4 Predicted onset temperatures……………………………………………. 72

4.1 Systematic approach for assessing reactive hazards……………………. 82

4.2 Deviations from the average heat of formation values for the methods

employed in Table 4.3…………………………………………………...

93

Page 10: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

x

Page

4.3 APTAC temperature-time profile for 50 wt% HA……………………… 95

4.4 Probable elementary reactions of HA in presence of Fe2+ …...………… 99

4.5 Fe2+ interacting with the nitrogen atom of HA ………………………… 99

4.6 Fe2+ interacting with the oxygen atom of HA …………………………. 100

4.7 Hydroxylamine production……………………………………………… 102

4.8 Quantitative risk analysis scheme………………………………………. 103

Page 11: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

1

CHAPTER I

INTRODUCTION

Hazards posed by chemical reactions have historically received considerable

attention from industries, government agencies, and communities. There have been

numerous publicized incidents related to reactive chemicals.1,2

A typical chemical plant routinely produces, stores, and transports a number of

highly energetic chemicals. However, a few of these chemicals may pose toxic, fire, or

explosion hazards. This work focuses on evaluation of energetic hazards due to chemical

reactions. The release of energy stored within substances can be triggered by

temperature, mechanical impact, shock, or electrostatic energy. Therefore, it is important

to understand the conditions and scenarios leading to rapid release of energy and

consequently explosion.

Based on investigations of past incidents, the U.S. Chemical Safety and Hazard

Investigation Board (CSB) has concluded that reactive hazards are a significant problem3

in the Chemical Process Industry (CPI) and has recommended regulation of reactive

hazards. The state of New Jersey has broadened its Toxic Catastrophe Prevention Act

(TCPA)4 to include reactive chemicals. Also, the Center for Chemical Process Safety

(CCPS) has recently published a book dealing with reactive hazard management.5 Thus

there is increased impetus on managing reactivity within a manufacturing unit and

understanding reactive behavior of chemicals to prevent incidents.

This dissertation follows the style and format of Industrial & Engineering Chemistry Research.

Page 12: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

2

The main aim of reactive chemicals legislation is to enforce reactive hazard

assessment. In an informal proposal, CSB has proposed a heat of reaction (-∆H) value of

100 cal/g as a threshold for performing reactive hazard assessment. This is an extremely

conservative criterion. For example, according to this criterion all the substances in

Table 1.1 would be subjected to reactive hazard assessment.

Table 1.1. Energies Associated with Typical Reactions 6

Chemical Reaction Associated Energy (-∆H, cal/g)

Methanol combustion 5430 Rusting of iron 1200

TNT decomposition 1100 Hydrogen peroxide decomposition 700 Ammonium nitrate decomposition 382

Sucrose (table sugar) decomposition 114

The reactive chemical section of the TCPA rule utilizes lists of chemicals and certain

functional groups, along with threshold values, as a trigger to perform reactive hazard

assessment. But neither the New Jersey regulation nor the CSB proposal consider

reaction kinetics.

It is important to point out the extreme difficulty to list properties characterizing

reactivity hazards. For example, commercial explosives are typically characterized by

2000 cal/g or more of energy; however, most of the chemicals leading to incidents in the

chemical industries have energies between 500 and 1500 cal/g. Thus, there are a variety

of reasons, besides the energy content, that can pose chemical reactivity hazards, and

recognizing chemicals and hazardous conditions is an area of considerable research. The

Page 13: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

3

objectives of this work are to study the characteristics of reactivity hazards, expedite

hazard assessment, develop guidance for recognizing, and evaluating more hazardous

compositions and thus enable better utilization of resources.

Chapter II discusses experimental characterization of reactive hazards and

proposes a classification to rank reactive hazards. Chapter III highlights the role of

molecular modeling to predict reactive hazards and presents results of molecular

modeling. Chapter IV discusses the hydroxylamine system, as an example of a highly

reactive system.

Page 14: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

4

CHAPTER II

EXPERIMENTAL CHARACTERIZATION OF REACTIVE HAZARDS

1. Experimental Techniques

There are a variety of experimental techniques to characterize and quantify

hazards due to chemical reactions. Experimentation provides a better understanding of

the energy content of a substance and its behavior under various conditions. Such

information is extremely useful for assessing reactive hazards and managing risks.

Several popular experimental techniques are discussed in Sections 1.1 and 1.2.

1.1 Thermal Analysis

A reactivity hazard involves conversion of stored chemical energy of the

components into mechanical or heat energy, and it is the uncontrolled release of this

stored energy that causes the damage in a reactive chemical incident. The reactivity of a

substance is normally assessed by performing calorimetric measurements.6 Information

about the amount of energy released and the rate of energy released for a process

chemical can be obtained by performing calorimetric tests. A small amount of the

sample is heated over a range of temperature (usually within 30 oC – 400 oC), and

temperature, pressure, and time data are recorded. This information is then used for

alarm settings, relief sizing, and process modeling. Overall thermodynamics and kinetics

of a reaction can be estimated from temperature-time data obtained from a calorimeter,

and this information is used to identify the material hazards posed by a composition and

risk of potential runaway reactions.

Page 15: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

5

1.11Types of Calorimeters

There are various calorimeters available for performing reactive hazard

assessments. Prior to detailed testing, screening tests are performed7 using calorimeters

such as a Differential Scanning Calorimeter (DSC) or the Reactive System Screening

Tool (RSST) from Fauske and Associates (http://www.fauske.com). Such screening tests

are relatively inexpensive and can be performed quickly. Detailed testing can be

performed using other calorimeters such as the Automated Pressure Tracking Adiabatic

Calorimeter (APTAC) from TIAX (http://www.tiax.biz) or the Vent Sizing Package

(VSP) from Fauske and Associates. A comparison of three available calorimeters is

presented in Table 2.1 and a brief discussion of the various calorimeters is provided in

the following paragraphs.

Table 2.1. Comparison of Available Calorimeters

Calori- meter

Capital Cost

Time for a run

Sample size

Scanning Rate

(oC/min)

Data obtained

Comments

DSC 1 $ 1 hr 1-10 mg 10 T vs. time

Popular method to screen

reactive hazards

RSST 1.5/2 $ 6 hrs Up to 10 ml 1-5 T,P vs.

time

Open cell; data can be used for

relief sizing

APTAC / VSP 5 $

12-16 hrs

Up to 130 ml 1-2

T,P vs. time

Maintains adiabatic

conditions; maintenance

intensive

Page 16: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

6

DSC

A DSC run can provide an overall indication of exothermic activity of the

composition being tested and can help assess potential reactive hazards. In a DSC, a

sample and a reference are subjected to a continuously increasing temperature and heat

is added to the reference to maintain it at the same temperature as the sample. This added

heat compensates for the heat lost or gained as a consequence of an overall endothermic

or exothermic reaction. When the rate of heat generation (Watts) in the sample exceeds a

particular value, the heat supply to the sample is cut off and this additional heat gain is

attributed to exothermic activity within the sample. This cut-off value depends on the

sensitivity of the particular instrument. For an exothermic reaction, a heat vs. time curve

exhibits a peak as shown in Figure 2.1. A base line is constructed from the initial heating

mode, and another line is drawn to coincide with the initial rise due to the exotherm. The

temperature, at the intersection of the two lines is called the onset temperature and

corresponds to a detectable level of heat due to a chemical reaction. The detected onset

temperature is thus a measure of the reaction kinetics and serves as a guideline for

selecting process or storage temperature. The energy released (-∆H) during the process is

calculated as the area under the heat-supplied (Watts) and time curve. DSC is a popular

screening tool because it is safer, since it involves a small amount of sample (1-10 mg),

and is faster, since with 10 oC/min scanning rate, a DSC run can be completed in an

hour. Normally, during a DSC experiment, pressure data are not recorded.

Page 17: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

7

Figure 2.1. A typical DSC run.

RSST

The RSST uses a 10-ml sample cell contained in an open, well-insulated glass

test cell. A cross-section of the RSST is shown in Figure 2.2. The RSST is a ramping

calorimeter and ramps the temperature of the sample at a fixed rate using an electric

heater. It allows scanning rates up to 2 oC/min and can generate temperature, pressure,

and time profiles. Output from the RSST for 50 wt% hydrogen peroxide-water system is

illustrated in Figure 2.3. The RSST is used for screening reactive hazards, since it

provides temperature, pressure vs. time data at a moderate cost compared to the APTAC

or VSP.

Baseline Onset temperature (To)

Temperature

Hea

t sup

plie

d (W

)

Page 18: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

8

Figure 2.2. RSST.

0

50

100

150

200

250

300

Time (min)

0

50

100

150

200

250

300

350

400

450

500

Figure 2.3. RSST run on 50 wt% hydrogen peroxide.

Sample cell

Container vessel

Temperature

Pressure

Page 19: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

9

APTAC

The DSC and the RSST, discussed earlier, are quasi-adiabatic calorimeters, since

the sample cell losses heat to the surroundings. Adiabatic calorimeter minimizes the heat

loss to the surrounding by maintaining the surrounding temperature as close to the

sample temperature, and has proven to be an extremely useful tool to assess thermal

hazards. Following the screening tests, detailed measurements are generally performed

for more hazardous compositions using an adiabatic calorimeter such as the Automated

Pressure Tracking Adiabatic Calorimeter (APTAC). A cross-section of the APTAC is

shown in Figure 2.4 and a typical output is illustrated in Figure 2.5.

Figure 2.4. APTAC.

Reaction vessel Side

Stirrer Pressure

vessel

Top

Bottom

heater

Tube heater

Page 20: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

10

The APTAC can be operated in a variety of test modes, such as heat-wait-search,

heat-ramping, and isothermal. If the self-heat rate of the sample is greater than a preset

threshold (0.1 °C/min), the apparatus tracks the reaction adiabatically until the reaction

is over or if one of the shutdown criteria is met. If no exotherm is detected, the sample is

heated to the next search temperature and the steps are repeated until one of the shut-

down criteria is met. Besides the temperature, the pressure outside the sample cell is

controlled to match the pressure inside the sample cell.

Figure 2.5. APTAC run on 50 wt.% hydroxylamine-water system

Hydroxylamine decomposition test

0

50

100

150

200

250

300

400 600 800 1000 1200 1400 1600

Time (min)

Onset TemperatureTo

Maximum TemperatureTmaxStep Heating

Wait

Search

Adiabatic Mode Begins

Adiabatic Mode Ends

Page 21: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

11

1.2 Sensitivity Tests

Calorimetric tests capture temperature-time response of a substance and are

performed to detect thermal instability. However, the energy stored within a substance

can be released by a variety of stimuli. Sensitivity is defined as the ease with which a

substance subjected to external stimuli, such as shock, impact or heat, can undergo

detonation.8 A few of the techniques used to determine sensitivity9 of a material are

discussed below.

1.21 Shock Sensitivity10

The Gap test determines initiation of an explosion of a substance due to

detonation in the vicinity. Two cartridges of the smallest commercially manufactured

diameter are coaxially attached on a rod made of soft iron, wood, or plastic, as illustrated

in Figure 2.6. The gap value is the distance between the two cartridges. The gap medium

is such that it stops flying particles and direct heat transmission completely, thus serving

as a heat filter. Consequently, the shock wave is the only energy transmitted to the

substance being tested. The donor charge is a well-characterized explosive, for example

50 g RDX, that generates a known pressure wave (shock wave), and is set off during the

test. The resulting shock wave, generated during this explosion, is transmitted to the

testing material and may trigger a detonation.

Page 22: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

12

Figure 2.6. Gap test.10

Detonation in the test sample is verified on the basis of observed mechanical effects. The

test results, based on degree of fragmentation observed, are typically reported as follows:

‘Yes’ or ‘+’ Tube completely fragmented or fragmented at both ends

‘Partial’ Tube only fragmented at booster end but fragmented length is more

than 1.5 times the average fragmented length found with an inert

material.

‘No’ or ‘−’ Fragmented length is less than 1.5 times the fragmented length with

an inert substance.

1.22 Impact Sensitivity11

During impact tests, the impact of a drop-weight on a substance is assessed. The

sample, placed between two flat, parallel, hardened steel surfaces, is subjected to an

impact by dropping a weight. The impact may result in initiation depending on

sensitivity of the material, weight mass, and its drop height (impact energy). Initiation is

Page 23: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

13

observed by sound, light effects, or smoke, or by inspection. The BAM impact

apparatus, known to give fairly reproducible results, is shown in Figure 2.7. Typically

drop weights having a mass of 1, 2, 5, or 10 kg are used and the lowest impact energy

required to create a detonation is recorded. Thus drop-weight and drop-height at which

the initiation of the sample occurs are the main parameters determined from impact

testing. The drop height at which detonation is observed is thus a measure of impact

sensitivity of an explosive.

Figure 2.7. BAM impact sensitivity apparatus.11

1. Guiding rods 2. Locking and unlocking devices 3. Drop weight 4. Calibrated scale 5. Indented rod 6. Piston device 7. Anvil 8. Steel block 9. Steel Base

Page 24: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

14

1.23 Heat Sensitivity10

Heat sensitivity tests, such as Koenen and time/pressure, are performed to assess

the role of heat in initiation of explosives.

Koenen test

The Koenen test measures the effect of strong heating under confinement. The

test sample is contained in a drawn steel tube (27 ml) equipped with a closure, which

allows orifice plates with various apertures of diameter 1.0, 1.5, 2.0, 2.5, 3.0, 5.0, 8.0,

12.0, or 20.0 mm. The tube is heated with four calibrated propane burners. The result

reported from such a test is the largest size orifice at which the tube is fragmented, and

the following guidelines are used for reporting:

‘Violent’ Limiting diameter greater than or equal to 2.0 mm.

‘Medium’ Limiting diameter is 1.5 mm.

‘Low’ Limiting diameter is equal to or less than 1 mm but an effect is observed on

the tube.

‘No’ Limiting diameter is less than 1 mm and in all tests the tube is unchanged.

A schematic of the apparatus used for performing Koenen tests is shown in Figure 2.8.

Page 25: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

15

Figure 2.8. Koenen Test.

Time/Pressure test

This test measures the ability of a material to deflagrate under confinement. A 5

g sample is subjected to a flame in a pressure vessel (20 ml) fitted with a pressure

recording device and a bursting disc (2200 kPa). Based on the shortest time in three runs

for the pressure to rise from 690 to 2070 kPa, test results are classified as follows:

‘Rapid’ Time is less than 30 ms.

‘Slow’ Time is 30 ms or more.

‘No’ A gauge pressure of 2070 kPa is not achieved.

A schematic of assembly used for the time/pressure test is shown in Figure 2.9.

Nozzle

Burner

Sample test tube

Page 26: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

16

Figure 2.9. Time/Pressure test assembly.

Sensitivity tests typically require more sample, elaborate testing facilities, and

are more expensive than calorimetric tests. The next section discusses hazard

classification of substances based on sensitivity tests, calorimetric tests, and

interrelationship between the two.

2. Classifying Reactive Hazards

Based on tests discussed in the earlier section, researchers have attempted to

develop a classification to enable ranking of chemicals based on material hazards. Such

a ranking can help can help to develop guidelines for handling, storage, and

transportation of materials.

Burst Disc

Sample (5 g.)

Pressure Transducer

Ignition system

Page 27: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

17

2.1 Classification Based on Sensitivity Tests

Table 2.2 provides an explosivity-ranking scheme based on the three

recommended UN sensitivity tests, namely Gap, Koenen, and time/pressure. It is

identical to that used by Whitmore12, except this scheme gives precedence to the UN

Gap result over the BAM 50/60. Ranks A and B identify potential Class 1 substances:

“A” indicates substances that detonated, and “B” indicates those substances that did not

detonate but were strongly positive in the Koenen and/or Time/Pressure tests. Rank C

substances, which had milder results in the Koenen and Time/Pressure tests, are not

Class 1 but are candidates for classification as Self-Reactives or Organic Peroxides.

Rank D substances exhibited no positive results but may still be Self-Reactives or

Organic Peroxides, based on the results of the other recommended tests. This

classification is popular for categorizing substances for transportation.

2.2 Classification Based on Calorimetric Data

Unlike sensitivity tests, hazard classification based on calorimetry is not

well-established. There is disagreement among researchers regarding parameters

characterizing reactive chemicals. A part of the problem is that results from calorimetric

studies are highly dependent on the calorimeter and other conditions during the

experiment. The remainder of this chapter discusses approaches for developing a hazard

classification and associated issues.

Page 28: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

18

Table 2.2. Explosivity Rank

Explosivity Rank

Severest Class 1 Property Correspondence to UN Classification

A Detonates (positive result in UN Gap, or BAM 50/60 or TNO 50/70 if UN Gap unavailable)

Potentially Class 1

B Heating under confinement: Violent (Koenen limiting diameter >2 mm), and/or Deflagration: Rapidly (pressure in Time/Pressure >2070 kPa in <30 ms)

Potentially Class 1, but does not detonate

C Heating under confinement: Medium or Low (Koenen limiting diameter <1.5 mm), and/or Deflagration: Slowly (pressure in Time/Pressure >2070 kPa in >30 ms)

Not Class 1

D No effect of heating under confinement, and does not deflagrate (pressure rise in Time/Pressure <2070 kPa)

No explosive properties with respect to transport classification

As discussed in Section 1, calorimetry or thermal analysis techniques represent

temperature, pressure vs. time behavior of a substance. From the temperature-time data,

the energy released (-∆H) during the process is calculated using the following formula:

)( max op TTmCH −Φ=∆−

where

p

ppss

mC

mCCm +=Φ – Phi factor

ms – Mass of the sample cell

Page 29: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

19

Cps – Heat capacity of the sample cell

m – Mass of the sample

Cp – Heat capacity of the sample

Tmax – Maximum temperature attained by the sample during the reaction

From the temperature-time data, the rate constant for the reaction is obtained using the

following formula13:

( )o

n

o

TTTTTT

dtdT

k

−−

=

maxmax

max

where

n – order of the reaction

Thus, the overall thermodynamics and kinetics of a reaction can be estimated from

temperature-time calorimetric data.

The National Fire Protection Association (NFPA) recommends a classification

for intrinsic thermal instability14 of a substance based on Instantaneous Power Density

(IPD), which is defined as

RateHIPD *∆−=

RTEordero AeACHIPD /*** −∆−=

where

-∆H – enthalpy of reaction (cal/g)

Rate – Rate of reaction = A*exp(-EA/RT)*Coorder (g/ml s)

A – Arrhenius pre-exponential factor

Page 30: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

20

EA – Activation energy

Co – Initial concentration of the material

R – Gas constant

Therefore IPD has units of W/ml, and the rate of reaction can be obtained from

calorimetric data. Based on the IPD, the classification illustrated in Table 2.3 is applied

for rating thermally unstable compounds.14

Table 2.3. Instability Rating Based on IPD

Instability Rating IPD at 250 oC

W/ml

Decomposition

initiation temperature (oC)

4 IPD ≥1000

3 100 ≤ IPD < 1000

2 10 ≤ IPD < 100 < 200

1 0.01 ≤ IPD < 10 200 ≤ IPD < 500

0 < 0.01 < 500

As an example, the IPD is calculated for the following system14:

Enthalpy of decomposition (-∆H) : -80.5 cal/gm

Arrhenius activation energy (EA) : 36.4 kcal/mol

Arrhenius pre-exponential (A) : 1.6 * 1015 /s

Reaction order : 1

Initial concentration or density of pure material: 0.8 g/ml

IPD = 0.8 * 1.6 * 1015*e {-36400/1.987*(250+273)}* 4.184 (W/ml)

= 270 W/ml

Page 31: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

21

Thus this material is given an instability rating of 3.

The calculation of IPD is sensitive to the activation energy (EA) values. For

example, if the EA value were to differ by 5 % and be 34.6 kcal/ml instead of 36.4

kcal/mol, the IPD value would increase to 1813 W/ml and the instability rating of the

material would be 4.

