Computational Analysis of Nitramine Derivatives &

Carbon Analogs as Potential HEDMs

 

Daniel LaMaster

Abstract

            The purpose of this project is to determine the thermochemical properties of the molecules of interest in order to evaluate them as explosive materials and determine if they meet the requirements to be classified as high energy-density materials (HEDMs) while comparing the carbon and nitrogen skeletons.  The six molecules of interest are trinitramide (N1), tetranitrotetrazetidine (N2), hexanitrohexazinane (N3), trinitromethane (C1), 1,2,3,4-tetranitrocyclobutane (C2), & 1,2,3,4,5,6-hexanitrocyclohexane (C3).  The Gaussian program is selected to run the computations for this project due to its popularity and use by experts in this field.2-4,16  Calculations are performed with the Density Functional Theory at the B3LYP level using the 6-311G** basis set as previous studies have found satisfactory results with this theory and basis set .2-4,8,9,16  Formation enthalpies and other thermal properties were obtained from the computational output.  The oxygen balance was used to determine the detonation products, and the heats of detonation were used to evaluate the detonation performances of the molecules.

 

1. Introduction. 

 

            Computational methods are procedures that use simple algorithms to evaluate complex problems and analyze large data sets.  They employ computers to reduce the expense of lengthy calculations, increase the reliability of analyzing various sizes of data sets, and provide easier accessibility by graphing large data sets.  The use of computational methods began when advancements in mathematics, and its various application fields produced more complex formulas to accurately describe the various phenomena.  Today, computational methods are used in a variety of fields including both the natural and social sciences as well as mathematics.  A few of the numerous applications are drug design, weather forecasting, evaluating proposed syntheses, protein or DNA/RNA sequencing and sequence comparisons, molecular modeling, and calculating atomic and molecular properties using quantum mechanics.  The accuracy of the computed values based on quantum mechanics compared to the experimental values depends on the level of theory and basis set used and varies from program to program for a property of interest.  In general, the freely available programs are accurate enough to provide qualitative data at relatively low levels of theory, but higher levels of theory are required to obtain quantitative data and not all of the programs have this capability.  In general, the commercially available programs seem to be more comprehensive and capable than the freeware programs.  

 

            Quantum mechanical computational methods are used in this project to investigate the explosive properties of three nitramine based molecules and their carbon skeleton analogs.  The principal interests of this research were to determine if any of the molecules fit the criteria of high-energy density materials, have potential as a propellant or an explosive, and whether a carbon or nitrogen skeleton yields better performance. 

 

            An explosion is a rapid and violent release of mechanical, chemical, or nuclear energy accompanied by a physical, and usually audible, shockwave, the generation of high temperatures, the release of gases, and sometimes light resulting from a rapid physical, chemical, or nuclear change.12,13  There are three types of explosions: physical, chemical, and atomic.  Of these three types, this research project involves chemical explosions.  A chemical explosion is a chemical reaction or change of state which occurs over an exceedingly short frame of time with the generation of a large amount of heat and generally a large quantity of gas.  Based on their performances and properties, chemical explosives are separated into three groups: primary explosives, secondary explosives, and propellants.1,10  Materials within the primary and secondary categories can also be classified as high energy-density materials (HEDMs) if they, ideally, have a density of 1.9 g cm-3, a detonation velocity of 9.0 km s-1, and a detonation pressure of 40.0 GPa (approximately).2 

 

            Primary explosives are quite sensitive to shock, friction, electric sparks, and high temperatures; consequently, they will detonate when subjected to any stress.  Upon detonation, the material’s molecules dissociate and produce a tremendous amount of heat and shock which, in turn, initiates a second more stable explosive.  In other words, they rapidly transition to detonation from burning and are capable of transmitting the detonation wave to less sensitive explosives (i.e. secondary explosives).  As a result of their capacity to initiate a secondary explosive in addition to their sensitive nature, their ability to explode regardless of confinement, and a general detonation velocity range of 3,500 to 5,500 m s-1, primary explosives are frequently used in ‘initiating devices’.1,10 

 

            Materials that fall into the secondary explosives (a.k.a. high explosives) category cannot be readily detonated by heat or shock and are generally more powerful than primary explosives.  The secondary explosives can only be detonated by the shock produced from the explosion of a primary explosive.  Upon initiation, they almost instantaneously dissociate into more stable components and detonate with a general velocity range of 5,500 to 9,000 m s-1.1,10 

 

