Anda di halaman 1dari 7

Journal of the Chinese Institute of Engineers, Vol. 29, No. 1, pp.

145-151 (2006)

145

Short Paper

CALCULATION OF DETONATION PRESSURES OF CONDENSED CHNOF EXPLOSIVES

Mohammad Hossein Keshavarz* and Hamid Reza Pouretedal

ABSTRACT
It is shown that a simplified theoretical approach exists to estimate the detonation pressure of CHNOF explosives at various loading densities. Estimated heat of formation of the explosive in gas phase is one of the essential parameters in the calculation. It is assumed that the detonation products for an oxygen-rich explosive are limited to CO, CO 2 , H 2 O, N2 , HF and O2 ; on the other hand solid carbon and H2 are also counted as major products for an oxygen-lean explosive. Calculated detonation pressure for some known CHNOF explosives show good agreement with the experimental C-J pressures as compared to pressures yielded by complicated BKW-EOS computer code. Key Words: detonation pressure, CHNOF explosives, confined and unconfined conditions, approximate detonation temperature.

I. INTRODUCTION The detonation of an explosive is the result of a complicated interplay between chemistry and hydrodynamics, which produces the extreme pressure and temperature immediately behind the detonation wave. Chapman-Jouguet (C-J) thermodynamic detonation theory has traditionally been used to study detonation of explosives (Cook, 1963). This theory assumes that thermodynamic equilibrium of the detonation products is reached instantaneously. The C-J point can be usually determined by the intersection of the Rayleigh line with the measured isentrope or Hugoniot. Only Chapman-Jouguet pressure, PCJ, and velocity, VCJ, are measured experimentally. Actual temperatures and compositions are almost unknown. Theoretical description of thermodynamic properties of condensed media at high pressure and temperature, such as those generated by very strong shock waves is a very complex problem. However, the application of hydrodynamic
*Corresponding author. (Tel: 0098-0312-522-5071; Fax: 00980312-522-5068; Email: mhkeshavarz@mut-es.ac.ir) The authors are with the Department of Chemistry, Malek-ashtar University of Technology, Shahin-shahr, P.O. Box 83145/115, Islamic Republic of Iran.

theory for determining the detonation properties usually requires an equation of state for the detonation products. Many equations of state are used to describe shock and detonation performance of condensed explosives. Some of the important equations of states include: the Becker-Kistiakosky-Wilson equation of state (BKW-EOS) (Mader, 1963), the Jacobs-CowperthwaiteZwisler equation of state (JCZ-EOS) (Cowperthwaite and Zwisler, 1976), Kihara-Hikita-Tanaka (KHTEOS) (Tanaka, 1985). The Kamlet and coworkers (Kamlet and Ablard, 1968; Kamlet and Dickinson, 1968; Kamlet and Hurwitz, 1968; Kamlet and Jacobs, 1968) simplified computation method, with the use of experimental heat of formation, can predict the detonation properties of CHNO explosives at loading density greater than 1.0 g/cm3. One of the important input parameters for calculation of the detonation performance in the mentioned methods is the condensed heat of formation which can be estimated for some classes of CHNO explosives (Keshavarz and Oftaded, 2003/2004). To a chemist concerned with the synthesis of a new high explosive, the ability to compute detonation performance without the use of experimental solid or liquid heat of formation is a very important consideration. Some relations have been recently introduced for

146

Journal of the Chinese Institute of Engineers, Vol. 29, No. 1 (2006)

determining C-J detonation pressure of pure and mixed CHNO explosives (Keshavarz and Oftadeh, 2002; Keshavarz and Oftadeh, 2003a; Keshavarz and Oftadeh, 2003b) via PM3 procedure. PM3 is one of the semi-empirical quantum mechanical methods, so that it is a second parametrization of MNDO, but with some significant improvements (Stewart, 1989). Since computation of heat of formation by PM3 requires software and has special complexity, calculation of C-J pressure takes much more time. The purpose of this work is the extension of previous works to correlate and to predict the detonation pressure of CHNOF explosives, which include fluorinated explosives, based on approximate detonation temperature yielded by confined and unconfined explosions, the number of gaseous products per unit of weight of the explosive and loading (initial) density. We can assume that on the basis of oxygen balance of explosive, the detonation products include CO, CO 2, H 2O, N 2, HF, O 2 as well as solid carbon and H2. An important criterion for this method is to use some group estimation gas phase heat of formation of the explosive related to the molecular structure, e.g. the methods of Benson (Benson et al., 1969), Yoneda (Reid et al., 1987), Joback (Reid et al., 1987), etc., that normally provides reliable estimates of enthalpy of formation. More importantly, there is no need to know the solid or liquid heat of formation or any experimental data of CHNOF explosives. II. METHOD OF OBTAINING THE PERFORMANCE CHNOF EXPLOSIVES Since combustion and detonation reactions of explosives are usually complicated and violent, they have such characteristics as high reaction rates, high temperatures, complicated product compositions and so on. High performance explosive can be obtained by maximizing the C-J particle density or the number of moles of gas per gram of explosive and the heat of detonation (Keshavarz and Pouretedal, 2004). It is general, for any explosive, that increasing the density or the hydrogen content increases PCJ and VCJ. The calculated detonation properties may be meaningful for any new energetic materials. It can be inferred that a high detonation performance is promoted by two fundamental parameters. The first is the formation of light gaseous products, since a greater number of moles are produced per unit weight of explosives. The next factor is to have a high positive heat of formation, since this leads to a greater release heat of detonation and a higher detonation temperature. Detonation products are generally complex mixtures of the large numbers of molecular species whose concentrations change with temperature and pressure.