2.3 Proposed Classification15

Although the calculation of IPD appears intuitive, it is difficult to obtain accurate

kinetic parameters based on calorimetric data. Because the calculation of kinetic

parameters requires additional work for the user and can be time consuming; a goal of

this work is to provide an easy method for classification.

2.31 Thermodynamic Parameter

The net energy released during reaction is a measure of stored potential energy of

the system. Therefore heat of reaction (-∆H) is recommended as one of the parameters

for characterizing energetic materials. The lower this energy the less energy that is

available for detonation.

2.32 Kinetic Parameter

The temperature at which a system first exhibits exothermic activity is called the

onset temperature (To) and denotes a rate of a chemical reaction significant enough to be

measured by the calorimeter. The detected onset temperature is thus a measure of the

reaction kinetics. Although there is considerable argument over its interpretation16 for

selecting appropriate process temperatures, the onset temperature is an important

parameter at the screening level of testing. In this regard, Ando et al. have proposed that

Page 32: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

22

onset temperature be used as a parameter to classify reactive chemicals.17 Heat of

reaction is representative of the energy release potential of a substance and the onset

temperature is a measure of the rate of energy release, therefore these two parameters

can be combined to develop a hazard index. The onset temperature (To) and heat of

reaction (-∆H) can be easily determined from the temperature-time data.

Based on the above discussion, the reactive chemicals can be divided into the

following four classes, as shown in Figure 2.10:

1. Class I: These compounds react at low temperatures liberating a large amount of

heat.

2. Class II: Compounds react with significant heat release at higher temperatures.

3. Class III: Compounds react at low temperatures, similar to compounds in Class I, but

are more exothermic than chemicals in Class I.

4. Class IV: Chemicals react at higher temperatures and are mildly exothermic.

Thus, reactive hazards decrease from Class I to IV. The most hazardous chemicals are in

Class I, since they decompose at lower temperatures and release large amounts of heat.

These are also the chemicals that are more likely to decompose violently and should be

carefully handled and thoroughly tested. Chemicals in Class II lie in the high hazard

category since they release large amounts of energy, but chemicals in Classes III and IV

pose medium and low risk, respectively. It should be noted that by neglecting pressure

effects, this classification scheme could miss mildly exothermic but pressure generating

reactions.

Page 33: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

23

Figure 2.10. Proposed classification for reactive chemicals.

2.33 Choice of Critical Values

The classification places rigid boundaries based on values for onset temperature

and heats of reaction. These rigid boundaries can be avoided by the use of fuzzy logic to

define the bounds, however our aim is to demonstrate a basis for reactive chemical

classification. With the static limits choices of threshold values are subject to judgment,

and choices of these values are discussed below.

A value of 200 oC is recommended for the critical onset temperature To,critical.

This value is in agreement with the NFPA intrinsic thermal stability rating, which

Hea

t of r

eact

ion

(-∆

H)

Onset temperature (To)

High -∆H, Low To

Low -∆H, Low To

Low -∆H, High To

High -∆H, High To

To,critical

-∆Hcritical

I

III IV

II

Very High Hazard High Hazard

Medium Hazard Low Hazard

Curve distinguishing detonating or deflagrating compounds (Explosion Potential line)

Page 34: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

24

classifies materials that exhibit adiabatic exothermic initiation temperatures below 200 oC as more hazardous and specifies a hazard rank of 2.

The ASTM CHETAH program18 calculates the maximum heat of decomposition

based on the heat of formation, and if the maximum heat of decomposition is more

exothermic than -2.929 kJ/g it classifies the material as hazardous. Thus – 3 kJ/g (~ 700

cal/g) can be selected as a critical threshold value for the heat of reaction, ?Hcritical.

It is important to realize that DSC data are not explicitly indicative of possible

detonation or deflagration hazards. Previous research has shown that the boundary

between explosion propagation can be represented by the following equation 19:

log(QDSC) = 0.38 log(TDSC – 25) + 1.67

Rearranging the above equation, one can define a variable EP such that,

Explosion Potential (EP) = log(QDSC) - 0.38 log(TDSC – 25) - 1.67

where

QDSC – Heat of reaction, -∆H, (cal/g)

TDSC – Extrapolated onset temperature, (oC)

Substances with EP > 0, are considered to have potential to explode, either by detonation

or deflagration. The above equation for EP further strengthens the argument that onset

temperature (To) and heat of reaction (-∆H) are key parameters for hazard classification.

The EP equation can be used to define a boundary to separate compounds or

compositions that can undergo detonation or deflagration as indicated by the dotted line

in Figure 2.10. This is a more natural classification, than the classification based on rigid

boundaries and one can envision a family of curves, similar to the ones shown in Figure

Page 35: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

25

2.10 for categorizing reactive hazards. An application of this classification is discussed

in the next section.

2.34 Application of Proposed Classification for Screening Reactive Hazards20

It is commonly emphasized to perform detailed analyses on the more reactive

systems. However there is no common consensus among researchers on the attributes of

a system that qualify it for a more sophisticated testing. The classification proposed

earlier can be used as a screen for recognizing more hazardous compositions. For

example, if a compound is categorized as Class I or II, further testing is entailed. This

approach is illustrated based on RSST data obtained for 30 wt % di-tert-butyl peroxide

(DTBP) in the nine organic solvents summarized in Table 2.4.

Table 2.4. Summary of RSST Data for 30 wt % DTBP in Various Solvents

Sr. no. Solvent Tonset Tmax (dT/dt)max (dP/dt)max Heat of reaction BDE

oC oC oC/min psi/min cal/g kcal/mol

1 methanol 117 180 10 7 126 33.7 2 chlorobenzene 120 260 1000 1000 117 33.8 3 n-butylamine 125 230 220 400 167 33.8 4 t-butanol 130 200 15 18 163 33.8 5 n-butanol 130 210 43 33 160 33.8 6 cyclohexane 133 240 300 633 113 37.2 7 toluene 135 250 700 900 153 37.4 8 tetrahydrofuran 135 243 425 1000 170 36.3 9 acetone 138 205 35 65 185 36.1

Page 36: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

26

A plot of heat of reaction vs. Tonset for DTBP in the nine solvents, Figure 2.11,

indicates that all the compositions lie in Class III, according to the proposed

classification. These results suggest that significant decomposition of 30 wt % DTBP in

organic solvent can be initiated at temperatures lower than 200 oC, but the heat released

during the reaction is much less than the heat released during TNT decomposition.

Based on the tests carried out at the screening level, detailed testing of these mixtures

may not be necessary unless the operating temperature is greater than ~100 oC.

0

20

40

60

80

100

120

140

160

180

200

115 120 125 130 135 140

Tonset (oC)

Figure 2.11. Heat of reaction and onset temperature for DTBP in various solvents.

Thus the proposed classification, based on To and -∆H provides a powerful tool to

recognize the hazardous compositions. The next section presents results of our study

aimed at improving prediction of sensitivity tests from calorimetry data and thereby

identify parameters to refine the existing classification.

Page 37: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

27

3. Further Investigation of DSC Parameters to Quantify Reactivity21

Domestic and international requirements mandate classification of hazardous

chemicals as a prerequisite for their transport. The interrelationship of DSC and shock

sensitivity data, represented by the Explosion Potential (EP) equation, suggests that there

may be an intrinsic correspondence of mechanism for detonation of energetic materials

initiated by different stimuli. Because DSC experiments are non-destructive, relatively

inexpensive, and utilize a small amount of sample, it is worthwhile to develop

correlations to predict sensitivity information from DSC data. These correlations can

also be used to predict energy release potential of substances and eventually serve as a

guide for development of a better classification.

Bodman22 recently investigated the interrelationship between To ,-∆H and

sensitivity tests with the aim of improving the EP Equation and proposed the following

correlation:

Explosion Potential (EP’) = log (-∆H) - 0.44 log(DSC To-25) – 2.11.

It is worth noting that the proposed relationships fail to identify all of the

detonating compounds, and further research is needed to develop more robust

correlations.

Onset temperature is a single point on a DSC curve and by including more

information from the thermal analysis curve a better correlation can be obtained. A

schematic of DSC runs on two energetic materials is illustrated in Figure 2.12. The DSC

peak for an energetic material that decomposes violently, in this case compound 2, is

much narrower than for a less energetic material, compound 1. In principle, the two

energetic materials can have comparable heats of reaction, calculated as area under the

Page 38: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

28

curve but different peak shapes depending on the rate of energy release. For example, an

extremely narrow peak indicates a rapid release of energy and consequently possible

explosive behavior. From a safety viewpoint, the rate of release is extremely important

and therefore heat of reaction alone is not a sufficient measure for a realistic reactive

hazard assessment. Other parameters such as peak height, peak width, and aspect ratio

obtained from DSC curve are utilized to investigate relationship between the thermal

data and explosion behavior of a substance.

Figure 2.12. Schematic of DSC curves for two energetic materials.

When other parameters were included from peak shape analysis, our results

indicate that DSC parameters correlate effectively with the UN tests. The methodology

of this research is discussed below.

Temperature

Hea

t sup

plie

d )

Onset temperature

Compound 1

Compound 2

Page 39: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

29

3.1 Experimental Data

DSC data on various compounds listed in Table 2.5 was obtained from Eastman

Kodak Company, Rochester, NY. Standard DSC experiments were conducted using a

TA Instruments Model 2920 calorimeter. A sample of ∼1 mg was sealed under nitrogen

in a glass capillary23, placed in a silver cradle, and heated in scanning mode at 10°C/min

from 25 to 420°C. The instrument was calibrated using a dedicated 1 mg indium

standard in a glass capillary at 10°C/min. The TA2920 DSC computer automatically

updated the cell constant, onset slope, and temperature correction factors. One DSC

scan was conducted for each of the 22 substances that are candidates for Class 1 testing

based on the UN Recommendations.

The test results from the three recommended UN Class 1 methods, listed in

Table 2.5, were acquired for the 22 substances from a number of sources and, therefore,

were generally run on different samples than those used for the DSC experiments, which

were conducted over a substantial period of time. The methods used were the UN Gap

Series 1(a) for detonation, Koenen for heating under confinement, and Time/Pressure for

deflagration. Results of the BAM 50/60 Steel Tube and TNO 50/70 Steel Tube tests,

which are considered to be reliable Class 1 detonation methods24, were used as

surrogates for the UN Gap test when the UN Gap results were not available.

Page 40: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

30

Table 2.5. Summary of DSC and Sensitivity Data

Sr. no. Material

DSC, To

DSC, -∆H

Aspect ratio -∆H/To -∆H /To0.5 UN Gap Koenen

Time/ Pressure

oC J/g W/g oC

1 2,4-Dinitrophenyhydrazine

275 3381 3.38 12.3 203.9 Yes+ Violent NA

2 2,4-Dinitrotoluene 345 3529 75.15 10.2 190.0 Yes+ Medium Slow

3 4-Nitrophenylhydrazine 100%

255 2377 0.34 9.3 149.0 Yes Violent Rapid

4 3,5-Dinitrobenzoic acid 383 2909 4.48 7.6 148.6 Yes Low NA

5 2-Bromo-2-nitropropane-1,3-diol

209 2154 3.25 10.3 149.0 Yes+ Low No

6 2,2’-Dithiobis(4-methyl-5-nitrothiazole)

223 2221 0.89 10.0 148.7 Yes Medium Slow

7 Benzoyl peroxide 100% 107 1211 0.50 11.3 117.1 Yes+ Violent Rapid

8 Benzoyl peroxide 70% with H2O

104 908 0.21 8.7 89.0 Yes (No+)

Violent Rapid

9 2-Chloro-5-nitrobenzoic acid

366 1722 1.16 4.7 90.0 Yes Medium No

10 t-Butyl peroxybenzoate 124 1333 0.15 10.8 119.7 No+ Violent Slow

11 2-Diazo-1-napthol-5-sulphochloride

133 859 0.30 6.5 74.5 No+ Violent NA

12 4-Nitrophenylhydrazine 76% with H2O

244 2224 0.18 9.1 142.5 No Medium No

13 2-Amino-4-chloro-5-nitrophenol

225 1685 10.39 7.5 112.3 No Low Slow

14 Di-t-butyl peroxide 161 1253 0.63 7.8 98.8 No+ No Slow

15 1-Phenyl-5-mercapto tetrazole

159 1235 5.83 7.8 97.9 No++ Low Slow

16 Organic perchlorate # 1 224 1240 0.09 5.5 82.9 No Low No

17

Benzenediazonium, 2-methoxy -4-(phenylamino)-, sulfate (1:1)

178 832 2.43 4.7 62.4 No Low No

18 3-Nitrobenzenesulfonic acid sodium salt

368 1099 2.99 3.0 57.3 No No Slow

19 Malononitrile 276 1848 4.43 6.7 111.2 No No No 20 Organic perchlorate # 2 229 1465 0.07 6.4 96.8 No No No 21 Dilauroyl peroxide 88 721 0.13 8.2 76.9 No+ No No 22 3-Thiosemicarbazide 176 908 0.24 5.2 68.4 No No No

+BAM 50/60, ++TNO 50/70 instead of UN Gap

Page 41: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

31

3.2 Results and Discussion

Earlier correlations failed to recognize a few of the Class A compounds. As

suggested above, additional information from calorimetric data, besides the onset

temperature and decomposition energy, should correlate more effectively with the UN-

recommended tests. The following parameters were obtained for each compound from

DSC experiments:

1. Decomposition energy (∆H or E, J/g)

2. Extrapolated onset temperature, (To, oC)

3. Peak height (W/g)

4. Peak width, at half the peak height (oC)

5. Aspect ratio (peak height/peak width at half peak height)

6. Initial slope (W/g/oC)

7. Maximum slope (W/g/oC)

Various combinations of the above set of variables were used to investigate the

correlation among DSC data and UN tests (Gap, Koenen, and Time/Pressure)

individually. A possible relationship was investigated by employing one of the above

parameters in combination with decomposition energy, which was employed for all the

trials. No more than two parameters were used during a trial, except for the proposed

correlation for UN Gap test. A visually satisfactory separation of data points on the

graphs was considered an acceptable solution. An energy threshold of 500 J/g is

employed for all the correlations, and any composition with energy less than 500 J/g is

not considered to be in Class 1.

Correlations among UN tests and DSC data, based on this analysis, are discussed

below:

Page 42: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

32

3.21 Gap Test

For the Gap test, it was found that ∆H/To0.5 and the aspect ratio, obtained from

the DSC data, separated the detonating compounds (“Yes” on the Gap test), as shown in

Figure 2.13. The decomposition energy is a measure of the heat content of a system and,

therefore, the energy available for detonation. A lesser onset temperature is indicative of

faster kinetics and, consequently, a more rapid energy release for high energetic

materials.

The aspect ratio (peak height/peak width) represents the decomposition behavior

of a substance. For an ideal explosive, the peak height would be infinite and the peak

width would be zero, and therefore the higher the aspect ratio, the higher the tendency

for a violent decomposition. Compounds having ∆H/To0.5 > 88 J/g oC0.5 and aspect ratio

> 0.2 W/g oC are expected to exhibit a positive Gap test.

10

100

1000

0 0 1 10 100

Aspect ratio (W/g oC)

Yes

No

y = 88

x = 0.2

Figure 2.13. Correlation between UN Gap test and DSC data

Page 43: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

33

3.22 Koenen test

The onset temperature and heat of reaction, as shown in Figure 2.14, yielded

optimum separation for compounds exhibiting “Violent” behavior on the Koenen test.

Based on limited data set, it is proposed that compounds above the line, ∆H = 12.4 To –

796, are expected to exhibit a violent behavior on the Koenen test. Thus a Koenen

potential (KP) can be defined,

KP = ∆H – (12.4 To – 796)

such that if KP > 0, then the compound will display a violent behavior on the Koenen

test.

0

500

1000

1500

2000

2500

3000

3500

4000

0 50 100 150 200 250 300 350 400 450

To(oC)

Violent

Medium

Low

No

y = 12.4 x - 796

Figure 2.14. Relationship between the Koenen test and DSC data.

Page 44: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

34

3.23 Time-pressure

As seen from Figure 2.15, a ∆H/To ratio of 8 (J/g oC) separates the compounds

exhibiting rapid time-pressure behavior.

0.0

2.0

4.0

6.0

8.0

10.0

12.0

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Serial no. of compounds listed in Table 2.5

Rapid

Slow

No

y = 8

Figure 2.15. Relationship between the Time/Pressure test and DSC data.

Based on the above recommendations, the predicted class are listed in Table 2.6. It is

apparent that by applying the proposed correlations, unlike previous correlations, all of

the Class A and B compounds can be identified. The correlations proposed in this work

screen out 5 of the 11 Class C and D compounds and all of the Class A and B

Page 45: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

35

compounds. A summary of the predictions using earlier correlations and this work is

summarized in Table 2.7.

UN-recommended tests for transportation are time consuming and expensive.

Earlier researchers have proposed correlations to predict UN test outcomes using

parameters from DSC data, such as onset temperature and heat of reaction. Based on the

above correlations, DSC data can be used effectively to recognize Class 1 compounds.

Therefore, these correlations can be employed as a screening tool to reduce testing and

focus resources on more hazardous chemicals.

4. Conclusions

Various experimental techniques are available for characterizing reactive

hazards, such as sensitivity tests and thermal analysis. A classification, based on To and

-∆H obtained from calorimetric data is proposed to facilitate ranking of reactive

chemicals. Although it is true that criteria utilizing multiple parameters, such as

pressure change and rate of heat generation, obtained from calorimetric data would lead

to a better classification scheme, but the main aim was to establish a basis for

classification. By neglecting pressure effects, the classification scheme could miss

mildly exothermic but pressure generating reactions. The pressure criterion was not

included because DSC is often employed to gather data and it does not gather pressure

vs. time data. The correlations discussed in Section 3 of this Chapter guide the selection

of parameters for classification. For example, the aspect ratio correlates well with the

Gap test, a measure of explosive tendency of a material and is therefore a potential

candidate for refining the proposed classification. Since the sensitivity tests are

typically more resource consuming, correlations among DSC and sensitivity tests, can

serve as an effective tool for screening compounds for sensitivity testing.

Page 46: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

36

Table 2.6. Explosion Propagation vs. Actual Rank

Sr. no.

Material Predicted Class

Explosion Propagation

Actual Class

UN Gap

Koenen Time/ Pressure

EP19 EP’22

1 2,4-Dinitrophenyhydrazine A 0.33 0.36 A Yes+ Violent NA 2 2,4-Dinitrotoluene A 0.30 0.33 A Yes+ Medium Slow 3 4-Nitrophenylhydrazine 100% A 0.20 0.24 A Yes Violent Rapid 4 3,5-Dinitrobenzoic acid A 0.20 0.23 A Yes Low NA

5 2-Bromo-2-nitropropane-1,3-diol A

0.18 0.22 A Yes+ Low No

6 2,2’-Dithiobis(4-methyl-5-nitrothiazole) A

0.18 0.22 A Yes Medium Slow

7 Benzoyl peroxide 100% A 0.06 0.13 A Yes+ Violent Rapid

8 Benzoyl peroxide 70% with H2O A

-0.05 0.01 A Yes (No+) Violent Rapid

9 2-Chloro-5-nitrobenzoic acid A -0.02 0.01 A Yes Medium No 10 t-Butyl peroxybenzoate B 0.07 0.13 B No+ Violent Slow

11 2-Diazo-1-napthol-5-sulphochloride B

-0.13 -0.07B No+ Violent NA

12 4-Nitrophenylhydrazine 76% with H2O B 0.18 0.23 C No Medium No

13 2-Amino-4-chloro-5-nitrophenol A 0.06

0.10 C No Low Slow

14 Di-t-butyl peroxide A -0.01 0.05 C No+ No Slow 15 1-Phenyl-5-mercapto tetrazole A -0.01 0.04 C No++ Low Slow 16 Organic perchlorate # 1 C/D -0.07 -0.03 C No Low No

17 Benzenediazonium, 2-methoxy-4-(phenylamino)-, sulfate (1:1)

C/D -0.20 -0.15 C No Low No

18 3-Nitrobenzenesulfonic acid sodium salt C/D -0.21

-0.19 D No No Slow

19 Malononitrile A 0.06 0.10 D No No No 20 Organic perchlorate # 2 C/D 0.00 0.04 D No No No

21 Dilauroyl peroxide B -0.12 -0.05 D No+ No No

22 3-Thiosemicarbazide C/D -0.16 -0.11 D No No No

+BAM 50/60, ++TNO 50/70 instead of UN Gap, Bold EP and EP’ values indicate incorrectly categorized Class A and B compounds

Page 47: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

37

Table 2.7. Summary of Sensitivity Tests Predictions

A further advancement to reduce experimentation and expedite hazard evaluation

is prediction of experimental data using computational techniques, which is discussed

in the next chapter.