            The third class of chemical explosives are propellants which are combustible materials containing enough oxygen to achieve complete combustion.  While they can be initiated by a spark or flame just like primary explosives, the key differences between propellants and the other explosives are that they burn but do not explode and change from a solid to a gaseous state relatively slowly (a few milliseconds compared to a millisecond or less).1,10 

 

            When an explosive reaction takes place, the explosive material atomizes in an extremely exothermic process and at an extremely high rate.  The atomization products immediately assemble into small stable molecules; generally forming carbon monoxide, carbon dioxide, water, and nitrogen gas.  The products of the decomposition reaction are determined by the material’s oxygen content which is described by its oxygen balance (%OB).  This quantity describes the amount of oxygen, in weight percent, liberated from the material by complete oxidation of the atomization products.1,5,10  If the %OB is greater than or equal to zero, the explosive undergoes complete combustion and the heat released during the formation of the gas products is the heat of explosion; however, if the %OB is less than zero, incomplete combustion takes place giving rise to products like carbon monoxide, solid carbon, and hydrogen gas with a less amount of energy released as the heat of explosion. 

 

            The material’s detonation reaction follows one of the two following sets of rules: the Kistiakowsky-Wilson Rules (K-W rules) and the Modified Kistiakowsky-Wilson Rules.1,10,25 

 

Table 1: Detonation Reaction Rules1,10,25

Kistiakowsky-Wilson Rules

Modified Kistiakowsky-Wilson Rules

Rule

Condition

Rule

Condition

1

Carbon atoms are converted to carbon monoxide

1

Hydrogen atoms are converted to water

2

If any oxygen remains then hydrogen is then oxidized to water

2

If any oxygen remains then carbon is converted to carbon monoxide

3

If any oxygen still remains then carbon monoxide is oxidized to carbon dioxide

3

If any oxygen still remains then carbon monoxide is oxidized to carbon dioxide

4

All the nitrogen is converted to nitrogen gas, N2

4

All the nitrogen is converted to nitrogen gas, N2

 

            The set of rules used is determined by the material’s %OB.  The K-W rules are used with materials ranging from oxygen abundant to moderately oxygen deficient (an %OB greater than – 40.0), while the Modified K-W rules are used for severely oxygen deficient materials (an %OB less than – 40.0).1-3,5,16,25

 

            The properties essential for the evaluation of the theoretical compounds are the detonation velocity, detonation pressure, and relative stability.  Because the reaction occurs in approximately 0.01 s and gases cannot instantly expand, there is a fraction of a second where the gas remains inside of the container with constant volume.  The extreme heat produced by the explosion and the small volume of the gaseous products yield a pressure great enough to create a shock/blast wave.1,10  The rate at which this wave propagates is called the detonation velocity.6  The detonation pressure is the pressure produced by the reaction in the fraction of a second before the volume changes.20-23  The relative stability is the compound’s stability based on the calculated Gibbs free energy of formation of the compound relative to that of a reference compound which is known to be stable.  The determination of these properties depends on a number of parameters including theoretical density, Gibbs free energy of formation, zero-point energy, enthalpies of formation and detonation, molecular volume, and oxygen balance.2-5,8,9,15 

 

            In this project, the theoretical parameters and properties essential for evaluation of explosive power are determined using the well-established equations and computed molecular properties for the compounds shown in Figure 1.

 

            This project was selected after inquiring about the structure of 1,3,5-trinitroperhydro-1,3,5-triazine (a.k.a. RDX), the explosive compound in the military grade plastic explosive C-4.  Because of the computational expense and the limited computing power available, a six-membered ring was determined to be the largest molecule to be studied.  Due to the amount of time spent attempting to initially build RDX, it was decided that the study would be limited to three molecule sizes.  With RDX being composed of methylene and nitramine groups, a fully nitrated nitramine was selected as the smallest aza molecule.  Four-membered rings were chosen to be the third group of molecules intermediate in size between the nitramine and the six-membered ring molecules. 

 

Group 1

Group 2

Group 3

Trinitramide

N1

Tetranitrotetrazetidine

N2

Hexanitrohexazinane

N3

Trinitromethane

C1

1,2,3,4-Tetranitrocyclobutane

C2

1,2,3,4,5,6-Hexanitrocyclohexane

C3

Nitromethane

R1

1,3,3-Trinitroazetidine

R2

1,3,5-Trinitroperhydro-1,3,5-triazine

R3

Figure 1: Structures of Cyclic Nitramide Derivatives (N), Carbon-Skeleton Analogues (C), and Similar Skeletal References (R).