Yoshida19 Bodman22 This work

Number of Class A and B compounds successfully categorized

8/11 10/11 11/11

Number of Class C and D compounds successfully screened 7/11 5/11 5/11

Page 48: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

38

CHAPTER III

STRCUTURE BASED PREDICITION OF REACTIVITY HAZARDS

As discussed earlier, a reliable experimental technique for assessing reactivity is

calorimetric analysis, which can be resource consuming and thus possible only for a

limited set of compounds. A computational screening tool can reduce experimentation

and screen out benign compounds. This chapter provides a brief review of computational

methods that can be employed quickly to estimate reactive hazards and a description of

efforts to develop methods for predicting calorimetric data. A goal of this research is to

develop a computerized program for screening reactive hazards.

Generally, rules of thumb based on prior experience and chemical knowledge are

used for screening and estimating reactive hazards of compounds. For example, the

presence of a ‘nitro’ group is regarded as an indicator of potential energy, as in

trinitrotoluene (TNT). Attempts have been made to develop a generalized framework for

estimating reactive hazards based on molecular structure, such as the oxygen balance

method25, Chemical Thermodynamic and Energy Release Evaluation (CHETAH)18, and

Calculated Adiabatic Reaction Temperature (CART).26 A brief description of these

methods is provided in Section 1 of this chapter. The above methodologies have

limitations, and considerable chemical intuition and experience are required for their

effective use. Also, the reliability of estimations for a range of compounds and process

conditions varies significantly.27 Sections 2 through 5 of this chapter discuss

Part of this chapter is reprinted with permission from, “Prediction of reactive hazards based on molecular structure”, by S.R. Saraf, W.J. Rogers, and M. Sam Mannan, 2003, J. Hazardous Materials, A98, 15-29. Copyright 2003 by the name of Elsevier. Part of this chapter is reprinted with permission from “Application of transition state theory for thermal stability prediction”, by S.R. Saraf, W.J. Rogers, and M. Sam Mannan, 2003, I & EC Res., 42, 1341. Copyright 2003 by the name of American Chemical Society (ACS).

Page 49: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

39

improvements and advancements in predictive techniques for reactive hazard

assessment.

1. Review of Available Methods

This section reviews some popular methods for reactivity hazard evaluation,

including their strengths and limitations, and it attempts to provide the reader with

enough information to choose a method for a particular application. Some of these

methods have been reviewed previously.26

1.1 Rules of Thumb

The presence of certain functional groups is considered an indicator of reactivity.

This is the simplest possible reactivity screening method and serves as a guideline for

further analysis. For example, chemicals containing the following functional groups can

be considered potentially reactive:

-NO2 : organic nitro compounds

-O-O-, -O-OH : organic/inorganic peroxide and hydroperoxide compounds

-C C- : triple bonded carbon atoms as in acetylene and acetylenic compounds

A comprehensive summary of reactive groups can be found in Bretherick’s handbook28,

and a few of the reactive groups are summarized in Table 3.1.

Page 50: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

40

Table 3.1. Functional Groups Indicative of Reactive Hazards 4

Groups containing Carbon -C≡C- Acetylenic compounds -C≡C-M Metal acetylides -C≡C-X N=N C

Haloacetylene derivatives Diazirines

-CN2 Diazo compounds -C-N=O, -N-N=O Nitroso compounds -C-NO2 Ar-NO2, Ar(NO2)n C(NO2)n O2NC-CNO2 HC[OCH2C(NO2)3]3, C[OCH2(NO2)3]4

Nitroalkanes, C-nitro and Nitroaryl and Polynitroaryl compounds Polynitroalkyl compounds Trinitroethyl orthoesters

-C-O-N=O Acyl or alkyl nitrites -C-O-NO2 Acyl or alkyl nitrates >CC< O

1,2-Epoxides

MC=N→O C=N-O-M

Metal fulminates or aci-nitro salts, oximates

-(CH-CH-)n- olymerization alkene monomers

Groups containing Oxygen -C-O-O-H, R-CO-OOH Alkylhydroperoxides,

Peroxyacids -C-O-O-C-, -CO-OOR Peroxides (cyclic, diacyl,

dialkyl,), peroxyesters -O-O-M, EOO-, MOO- Metal peroxides, peroxoacid

salts -O-O-E Peroxoacids, peroxyesters H3N.Cr-OO- Amminechromium

peroxocomplexes -O-X XOn -Cl-O3 ClO2

- R-O-Cl-O3 RN+H3ClO4

Hypohalites Halogen oxides Perchloryl compounds Chlorite salts Alkyl perchlorates Aminium perchlorates

S2O4 Dithionites -(C-C-O-)n O

Polymerization ester monomers

-(C-C-N-)n O

Polymerization amide monomers

Page 51: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

41

Groups Containing Nitrogen F-C- (NO2)2 Fluorodinitromethyl

compounds -N-M N-metal derivatives -N=Hg+=N- Poly(dimercuryimmonium

salts) -N-NO2 N-nitro compounds =N+-N-NO2 N-Azolium nitroimidates -C-N=N-C- Azo compounds Ar-N=N-O-R Arenediazoates ArN=N-S-Ar Arenediazo aryl sulfides Ar-N=N-O-N=N-Ar Bis(arenediazo) oxides Ar-N=N-S-N=N-Ar Bis(arenediazo) sulfides -C-N=N-N-C- R (R=H, CN, OH, NO)

Trizenes

-N=N-N=N- -N=N-N=C-

High-nitrogen compounds Tetrazoles

-N3 Azides (acyl, halogen, nonmetal, organic)

C-N2±O- Arenediazonium oxides -C-N2 +S- Diazonium sulfides and

derivatives, .Xanthates N+-HZ-, N+EOn

- Hydrazinium salts, Oxosalts of nitrogenous bases

-N+-OH Z- Hydroxylaminium salts -C-N2+Z- Diazonium carboxylates or

salts [N→Metal]+ Z- Amminemetal oxosalts Ar-Metal-X X-Ar-Metal

Halo-arylmetals, Haloarenemetal p-complexes

-N-X XN3 -C-N-C- O X O

Halogen azides N-halogen compounds N-haloamides

-N-F2 -C(NF)NF2

Difluoroamino compounds N,N,N-trifluoroalkylamidines

N-O- N-O compounds

Abbreviations: Ar = aromatic (benzene); M = metal; R = organic chain; X = halogen; E = nonmetal; Z = anion; n = integer variable; all other abbreviations are for the element symbols from the periodic table of elements Note: Not all chemical bond symbols are shown .

Table 3.1. (Contd.)

Page 52: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

42

1.2 Oxygen Balance Method

A quantitative correlation has been demonstrated between oxygen balance and

various measures of explosive effectiveness for several classes of more than 300

compounds organic explosives25, and the following formula was recommended for

calculating oxygen balance for a compound:

where

X – Number of atoms of carbon

Y – Number of atoms of hydrogen

Z – Number of atoms of oxygen

MW – Molecular weight

The above formula yields a value of zero for oxygen-balanced compounds, negative for

oxygen-poor, and positive for oxygen-rich compositions. This method is a criterion for

evaluating self-reactivity in the CHETAH program, and the classification indicated in

Table 3.2 is recommended for estimating hazard potential based on oxygen balance.29

Table 3.2. Oxygen Balance and Hazard Rank

Oxygen Balance Hazard Rank

More positive than +160 Low

+160 to +80 Medium

+80 to –120 High

–120 to –240 Medium

More negative than –240 Low

MWZYX

BalanceOxygen)2/2(1600 −+−

=

Page 53: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

43

1.21 Strengths

The oxygen balance method is useful for estimating hazards of organic nitro

compounds and is universally employed in the explosive industry. In general, this

method is applicable to compounds containing C, H, N, and O.29

1.22 Weaknesses

However, it has been shown that there is no necessary connection between

oxygen balance and self-reactivity29. For example, water (H2O) has an oxygen-balance

value of 0 and is given a ‘high’ hazard ranking by this criterion. Also, the method cannot

be applied to oxygen free but hazardous compounds such as acetylene. Application of

the above oxygen balance equation to low-oxygen content or oxygen-free compounds

produces a highly negative, non-hazardous ranking regardless of the actual hazard

potential.

1.3 CHETAH18

CHETAH is a popular program available from the American Society for Testing

and Materials (ASTM) for prediction of reactivity hazards. The software uses the

‘Benson group contribution method’ 30 to estimate heat capacity, heat of formation, and

heat of combustion for a multitude of compounds. Also, the program includes a database

of thermo-chemical properties for selected organic and inorganic compounds. CHETAH

classifies chemicals based on their potential for violent explosion and includes the

following hazard evaluation criteria:

§ Maximum heat of decomposition

§ Oxygen balance

Page 54: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

44

According to the first criterion, compositions with heats of reaction more

negative and therefore more exothermic than – 2.929 kJ/g are placed in a ‘high’ hazard

category. A detailed explanation of the above evaluation criteria and three additional

ones can be found in the CHETAH reference manual, and a critical review of CHETAH

for predicting reactivity hazards is available.31 The first criterion has proved to be a

reliable indicator of potential reactive hazards. The other criteria, however, are not

effective for all chemicals and compositions.31

1.31 Input to the Program

The molecular formula of a compound is the only input to the program. From the

included database, the thermodynamic properties are estimated and the hazard criteria

are determined from these values. An energy release evaluation sheet from

CHETAH 7.2 is shown in Figure 3.1. Based on its maximum heat of decomposition,

H2O2 is given a “high” hazard classification. However, it should be noted that the

estimated product spectrum might be incorrect. Thus, one problem is the thermodynamic

feasibility of a proposed stoichiometry under process conditions. Further, the program

gives no indication about the sensitivity to reaction initiation or process conditions to be

avoided. It is difficult to determine conditions under which H2O2 may pose reactive

hazards based only on such an analysis. Thus, the problem of reactivity is not just a

combinatorial problem (stoichiometric analysis) as implicitly suggested by this method.

It is important to realize that CHETAH provides an estimate of a material hazard but not

an estimate of the process risk in using the material.

1.32 Strengths

The software is user friendly and offers the flexibility to include user-defined

group values. It is computationally inexpensive and can be installed on a standard PC.

Page 55: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

45

ENERGY RELEASE EVALUATION

Compound Name: Hydrogen Peroxide Formula: H2O2 Molecular weight: 34.015 Amount: 1 Mole(s) Heat of formation at 25 oC: -32.530 kcal/mol

PLOSIVE HAZARD CLASSIFICATION: Over-all Energy Release Potential is HIGH Value: -1.375

Contributing Details: Criterion Value Units Hazard Classification Maximum Heat of Decomposition (#2) -0.743 kcal/g HIGH Fuel Value – Heat of Decomposition 0.000 kcal/g HIGH Oxygen Balance (#3) 47.037 percent HIGH CHETAH ERE Criterion 4 46.934 kcal2/mol MEDIUM Total Number of Peroxide Bonds 1.000 Net Plosive Density (#4) 0.435 PLOSIVE Warning: These ratings only apply to hazards associated with strong mechanical shock. This does not

imply the absence of other hazards. Notes #1 This evaluation was developed to classify a composition as able or not able to decompose with

violence, if subjected to the proper conditions. Information on the interpretation of hazard classification criteria used by CHETAH may be found in the CHETAH documentation. (ASTM publication DS-51A) and in J. Chem. Ed., v66, A137 (1989).

#2 For decomposition products shown. #3 Experience has shown that the oxygen balance criterion is useful only for compounds composed

of the elements C, H, N, and O. #4 Sum of auxoplosive and plosphoric weights per gram of mixture.

Decomposition Products (chosen to maximize heat of decomposition) Moles State Species 0.500 ref-gas O2 Oxygen 1.000 gas H2O Water

HEAT OF COMBUSTION SECTION Fuel Value (Net Heat of Combustion)

MASS BASIS MOLE BASIS -0.743 kcal/g - 25.270 kcal/mol -3.108 kJ/g -105.730 kJ/mol -1337.246 Btu/lb

Combustion Products (Chosen for Fuel Value and Net Heat of Combustion)

MOLES STATE SPECIES 1.000 gas H2O Water

Figure 3.1. Hazard evaluation output from CHETAH 7.2.

Page 56: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

46

1.33 Weaknesses

The Benson method can fail for group values that are not available in the

database or are incorrect. In the evaluation criteria, the program classifies compounds or

compositions into specific hazard categories. Thus the program places a boundary based

on a threshold value and thereby blurs the distinction between hazardous and non-

hazardous chemicals. Also, the program provides no insight into process conditions to be

avoided or information about the sensitivity of compounds to initiation of a reaction.

1.4 CART26

Adiabatic temperature rise due to a reaction is defined as:

CpH

Tadiabatic∆−

=∆

where

∆H – heat of reaction

Cp – average heat capacity of the reacting mixture

∆Tadiabatic – adiabatic temperature rise

The code developed for CART performs multiphase Gibbs free energy minimization and

adiabatic reaction temperature calculations. Based on the calculated ∆Tadiabatic, the

substance is classified as26:

E - Can explode when unconfined

N- No known explosion hazard when unconfined

Page 57: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

47

An adiabatic temperature rise of 1400 K is considered a cut-off value for the above

classification. Thus, substances with an adiabatic temperature rise of more than 1400 K

are classified as E and lower than 1400 K as N. This value is based on the fact that most

combustion reactions leading to formation of CO2 and H2O have a threshold temperature

value near 1400 K, which is the minimum temperature required for carbon monoxide to

propagate a self-sustaining flame. A cut-off value of 1200 K for conservative estimates

is recommended.26 The heat of reaction can be approximated by the maximum heat of

decomposition that is calculated by CHETAH. For H2O2, CHETAH estimates a

maximum heat of decomposition of -25.27 kcal/mol and Cp as 10.31 cal/mol K, yielding

an adiabatic temperature rise of 2450 K. Based on the above discussion, H2O2 would be

classified as E. Again, this value is useful but it does not provide information about

process conditions to be avoided. It is worth noting that for H2O2, the CHETAH criteria

and the CART adiabatic temperature value indicate a potential reactive hazard.

1.41 Input to the CART Method

The heat of reaction, heat capacity of the reaction mixture, and other system

thermo-chemical values are required program inputs. All of these values can be obtained

from the literature or estimated using the CHETAH program.

1.42 Strengths

The CART criterion takes into account the heat capacity of the reaction mixture

and is therefore more effective than employing only the reaction energy of the first

CHETAH criterion. Higher CART values are associated with greater sensitivities to

initiation and higher propagation rates. CHETAH programmers should include

‘adiabatic temperature rise’ as one of the hazard evaluation criteria.

Page 58: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

48

1.43 Weaknesses

Thermo-physical values, such as ∆H and Cp, must be used to estimate the

adiabatic rise in temperature. The heat of reaction can be approximated by the maximum

heat of decomposition calculated by CHETAH, and CHETAH can also be used to

estimate the average Cp value. If the CHETAH program is employed to calculate

thermo-physical values, limitations similar to the group contribution method are

encountered, as discussed above. The classification based on adiabatic temperature rise

works well for hazard estimation of compounds undergoing combustion reactions. For

compounds that do not undergo combustion type reactions, determining a threshold

value of ∆Tadiabatic is difficult. As pointed out in the same reference 26, the CART

classification and heats of reaction values fail for hazard ranking of organic peroxides.

Also, like CHETAH, this classification places a distinct boundary based on a threshold

value. It is also worth noting that for a realistic calculation of ∆Tadiabatic , an average value

of heat capacity for the reaction mixture is required, and accurate heat capacity values

are difficult to estimate.

2. Advanced Prediction Techniques

None of the available theoretical methods address the issue of chemical kinetics,

which can significantly affect the rate of energy release and consequently the hazards

posed by the substance. Therefore, future work is focused on developing theoretical

methods to quantify both kinetics and thermodynamics based on molecular structure

alone.

The calorimetric experiments are designed to gather temperature, pressure, and

time data and typically very little or nothing is known about the reaction mechanism and

species involved. It is difficult to extract exact kinetics from calorimetric data, since the

calorimetric data on a substance reflects temperature vs. time behavior. This limitation

Page 59: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

49

of calorimetric information demonstrates the need for property predictions that can be

compared to experimental temperature-time data.

It is well known that the structure of a substance affects its reactivity6 and this

reactive nature of a compound is reflected in the calorimetric data. With advances in

computation speed and associated growth in computational chemistry algorithms,

molecular modeling has evolved into a popular resource for predicting material

properties. A brief overview of molecular modeling techniques, computational methods,

and available softwares is provided in Appendix A. Molecular modeling techniques,

principally quantum chemical calculations, were employed to gain information at

molecular (microscopic) level and develop predict observed data (macroscopic level).

Two different approaches, employing computational chemistry calculations were probed

to predict calorimetric data:

§ Applying Transition State Theory (TST) to the rate-limiting step.

§ Developing correlations using the Quantitative Structure Property Relationship

(QSPR) technique.

These two approaches are discussed in the following sections.

3. Application of Transition State Theory for Thermal Stability Prediction

Prediction of potential hazards requires the knowledge of both the kinetics and

thermodynamics. In this work, computational chemistry techniques are combined with

Transition State Theory32 (TST) for predicting calorimetric data for aromatic nitro

compounds. The heat of reaction (- ∆Hrxn) is based on the heat of formation data and is

approximated by the enthalpy of maximum decomposition as calculated by the

CHETAH18 program. The C−NO2 bond fission is proposed to be the rate-limiting step

for decomposition of aromatic nitro compounds (R−NO2), and TST is applied to

approximate the kinetics of this single step. The activation energy is estimated from the

bond strength values obtained at the density functional level of theory, B3P8633, and

Page 60: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

50

cc-pVDZ34 basis set. The resulting predictions are compared with DSC data.

3.1 Model Development

To predict calorimetric data, the thermodynamic and kinetic parameters for a

reacting system must be combined with an unsteady state model for an adiabatic batch

reactor35. It has been shown that such a model for a batch reactor can fairly well

reproduce calorimetric data36 and is based on the following assumptions:

a. The reacting environment is adiabatic and therefore heat losses are negligible.

b. The mass and the volume of the liquid or solid phase remain constant (i.e.,

evaporation losses can be neglected).

c. The specific heat of the material is assumed constant during the reaction.

d. The reacting phase is assumed to be at uniform temperature.

With the above approximations, the appropriate mass and energy balance equations for a

first-order reaction are:

AA kC

dtdC

=−

(3.1)

V

Arxn

CkCH

dtdT

φρ∆−

=− (3.2)

where

CA - Concentration of the reactant (gmol/m3)

-∆Hrxn - Heat of reaction (cal/mol)

k - Rate constant for a first-order reaction (/s)

V

VVss

mCmCCm +

=φ - Phi factor ( ≥ 1 for a typical industrial vessel)

CV - Heat capacity of the reacting mixture (cal/mol K)

CVs - Heat capacity of the sample cell (cal/mol K)

Page 61: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

51

ρ - Density of the reacting mixture (kg/m3)

ms - Mass of the sample cell

m - Mass of the test sample

Simulation of the temperature-time curve requires the prediction of kinetic (k) and

thermodynamic (-∆Hrxn) parameters, and the ρ, Cv data for the system. The choice of the

values for these parameters is discussed in the next section.

3.11 Recommended Parameters

Thermodynamic parameter estimation

The heat of reaction or the energy of reaction can be theoretically estimated using

the thermodynamic identity:

∆Hrxn = ∆Hf,Products - ∆Hf,Reactants

where

∆Hf,Products - Heat of formation of products

∆Hf,Reactants - Heat of formation of reactants

However, during the calorimetric analysis the end products are not determined, and the

multitudes of reactions and products within the reacting system are difficult to predict.

For a given chemical formula, the CHETAH18 program calculates the enthalpy of

reaction or enthalpy of maximum decomposition from the heat of formation (∆Hf). This

value is thus an upper bound on the energy of reaction and will be used in the

simulations discussed in this paper to approximate -∆Hrxn. Heat of formation values can

be easily obtained from the literature37,38 or can be calculated by employing

Page 62: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

52

computational techniques39. The heat of formation and heat of maximum reaction values

used in these simulations are tabulated in Table 3.3. The variation of heat of reaction

with temperature is neglected. For compounds with available heat of formation values

for the gas, liquid, and solid phases, the numerically largest heat of reaction value was

used in each case.

Kinetic parameter estimation

A possible set of elementary steps is required to estimate the kinetics for a

reaction pathway. In this case, we assume that the nitro compounds undergo

unimolecular decomposition as shown below and potential energy for the reaction

pathway is depicted qualitatively in Figure 3.2.

R – NO2 R• + NO2• ………………. Product

This reaction is assumed to follow a radical mechanism. The first step is the rupture of

the weakest bond. We assume that the remaining steps are relatively fast, therefore the

bond scission is the rate-limiting step.

Page 63: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

53

Table 3.3. Summary of Heat of Formation and Maximum Heat of Decomposition

Sr. no. Compound Standard heat of

formation* kcal/mol

Maximum heat of decomposition†

kcal/mol

gas liq. solid gas liq. solid

1 nitrobenzene 16.38 2.98 - -136.4 -123.0 -

2 2 nitrotoluene - - - - -127.1‡ -

3 3 nitrotoluene - -11.0 - - -117.9 -

4 4 nitrotoluene 7.38 - - -136.3 - -

5 1,2 dinitrobenzene - - -0.4 - - -209.3

6 1,3 dinitrobenzene - - -6.5 - - -203.1

7 1,4 dinitrobenzene - - -9.2 - - -200.4

8 2,6 dinitrotolueme - - -13.2 - - -207.3

9 3,4 dinitrotoluene - - - - -214.4§ -

10 2,4 dinitrotoluene 7.93 - -15.87 -228.4 - -204.5

11 2 nitroaniline - - -6.3 - - -118.2

12 3 nitroaniline 14.9 - - -139.4 - -

13 4 nitroaniline 13.2 - - -137.7 - -

14 2 nitrobenzoic acid - - -95.3 - - -119.6

15 3 nitrobenzoic acid - - -98.9 - - -116.1

16 4 nitrobenzoic acid - - -102.1 - - -112.9

17 2 nitrophenol -31.62 - - -136.3 - -

18 3 nitrophenol -26.12 - - -141.9 - -

19 4 nitrophenol -27.41 - - -140.6 - -

* Standard heat of formation values at 1 atm and 298.15 K are from the NIST Chemistry WebBook. † Calculated by CHETAH. ‡ Because the heat of formation value for 2 nitrotoluene was not available, an average of decomposition

values for 3 nitrotoluene and 4 nitrotoluene was used. § Because the heat of formation value for 3,4 dinitrotoluene was not available, an average of

decomposition values for 2,4 and 2,6 dinitrotoluene was used.

Page 64: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

54

Figure 3.2. Hypothesized potential energy surface (PES) for a runaway reaction.

Reaction co-ordinate

Ene

rgy

EA

1st step

Page 65: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

55

The rate constant for an unimolecular reactions under high pressure conditions is given

by the equation40:

RTEB A

A

TS eqq

hTk

k /*

−= (3.3)

where

kB - Boltzmann constant (J/K) = 1.38 x 10-23

h - Planck constant (Js) = 6.63 x 10-34

R - Gas constant (cal/mol K) = 1.987

T - Temperature (K)

qTS* - Partition function for the transition state (with 1 degree of freedom, along the

reaction coordinate, removed)

qA - Partition function for the reactant

EA - Activation energy (cal/mol)

In all further equations, the activation enthalpy is represented by EA and its

approximated value is based on bond dissociation energy. Each partition function is a

product of translational, vibrational, rotational, and electronic partition functions. For

unimolecular decomposition, the ratio of qTS and qA differs by one degree of vibrational

freedom - the one along which reaction occurs, and this ratio can vary between ~ 0.1 and

~1. For the simulations discussed here, this ratio is approximated by 1 to obtain a

conservative estimate for k. Substituting the values for kB and h, the rate constant can be

written as:

min)(/10*25.1)(/10*08.2 1210 RTE

RTE

RTE

BAAA

eTseTehTk

k−−−

=== (3.4)

Thus, EA is the only missing parameter and is estimated as discussed below.

Following the Polanyi type equation41,

Page 66: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

56

EA = EA0 + γP?Hrxn

where

EA - Activation enthalpy for an elementary step

EA0 - Intrinsic activation enthalpy for a reaction class

γP - Proportionality constant, called the transfer coefficient, for a reaction class

?Hrxn - Heat of reaction for the elementary step

For a bond scission reaction the above equation reduces to42:

EA = γP * BDE (3.5)

where EA0 ~ 0

We further assume that the overall kinetics can be approximated by applying the

TST to evaluate the kinetic parameters for the rate-limiting first step. The activation

energy is therefore a fraction of the bond dissociation energy (BDE) of the weakest

bond, C−NO243. Computational chemistry calculations were performed to calculate the

BDE at the B3P8633 level of theory with the cc-pVDZ34 basis set, and the quantum

mechanical calculations were performed using the Gaussian 9844 suite of programs on

the Texas A&M supercomputer. The Gaussian software calculates energies, optimized

molecular structures, and vibrational frequencies, together with molecular properties that

are derived from these three basic computation types for a chemical formula. Optimized

geometries were obtained for the reactant (R−NO2) and the two fragments

(R• and NO2• ). The BDE is then calculated as

BDE = ENO2• + ER• - ER-NO2 (3.6)

BDE values can be calculated if the experimental heats of reaction are available.

Equation (3.4) reduces to

Page 67: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

57

min)(/10*25.1 12 RTBDEP

eTkγ

−= (3.7)

Although the above TST equations apply for reactions in the gas phase, it is assumed

that the rate in the gas phase approximates the rate in a condensed phase. With the

current reaction-solvation theories, this is the best working assumption at this level of

analysis.

3.12 Physico-chemical Properties

The concerned physical properties, density (ρ) and heat capacity (Cv) are

available in the open literature37,38 or can be estimated with reasonable accuracy. For the

compounds considered in this research the Cv varied between 0.25 – 0.35 cal/gm K, and

the density varied between 800 – 1200 kg/m3.

Equations (3.1) and (3.2) can be further simplified to calculate onset temperature.

We assume that the concentration of the reactant (CA) is equal to the initial concentration

(CA0) until the temperature equals the onset temperature. This assumption decouples

Equation (3.1) and (3.2), and Equation (3.2) reduces to

MWCkH

CkCH

CkCH

dtdT

V

rxn

V

Arxn

V

Arxn3

0 10∆−=

∆−≈

∆−=

ρφρ (3.8)

MWC A

3

010*ρ

=Q

Substituting the expression for k from Equation (3.7) in Equation (3.8), we obtain

MWCeTH

dtdT

V

RTBDErxn

3/12 10***10*25.1* γ−∆−= (3.9)

Page 68: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

58

Thus, γP is the only undetermined variable, and it serves as an adjustable parameter for

the simulations. When the rate of temperature increase exceeds a particular amount (ε),

the calorimeter detects the exothermic reaction. Thus, when dT/dt ≥ ε, T = To, and the

value of ε depends on the sensitivity of the calorimeter. For a given compound the above

equation has the form

TBeTAdt

dT /** −= (3.10)

where A and B are constants. Therefore at higher temperatures the exponential term

dominates, and values of dT / dt from Equation (3.9) are sensitive to the BDE values.

3.2 Experimental Details

For this work, we used available DSC data on aromatic nitro compounds. Pure

organic nitro compounds, aliphatic and aromatic, decompose at high temperatures and

exhibit large exotherms45. These compounds are identified as energetic materials, and

trinitrotoluene (TNT) is used as an explosive. Since a graphical detection procedure is

employed to obtain To, a variation of 5-30 oC is possible in the reported To values for the

same compound. The energy of reaction (-∆Hrxn) is the net heat released during the

reaction and is not the thermodynamic heat of reaction but includes other effects such as

sublimation, evaporation, adsorption, and enthalpy of mixing. Therefore, the

experimentally determined parameters, To and -∆Hrxn, depend on the type of calorimeter,

sample size, sample phase, and scanning rate. To compare the theoretically predicted

values, we chose experimental data from a single reference46 to maintain consistency the

experimental procedure. In this reference46, authors employed a Mettler TA4000 DSC

(0.2 W/gm sensitivity) with DSC25 measuring cell, scanning rate of 4 oC/min to

determine the To and -∆Hrxn values for 19 nitro compounds.

Page 69: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

59

3.3 Results and Discussion

3.31 Choice of γp

Assuming a sensitivity of 0.1 oC/min and substituting the experimental onset

temperature for T and BDE for the 19 compounds listed in Table 3.4, we can determine

γp for each compound from Equation (3.9). The average of the γp values calculated for

mono-nitro compounds is 0.67 ± 0.06 and for di-nitro compounds is 0.71 ± 0.06. For a

sensitivity of 0.01 oC/min at To, the average γp value is 0.70 ± 0.06 for mono-nitro and

0.76 ± 0.06 for di-nitro compounds, respectively. Note that increasing the sensitivity of

the apparatus to detect a lower onset temperature increases the γp but the variation in γp

remains the same. Therefore, there is a correlation between the calculated BDE and the

activation energy required to predict the experimentally determined onset temperature,

and irrespective of the sensitivity of the DSC, the γp parameter can be adjusted to

reproduce the experimental data.

3.32 Results

Based on the above discussion we assumed a value of 0.70 and 0.76 for γp for

mono and dinitro compounds, respectively. A FORTRAN program was written to

calculate dT/dt, as given by Equation (3.9), for temperatures starting with 30 oC. The

temperature was increased by 1 oC if the rate of increase of temperature was less than

0.01 oC/min. The temperature at which the gradient of temperature with time increased

at the specified rate of 0.01 oC/min was taken as the onset temperature. The predicted

onset temperatures for 19 different compounds are presented in Table 3.4 with an

average aggregate error of 11% and a bias of -2%. Typical errors associated with the

DSC detected onset temperatures are within ± 5%.

Page 70: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

60

B3P86 AM1

Sr. Structure To (oC) BDE To (oC) ? To (oC) error BDE To (oC) * ? To (oC) error

no. Expt.46 (kcal/mol) predicted exp - pre % (kcal/mol) predicted exp- pre %

1 nitrobenzene 380 75.6 297 83 22 27.6 302 78 21

2 2 nitrotoluene 290 73.4 282 8 3 26.2 290 0 0

3 3 nitrotoluene 310 75.9 301 9 3 27.6 304 6 2

4 4 nitrotoluene 320 76.7 305 15 5 28.5 312 8 3

5 2 nitroaniline 280 80.1 330 -50 -18 30.4 329 -49 -17

6 3 nitroaniline 300 76.5 303 -3 -1 26.8 294 6 2

7 4 nitroaniline 310 80.9 335 -25 -8 30.8 332 -22 -7

8 2 nitrobenzoic acid 270 66.4 231 39 14 22.9 259 11 4

9 3 nitrobenzoic acid 300 74.7 292 8 3 26.4 293 7 2

10 4 nitrobenzoic acid 310 76.5 306 4 1 27.3 302 8 3

11 2 nitrophenol 250 82.7 346 -96 -38 30.3 324 -74 -30

12 3 nitrophenol 310 75.77 294 16 5 26.5 288 22 7

13 4 nitrophenol 270 78.3 313 -43 -16 28.3 305 -35 -13

14 1,2 dinitrobenzene 280 64.5 251 29 10 18.1 246 34 12

15 1,3 dinitrobenzene 270 73.2 320 -50 -19 26.9 338 -68 -25

16 1,4 dinitrobenzene 350 72.9 319 31 9 26.8 338 12 3

17 2,6 dinitrotoluene 290 68.2 282 8 3 21.2 280 10 3

18 3,4 dinitrotoluene 280 64.7 254 26 9 18.0 247 33 12 19 2,4 dintitrotoluene 250 74.1 322 -72 -29 23.3 302 -52 -21

* Using scaled AM1 BDE values

Table 3.4. Summary of Experimental and Predicted Values for To (Tonset)

Page 71: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

61

200

220

240

260

280

300

320

340

360

380

400

8 14 18 2 17 9 12 1 3 6 4 10 13 16 15 19 5 7 11

Serial number of the compound as in Table 3.4

Ref 45

Ref 17

Ref 6

Predicted

Figure 3.3. Comparison of onset temperatures.

Page 72: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

62

The predictions are compared with experimental values from three different

sources6, 17,46 in Table 3.5 and are plotted in Figure 3.3. To data in Table 3.5, column 3,

were measured with a DSC at a scanning rate of 4°C/min46, and in column 4 with a DSC

at 10°C/min10, where the higher scanning rate, with other conditions constant,

consistently results in higher detected onset temperatures. Tonset data in Table 3.5,

column 5, are from a variety of sources.6 As seen from Figure 3.3, the experimental data

for each compound are scattered and the predictions appear to be reasonable. One reason

for the scatter in the experimental data is the graphical detection method for the onset

temperatures. Predictions from this model should better match measured values if data

from a more sensitive calorimeter such as an APTAC (Automated Pressure Tracking

Adiabatic Calorimeter) with larger sample sizes were used.

BDE values calculated with B3P86//cc-pVDZ were used for obtaining the γp. We

chose a density functional theory, B3P86, to determine the BDE values within

experimental accuracy but a typical optimization calculation can take about an hour on a

supercomputer. We also calculated BDE values using the quicker and less expensive

semi-empirical AM1 method47, and BDE values using each method are displayed in

Table 3.4. A typical optimization calculation using the AM1 method takes few seconds.

Although the AM1 calculated values of BDE are significantly lower than the

experimental values, they show a similar trend as exhibited by the higher level theory.

Consequently, the AM1 and B3P86//cc-pVDZ bond dissociation values were correlated

at a confidence level of ~ 0.9 and were related by the following equation:

BDE B3P86 = 1.3* BDEAM1 + 40.2

Thus the AM1 BDE values were scaled to be consistent with the higher level density

functional values, and the predicted onset temperatures exhibit a similar average

aggregate error of 10 % and a bias of -2%. The predicted onset temperature values from

the density functional method and from the scaled AM1 method are shown in Table 3.4.

Page 73: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

63

In some cases where the offset is large between predicted and experimental Tonset,

thermal effects other than the reaction heat, especially heat of vaporization or

sublimation, as discussed earlier, could significantly affect the predictions. Measured

data for a range of compounds employing a more sensitive calorimeter at slower

scanning rates with larger sample sizes should provide a good test of this prediction

method because thermal effects other than heat of reaction should be less significant.

Table 3.5. Comparison of Onset Temperatures

Sr. no.

Compound

To46

oC To

17 oC

To6

oC Predicted

oC 8 2 nitrobenzoic acid 270 305 230 231 14 1,2 dinitrobenzene 280 - - 251 18 3,4 dinitrotoluene 280 322 - 254 2 2 nitrotoluene 290 338 280 282 17 2,6 dinitrotoluene 290 - - 282 9 3 nitrobenzoic acid 300 375 320 292 12 3 nitrophenol 310 353 320 294 1 nitrobenzene 380 400 360 297 3 3 nitrotoluene 310 361 300 301 6 3 nitroaniline 300 347 280 303 4 4 nitrotoluene 320 329 - 305 10 4 nitrobenzoic acid 310 379 - 306 13 4 nitrophenol 270 302 250 313 16 1,4 dinitrobenzene 350 - - 319 15 1,3 dinitrobenzene 270 - 380 320 19 2,4 dintitrotoluene 250 312 312 322 5 2 nitroaniline 280 341 280 330 7 4 nitroaniline 310 345 280 335 11 2 nitrophenol 250 300 250 346

Page 74: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

64

4. Correlating Calorimetric Data with Molecular Descriptors

To predict reactivity of a substance it is necessary to deduce probable reaction

pathways, but it is difficult to predict the reaction pathways for a compound, especially

at high temperatures. The exothermic behavior of a substance is influenced by the

presence of functional groups 6,17,48, which also form a basis for reactivity classification.

This approach suggests that there is an inherent structure-property relationship between

the observed calorimetric properties and molecular structure. For example, as mentioned

earlier, the presence of the nitro group (NO2) can be a potential source of significant

reactivity. The reactive nature is manifested in calorimetric data, but this dependence of

observed behavior and molecular structure has not yet been quantified.

The Quantitative Structure-Property Relationship (QSPR) is a popular tool for

correlating observed values based on molecular properties. QSPR techniques have been

successfully employed for drug design49 and for correlating physical properties such as

boiling point50, autoignition temperature51, and molecular properties. In addition to

providing a means of predicting properties, a QSPR study may also lead to better

understanding of structural features affecting the observed data. The objective is to

correlate and predict DSC calorimetric data, namely onset temperature and energy of

reaction.

As discussed earlier, Tonset is indeed a ‘detected’ onset temperature because the

value depends on the sensitivity of the instrument and experimental technique.

Depending on the type of calorimeter, sample size, and scanning rate, Tonset can vary

within 5-50 oC for the same compound. For DSC data there is further distinction

between To (first deviation from base-line) and extrapolated To (intersection of base line

and tangent to the peak).

Page 75: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

65

The energy of reaction, often due to decomposition or polymerization, is the net

heat released during a reaction. As discussed, this energy is not the thermodynamic heat

of reaction, because it includes other effects such as evaporation and enthalpy of mixing,

and heat absorbed by the sample cell.

We built a Quantitative Property Structure Relationship (QSPR) study table to

investigate correlations between calorimetric data and molecular descriptors. The first

column of the study table is either Tonset or energy of reaction obtained from calorimetric

experiments and is called the dependent variable. The remaining columns are the

independent variables (characteristic of the molecules) called descriptors, which are

characteristics of a molecule and account for the electronic and chemical structure of the

molecule. A descriptor value can be obtained by experimental measurement or

calculated based on molecular structure. Theoretical descriptors were used to facilitate

property predictions for unknown molecules.

4.1 Data Set Selection

We chose compounds belonging to the organic nitro family since a better

correlation of properties is expected within a family of similar compounds. This set of

compounds that was used to develop the correlation is called a ‘training set’. The choice

of data is critical for an effective correlation, and we chose data from a single reference46

to maintain consistency in experimental procedure and calorimetric sensitivity. With a

larger data set a predictive model could be developed, but we could not find a large set

of consistent data in the open literature was not found.

Page 76: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

66

4.2 Discussion of a Few Descriptors

The choice of descriptors depends on the property to be correlated. Here

descriptors were selected that were expected to correlate with the detected onset

temperature or determined energy of reaction.