 

2. Computational Methods

 

            For this project, a number of calculations were executed including Geometry Optimization2-4, Bond Order2,4,26, Monte Carlo integration2-4,16, Frequency (thermochemical analysis) and Vibrational Analysis2,3,26.  A variety of computational methods have been used for these calculations by other researchers, but the density functional theory (DFT) at the B3LYP level2-4,8,9 (DFT-B3LYP), and the quantum chemical composite methods G2 (Gaussian 2) & CBS-Q (Complete Basis Set Q)2 seem to be the most common.  Other researchers have successfully used the basis sets 6-31G(d)3,16 and 6-311G(d,p)2-4,8,9 for geometry and thermodynamic data.  Computational engines such as Gaussian2-4, LOTUSES (Linear Output Thermodynamic User-friendly Software for Energetic Systems)5, ChemBasis3D, and MOPAC are commonly used for this type of research. 

 

2.1  Geometry Optimization

 

            All geometry optimizations were carried out at the DFT–B3LYP level using the STO-3G, 3-21G, & 6-31G(d) basis sets sequentially.  The last job executed for each final structure was a DFT–B3LYP optimization at 6-311G(d,p).  To simplify and shorten the process of building the molecules, an appropriate base structure was modified stepwise to construct the target molecule.  For example, when converting hydrogen atoms of the base structure into amino groups, the hydrogen atom was changed to a nitrogen atom and the ‘Add Hydrogens’ tool was used to add the hydrogen atoms and form the amino group.  Conversion of amino groups into nitro groups was done by changing the hydrogens to oxygens and creating the double bonds.  When carbon atoms were converted to nitrogen atoms, the equatorial hydrogen was deleted and when a carbon‘s hydrogen was changed into a nitrogen, the axial hydrogen was changed to nitrogen. 

 

2.1.a Group 1 Molecules. 

 

            Starting with the central atom, the base molecule was created using the ‘Add Hydrogens’ button.  The appropriate amount of hydrogen atoms was then changed into nitro groups. 

 

2.1.b Group 2 Molecules.

 

            Cyclobutane was imported from WebMO's structure library.  The appropriate carbon atoms were changed into nitrogens and the structures were optimized.  The appropriate hydrogens were then changed into nitrogen atoms and the average central atom to NO2 bond parameters of the respective Group 1 molecule were used to make the nitro groups.  Before the structures’ final optimization, the dihedral angles formed by the two ring atoms and the N=O for each nitro group were balanced by setting them equal to the average of their absolute magnitudes with the appropriate sign.

 

2.1.c Group 3 Molecules.

 

            Cyclohexane was imported from WebMO’s structure library.  After each alteration, the structures were optimized.  First, the appropriate carbon atoms were changed into nitrogens, then the hydrogens on three non-adjacent atoms in the rings were changed into amino groups.  Next, the structural parameters of the amino groups were averaged for each molecule and used to create hexamines.  In the last two steps, the hexamines were converted into 1,3,5-trinitro-2,4,6-triamines using the structural parameters of the nitro groups in the respective Group 2 molecules followed by conversion to hexanitrated rings.  RDX (1,3,5-trinitroperhydro-1,3,5-triazine) was built via the repetitive conversion of a ring nitrogen's hydrogen atom into an amino group and use of the structural parameters of the N-NO2 group in TNAZ (1,3,3-trinitroazetidine).

 

2.2  Bond Order

 

            All bond orders were calculated at the DFT B3LYP level with the 6-31G(d) basis set.  This calculation predicted which bond in the molecule is the weakest and would be severed first.  This information was then used to eventually evaluate the relative thermal stability of the molecules via their bond dissociation energy (BDE).  The BDE is defined as the reaction enthalpy of the bond scission reaction at 1 atm and 298 K.  The general formula for the bond scission reaction of the molecule A–B is as follows:

 

 

A–B(g) → A∙(g) + B∙(g)

1

 

The enthalpy of this reaction is found using equation 2.

 

 

2

 

where  is the formation enthalpy of the radical A∙,  is the formation enthalpy of the radical B∙, and  is the formation enthalpy of the molecule.  These enthalpy values were calculated by the Thermochemistry calculation in the MOPAC engine of WebMO using the PM3 (Parametric Basis Set 3) basis set which has been found to produce acceptable thermodynamic data.2,3,4,26 

 

2.3  Monte Carlo Integration

 

            The Volume job type uses a Monte Carlo integration to calculate the molecular volume of the isolated gas phase molecule, which is defined as the space enclosed within the 0.001 electrons/bohr3 electronic isodensity surface.  The volume calculated by Gaussian is only accurate to two significant figures, but by using the option Tight with the Volume calculation, an increased density of points is used for the integration which increases the accuracy to ~10%.  As a result, the generally used method takes the average of 100 single point calculations.2,3,4,16,20,27

 

2.4  Frequency and Vibrational Analysis

 

            The Vibrational Analysis was used to determine if the optimized structure was at a local minimum or at a global minimum on the molecule’s potential energy surface.  Negative frequencies found by the vibrational analysis indicated that the structure was at a local minimum; If these were found, adjustments would have to be made to the structure followed by another optimization and vibrational analysis.

 

            The Frequency job was then used to calculate the thermochemical properties of the molecules.  The quantities used to evaluate HEDMs were then calculated from the output data of this job.

 

            After geometry optimization using DFT-B3LYP/6-311G(d,p) and a vibrational analysis to ensure that the optimized structures are at the global energy minimum on the potential energy surface2-4, the HEDM parameters are calculated.  The results are evaluated to classify the molecules of interest into one of the three categories of chemical explosives and their properties were considered to determine whether or not they qualify as HEDMs.

 

3. HEDM Parameters

 

            The properties used to evaluate the theoretical materials are the theoretical density, detonation velocity, and detonation pressure as well as the relative thermal stability.  Figure 2 shows that the heat of formation is the most important quantity used in the evaluation, but the evaluation depends on a number of other parameters including total molecular energy, zero-point energy, enthalpy and detonation, molecular volume, and oxygen balance.2-5,8,9,15

 


Figure 2: HEDM Parameter Relations18

 


 

3.1  Heat of Formation

 

3.1.a Methods

 

            The heats of formation of reference compounds (RDX, TNAZ, and nitromethane) were calculated and used to find the heats of detonation.  Several different methods can be used to determine a theoretical molecule’s formation enthalpy.  These methods use different types of reactions that aid in the cancellation of systematic errors that arise in quantum mechanical computations.  Atomization reactions compare the absolute energy of the molecule to the absolute energy of its constituent atoms with known ∆Hf.  Isodesmic reactions conserve the number of types of bonds.  Homodesmic reactions conserve both the number of types of bonds and the hybridization of the atoms.17  Isogyric reactions conserve the number of electron pairs.19 

 

            It is acknowledged that the use of homodesmic reactions would yield more accurate results.  However, since the procedure for their use from the output data was unclear, the atomization method, described by the White Papers and Technical Notes provided by Gaussian Inc.29, was used. 

 

3.1.b Atomization Method Calculations

 

            The first step was to calculate the atomization energy of the molecule, , by the following equation for a molecule composed of x many atoms of X:

 

 

3

 

where  is the molecule’s total electronic energy,  is the molecule’s zero-point energy, and  is the total electronic energy of the atoms. 

 

            In the next step, the formation enthalpy of the molecule at 0 K, , was calculated by combining the formation enthalpy of the atoms at zero Kelvin, , with the molecule’s atomization energy.

 

 

4

 

            Finally, the formation enthalpy of the molecule at 298 K, , was calculated using equation 5:

 

5

 

where  is the correction to the thermal enthalpy of the molecule from 0 to 298 K, and  is the thermal correction to the enthalpy of the atoms from 0 to 298 K.

 

3.2 Oxygen Balance

 

            A material’s oxygen balance describes the amount of oxygen, in weight percent, liberated from the material by complete oxidation of the atomization products1,5,10,25 and is an important criteria used in classifying potential HEDMs15.  The %OB of an explosive of the generic form CwHxNyOz is calculated using equation 6:

 

 

6

 

where  is the molar mass of oxygen, which is set to 16 g/mol, and  is the molar mass of the explosive.1,10  The %OB value was then used to determine the ideal detonation reaction each material would undergo.1-3,5,16, 

 

3.3 Detonation Performances

 

            The detonation velocity (D) and detonation pressure (P) were calculated using the Kamlet-Jacobs equations,2,3,4,14,20,26

 

 

7

 

 

 

 

8

 

where N is the moles of ideal gaseous detonation products per gram of explosive, M is the average molecular mass of the gaseous products (g/mol), Q is the enthalpy change of the detonation reaction (cal/g), and ρ is the density of explosive (g/cm3).  The theoretical density was obtained from the molecular weight and the theoretical molecular volume which is generally defined as the space within an electronic isodensity contour of 0.001 electron/bohr evaluated using a Monte Carlo integrator.2-4,8,9,14,16  The generally accepted method for obtaining the molecular volume is to use the average molecular volume of 100 single point calculations2,16; however, due to time constraints and resource availability, the volume produced by a single calculation was used. 