4.21 Highest occupied molecular orbitals (HOMO)

HOMO (highest occupied molecular orbital) is the highest energy level in the

molecule that contains electrons. It is crucially important in governing molecular

reactivity and properties. When a molecule acts as a Lewis base (an electron-pair donor)

in bond formation, the electrons are supplied from the HOMO. The ease of donation is

reflected in the energy of the HOMO. Molecules with high HOMOs are more able to

donate their electrons and are hence relatively reactive compared to molecules with low-

lying HOMOs.

4.22 Lowest unoccupied molecular orbitals (LUMO)

LUMO (lowest unoccupied molecular orbital) is the lowest energy level in the

molecule that contains no electrons. It is important in governing molecular reactivity and

properties.When a molecule acts as a Lewis acid (an electron-pair acceptor) in bond

formation, incoming electron pairs are received in its LUMO. Molecules with low-lying

LUMOs are more able to accept electrons than those with high LUMO.

HOMO and LUMO, collectively called as frontier orbitals, have been regarded as

major determinants of chemical reactivity.52 It has been shown that the electron

delocalization between HOMO and LUMO is generally indicative of easiness of

chemical reaction and the stereo-selective path, irrespective of intra and intermolecular

processes. For two approaching molecules larger the orbital overlapping for HOMO and

Page 77: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

67

LUMO and smaller the level of separation of two overlapping orbitals the larger the

contribution of the orbital pair to the stabilization of an interacting system. Thus a likely

explanation of the importance of chemical reactions would be the fact that molecular

orbital perturbation theory requires that interactions between molecules, leading to some

form of reaction, have to possess the appropriate frontier orbital energy values for

electron transfer.53

4.23 Highest positive charge (HPC) and highest negative charge (HNC)

These descriptors are probable indicators of electrophillic or nucleophillic attacks

and are expected to correlate with the Tonset, which is a measure of kinetics.

4.24 Weakest bond (WB)

The weakest bond, here the C-NO2 bond, in a molecule is a kinetic descriptor,

since it represents the minimum activation energy required for initiating the reaction.

4.25 Mid-point potential (Vmid)

Previous work54 has shown that the electrostatic potential produced by carbon

and nitrogen charges at the C-NO2 bond mid-point correlates with the impact sensitivity.

RQ

RQ

Vmid NC

5.05.0+=

where

QC – Atomic charge on carbon

QN – Atomic charge on nitrogen

R – C-NO2 bond distance

Page 78: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

68

It is believed that the buildup of positive electrostatic potential above the C-N bond is

responsible for destabilizing the bond.

4.26 Delocalizability index (Sr)

Here the descriptor (Sr) is defined as

∑=

=HOMO

i i

Sr1

where

ε i– Eigenvalue for the molecular orbital

The Sr index is similar to Fukui’s superdelocalizability index55. In the summation, the

eigenvalues for higher molecular orbitals (which are numerically smaller) will dominate,

and Sr is therefore a measure of delocalizability of electrons and a probable measure of

reactivity.

4.27 Dipole moment

The degree of polarity of a molecule is expressed in terms of dipole moment. The

electric dipole moment for a pair of opposite charges is defined as the magnitude of the

charge times the distance between them, and the defined direction is toward the positive

charge. It is useful in atoms and molecules where the effects of charge separation are

measurable, but the distances between the charges are too small to be easily measured.

The dipole moment of the molecule was also included as one of the descriptors.

4.28 Charge – bond strength descriptor (x)

This descriptor was calculated as follows

WBHNCHPC

x*5.0+

=

Page 79: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

69

and is an indicator of charge to bond strength ratio in a molecule.

The descriptors discussed above are aimed at capturing reactivity characteristics

of substances. Most of the reactions involving reactive chemicals follow radical

mechanism; these descriptors are suitable for elucidating reactivity since radical

reactions are less influenced by the environment of reactions than ionic reactions. The

reactivity indices, frontier electron density, superdelocalizability, and delocalizability are

defined to discuss the reactivities of unsaturated and saturated molecules.56 It can be

argued, based on electronic theory, that behavior of molecules, especially reactivity, can

be explained based on distribution of electrons in a molecule.

Values of these descriptors were obtained using the Gaussian44 suite of programs

with the B3P8633 density functional model and the cc-pVDZ34 basis set and are

summarized in Appendix B. To develop a computationally inexpensive method, more

sophisticated models were not tested at this stage of the project. Further statistical

calculations were performed using the Statistical Analysis System (SAS).57

4.3 Correlations

The correlations were obtained by performing least square regression analysis on

the training set of molecules. The developed correlation has the form

Y = A1X1 + A2X2 + A3X3 + ….+ AnXn

where

Y – Dependent variable

X – Independent variable (descriptor)

A – Regression constant

Page 80: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

70

4.31 Tonset

The details of the experimental onset temperatures and predicted values used are

summarized in Table 3.6. A statistical analysis was performed on the all descriptors and

their combinations, and the chosen descriptors were the ones that maximized the R2

value and minimized the error square terms58. As a result of this analysis the variables

that exhibited significant correlation with the dependent variable (Tonset) were retained,

and the statistically insignificant ones were discarded.

The following correlation based on the training set of 19 nitro compounds is

recommended:

Tonset (deg C) = 827.0 – 1035.79 * HPC – 4.4275 * Sr – 5.0654 * Dipole

A standard overall F-test (α = 0.05) indicates that the fitted correlation is significant and

not a chance correlation. The predicted onset temperature values with an absolute

average aggregate error of 6% and a bias of -0.5 % are listed in Table 3.6 and are plotted

against the experimental values in Figure 3.4. A correlation of 0.6 is obtained between

the predicted and the observed values. This level of correlation is reasonable given

significant variations in the experimentally determined onset temperatures due to the use

of a graphical detection procedure, as discussed above, and associated errors.

Page 81: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

71

Table 3.6. Onset Temperatures (observed and predicted)

Structure To (oC) Residual % error

Expt Predicted (Expt-Predicted)

1 nitrobenzene 380 348 32 8 2 1,2 dinitrobenzene 280 297 -17 -6 3 1,3 dinitrobenzene 270 304 -34 -12 4 1,4 dinitrobenzene 350 328 22 6 5 2 nitrotoluene 290 309 -19 -7 6 3 nitrotoluene 310 315 -5 -2 7 4 nitrotoluene 320 311 9 3 8 2,6 dinitrotoluene 290 280 10 3 9 3,4 dinitrotoluene 280 261 19 7 10 2,4 dintitrotoluene 250 260 -10 -4 11 2 nitroaniline 280 256 24 9 12 3 nitroaniline 300 302 -2 -1 13 4 nitroaniline 310 279 31 10 14 2 nitrobenzoic acid 270 281 -11 -4 15 3 nitrobenzoic acid 300 296 4 1 16 4 nitrobenzoic acid 310 295 15 5 17 2 nitrophenol 250 274 -24 -10 18 3 nitrophenol 310 316 -6 -2 19 4 nitrophenol 270 309 -39 -14

Page 82: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

72

240

260

280

300

320

340

360

380

400

240 260 280 300 320 340 360 380 400

Experimental (oC)

Predicted Agreement

Figure 3.4. Predicted onset temperatures.

Page 83: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

73

4.32 Energy of reaction

It is observed that the energy of reaction values correlate strongly with the count

of –NO2 group in the nitro compounds59, which is consistent with NO2 as an indicator of

reactivity. The experimental heat of reaction values for all the 19 compounds are listed

in Table 3.7 and summarized for the nitro and dinitro compounds in Table 3.8. The

correlation for the energy of reaction is:

Energy of reaction [-∆H] (kcal/gmol) = 75 * Number of nitro groups

Thus for TNT (3 nitro groups) the predicted energy of reaction is 225 kcal/mol, which is

consistent with the experimentally determined value of 2396 kcal/mol.

4.4 Correlations Using the Semi-empirical Method, AM1

The descriptors for the above study were generated using the B3P86/cc-pVDZ

model. Typically an optimization for an aromatic nitro molecule using this model

requires about an hour of CPU time on the supercomputer. Therefore, for the descriptors

to be easily calculable it is important that predictions be possible using a

computationally inexpensive semi-empirical theory, such as AM1.47

An optimization using AM1 can be performed in few seconds. To use the

correlations generated earlier we must scale the AM1 descriptor values to the

B3P86/cc-pVDZ values. The Sr and dipole descriptors calculated using the AM1 model

correlated with the B3P86/cc-pVDZ values to yield R2 values of 0.98 and 0.82,

respectively. However the HPC values calculated using the two models did not show a

good correlation. We recommend the following equations to scale the Sr and dipole

Page 84: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

74

Table 3.7. Experimental Energy of Reaction

Structure Energy of reaction (- ∆H) J/g kcal/mol

1 nitrobenzene 2757 81.1 2 1,2 dinitrobenzene 3310 133.0 3 1,3 dinitrobenzene 3488 140.1 4 1,4 dinitrobenzene 3701 148.7 5 2 nitrotoluene 2404 78.8 6 3 nitrotoluene 2070 67.9 7 4 nitrotoluene 2322 76.1 8 2,6 dinitrotoluene 3451 150.2 9 3,4 dinitrotoluene 3574 155.6 10 2,4 dintitrotoluene 3987 173.6 11 2 nitroaniline 2225 73.5 12 3 nitroaniline 2269 74.9 13 4 nitroaniline 2026 66.9 14 2 nitrobenzoic acid 1894 75.6 15 3 nitrobenzoic acid 1899 75.8 16 4 nitrobenzoic acid 1934 77.2 17 2 nitrophenol 2481 82.5 18 3 nitrophenol 2269 75.4 19 4 nitrophenol 2155 71.7

Page 85: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

75

Table 3.8. Summary of Energy of Reaction Values

Compounds Energy of reaction (- ∆H) (kcal/mol)

Average Mononitro* 75 ± 5

Dinitro† 150 ± 14

moment:

SrB3P86 = 1.3093 SrAM1 - 0.1105

DipoleB3P86 = 0.9329 DipoleAM1- 0.2047

With the scaled descriptors, the following three-parameter correlation can be employed

for calculating Tonset:

Tonset (deg C) = 828.5 – 5.7969 * SrAM1 – 4.7255*DipoleAM1

Using the AM1 scaled descriptors and the above two-parameter correlation, predictions

for onset temperatures for the 19 compounds yielded an average absolute aggregate error

of 7% and a bias of -1%, which is in good agreement with the predictions obtained using

the more expensive B3P86 descriptors. Values of descriptors and predicted onset

temperatures using AM1 are summarized in Appendix C.

* Statistical analysis was performed on 1,5,6,7,11,12,13,14,15,16,17,18,19 compounds in Table 3.7. † Statistical analysis was performed on 2,3,4,8,9,10 compounds in Table 3.7.

Page 86: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

76

5. Conclusions and Future Work

Computational screening to reactive hazards presents fundamental challenges in

predicting material properties and is of great interest to industry personnel. The QSPR

approach was also successfully used to correlate impact sensitivities to molecular

descriptors in our research group.60 Such correlations can be refined into a computerized

predictive tool to be run on a desktop computer.

Mary Kay O’Connor Process Safety Center (MKOPSC) is collaborating with

Dow Chemical Company, Midland, MI, and Eastman Kodak Company, Rochester, NY

to access their reactivity databases and develop correlations for a variety of families of

compounds. An initial analysis of available data indicates lack of data on pure

compounds, and computationally amenable compositions. A data set of 48 mono nitro

molecules, with average molecular weight of 234±85 g/mol was used to investigate

possible correlations with molecular descriptors. The energy of reaction yielded an

average value of 85±14 kcal/mol, consistent with results discussed earlier. However, the

onset temperatures did not yield satisfactory correlations with the descriptors discussed

earlier and lack of data on other families of compounds prevents testing of QSPR

descriptors. It is worth pointing out that calorimetric properties can be grouped

according to the functional groups, and values from different sources are summarized in

Table 3.9, 3.10, and 3.11. Based on values in the tables and the earlier classification, it is

possible to perform an easy screening of reactive hazards. For example, if the process

under consideration is ethylene (CH2=CH2) polymerization to form polyethylene, the

expected amount of heat released 20 kcal/mol, as indicated in Table 3.10. Since the

molecular weight of ethylene is 28 g/mol, the energy released per gram is ~ 700 cal,

indicative of a potential reactive hazard. If a molecule has a combination of functional

groups, the total energy released can be assumed to be sum of energy released by

individual functional groups, and the lowest of the onset temperatures among individual

groups can be taken as onset temperature for the combination.

Page 87: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

77

Table 3.9. Decomposition Energies for Typical Functional Groups 48

Sr. no.

Functional group Decomposition energy* (-∆H)

(kcal/mol) Formula Name

1. >C=C< 12 – 22 2. -C≡C- 29 – 36 3. -CC-

O

17 – 24

4. -COOH 55 – 67 5. -C=O

OOH

57 – 69

6. -C-O-O-C

81 – 86

7. >S=O 10 – 17 8. SO2Cl 12 – 17 9. -NH-NH- 17 – 22 10. -N=N- 24 – 43 11. -N≡N+- 38 – 43 12. >C=N≡N 41 – 45 13. -N=N≡N 48 – 57 14. -N=N-N< 60 – 65 15. >C=NOH Oxime 26 – 33 16. >N-OH N-hydroxide 43 – 57 17. >N:O N-oxide 24 – 31 18. >C-N=O Nitroso 36 – 69 19. -N=C=O Isocyanate 12 – 18 20. >C-NO2 Nitro 74 – 86 21. >N-NO2 N-Nitro 96 – 103 22. -O-NO2 Acyl nitrate 96 – 115 23. NH2.HNO3 Amine nitrate 72 – 84

*. To convert to cal/g multiply the heat of decomposition by 1000/MW

O O

Page 88: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

78

Table 3.10 Typical Values for Energetic Polymerization61

Monomer functional

group

Polymer bond

Polymerization Energy (-∆H)

kcal/mol 1. -C=C- -C-C- 20 2. -C=O- -C-O- 5 3. -C=N- -C-N- 1.4 4. -C≡N- -C=N- 7.2 5. -C=S- -C-S- 2 6. -S=O- -S-O- 7

Page 89: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

79

Table 3.11 -∆H, To Values Based on Functional Groups 59

Chemical Class

Unstable Group

Decomposition Energy (-∆H)

kcal/mol

DSC (To)

°C

Mean Standard dev. Mean Standa

rd dev. 1. Perchlorate -ClO4 830 60 200 20 2. Nitro -NO2 340 60 240 60 3. Polynitro 350* 40 240 60 4. Nitroso -N=O 200 30 120 30

5. N-oxide

N O

140 40 210 30

6. Hydroxylamine, Oxime >N-OH 140 50 160 40

7. Tetrazole

N

N N

N

210 40 160 10

8. Triazene, triazole -N=N-N< 210 50 230 60

9. Diazo

N+

N

N+

N

220 60 110 20

10. Azo -N=N- 200 60 230 40

11. Hydrazine -NH-NH2 >N-NH2

100 30 210 40

12. Substituted Hydrazine

-NH-NH- >N-N< 100 30 220 50

13. -N-N- in a ring 100 60 250 50

14. 2 -N-N- in a ring 190 90 210 50

15. Imidazole N-C-N (ring) 130 60 270 50 16. Oxazole N-C-O (ring) 80 40 260 50 17. Thiazole N-C-S (ring) 80 50 260 40

18. Tetrazole + nitro 540 120 160 20

19. Substituted Hydrazine+ nitro

430 50 180 20

Page 90: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

80

Future experimental work in our research group will focus on building libraries

of data for compounds, which would then be used to develop correlations based on

molecular level descriptors, and development of robust correlations to enable structure-

based predictions. Based on this analysis one can also envision descriptors (or imprints

of energetic materials) that can be developed into a qualitative hazard index.

The discussion so far has focused on characterizing and screening reactive

hazards. But systems where hazards are identified additional considerations and

resources are required. The next chapter focuses on a detailed investigation of

hydroxylamine (HA), as an example of a hazardous system.

Page 91: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

81

CHAPTER IV

DETAILED INVESTIGATION OF A REACTIVE SYSTEM

Almost all mid-size and large companies have a reactive hazard management

program to assess potential reactive hazards during storage, transport, and processing of

reactants, intermediates, and products. For evaluating reactive hazards, it is rational to

develop a systematic protocol with a screening step, utilizing available information,

computations and experiments, as discussed earlier, followed by detailed testing. The

classification discussed above can be used as a screen to select the more hazardous

compositions for detailed testing. An overall assessment approach is shown in the

Figure 4.1. Earlier chapters focused on development of tools that can expedite reactive

hazard assessment. This chapter focuses on an investigation of hydroxylamine system, as

an example of highly reactive system.

1. Background

Hydroxylamine (HA), NH2OH, has recently been involved in two major

industrial incidents with disastrous consequences.62,63 Calorimetric studies on aqueous

solutions of HA indicate that it is a highly reactive compound64, but its properties are

Part of this chapter is reprinted with permission from, “Theoretical thermochemistry: ab initio heat of formation for hydroxylamine”, by S.R. Saraf, W.J. Rogers, M. Sam Mannan, M.B. Hall, and L.M. Thomson, J. Phys. Chem., Vol. 107, No. 8, 2003. Copyright 2003 by the name of American Chemical Society (ACS). Part of this chapter is reprinted with permission from “Hydroxylamine production: Will a QRA help you decide?”, by K. Krishna, S.R. Saraf, Y.J. Wang, J.T. Baldwin, W.J. Rogers, J.P. Gupta, and M. Sam Mannan, Reliability Engineering & System Safety, 2003, 81, 215-224,. Copyright 2003 by the name of Elsevier.

Page 92: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

82

Figure 4.1. Systematic approach for assessing reactive hazards.

poorly characterized. Pure HA is known to explode at room temperature, and the

decomposition of HA is extremely sensitive to metal contamination.64

2. Ab initio Heat of Formation for HA

For chemicals with validated experimental data, estimations may not be

necessary, but for reactive substances with insufficient experimental data, such as

hydroxylamine, estimation methods are of prime importance.

Screening (Literature, Experimental and/or computational)

Selection of more hazardous compositions for further investigation

(Criteria based on classification discussed earlier)

Detailed testing (Sophisticated calorimetric tests and

detailed quantum calculations)

Page 93: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

83

The reported experimental value for the heat of formation of gaseous HA is

-12.0 ± 2.4 kcal/mol65, which was derived by an indirect calculation from the

experimentally determined heat of formation of solid HA and the heat of

sublimation66,67,68. Data from measurements of solid HA should be more reliable than

data from liquid HA, because pure HA decomposes as it melts near 32 °C.66 However,

the reliability of the listed heat of formation for solid HA could not be assessed, because

the experimental procedure used to determine it was not found in the original

reference.68

Previous calculations by Sana, et al. 69 yielded -11.7 kcal/mol for the heat of

formation of gaseous HA using an isodesmic reaction at the MP4 level of theory.

Anderson70 determined a value of -10.6 kcal/mol by combining the calculated H2N–OH

bond energy71 by the G1 method with the experimental heats of formation for NH2 and

OH. Also, a heat of formation70 of -7.9 ± 1.5 kcal/mol was derived from the appearance

potential for NH2OH72 and heat of formation of HNO. Based on a statistical average of

the reported theoretical and experimental values for HA, Anderson70 recommended -9.6

± 2.2 kcal/mol for the gaseous HA heat of formation at 1 atm and 298.17 K.

The purpose of this work is to compare theoretical methods combined with

isodesmic reactions to obtain a reliable heat of formation for gaseous HA. For

comparison with an estimation approach often used by industry, the HA heat of

formation is calculated by the traditional Benson group contribution method. These

methods together with the details of the calculations are discussed in the next section and

are followed by a discussion of the results. The heat of formation values reported from

this work are for gaseous species at 1 atm and 298.17 K.