 

4. Results & Discussion

 

4.1 Heats of Formation

 

            The magnitude of the heat of formation is a good indicator of the amount of energy a HEDM contains relative to its free elements and its accurate prediction is very important in calculating the detonation pressures and velocities.  Table 1 contains the relevant thermodynamic data computed by Gaussian and the calculated thermochemical properties for the six molecules of interest and the three reference compounds.  Table 2 contains the estimated formation enthalpies of the molecules.

 

            In each group, the ΔHf increases with the number of NO2 groups.  At 298 K, the nitrogen skeleton molecules have higher formation enthalpies than the carbon skeleton molecules and the reference molecules.  Thus, N–NO2 bonds contain more energy than C–NO2 bonds and yield significantly higher ΔHf values. 

 

Table 1. Computed and Calculated Thermodynamic Values

 

Computed Values

Calculated Values

Hartree/Part.

kcal/mol

ZPEcorr

Hcorr

∑(Ԑ0+Hcorr)

Ԑ0

Atomization

∆Hf(M, 298 K)

N1

0.040261

0.050313

-670.0919

-670.1422

732.1888

93.0453

N2

0.068045

0.083213

-1039.3641

-1039.4473

1170.7376

237.0028

N3

0.103915

0.126683

-1559.0867

-1559.2134

1781.5534

331.2266

C1

0.054277

0.064133

-654.0998

-654.1639

889.6246

53.1414

C2

0.119986

0.134332

-975.3522

-975.4865

1773.2088

102.0482

C3

0.183990

0.206085

-1463.0534

-1463.2595

2675.9565

139.8074

R1

0.049637

0.053833

-245.0367

-245.0917

567.1550

16.2960

R2

0.106048

0.118245

-786.8300

-786.9483

1465.3301

114.2016

R3

0.141725

0.155324

-897.5234

-897.6787

1773.8602

153.1587

Zero-Point Corrections (ZPEcorr), Thermal Corrections to Enthalpy (Hcorr), Total Electronic Energies (Ԑ0), Sum of Total Electronic Energies and Thermal Corrections to Enthalpy (∑(Ԑ0+Hcorr)),  Atomization Energies, & Formation Enthalpy(∆Hf)

 

Table 2: Estimated Formation Enthalpies in the Gas Phase, in kcal/mol

 

∆Hf (0 K)

∆Hf (298 K)

N1

58.9

68.6

N2

180.3

197.1

N3

244.9

270.3

C1

-10.9

-1.7

C2

31.1

45.4

C3

40.3

62.6

R1

-19.02

-14.5

R2

15.4

27.4

R3

45.3

59.3

Corrections were made using scaling factors based on the estimated difference between ∆Hf, 0K and ∆Hf, 298K as determined from computed Cv values.

 

4.2 Detonation Properties

 

            The detonation velocity, detonation pressure, and density are the three properties essential for evaluating energetic materials.  If the values of these parameters meets or exceeds the theoretical ideal values for an HEDM, then the material is classified as such.  The detonation velocities and pressures, along with the parameters involved in their calculation are shown in Table 3.  The average molar mass of the gaseous detonation products (M), the moles of gaseous detonation products produced per gram of explosive (N), and the enthalpy of detonation (Q) were determined from the detonation reactions based on the material’s oxygen balances in Table 4.


 

            While the carbon group had lower formation enthalpies than the nitrogen group, its detonations out-performed the nitrogen group.  The detonation velocities and pressures increased with the number of NO2 groups resulting in N3 and C3 giving the best performances within their groups.  It is worthy of noting that despite R3 having an equal number of NO2 groups as N1 and C1, its performance rivals N3 and C3.  This indicates that the presence of C–N bonds within the base structure greatly enhance the performance, enough to counteract the effects of only having three NO2 groups. 

 

Table 3: Kamlet-Jacobs Quantities

 

ρ

N

M

Q

D

P

 

g/cm3

Mmole

g/mole

cal/g

km/s

GPa

N1

1.97

32.90

30.3964

352.29

6.627

20.51

N2

2.05

33.34

29.9980

714.20

8.187

32.06

N3

2.27

33.34

29.9980

643.78

8.577

37.06

C1

2.06

31.79

28.3109

276.06

6.223

18.55

C2

1.81

33.90

29.5004

1547.79

9.104

36.86

C3

2.00

33.90

29.5004

1532.25

9.751

44.78

R1

1.49

49.17

20.3388

957.51

7.763

23.67

R2

1.77

33.85

29.3854

1383.79

8.699

33.19

R3

1.79 (1.82)

40.54

24.6705

1227.09

8.924 (8.75)

35.21 (34.00)

The densities were scaled using a factor of 1.089, which was calculated from the calculated ρ of R3 and the calculated density of RDX in Reference 2.  The values in parentheses are literature values from Reference 2.

 

Oxygen Balance (%OB) and Detonation Reactions

 

%OB

Detonation Reactions

N1

63.15

N2

53.32

N3

53.32

C1

37.08

C2

-13.55

C3

-13.55

R1

-26.21

R2

-16.66

R3

-21.61

            It should also be noted that the mere presence of NO2 groups on an atom does not result in greatly improved performance as can be seen by comparing C1 and R1.  The presence of a ring in the base structure, however, does significantly improve a material’s performance.  This can be seen by comparing the increase in performance from Group 2 to Group 3 to the increase from Group 1 to Group 2.  The increase from Group 1 to 2 is far greater than the increase from Group 2 to 3.

 

            Based on the ideal HEDM criteria, C3 is the only molecule worth looking into further.  However, based on the performances of the reference materials (R1-3), C2, N3, and N2 should also be considered for future study.

 

4.3 Relative Thermal Stability

 

            The bond dissociation energies (BDE) are a good indicator of how stable the compounds are; the larger the BDE, the greater the stability.  The BDE values on their own, however, do not tell us anything; they have to be compared to the calculated BDEs of materials known to be stable.  Previous studies have shown PM3 to give reasonable formation enthalpies at a much lower computational expense, so it was used for this evaluation.3  For both groups of molecules, the bond dissociation energies, shown with the reference materials in Table 5, increase with the number of NO2 groups.  This indicates that the larger molecules are more stable. 

 

Table 5: PM3 Enthalpies & Bond Dissociation Energies (BDE) For X-NO2 Molecules at 300K, in kcal/mol

 

Formation Enthalpy

BDE

 

Molecule

X

NO2

N1

6.470

5.467

2.467

1.464

N2

9.414

9.402

2.470

2.458

N3

14.145

16.568

2.471

4.894

C1

5.808

4.880

2.461

1.533

C2

8.650

8.815

2.458

2.623

C3

13.110

15.039

2.458

4.387

R1

3.152

1.073

2.457

0.378

R2

7.169

6.476

2.459

1.766

R3

8.940

8.018

2.458

1.536

 

            The BDE of both N1 and C1 are about four times that of R1 indicating that their stability is roughly 4 times greater.  The BDE of both N2 and C2 are almost 1.5 times that of R2.  The BDE of N3 is more than three times greater than that of R3 while the BDE of C3 is almost three times greater.

 

            While all three of the reference materials are known to be stable, R3 is the most stable and is currently in military applications, so the Group 1 and Group 2 molecules’ BDEs were also compared to that of R3.  The BDE of C1 is almost equal to the BDE of R3 and that the BDE of N1 is slightly lower than that of R3 indicating that C1 and R3 have approximately equal stability while N1 is slightly less stable.  The BDEs of both N2 and C2 were found to be more than 1.5 times that of R3. 

            This data shows that the thermal stability of the six molecules of interest ranges from slightly less stable than, up to three times as stable as R3 (RDX).

 

4.4 Sources of Error

 

            Due to resource restrictions, there are several sources of error in the computational analysis of these six compounds.  The main source of error is the structures of the molecules.  While they were optimized at a high level of theory, the vibrational analysis performed was done at a step lower.  From looking at the geometry of C3, it was clear that it had been optimized to a local minimum on the potential energy surface; however, the vibrational analysis that was able to be performed indicated that it was at a global minimum.

 

            Also related to the structures, in each Frequency output file there was a warning of the possible introduction of significant error from explicit consideration of X degrees of freedom as vibrations where X ranged from 3 for R1 to 30 for N3.  With the exception of R1, each Frequency output file also contained a warning of the possible introduction of significant error from the assumption of classical behavior for rotation.

 

            The method used to calculate the theoretical density, as well as the structures, also introduced error into the numerical results.  The keyword Tight was used to increase the density of points used in the Monte Carlo integration, but this only increases the accuracy of the calculation to 10%.  As a result, the generally accepted method for obtaining an accurate molecular volume is to take the average of 100 single point calculations.  However, not having personal access to the Gaussian software, it was not plausible to run 100 calculations for all nine molecules in addition to the other, at a minimum, 42 collective jobs for the molecules and their five detonation products. 

 

            Issues with the output data from Gaussian also arose, mainly surrounding the thermal corrections to enthalpy.  The electronic energies and the formation enthalpies at 0 K for the references agreed with the literature values found for the reference materials, but the formation enthalpies at 298 K did not agree with the literature values.  As a result, the heat capacity (Cv) from the Frequency output files was used to estimate the ΔHf values.