Page 94: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

84

2.1 Computational Methods

2.11 Benson Group Contribution Method

Benson and Buss proposed a hierarchy of additivity methods for molecular

property estimations and established a theoretical framework to estimate heats of

formation based on ‘molecular groups’73. Employing the commercially available

CHETAH18 software, which includes the Benson group contribution method. Because

the H2N-(O) and HO-(N) groups were not available, the group values for NH2-(N), 11.4

kcal/mol, and OH-(O), -16.27 kcal/mol, were used as substitutes for the missing ones, as

recommended in the CHETAH18 manual. This procedure yielded -4.87 kcal/mol for the

gaseous HA heat of formation. However, substitutions for missing group values often

leads to deviations from the experimental values.

2.12 Theoretical Methods

A variety of theoretical methods, semi-empirical (AM147), density functional

theory (B3P8633 and B3LYP74), composite (G2,75 G3,76 G2MP2,77 G2MP2B3,78 G3B3,78

and CBS-Q79), and ab initio (MP2,80 MP3,81 MP4(SDTQ),82 CCSD,83 CCSD(T),84 and

QCISD(T)85) as implemented in the Gaussian 98 suite of programs44, were used for

geometry optimizations and frequency calculations. These calculations were performed

with Dunning correlation consistent polarized valence basis sets (cc-pVDZ34,

cc-pVTZ86, cc-pVQZ87, and cc-pV5Z88, where D, T, Q, and 5 refer to the number of

contracted functions in each valence sub-shell), and Dunning correlation consistent

polarized valence basis sets with diffuse functions for radial flexibility to represent

electron density far from the nuclei (AUG-cc-pVDZ and AUG-cc-pVTZ). Pople-style

basis sets 89,90 (6-31G, 6-31+G(d), 6-31G(d,p), 6-31+G(2df,p), 6-311G(d), 6-

311+G(2df,p), 6-31+G(3df,2p), 6-311+G(3df,2p), 6-311++G(3df,2p)) including

diffuse91,87 (denoted by “+” for Pople-style) and polarization functions92 (denoted by

Page 95: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

85

“d”, “p”, ‘f”, for angular flexibility to represent regions of high electron density among

bonded atoms) were also employed. Finally, the Bond Additivity Correction (BAC)-

MP4 methodology was employed using the parameters listed by Melius and Zachariah93.

Errors in absolute quantities from quantum chemical calculations are often systematic.

To compensate for some of the systematic errors, isodesmic reactions, which conserve

the number of each type of bond in reactants and products, are used to obtain more

accurate heats of formation94. Here, the following isodesmic reactions were employed

for HA:

H2 + NH2OH ? H2O + NH3 (4.1)

H2O + NH2OH ? H2O2 + NH3 (4.2)

To benchmark the computed HA values, the heat of formation for hydrogen peroxide, a

similar species for which reliable experimental data are available, was calculated by the

same methods and with the following isodesmic reaction:

H2 + H2O2 ? 2 H2O (4.3)

The usual procedure for calculating the heat of formation value of an unknown

compound is to combine the heat of reaction obtained from an isodesmic reaction with

the experimental heat of formation values for the known compounds94. The HA heat of

formation using Reactions (4.1), (4.2), and (4.3) were determined from the equations

(4.4), (4.5), and (4.6) respectively using the calculated heat of reaction, ∆HCalcRxn, and

the experimental heats of formation values at 1 atm and 298.17 K for ammonia37,

water37, and hydrogen peroxide38 listed in Table 4.1.

∆Hf, NH2OH = ∆HExptf, NH3 + ∆HExpt

f, H2O - ∆HExptf, H2 - ∆HCalc

Rxn (1) (4.4)

∆Hf, NH2OH = ∆HExptf, NH3 + ∆HExpt

f, H2O2 - ∆HExptf, H2O - ∆HCalc

Rxn (2) (4.5)

Page 96: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

86

∆Hf, H2O2 = 2 ∆HExptf, H2O - ∆HExpt

f, H2 - ∆HCalcRxn (3) (4.6)

The choice of isodesmic reaction is important to obtain accurate values. Although there

are 5 single bonds on the reactant side (1 H-H, 1 O-H, 1 N-O, 2 N-H) and on the product

side (3 N-H, 2 O-H) in Reaction (4.1), the N-O bond on the reactant side is not balanced

by a similar σ bond on the product side. Reaction (4.3) is similar to (4.1) in terms of

bond balance with the O-O bond unbalanced on the reactant side. In Reaction (4.2), there

are 6 single bonds on the reactant side (3 O-H, 1 N-O,2 N-H) and on the product side (3

N-H, 2 O-H, 1 O-O), but here the N-O bond is balanced better by the O-O bond on the

product side. A better bond balance should result in a more effective cancellation of

errors, therefore, Reaction (4.2) should yield a more accurate value for ∆HCalcrxn than

Reaction (4.1) at the same level of theory. Thus similar errors are expected in the heat of

formation values calculated using Reactions (4.1) and (4.3), and faster convergence with

increasing level of theory for Reaction (4.2). In addition, agreement between values

obtained from Reactions (4.1) and (4.2) can serve as an indicator that the theory is

adequate to model the system.

Table 4.1: Experimental Heats of Formation (1atm and 298.17 K)

Compound Molecular Formula Heat of Formation (kcal /mol)

Ammoniaa NH3 – 10.98 ± 0.084

Watera H2O – 57.7978 ± 0.0096

Hydrogen Peroxideb H2O2 – 32.58 ± 0.05c

a. Ref. 37. b. Ref. 38. c. Based on the listed experimental errors.

Page 97: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

87

2.2 Results and Discussion

Values for the HA heat of formation calculated using the various levels of theory

and basis sets are presented in Table 4.2, and computed N-O bond lengths (HA) and O-O

bond lengths (hydrogen peroxide) are listed in Appendix D.

The Austin Model 1 (AM1) yielded a good prediction for the hydrogen peroxide

heat of formation, but the heat of formation value obtained for HA differed significantly

from the values obtained via ab initio, density functional, or the composite methods.

Semi-empirical methods, like AM1, perform equally well for similar compounds for

which parameters are available. However, in this case, AM1 models the O-O bond in

hydrogen peroxide but does not appear to work well for the N-O bond in HA.

Hartree-Fock (HF) is the lowest level ab initio theory employed in this work. The HF

theory is expected to yield fair to good results, despite the fact that it does not include a

full treatment of electron correlation, because errors are partially cancelled with the use

of isodesmic reactions. Heats of formation calculated with the Hartree-Fock model did

not exhibit consistent improvement with increasing basis sets, but they generally yielded

more consistent results for Reaction (4.2).

The density functional methods, although not truly ab initio, include electron

correlation at only a moderate increase in computing cost, as compared to HF, by using

functionals of electron density. Among the density functional methods, B3P86 yielded

slightly better results for hydrogen peroxide than B3LYP for identical basis sets.

However, even B3P86 with a 5Z basis set has an error of nearly 2 kcal/mol, as compared

to the experimental value, for hydrogen peroxide with Reaction (4.3). Unlike HF theory,

increasing basis functions (cc-pVDZ, cc-pVTZ, cc-pVQZ, and cc-pV5Z) and adding

diffuse functions in the density functional methods leads toward consistent values of the

heat of formation. Similar values were obtained using the 6-311+G (3df, 2p) and 6-31+G

(3df, 2p) basis sets. At the density functional level of theory, there was a significant

Page 98: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

88

Table 4.2. Summary of Calculated Heats of Formation (∆Hf)

Hydroxylamine Hydrogen Peroxide

Method Basis Set Heat of formation (kcal/mol)

Difference between rxn (1) and (2)

Heat of formation (kcal/mol)

Reaction (1) Reaction (2) Reaction (3)

AM1 -32.34 -31.31 1.0 -33.61

HF cc-pVDZ -12.14 -12.02 0.1 -32.69

cc-pVTZ -9.76 -12.48 2.7 -29.85

cc-pVQZ -8.83 -13.06 4.3 -28.34

6-31G -10.65a -7.14 3.5 -35.59a

6-31G(d) -16.10 -10.69 5.4 -37.98

6-31+G(d) -17.47 -8.07 9.4 -33.11

6-31G (d,p) -11.81a -12.06 0.3 -32.42a

6-31+G (2df,p) -7.65a -11.71 4.1 -24.93a

B3P86 cc-pVDZ -18.67 -7.79 10.9 -43.45

cc-pVTZ -13.99 -9.31 4.7 -37.57

cc-pVQZ -12.73 -10.08 2.7 -35.03

cc-pV5Z -12.01 -10.39 1.6 -34.20

AUG-cc-pVDZ -10.62 -10.23 0.4 -33.00

AUG-cc-pVTZ -11.83 -10.39 1.4 -34.02

6-311G (d) -21.70 -7.12 14.5 -47.16

6-31+G (3df,2p) -12.63 -10.88 1.8 -34.32

6-311+G (3df,2p) -12.14 -10.89 1.3 -33.83

6-311++G (3df,2p) -12.12 -10.89 1.2 -33.81

B3LYP cc-pVDZ -18.76 -5.24 13.5 -46.10

cc-pVTZ -14.92 -8.38 6.5 -39.12

AUG-cc-pVDZ -10.48 -9.26 1.2 -33.80

AUG-cc-pVTZ -12.18 -9.69 2.5 -35.07

6-311+G(3df,2p) -8.59 -6.17 2.4 -34.99

MP2 cc-pVDZ -14.39 -9.76 4.6 -37.21

cc-pVTZ -10.24 -11.01 0.8 -31.81

Page 99: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

89

Table 4.2. (Contd.)

Hydroxylamine Hydrogen Peroxide

Method Basis Set Heat of formation (kcal/mol)

Difference between rxn (1) and (2)

Heat of formation (kcal/mol

Reaction (1) Reaction (2) Reaction (3)

cc-pVQZ -8.61 -12.09 3.5 -29.10

MP3 6-31+G(2df,p) -8.76a - - -28.38a

cc-pVDZ -14.44b -9.82 b 4.6 -37.20 b

cc-pVTZ -10.46 b -10.77 b 0.3 -32.27 b

MP4 6-31+G(2df,p) -11.26 -11.89 0.6 -31.96

MP4(SDT

Q) cc-pVDZ -16.80 b -8.21 b 8.6 -41.17 b

CCSD(T) cc-pVDZ -17.05 -8.11 8.9 -41.52

cc-pVTZ -13.02 c -9.52 c 3.5 -36.07 c

cc-pVQZ -11.56 c -10.61 c 1.0 -33.52 c

QCISD(T) cc-pVDZ -17.08d -8.12d 9.0 -41.54d

BAC-MP4 MP4//HF -12.98 -11.09 1.9 -34.46

G2 -11.78 -11.53 0.3 -32.83

G2MP2 -11.69 -11.67 0.0 -32.60

G3 -11.15 -11.28 0.13 -32.46

G3MP2B3 -11.88 -11.45 0.4 -33.01

G3B3 -11.51 -11.35 0.2 -32.74

CBS-Q -12.18 -11.16 1.0 -33.60

GAe - 4.87 - - -32.50

Experiment

al

-12.0f

-7.9h - - -32.58g

a. Ref. 69 b. Single point energies and thermal corrections for the enthalpies for MP3 (SDTQ) and MP4 (SDTQ) were

calculated at MP2/cc-pVDZ geometry. c. Single point energies and thermal correction for the enthalpy for the CCSD(T)/cc-pVTZ and CCSD(T)/cc-pVQZ

were calculated at CCSD(T)/cc-pVDZ geometry. d. Single point energy and thermal correction for the enthalpy for the QCISD(T)/cc-pVTZ were calculated at

CCSD(T)/cc-pVDZ geometry. e. Group Additvity

f. Based on an indirect calculation as discussed in the text (Ref. 65, 66). g. Ref.37. h. Ref. 70 .

Page 100: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

90

change in the calculated heat of formation from the cc-pVDZ to cc-pVTZ basis set. The

magnitudes of the calculated values decreased with Reaction (4.1) as the quality of the

basis set was increased, but they increased for Reaction (4.2). Thus, Reaction (4.1) and

(4.2) approached the basis set limit for the heat of formation from opposite directions.

The composite theories (G2, G3, G2MP2, G3MP2B3, G3B3 and CBS-Q) are

expected to yield the best results since they have been developed to model accurately

thermochemical quantities for small, light-atom, main group molecules. The mean

absolute deviations (MAD) associated with heat of formation value obtained using G2

and G2MP2 theories (with the G2 test set) are 1.2 and 1.6 kcal/mol, respectively95. The

G3 theory is a further improvement over G2 and reduces the MAD to 1.0 kcal/mol76.

The CBS-Q accounts for errors due to basis set truncation by an extrapolation, and the

MAD associated with the method is 1.0 kcal/mol95. The MAD associated for G3,

G3MP2B3, and G3B3, based on heat of formation values for 148 different molecules,

are 0.94, 1.13, and 0.93 kcal/mol, respectively78. All of these composite theories

performed well and predicted accurate energies for hydrogen peroxide.

The MP3 (SDTQ) and MP4 (SDTQ) results were poor for cc-pVDZ, but the

MP3 prediction improved with the cc-pVTZ basis set. CCSD(T)/cc-pVDZ and

QCISD(T)/cc-pVDZ geometries agree well with the experimental values, but realistic

energy predictions were obtained only with the cc-pVQZ basis set.

For Reactions (4.1) and (4.3), as expected, the heat of formation values obtained

for HA and hydrogen peroxide respectively exhibited trends in similar direction for the

various levels of theory and basis sets. For the HA heat of formation values using

Reaction (4.2), there was faster convergence with the basis sets for the same level of

theory. As can be seen from the heat of formation values from Reaction (4.2), accurate

values were obtained at lower levels of theory and with smaller basis sets.

Page 101: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

91

2.3 Choice of Best Values

The difference between the values calculated using Reaction (4.1) and (4.2) can

be taken as a guide for selecting theories performing well for the system. The calculated

values, in Table 4.2, that exhibited a difference of 1 kcal/mol or less are shown in bold.

It is worth noting that these theories also predict a reasonable value for hydrogen

peroxide heat of formation (within 1 kcal/mol). However, not all of these methods are

reliable in other respects. Omitted from the final values to be averaged were the less

reliable semi-empirical AM1 predictions, since the predicted value for ∆Hf, NH2OH were

significantly different from the values obtained by other methods. The values obtained

using theories (HF, B3P86, B3LYP) that did not demonstrate an improvement in the

prediction with increasing basis set were also left out. The MP values were also left out

since these values did not exhibit convergence with increases in the basis set or the

perturbation level.

Table 4.3 summarizes the 7 best-predicted heat of formation values and their

averages, where the calculated values using Reactions (4.1) and (4.2) differed by no

more than 1 kcal/mol. The average calculated heat of formation for hydrogen peroxide

was -32.9 kcal/mol with a standard deviation of 0.4 kcal/mol. With Reaction (4.1), an

average value of -11.7 kcal/mol with a standard deviation of 0.3 kcal/mol was

calculated. With the more balanced Reaction (4.2), the average value was -11.4 kcal/mol

with a standard deviation of 0.3 kcal/mol.

Page 102: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

92

Table 4.3. Accurate Values for HA Heat of Formation

NH2OH H2O2

Heat of formation (kcal/mol)

Theory Basis set

Mean

Average

Deviation

Rxn (1) Rxn (2) Rxn (3)

G2 1.295 -11.78 -11.53 -32.83

G2MP2 1.695 -11.69 -11.67 -32.60

G3 1.078 -11.15 -11.28 -32.46

G3MP2B3 1.1378 -11.88 -11.45 -33.01

G3B3 0.9378 -11.51 -11.35 -32.74

CBS-Q 1.095 -12.18 -11.16 -33.60

CCSD(T) cc-pVQZ - -11.56 -10.61 -33.52

Average 1.1 -11.7 -11.4 -32.9

St. dev. 0.3 0.3 0.3 0.4

Experimental -12.065 -32.5837

-7.970

The deviations from the average heat of formation for the various methods in

Table 4.3 are shown in Figure 4.2. From this pattern, it is apparent that the deviations

from average for heat of formation values from Reaction (4.1) and (4.3) track each other,

whereas the deviations for Reaction (4.2) do not follow the same trend. The average

calculated value of ∆Hf value for hydrogen peroxide is greater than the experimental

value by 0.3 kcal/mol. Since Reaction (4.1) and (4.3) are expected to yield similar errors,

Page 103: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

93

we believe that ∆Hf value obtained using Reaction (4.1) will differ from the true value in

a similar manner and therefore recommend -11.4 kcal/mol for the ∆Hf of NH2OH as the

best estimate from both Reaction (4.1) and (4.2). The mean absolute deviation (MAD)

for each of the methods employed is listed in Table 4.3 and the average MAD value is

approximately 1.1 kcal/mol. However, the HA ∆Hf values are computed from isodesmic

reactions which should yield values with smaller errors, perhaps down to twice the

standard deviation of the various method. Thus the recommended ∆Hf value for HA,

including our precision, judgment of methodology, and accuracy is -11.4 ± 0.6 kcal/mol.

Furthurmore, the agreement and the consistency of the calculated hydrogen peroxide

average value with the experimental value suggests that our calculated average value for

∆Hf of HA is more reliable than the available experimental values, which as discussed in

the introduction, cannot be properly assessed.

-0.80

-0.60

-0.40

-0.20

0.00

0.20

0.40

0.60

0.80

G2 G2MP2 G3 G3MP2B3 G3B3 CBS-Q CCSD(T)

Method

Dev

iati

ons

from

ave

rage

hea

t of

fo

rmat

ion

(kca

l/mol

)

Rxn (1)

Rxn (2)

Rxn (3)

Figure 4.2. Deviations from the average heat of formation values for the

methods employed in Table 4.3.

Page 104: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

94

Also, we have illustrated the importance of well-balanced isodesmic reactions for

determining accurate heats of formation, especially at lower levels of theory. Depending

on the level of theory, triple-ζ (6-311G or cc-pVTZ) or larger basis sets are necessary to

predict accurate HA heat of formation values. At all levels of theory the double-ζ (6-31G

or cc-pVDZ ) basis set yielded poor energies, but CCSD(T) and QCISD(T) predicted

accurate geometries in this basis set. The methods employed in obtaining the average

heat of formation value have an absolute accuracy of approximately 1.1 kcal/mol, but the

value obtained using isodesmic reactions is expected to have a smaller error and

therefore 1.1 kcal/mol represents the maximum absolute error in the calculation. As

expected the highly parameterized composite methods (G2, G3, G2MP2, G3MP2B3,

G3B3, CBS-Q) yielded the most accurate values. However, the unparametrized ab intio

CCSD(T)/cc-pVQZ yielded nearly as accurate values and in some cases, depending on

the accuracy needed, density functional methods, MP3, and MP4 may be adequate.

Heat of formation values can also be used for estimating decomposition energies.

If we use the value of –11.4 kcal/mol, CHETAH calculates the maximum heat of

decomposition as –1515 cal/g; thereby indicating a potential reactive hazard. According

to the program HA should decompose with the following stoichiometry:

NH2OH (g) → 0.33 N2 (g) + H2O (g) + 0.33 NH3 (g)

The experimentally observed heat of reaction is around 850 cal/g64, which includes all

thermal effects discussed above. The next section discusses a plausible decomposition

pathway for HA.

Page 105: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

95

3. Investigation of Hydroxylamine Runaway Behavior

3.1 Experimental Observations

Pure HA is known to explode at room temperature. Calorimetric studies on

aqueous solutions of HA indicate that it is a highly reactive compound and the

decomposition of HA is extremely sensitive to metal contamination.64 Thermal

instability of hydroxylamine is clear from the APTAC results illustrated in Figure 4.3.

Figure 4.3. APTAC temperature-time profile for 50 wt% HA.

As shown in the above figure a significant reaction for 50 wt% HA is detected at around

120 oC. But when a small quantity of Fe2SO4 solution (8 ppm Fe2+) is added, HA reacts

rapidly at room temperature. Also, the contaminated sample shows a larger exotherm, as

seen from the red curve in the graph. The mechanism of Ha decomposition was

0

50

100

150

200

250

0 500 1000 1500

Time (min)

50 wt% HA + 0.8 ppm Fe2+

50 wt% HA

Page 106: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

96

investigated, in presence and absence of metals. The aim of studying elementary

reactions is to understand initiation steps leading to the HA runaway reactions. Such

understanding of elementary reactions will aid in understanding the fundamental

behavior of HA, development of better inhibitors to prevent metal catalysis, and

consequently thermal runaway reactions. The results of our calculations are discussed in

the following section.

3.2 Theoretical Calculations

Runaway reactions, leading to explosion, generally follow a radical mechanism

and the mechanism can be categorized into initiation, propagation, and termination steps.

A key component of such explosive reactions are branching reactions, wherein two or

more radicals are created from a single radical, that accelerate the reaction.40 With the

above considerations, the following rules were observed to generate reaction network for

hydroxylamine decomposition reactions:

1. The first step is cleaving of bonds in the HA molecule. The dimerization of

hydroxylamine was not considered because experimental results do not suggest a

polymerization reaction.

2. There are no ionic reactions.

3. Every specie reacts with every other specie, but for any elementary reaction,

there are no more than two reactants. This is a reasonable assumption since three

or more species coming together is a much rarer event.

4. Experimentally observed products96 or stable species, such as NH3, H2O, N2O do

not participate in the chain propagation step.

5. Solvation effects have been neglected.

The following elementary reactions of hydroxylamine decomposition calculated

using the B3P86/cc-PVDZ level of theory, since good energy values were obtained for

the heat of formation calculations using this theory. Based on earlier discussion, the

Page 107: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

97

proposed reactions are divided into three classes – initiation, propagation, and

termination, and the favorable reactions are summarized below. Associated enthalpies in

kcal/mol are indicated by ‘∆’ and products (stable species) are underlined.

Initiation

NH2OH → •NH2 + •OH ∆ = 61.6 kcal/mol

NH2OH → NH2O• + •H ∆ = 66.1

NH2OH → •NHOH + •H ∆ = 75.2

Propagation •NH2 + NH2OH → NH3 + NH2O• ∆ = -38.0 •OH + NH2OH → H2O + NH2O• ∆ = -47.2 •H + NH2O → •NH2 + H2O ∆ = -51.7 •H + NH2OH → •OH + NH3 ∆ = -42.2 •OH + NH2O• → HNO + H2O ∆ = -46.3 •NH2 + NH2O• → NH3 + HNO ∆ = -37.0 •H + NH2O• → HNO + H2 ∆ = -34.4 •H + NH2O• → NH + H2O ∆ = -26.1 •H + NH2O• → NH3 + O• ∆ = -8.6

HNO + NH2OH → NH2O• + NH2O• ∆ = -0.9 (Branching reaction)

NH + NH2OH → •NH2 + NH2O• ∆ = -25.6 (Branching reaction)

•O + NH2OH → NH2O• + •OH ∆ = -33.8

HNO + •NH2 → NH3 + •NO ∆ = -58.2

HNO + •OH → •NO + H2O ∆ = -67.5 •NO + NH → N2 + •OH ∆ = -87.6

Page 108: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

98

Termination •NH2 + •H → NH3 ∆= -104.0 •OH + •H → H2O ∆= -114.0 •NH + NHO → N2O + H2 ∆= -26.0

NHO + NHO → N2O + H2O ∆= -87.7 •NO + •NH2 → N2 + H2O ∆= -109.2

During the initiation of a reaction, the newly formed radicals have a higher probability of

reacting with the reactant since the concentration of reactants is significantly higher. So

for the propagation steps the reactions involving NH2OH would be responsible for

accelerating the rate initially. The product spectrum, N2, N2O, H2O, NH3, H2, agrees

with experimentally observed products. However, from a safety viewpoint, it is

important to understand the initiation steps leading to the HA runaway reactions. Such

understanding of elementary reactions will aid in understanding the fundamental

behavior of HA, development of better inhibitors to prevent metal catalysis, and

consequently thermal runaway reactions. The catalytic behavior of HA was investigated

by probing potential reaction pathways in the presence of ferrous (Fe2+) ions. The

Stuttgart basis set97 was employed for modeling the Fe2+ ion. A few of the possible

reactions are shown in Figures 4.4, 4.5, and 4.6.

Page 109: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

99

N

Fe

O

Fe

NO

Figure4.4. Probable elementary reactions of HA in the presence of Fe2+.

N

FeFe

O

Fe

N

N

O

Figure 4.5. Fe2+ interacting with the nitrogen atom of HA.

16 kcal/mol

- 2 H2O

- 2 kcal/mol

- H2O

63 kcal/mol

- OH

Page 110: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

100

Fe

N

Fe

O

Fe

O

NO

Figure 4.6. Fe2+ interacting with the oxygen atom of HA.

Theoretical studies have provided insight into the HA - metal interactions, an

area of considerable interest in the field of catalysis modeling. Future experimental work

is aimed at gathering species-time data for developing better kinetic models and

validate some of the proposed elementary reactions.

4. Integrating Reactivity Data and Risk Analysis for Improved Process Design98

From the earlier sections of this chapter, it is evident that HA poses reactive

hazards. Although studies have characterized some of the hazards, risk involved from an

operational perspective is not obvious. The aim of this work is to integrate the available

knowledge on HA stability and perform a quantitative risk assessment on a generic HA

production plant as shown in Figure 4.7. Quantitative risk analysis aids decision-making

in two ways: it identifies the dominant contributors to the total risk, and it quantifies the

benefits of possible changes. These findings lead naturally to the specification of

- 6 kcal/mol

- H2O

59 kcal/mol

- NH2

Page 111: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

101

possible measures to improve reliability or reduce the damage potential. Figure 4.8

depicts the procedure involved in quantitative risk analysis.

Based on available knowledge, major risks lie in the CSTR and distillation tower

besides the design fault, construction error, and other external factors. Fault tree

techniques are used to estimate the probability of these events with the existing

safeguards. The results of the fault tree are analyzed and conclusions and

recommendations are determined. The benefit of design changes and safeguards, such as

a temperature interlock on the CSTR and quench valve on the distillation column, can be

easily verified and compared by the fault tree results. These guidelines are also

applicable to the production of other hazardous chemicals.

Page 112: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

102

Figure 4.7. Hydroxylamine production.

Page 113: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

103

Figure 4.8. Quantitative risk analysis scheme.

Page 114: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

104

5. Conclusions

Earlier chapters discussed the challenges in characterizing energetic materials.

This chapter focuses on hydroxylamine (HA) system, since HA was involved in two of

major industrial incidents, it has received considerable attention from researchers in last

few years. A systematic protocol for reactive hazard assessment would have recognized

the thermal hazards involved in the HA production unit. Based on calculations, the

recommended heat of formation for HA is -11.4 ± 0.6 kcal/mol. This is especially

important since experimentation on pure HA is not possible due to explosive hazards,

and demonstrates the importance of predictive techniques. A summary of plausible

reaction steps is proposed for HA decomposition in agreement with experimentally

observed products and heat of reaction. Continuing work focuses on interaction of HA

with metal atoms. The main objective is to understand the initiation and critical reactions

involved in HA runaway and use this information to develop inhibitors, radical

scavengers or chelating agents, to restrain progress of the reaction. Further, a risk

assessment study was performed to utilize the available knowledge of HA reactivity for

designing a safer production unit.

It should be noted that HA is an example of extremely reactive system and such

detailed analysis cannot be performed on all chemicals. Therefore development of a

systematic approach, as shown in Figure 4.1, is emphasized.

Page 115: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

105

CHAPTER V

CONCLUSIONS

The field of reactive chemicals has attracted attention of researchers and industry

in the last few years in anticipation of possible regulations. This work is intended to

resolve some of the current issues and advance reactive hazard assessment.

A classification based on To and -∆H, obtained from calorimetric data, is

proposed to help recognize the more hazardous compositions. Previous researchers have

published correlations to predict explosive properties from calorimetric parameters

namely, To and -∆H. Using an additional parameter aspect ratio as an indicator of rate of

energy released, these correlations were refined. This has significantly improved

screening potential of DSC data for identifying explosion hazards.

A further advancement would be a computerized program for predicting DSC

properties, which can be developed into a screening tool. Based on experimental data on

aromatic mono nitro compounds, it was shown that calorimetrically determined

parameters can be predicted based on properties calculated at molecular level. A simple

potential energy surface with the first step of bond breaking as rate limiting was

proposed and onset temperatures for 19 mono nitro compounds were predicted with an

average error of 11%. A reduction in the predicted error to 6% was achieved by

correlating observed onset temperatures to molecular descriptors, similar to Quantitative

Structure Property Relationships (QSPR). Mary Kay O’Connor Process Safety Center

(MKOPSC) is currently collaborating with Dow Chemical Company and Eastman

Kodak to extend QSPR methodology to different families of compounds.

Page 116: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

106

These correlations can be effectively developed into an automated computerized

screening tool to screen out the compositions. However the more hazardous compounds

necessitate additional testing and resources. Detailed investigation of HA system, as an

example of highly hazardous system, is provided. Pure HA is reported to explode at

room temperature and is therefore not well characterized. Based on theoretical

calculations, HA heat of formation is estimated as –11.4 kcal/mol and is believed to be

within 1 kcal/mol of experimental value. Further a mechanism for HA decomposition is

proposed with the aim of understanding the instability with respect temperature and

contamination, and propose effective inhibitors for the runaway reaction. Finally, a QRA

study was performed to integrate available reactivity and manufacturing information for

risk assessment.

The area of reactive chemicals presents unique challenges and this work has

demonstrated the significance of molecular characterization of energetic materials.

Page 117: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

107

LITERATURE CITED

(1) Barton, J. A.; Nolan, P. F. Incidents in the chemical industry due to thermal runaway

chemical reactions. I ChemE Symposium Series No. 115 1989, 3.

(2) Etchells, J.C. Why reactions run-away. Organic Process Research and Development

1997,1, 435.

(3) U.S. Chemical Safety and Hazard Investigation Board (CSB) reactive hazard

investigation report Improving reactive hazards 2002. (Available on CSB website,

http://www.csb.gov/reports/)

(4) Toxic Catastrophe Prevention Act (TCPA) Program, Proposal number: PRN 2003-

76, New Jersey, 2003.

(5) Johnson, R. W.; Rudy, S. W.; Unwin, S. D. Essential Practices for Managing

Chemical Reactivity Hazards; AIChE Center for Chemical Process Safety (CCPS):

New York,2003.

(6) Grewer, T. Thermal Hazards of Chemical Reactions, Elsevier Science: Amsterdam,

Netherlands, 1994.

(7) Barton J.; Rogers, R. Chemical Reaction Hazards, Second edition, Institute of

Chemical Engineers: Rugby, UK, 1997.

(8) Iyer, S.; Slagg, N. Molecular Aspects in Energetic Materials. Molecular Structure

and Energetics, Structure and Reactivity, Chap. 7, edited by Liebman, J.;

Greenberg, A., VCH publishers: New York, 1988.

(9) United Nations, Recommendations on the Transport of Dangerous Goods, Manual of

Tests and Criteria, 3rd Revised Edition; UN, 1999.

(10) Meyer, R.; Kohler, J.; Homburg, A. Explosives, 5th edition, Wiley: New York,

2002.

(11) Sucesha, M. Test Methods for Explosives; Springer: New York, 1995.

(12) Whitmore, M.W.; Baker, G.P. Investigation of the use of a closed pressure vessel

test for estimating condensed phase explosive properties of organic compounds. J.

Loss Prev. Process Ind. 1999, 12(3), 207.

Page 118: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

108

(13) Townsend, D; Tou, J.C. Thermal hazard evaluation by an accelerating rate

calorimeter. Thermochimica Acta 1980, 37, 1.

(14) National Fire and Protection Association, NFPA 704. Standard System for the

Identification of the Hazards of Materials for Emergency Response. NFPA: Quincy,

MA, 2001.

(15) Saraf, S. R.; Rogers, W.J.; Mannan, M.S. Challenges in the Classification of

Reactive Chemicals. CEP 2004.(accepted for publication)

(16) Hoeflich, T.C.; LaBarge, M.S. On the use and misuse of detected onset temperature

of calorimetric experiments for reactive chemicals. Journal of Loss Prevention in

the Process Industries 2002, 15, 163.

(17) Ando, T.; Fujimato, Y.; Morisaki S. Analysis of differential scanning calorimetric

data for reactive chemicals. Journal of Hazardous Materials 1991, 28 (3), 251.

(18) CHETAHTM, Version 7.2, The ASTM Computer Program for Chemical

Thermodynamic and Energy Release Evaluation (NIST Special Database 16),

ASTM Subcommittee E27.07, ASTM, West Conshoshocken, PA, 1998.

(19) Yoshida, T.; Yoshizawa, F.; Itoh, M.; Matsunaga, T.; Watanabe, M.; Tamura, M.

Prediction of fire and explosion hazards of reactive chemicals. I. Estimation of

explosive properties of self-reactive chemicals from SC-DSC data. Kpgyo Kayaku

1987, Vol. 48, No. 5, 311.

(20)Saraf, S.R.; Rogers, W.J.; Mannan, M. S. Using screening test data to classify

reactive hazards. Mary Kay O’Connor Process Safety Center Symposium, College

Station, TX, Oct. 29-30, 2002, 613.

(21) Saraf, S.R.; Rogers, W.J.; Mannan, M. S; Chervin, S.; Bodman, G.T. Correlating

Explosive Properties to DSC Parameters. 31st Annual Conference of the North

American Thermal Analysis Society (NATAS) conference, Albuquerque, NM, Sept.

22 – 24, 2003.

Page 119: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

109

(22) Bodman, G.T. Use of DSC in screening of explosive properties. 30th North

American Thermal Analysis Society (NATAS) conference, Pittsburgh, PA, Sept. 23-

25, 2002, 605.

(23) Whiting, L.F.; Labean, M.S.; Eadie, S.S. Evaluation of a capillary tube sample

container for differential scanning calorimetry. Thermochim. Acta, 1988, 136, 231.

(24) United Nations Recommendations on the Transport of Dangerous Goods, Tests and

Criteria; 2nd Edition, UN, 1990, 10, 42.

(25) Lothrop, W.C.; Handrick, G.R. Relationship between performance and constitution

of pure organic explosive compounds. Chemical Reviews 1949, 44, 419.

(26) Melhem, G.A.; Shanley, E.S. On the estimation of hazard potential for chemical

substances. Process Safety Progress 1996, 15 (3), 69.

(27) Frurip, D.J.; Chakrabarti, A.; Downey, J.R.; Ferguson, H.D.; Gupta, S.K.; Hoeflich,

T.C.; LaBarge, M.S.; Pasztor, A.J.; Peerey, L.M.; Eveland, S.M. ; Suckow, R.A.;

Determination of chemical process heats by experiment and prediction.

International Symposium on Runaway Reactions and Pressure Relief Design,

(AIChE), Boston, MA, Aug. 2-4, 1995, 95.

(28) Urben, P.G. Bretherick’s Handbook of Reactive Chemicals; 5th edition,

Butterworth-Heineamann: London, UK, 1995.

(29) Melhem, G.A.; Shanley, E.S. The oxygen balance criterion for thermal hazards

assessment. Process Safety Progress 1995, 14 (1), 29.

(30) Benson, S.W. Thermochemical Kinetics: Methods for the Estimation of

Thermochemical Data and Rate Parameters; 2nd edition, Wiley: New York, 1976.

(31) Shanley, E.S.; Melhem, G.A. A review of ASTM CHETAH 7.0 hazard evaluation

criteria. J. Loss Preven. Process Ind. 1995, 8(5), 261.

(32) Houston, P. L. Chemical Kinetics and Reaction Dynamics; 1st edition, McGraw Hill

Science/Engineering/Math: New York, 2001.

(33) Becke 3 Perdew-Wang 86 (B3P86): Becke, A.D. Density functional

thermochemistry. III. The role of exact exchange J. Chem. Phys. 1993, 98, 5648.

Page 120: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

110

(34) Dunning, Jr., T.H. Gaussian basis sets for use in correlated molecular calculations.

I. The atoms Boron through Neon and Hydrogen. Chem. Phys. 1989, 90, 1007.

(35) Fogler, H.S. Elements of Chemical Reaction Engineering; Second edition,

Prentice – Hall Inc.: Upper Saddle River, NJ, 1992.

(36) Townsend, D.I., and Tou, J.C. Thermal hazard evaluation by an accelerating rate

calorimeter Thermochimica Acta 1980, 37, 1.

(37) National Institute of Standards and Technology, Chemistry Web Book,

http://webbook.nist.gov, Last accessed on 10/18/2003.

(38) Lide, D.R. CRC Handbook of Chemistry and Physics; 79th ed., CRC press: Boca

Raton, FL, 1998.

(39) Irikura, K.; Frurip, D.J. Computational Thermochemistry ACS Symposium Series

677, American Chemical Society, Washington DC, 1996.

(40) Pilling, M.J.; Seakins, P.W. Reaction Kinetics, 2nd edition, Oxford University

Press: New York, 1996.

(41) Evans, M. G.; Polanyi, M. Inertia and driving force of chemical reactions. Trans.

Faraday Soc. 1938, 34, 11.

(42) Masel, R. I. Chemical Kinetics and Catalysis; John Wiley and Sons: New York,

2001.

(43) Akutsu, Y.; Tamura, M. A Study on the Thermal Stability of Energetic Materials by

DSC and ARC. Proceedings of the International Pyrotechnics Seminar 1999, 26, 1.

(44) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;

Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A; Stratmann, R. E.; Burant,

J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.;

Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli,

C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.;

Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Cioslowski, J.; Ortiz, J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.;

Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.;

Page 121: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

111

Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.;

Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.;

Replogle, E. S.; Pople, J. A. Gaussian 98 (Revision A.9), Gaussian Inc., Pittsburgh

PA, 1998.

(45) Gustin, J. Runaway reaction hazards in processing organic nitro compounds.

Organic Process Research and Development 1998, 2, 27.

(46) Duh, Y.; Lee, C.; Hsu, C.; Hwang, D.; Kao, C. Chemical incompatibility of

nitrocompounds. Journal of Hazardous Materials 1997, 53, 183.

(47) Austin Model 1: Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. F.; Stewart, J.P.

Development and use of quantum mechanical molecular models. 76. AM1: a new

general purpose quantum mechanical molecular model. J. Amer. Chem. Soc. 1985,

107(13), 3902.

(48) Grewer, T. The influence of chemical structure on exothermic decomposition

Thermochimica Acta 1991, 187, 133.

(49) Young, D.C. Computational Chemistry: A Practical Guide for Applying Techniques

to Real World Problems. John Wiley and Sons: New York, 2001.

(50) Wessel; M.D.; Jurs, P.C. Prediction of normal boiling points for a diverse set of

industrially important organic compounds from molecular structure. J. Chem. Inf.

Comput. Sci. 1995, 35(5), 841.

(51) Mitchell, B.E.; Jurs, P.C. Prediction of autoignition temperatures of organic

compounds from molecular structure. J. Chem. Inf. Comput. Sci. 1997, 37(3), 538.

(52) Fukui, K. Role of frontier orbitals in chemical reactions. Science 1982, 218, 4574.

(53) Fleming, I. Frontier Orbitals and Organic Reactions. Wiley: London, 1976.

(54) Politzer, P.; Murray, J.S. Relationship between dissociation energies and

electrostatic potentials of C-NO2 bonds: application to impact sensitivities. J. Mol.

Structure 1996, 376, 419.