 

5. Conclusion

 

            The objective of this project was to: evaluate the molecules of interest in order to determine their potential as HEDMs; classify them as propellants, secondary explosives, or primary explosives; and compare the detonation performances of the nitrogen based molecules to those of the carbon analogues.  This was accomplished by using the DFT-B3LYP method with the 6-311G(d,p) basis set and the PM3 method.  The results show that the formation enthalpies increase with the number of NO2 groups and by making the base structure cyclic.  The nitrogen based structures were found to have larger positive formation enthalpies than the carbon analogs; however, the predicted detonation performances indicate that the carbon molecules have better HEDM potentials than the nitrogen molecules. 

 

            All six molecules investigated meet the qualifications to be classified as secondary explosives.  The only molecule found to perform at or above the threshold for ideal HEDMs was C3.  C2 exceeded the detonation velocity threshold, but fell slightly below the detonation pressure and density minimums.  Since both its density and detonation pressure were still above those of R3 (RDX), which is currently being used, it is worth looking into further.  From the nitrogen group, all three members exceeded the density threshold, but only N2 and N3 have detonation pressures and velocities close to those of the references R2 and R3.  The detonation pressure of N3 exceeds that of R3 but the detonation velocity is smaller by 347 m s-1.  Both the detonation velocity and pressure of N2 fall short of R2’s performance.  However, considering how close these values are to those of the references, both N2 and N3 should be looked into further along with C2. 

 

            The relative thermal stabilities were evaluated using the bond dissociation energies calculated using the PM3 method.  For molecules with equal numbers of NO2 groups, the stability decreases with the number of carbons replaced by nitrogens.  It is known that stability decreases with increasing NO2 groups, but it increases with the size of the base structure.  The BDEs of C2, C3, N2, and N3 exceed those of all three references.  Considering the thermal stability and the detonation performances, four of the six materials investigated show promise as HEDMs and should be studied further.

 

Acknowledgements

 

            I would like to thank my advisor, Dr. Ozturk, for overseeing this project and helping when needed; Dr. Charles Kirkpatrick of St. Louis University’s Chemistry department for running my input files through Gaussian 09; Drs Matt Tuchler and Shawn Sendlinger, along with the Shodor Education Foundation, Inc., for accepting my application to attend Shodor’s Computational Chemistry for Chemistry Educators workshop, offered through the National Computational Science Institute (NCSI) and funded by the National Science Foundation (NSF), which introduced me to computational chemistry; and my boss, Gail Hollis, for moral support.


 

References

[1] Akhavan, J.  The Chemistry of Explosives, 3rd ed.; Royal Society of Chemistry: Cambrige, 2011.

[2] Wang, F.; Wang, G.; Du, H.; Zhang, J.; Gong, X.  Theoretical Studies on the Heats of Formation, Detonation Properties, and Pyrolysis Mechanisms of Energetic Cyclic Nitramines,  J. Phys. Chem. A  2011, 115, 13858.

[3] Wang, G. X.; Xiao, H. M.; Ju, X. H.; Gong, X. D.  Calculation of Detonation Velocity, Pressure, and Electric Sensitivity of Nitro Arenes Based on Quantum Chemistry,  Propellants, Explos., Pyrotech.  2006, 31, 361.

[4] Wei, T.; Zhu, W. H.; Zhang, X. W.; Li, Y. F.; Xiao, H. M.  Molecular Design of 1,2,4,5-Tetrazine-Based High-Energy Density Materials,  J. Phys. Chem. A  2009, 113, 9404.

[5] Muthurajan, H.; Sivabalan, R.; Talawar, M.B.; Venugopalan S.; Gandheb, B.R.  Computer Code for the Optimization of Performance Parameters of Mixed Explosive Formulations,  J. Hazard. Mater.  2006, 136, 475.

[6] Meyer, R.; Köhler, J.; Homberg, A. Explosives, 6th ed.; Wiley-VCH: Germany, 2007.

[7] Fan, X. W.; Ju, X. H.; Xiao, H. M.; Qiu, L.  Theoretical Studies on Heats of Formation, Group Interactions, and Bond Dissociation Energies in Neopentyl Difluoroamino Compounds,  J. Mol. Struct.: THEOCHEM  2006, 801, 55.

[8] Wei, T.; Zhu, W. H.; Zhang, J.; Xiao, H. M.  DFT Study on Energetic Tetrazolo-[1,5-b]-1,2,4,5-tetrazine and 1,2,4-Triazolo-[4,3-b]-1,2,4,5-tetrazine Derivatives,  J. Hazard. Mater.  2010, 179, 581.

[9] Fan, X. W.; Ju, X. H.; Xiao, H. M.  Density Functional Theory Study of Piperidine and Diazocine Compounds,  J. Hazard. Mater.  2008, 156, 342.

[10] Davis, T.L.  The Chemistry of Powders and Explosives; Angriff Press: Los Angeles, 1943.

[11] Scott, A.P.; Radom, L.  Harmonic Vibrational Frequencies  An Evaluation of Hartree-Fock, Moller Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors,  J. Phys. Chem.  1996, 100, 16502.

[12] Explosion.  The American Heritage Dictionary of the English Language, 4th ed.; Houghton Mifflin: Boston, 2009; pp. 626.

[13] Explosion.  Collins English Dictionary – Complete and Unabridged [Online]; William Collins Sons, Updated prior to July 21, 2012.

[14] Kamlet, M. J.; Jacobs, S. J.  Chemistry of Detonations. I. A Simple Method for Calculating Detonation Properties of C–H–N–O Explosives,  J. Chem. Phys.  1968, 48, 23-35.

[15] Kamlet, M. J. and Adolph, H. G.  The relationship of Impact Sensitivity with Structure of Organic High Explosives. II. Polynitroaromatic explosives,  Propellants, Explos., Pyrotech.  1979, 4, 30.

[16] Rice, B. M.; Hare, J. J.; Byrd, E. F. C.  Accurate Predictions of Crystal Densities Using Quantum Mechanical Molecular Volumes,  J. Phys. Chem. A  2007, 111, 10874.

[17] Yu, J.; Sumathi, R.; and Green, W. H., Jr.  Accurate and Efficient Method for Predicting Thermochemistry of Polycyclic Aromatic Hydrocarbons - Bond-Centered Group Additivity,  J. Am. Chem. Soc.  2004, 126.39, 12685–12700.

[18] Muthurajan, H.; Sivabalan, R.; Talawar, M.B.; Anniyappan, M.; Venugopalan S.  Prediction of Heat of Formation and Related Parameters of High Energy Materials,  J. Hazard. Mater.  2006, 133, 30-45.

[19] Standard, J. M.  Better Computational Thermochemistry: Isogyric and Isodesmic Reactions, 2011.  Illinois State University: Department of Chemistry.  http://chemistry.illinoisstate.edu/standard/che38037/handouts/380.37thermo.pdf  (accessed March 19, 2012).

[20] Politzer, P.; Lane, P.; Murray, J. S.  Computational Characterization of a Potential Energetic Compound: 1,3,5,7-Tetranitro-2,4,6,8-tetraazacubane,  Cent. Eur. J. Energ. Mater.  2011, 8.1, 39-52.

[21] Iyer S., Slagg N., Molecular Aspects in Energetic Materials, in: Structure and Reactivity, (Liebman J.F., Greenberg A., Eds.), VCH Publishers, New York, 1988, ch. 7.

[22] Dlott D.D., Fast Molecular Processes in Energetic Materials, in: Energetic Materials. Part 2. Detonation, Combustion. (Politzer P., Murray J.S, Eds.), Elsevier, Amsterdam 2003, ch. 6, 125-191.

[23] Meyer R., Köhler J., Homburg A. Explosives, 6th ed., Wiley-VCH, Weinheim 2007.

[24] Politzer, P.; Murray, J. S.  Some Perspectives on Estimating Detonation Properties of C, H, N, O Compounds,  Cent. Eur. J. Energ. Mater.  2011, 8.3, 209-220.

[25] Sochet, I.  Blast effects of external explosions.  In Eighth International Symposium on Hazards, Prevention, and Mitigation of Industrial Explosions.  Yokohama, Japan, Sept 6-10, 2010.  2010; 1-32.

[26] Qui, L.; Xiao, H.; Gong, X.; Ju, X.; Zhu, W.  Theoretical Studies on the Structures, Thermodynamic Properties, Detonation Properties, and Pyrolysis Mechanisms of Spiro Nitramines,  J. Phys. Chem. A  2006, 110, 3797-3807.

[27] Bader, R.F.W.; Carroll, M.T.; Cheeseman, J.R.; Chang, C.  Properties of Atoms in Molecules: Atomic Volumes,  J. Am. Chem. Soc.  1987, 109, 7968-7979.

[28] Frisch, M. J. et al.  Gaussian 09, Revision A.02; Gaussian, Inc., Wallingford CT, 2009.

[29] Gaussian 09, Revision C.01, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian, Inc., Wallingford CT, 2009.