(55) Hosoya, H.; Iwata, S. Revisiting superdelocalizabiltiy. Mathematical stability of

reactivity indices. Theor. Chem. Acc. 1999, 102, 293.

Page 122: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

112

(56) Fujimoto, H; Yamabe, S; Minato, T.; Fukui, K. Molecular orbital calculation of

chemically interacting systems. Interaction between radical and closed-shell

molecules. JACS 1972, 94(26), 9205.

(57) The SAS System for Windows, Release 8.01, copyright(c) 1999-2000 by SAS

Institute Inc., Cary, NC, USA.

(58) Montgomery, D.C.; Runger, G.C. Applied Statistics and Probability for Engineers.

2nd Edition, John Wiley and Sons: New York, 1999.

(59) Chervin, S.; Bodman, G.T. Method for estimating decomposition characteristics of

energetic chemicals. 29th North American Thermal Analysis Society (NATAS), St.

Louis, Missouri, Sept 24-26, 2001.

(60) Badders, N. R.; Mannan, M.S. Predicting impact sensitivities of energetic materials.

Summer report submitted to Chemical Eng. Dept., Texas A&M University, 2003.

(61) Melhem, G.A. Polymerization modeling for relief system design. DIERS User

Group, Philadelphia, PA, 28 - 30 April, 2003.

(62) BUSINESS CONCENTRATES: Chemical & Engineering News 1999, 77(9), 11.

(63) BUSINESS CONCENTRATES: Chemical & Engineering News 2000, 78(25), 15.

(64) Cisneros, L.O.; Rogers, W.J.; Mannan, M.S. Adiabatic calorimetric decomposition

studies of 50 wt.% hydroxylamine/water. J. Hazard. Mater. 2001, 82, 1.

(65) Gurvich, L.V.; Veyts, I.V.; Alcock, C.B. Thermodynamic Properties of Individual

Substances. Vol 1 and 2, 4th edition, Hemisphere Publishing Corporation: New

York, 1989.

(66) Back, R.A.; Betts, J. The determination of the saturation vapor pressure of solid

hydroxylamine using a piston pressure gauge. Canadian J. Chem. 1965, 43, 2157.

In this reference, the authors used a value for the heat of formation of solid HA

(-25.5 kcal/mol) attributed to reference (66 a), to obtain -10.2 kcal/mol for the

standard heat of formation of gaseous HA. Here, the authors use their HA vapor

pressure data to calculate 15.3 kcal/mol for the enthalpy of sublimation, and they

extrapolate the vapor pressure data to obtain 11.4 kcal/mol for the enthalpy of

Page 123: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

113

vaporization. However, -25.5 kcal/mol is listed in reference (66 a) as the heat of

formation of liquid HA and not solid HA.

(67) a. Bichowsky, F.R.; Rossini, F.D. Thermochemistry of Chemical Substances;

Reinhold Publishing Corporation: New York, 1936.

b. Thomsen, Thermochemische Untersuchungen; Vols. I – IV: Barth, Leipzig,

1882 - 1886.

c. Berthelot and Matignon, Ann. Chim. Phy. 1892, 27, 289.

(68) Bailer Jr., J.C.; Emeleus, H.J.; Nyholm, R.; Trotman-Dickenson, A.F.

Comprehensive Inorganic Chemistry; Vol 2, Ch. 19, 1st ed., Pergamon Press: UK,

1973.

(69) Sana, M.; Leroy, G. Theoretical thermochemistry. The enthalpies of formation of

some XHn and XYHn compounds. THEOCHEM 1991, 72(3-4), 307.

(70) Anderson, W.R. Heats of formation of HNO and some related species. Combustion

and Flame 1999, 117, 394.

(71) Wiberg, K.B. Bond dissociation energies of H2NX compounds: comparison with

CH3X, HOX, and FX compounds. J. Phys. Chem. 1992, 96, 5800.

(72) Kutina, R.E.; Goodman, G.L.; Berkowitz, J. Photoionization mass spectrometry of

hydroxylamine: heats of formation of nitroxyl ion (HNO+ and NOH+). J. Chem.

Phy. 1982, 77, 1664.

(73) Benson, S.W., and Buss, J.H. Additivity rules for the estimation of molecular

properties. Thermodynamic properties. J. Chem. Phy. 1958, 29, 546.

(74) Becke 3 Lee, Yang, and Parr (B3LYP):

a. Becke, A. D. Density-functional thermochemistry. III. The role of exact

exchange. J. Chem. Phys. 1993, 98, 5648.

b. Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-

energy formula into a functional of the electron density. Physical Review 1988, B

37, 785.

Page 124: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

114

(75) Gaussian 2 (G2) : Curtiss, L. A.; Raghavachari, K.; Trucks, G. W.; Pople, J. A.

Gaussian-2 theory for molecular energies of first- and second-row compounds. J.

Chem. Phys. 1991, 94, 7221.

(76) Gaussian 3 (G3): Curtiss, L.A; Raghavachari, K.; Redfern, P.C.; Rassolov, V.;

Pople, J.A. Gaussian-3 (G3) theory for molecules containing first and second-row

atoms. J. Chem. Phys. 1998, 109(18), 7764.

(77) Gaussian-2, second order variant (G2MP2): Curtiss, L. A.; Raghavachari, K.;

Pople, J. A. Gaussian-2 theory using reduced Moeller-Plesset orders. J. Chem. Phys.

1993, 98(2), 1293.

(78) Gaussian-2, second order variant Becke’s 3 (G2MP2B3), Gaussian-3 Becke’s 3

(G3B3): Baboul, A.G.; Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-3

theory using density functional geometries and zero-point energies. J. Chem. Phys.

1999, 110, 7650.

(79) Complete Basis Set – Q(CBS-Q) : Petersson, G. A.; Bennett, A.; Tensfeldt, T. G.;

Al-Laham, M. A.; Shirley, W. A.; Mantzaris, J. A complete basis set model

chemistry. I. The total energies of closed-shell atoms and hydrides of the first-row

elements. J. Chem. Phys. 1988, 89, 2193.

(80) Second Order MØller-Plesset perturbation theory (MP2): Gordon, M.H.; Pople, J.

A.; Frisch, M. J. MP2 energy evaluation by direct methods. Chem. Phys. Lett. 1988,

153(6), 503.

(81) Third Order MØller-Plesset perturbation theory (MP3) : Pople, J. A.; Seeger, R.;

Krishnan, R. Variational configuration interaction methods and comparison with

perturbation theory. Int. J. Quant. Chem. Symp. 1977, 11, 149.

(82) Fourth Order MØller-Plesset perturbation theory (MP4 (SDTQ) ): Krishnan, R.;

Pople, J. A. Approximate fourth-order perturbation theory of the electron

correlation energy. Int. J. Quant. Chem. 1978, 14(1), 91.

Page 125: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

115

(83) Coupled-cluster, singles and doubles (CCSD): Pople, J. A.; Krishnan, R.; Schlegel,

H. B.; Binkley, J. S. Electron correlation theories and their application to the study

of simple reaction potential surfaces. Int. J. Quant. Chem. 1978, 14(5), 545.

(84) Coupled-cluster, singles and doubles with perturbative triples CCSD(T): Pople, J.

A.; Gordon, M.H.; Raghavachari, K. Quadratic configuration interaction. A general

technique for determining electron correlation energies. J. Chem. Phys. 1987, 87,

5968.

(85) Quadratic configuration interaction, singles and doubles with perturbative triples

QCISD(T): Gauss, J.; Cramer, D. Analytical evaluation of energetic gradients in

quadratic configuration interaction theory. Chem. Phys. Lett. 1988, 150, 280.

(86) Kendall, R. A.; Dunning, Jr., T. H.; Harrison, R. J., Harrison, R. J. Electron

affinities of the first-row atoms revisited. Systematic basis sets and wave functions.

J. Chem. Phys. 1992, 96(9), 6796.

(87) Woon, D.E.; Dunning, Jr., T. H. Gaussian basis sets for use in correlated molecular

calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98(2),

1358.

(88) Peterson, K.A.; Woon, D. E.; Dunning Jr., T. H. Benchmark calculations with

correlated molecular wave functions. IV. The classical barrier height of the H + H2

→ H2 + H reaction. J. Chem. Phys. 1994, 100(10), 7410.

(89) Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods.

XII. Further extensions of Gaussian-type basis sets for use in molecular orbital

studies of organic molecules. J. Chem. Phys. 1972, 56, 2257.

(90) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-consistent molecular

orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys.

1980, 72, 650.

(91) Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. v. R. Efficient diffuse

function-augmented basis sets for anion calculations. III. The 3-21 + G basis set

for first-row elements, lithium to fluorine. J. Comp. Chem. 1983, 4, 294.

Page 126: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

116

(92) Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-consistent molecular orbital methods.

25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80(7),

3265.

(93) Melius, C.F., and Zachariah, M.R. Computational Thermochemistry, ACS

Symposium Series 677, Chapter 9, American Chemical Society: Washington, D.C.,

1996.

(94) a. Hinchliffe, A. Modeling Molecular Structures, 2nd edition; Wiley: New York,

2000.

b. Hehre, W.J.; Radom, L.; Schleyer, P.; Pople, J.A. Ab initio Molecular Orbital

Theory, 1st Edition; Wiley: New York, 1986.

(95) Peterson, G.A. Computational Thermochemistry, ACS Symposium Series 677,

American Chemical Society: Washington, D.C., Chapter 13, 1996.

(96) Cisneros, L.O. 2002 Adiabatic Calorimetric Studies of Hydroxylamine Compounds.

PhD Dissertation, Texas A&M University, College Station, TX.

(97) Dolg, M.; Wedlig, U.; Stoll, H.; Preuss, H. Energy-adjusted ab initio

psudopotentials for the first row transition elements. J. Chem. Phy. 1987, 86(2),

866.

(98) Krishna, K.; Wang, Y.J.; Saraf, S.R.; Baldwin, J.T.; Rogers, W.J.; Gupta, J.P.;

Mannan, M. S. Hydroxylamine production: Will a QRA help you decide?.

Reliability Engineering & System Safety 2003, 81, 215.

Page 127: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

117

APPENDIX A. DISCUSSION OF MOLECULAR MODELING METHODS

Molecular modeling techniques can be broadly classified into electronic structure

theories and force-field based methods. Electronic structure methods use the laws of

quantum mechanics to calculate geometries, energies, and other related properties of a

molecule. For the work in this dissertation, electronic structure or quantum chemistry

calculations were employed for obtaining information at molecular level. The

application of quantum mechanical laws to a system involves approximate numerical

solution to Schrödinger equation. Thus the effectiveness of the calculations depends on

the approximation involved and is a balance between computational resources and

desired accuracy. Generally, a theoretical calculation requires specification of level of

theory (approximations involved in solving Schrödinger equation) and

basis set (mathematical representation of molecular orbitals). It can also be argued that a

combination of level of theory and basis set represents approximation to Schrödinger

equation. The chart below illustrates various methods:

Page 128: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

118

Table 1. Various theoretical methods

HF MP2

MP3 MP4 QCISD(T) .. .. Full CI

Minimal STO-3G

Split valence 3-21 G

Polarized 6-31 G

6-311 G (d,p) Diffuse 6-311+G (d,p)

6-311+G(2d,p) 6-311++G(3df,3pd) .. …

Bas

is s

et

∞ HF limit

Schrödinger Equation

The columns correspond to various theoretical methods and the rows to different basis

sets. With increasing basis set and level of theory, one approaches the exact solution to

Schrödinger equation and is accompanied by an increase in computational time. In

general, computational cost and accuracy increases as you move to the right or below in

the above chart.

There are commercial programs available for theoretical calculations and a few

of them are listed below:

§ Gaussian Inc. (http://www.gaussian.com)

§ Schrödinger Inc. (http://www.schrodinger.com/)

§ Accelrys (http://www.accelrys.com/)

Page 129: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

119

APPENDIX B. DESCRIPTORS CALCULATED WITH B3P86/cc-pVDZ

HOMO LUMO HPC HNC WB Vmid Sr Dipole (HPC+HNC)/WB

a.u a.u kcal /mol 1/Ao 1/hartree mol/kcal

1 nitrobenzene -0.38818 -0.03931 0.519787 -0.35166 27.63 0.5249 -35.9097 5.2426 0.0061

2 1,2 dinitrobenzene -0.41432 -0.07306 0.515704 -0.35836 18.09 0.6140 -46.078 7.5324 0.0087

3 1,3 dinitrobenzene -0.41999 -0.07024 0.525467 -0.33995 26.93 0.5235 -45.5774 4.8378 0.0069

4 1,4 dinitrobenzene -0.41691 -0.08116 0.518757 -0.33439 26.80 0.5629 -45.5071 0.0014 0.0069

5 2 nitrotoluene -0.37379 -0.03716 0.520176 -0.3279 26.22 0.5247 -41.1747 5.0368 0.0073

6 3 nitrotoluene -0.37486 -0.03736 0.51954 -0.35262 27.60 0.5380 -41.0261 5.4371 0.0060

7 4 nitrotoluene -0.37868 -0.0384 0.521164 -0.3537 28.54 0.5118 -41.0197 5.7281 0.0059

8 2,6 dinitrotoluene -0.40231 -0.06426 0.522868 -0.35192 21.21 0.5094 -50.7843 3.6738 0.0081

9 3,4 dinitrotoluene -0.40235 -0.07084 0.517086 -0.36083 18.03 0.6272 -51.1144 7.8496 0.0087

10 2,4dinitrotluene -0.40535 -0.06766 0.526903 -0.34472 23.29 0.5102 -50.7286 5.4311 0.0078

11 2 nitroaniline -0.33332 -0.02919 0.533234 -0.53735 30.41 0.6039 -41.038 5.2902 -0.0001

12 3 nitroaniline -0.33065 -0.03154 0.51562 -0.56357 26.77 0.5118 -41.0257 5.4371 -0.0018

13 4 nitroaniline -0.3366 -0.0259 0.530443 -0.55023 30.78 0.4469 -41.0743 5.7281 -0.0006

14 2 nitrobenzoic acid -0.40089 -0.05264 0.524503 -0.36507 22.90 0.5965 -46.87 5.31 0.0070

15 3 nitrobenzoic acid -0.39561 -0.05653 0.480229 -0.33915 26.39 0.4975 -46.6519 3.318 0.0053

16 4 nitrobenzoic acid -0.40046 -0.0636 0.517231 -0.37614 27.31 0.5636 -46.5918 3.4293 0.0052

17 2nitrophenol -0.36422 -0.04351 0.531201 -0.39326 30.26 0.4201 -39.7609 4.3342 0.0046

18 3nitrophenol -0.36623 -0.04285 0.51869 -0.35112 26.48 0.5776 -39.8012 4.0073 0.0063

19 4nitrophenol -0.37015 -0.03914 0.526489 -0.35643 28.29 0.4857 -39.8302 5.2641 0.0060

Page 130: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

120

APPENDIX C. AM1 DESCRIPTORS AND PREDICTED ONSET

TEMPERATURES

HOMO LUMO HPC HNC WB Vmid Sr Dipole AM1

predicted To

a.u a.u kcal /mol 1/Ao 1/hartree oC

1 nitrobenzene -0.30771 -0.11325 0.162701 -0.24179 75.63 0.3398 46.6603 4.3264 317

2 1,2 dinitrobenzene -0.31636 -0.13446 0.167717 -0.22571 64.50 0.3941 59.6145 6.3759 271

3 1,3 dinitrobenzene -0.3354 -0.13839 0.162783 -0.23217 73.15 0.3661 59.2751 4.0095 291

4 1,4 dinitrobenzene -0.33376 -0.15103 0.162391 -0.22926 72.92 0.3793 59.1759 0.0002 324

5 2 nitrotoluene -0.2949 -0.10907 0.152932 -0.25153 73.41 0.3259 53.4679 4.0731 303

6 3 nitrotoluene -0.29441 -0.11029 0.161248 -0.24312 75.86 0.3533 53.3280 4.6931 300

7 4 nitrotoluene -0.29757 -0.10866 0.159664 -0.24536 76.68 0.3545 53.3267 5.0266 298

8 2,6 dinitrotoluene -0.31534 -0.12849 0.146081 -0.23224 68.25 0.3515 66.0496 2.7574 283

9 3,4 dinitrotoluene -0.30856 -0.1297 0.164868 -0.22902 64.69 0.3926 66.2135 7.0637 253

10 2,4dinitrotluene -0.32353 -0.13239 0.158883 -0.24237 74.13 0.3671 65.9801 4.6759 271

11 2 nitroaniline -0.24817 -0.10364 0.162259 -0.30019 80.09 0.4423 53.6182 4.4874 301

12 3 nitroaniline -0.24378 -0.10277 0.159222 -0.24785 76.50 0.3530 53.8718 5.9096 300

13 4 nitroaniline -0.25054 -0.09118 0.109142 -0.26109 80.86 0.3305 54.0180 7.4857 298

14 2 nitrobenzoic

acid -0.31702 -0.13099 0.18627 -0.2355 66.41 0.3670 63.9692 3.7116 284

15 3 nitrobenzoic

acid -0.3187 -0.12183 0.229646 -0.23905 74.65 0.3530 61.0638 2.5373 298

16 4 nitrobenzoic

acid -0.31553 -0.13874 0.228782 -0.2358 74.12 0.3663 60.9865 3.5715 297

17 2nitrophenol -0.27598 -0.12394 0.180617 -0.29723 82.74 0.4017 51.5800 3.3382 312

18 3nitrophenol -0.25239 -0.11173 0.15679 -0.25364 75.75 0.3532 52.1654 3.4879 314

19 4nitrophenol -0.28044 -0.10407 0.163891 -0.25138 78.33 0.3422 52.2803 5.2343 305

Page 131: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

121

APPENDIX D. SUMMARY OF BOND LENGTHS

NH2OH H2O2

Method Basis Set N-O (Å) O-O (Å)

AM1 1.342 1.303

HF cc-pVDZ 1.400 1.393

cc-pVTZ 1.398 1.387

cc-pVQZ 1.396 1.384

B3P86 6-311+G(3df,2p) 1.426 1.430

6-31+G(3df,2p) 1.428 1.431

6-311++G(3df,2p) 1.426 1.430

6-311G* 1.430 1.436

cc-pVDZ 1.441 1.437

cc-pVTZ 1.432 1.435

cc-pVQZ 1.428 1.432

cc-pV5Z 1.427 1.432

Aug-cc-pVDZ 1.415 1.435

Aug-cc-pVTZ 1.430 1.435

B3LYP 6-311+G(3df,2p) 1.441 1.446

cc-pVDZ 1.415 1.453

cc-pVTZ 1.447 1.452

Aug-cc-pVDZ 1.415 1.451

Page 132: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

122

Aug-cc-pVTZ 1.445 1.451

MP2 cc-pVDZ 1.444 1.457

cc-pVTZ 1.442 1.451

cc-pVQZ 1.437 1.457

MP4 6-31+G(3df,2p) 1.449 1.465

CCSD(T) cc-pVDZ 1.454 1.470

CCSD(T) cc-pVDZ 1.454 1.470

G2 1.451 1.468

G2MP2 1.451 1.468

CBS-Q 1.438 1.453

G3 1.451 1.468

G3MP2B3 1.448 1.456

G3B3 1.448 1.456

BAC-MP4 1.404 1.396

Experimental 1.45338 1.47538

Page 133: MOLECULAR CHARACTERIZATION OF ENERGETIC MATERIALS

123

VITA

Sanjeev R. Saraf was born on 6th Feb. 1977 in Nagpur, India. He received his

Bachelor of Science degree in chemical engineering from University Department of

Chemical Technology (U.D.C.T), Mumbai in May 1999. He joined the Chemical

Engineering Department at Texas A&M University to pursue his Ph.D. in Aug. ’99. His

permanent address is 3 Savarkar Nagar, Khamla Road, Nagpur - 440015, India.