Furthermore, they may consist of more than one mixture in more than one phase. Depending upon the composition of CHNOF explosives, the major detonation products may contain CO, CO 2, H 2O, N 2, HF and solid carbon as well as minor amounts of H2, NH3, O 2, NO and other chemical species. The amounts of these various products depends upon the stoichiometry of the detonation process and the effects of whatever other equilibria are in effect, such as 2CO H 2 + CO CO 2 + C H2O + C

The calculations of P CJ and V CJ of CHNOF explosives by the BKW-EOS are very sensitive to the equilibrium between HF, carbon and carbon tetrafluoride (Mader, 1998). Since the molecular weight and detrimental effect of CF 4 on the particle density is large, the formation of CF4 is less desirable than HF. Adding elemental fluorine, or boron and aluminum elements, has been of interest for chemists, because the heat of detonation is as much as doubled. To apply the effect of gaseous products in the detonation process, we can simply select the procedure for stoichiometric decomposition reaction so that the above mentioned detonation products for both oxygen-lean and oxygen-rich detonations are typical. All nitrogens are assumed to go to N 2 and fluorines to HF, if hydrogens are available, a portion of oxygens preferentially to form H 2O, otherwise the carbon, at first, will produce CO rather than CO 2 . In order to find a correlation for detonation pressure as a function of the number of detonation products, we can assume that the products are limited to HF, CO, CO2, H 2O, N 2 and O 2 for oxygen-rich explosive. On the other hand solid carbon and H 2 may also counted as major products for an oxygen-lean explosive. This approximation simplifies our procedure, because there is no need to know the precise composition of detonation products. We can use such decomposition reactions to calculate the parameter , which provides a rough estimate of the number of moles of gaseous products available per unit weight of the CHNOF explosives. An explosive can be initiated either by rapid burning or by detonation and its energy is released in the form of heat. The heat so released under adiabatic conditions determines the work capacity of the explosive (Akhavan, 1998; Bailey and Murray, 1989). The temperature of detonation or explosion is the maximum temperature that the detonation products can attain under adiabatic conditions and is often used when calculating ability of an explosive or propellant to do work (Bailey and Murray, 1989). Based on the decomposition reactions a simple

M. H. Keshavarz and H. R. Pouretedal: Calculation of Detonation Pressures of Condensed CHNOF Explosives

147

approach for obtaining a rough approximation of detonation pressure involves assuming that the heat of detonation of the explosive is used entirely to heat the detonation products to the detonation temperature. Since the detonation reaction is extremely fast, the heat of detonation reaction raises the temperature of detonation products in an adiabatic condition to explosion or detonation temperature (Akhavan, 1998; Bailey and Murray, 1989; Meyer et al., 2002). It can be assumed that the heat of detonation is constant over the temperature range between the initial temperature and detonation temperature. Heat of detonation can be calculated from knowledge of the molar heat of formation of the explosive and detonation products. A detonation is usually a confined explosion, which occurs in a closed chamber where volume is constant. Meanwhile an unconfined explosion is an explosion occurring in the open air where the atmospheric pressure is constant (Bailey and Murray, 1989; Meyer et al., 2002). We can calculate the approximate detonation temperature by using heat of formation of the explosive in gas phase instead of solid or liquid state. To calculate approximate detonation temperature, considering the two above assumptions for decomposition reaction and heat of detonation, we have used heats of formation of explosive in gas phase instead of solid or liquid state. This condition is applied in order to obtain a generalized correlation for predicting detonation pressure without the use of any experimental data. It can be assumed that the approximate detonation temperature is estimated by the following equations under confined and unconfined conditions (Bailey and Murray, 1989; Kubota, 2002; Politzer et al., 1991): T ad = 298.15 Q ad / C v Tad = 298.15 Q ad / C p (1) (2)

and Ciller, 1993; Levine, 1983; Cook, 1998). The additivity rules of the group contributions such as the methods of Benson (Benson et al., 1969), Yoneda (Reid et al., 1987) and Joback (Reid et al., 1987) have been better estimation methods for predicting the heat of formation of organic compounds in gas phase. Jobacks method is the simplest technique, which is broadly applicable for quite complex organic compounds and provide reliable estimates (Reid et al., 1987). Moreover, only the more complex Bensons method is more accurate (Benson et al., 1969; Reid et al., 1987). III. SIMPLE EQUATION FOR C-J DETONATION PRESSURE Due to the nonsteady state nature of the detonation waves, there is generally a lower accuracy (10 to 20%) in the experimental measurements of PCJ (Mader, 1998). Experiments for determining the performance of explosives reveal that P CJ is roughly proportional to the cubic of loading density, 0 (Chiart and PittionRossillon, 1981; Johansson and Persson, 1970). The behavior of P CJ versus loading density should be obtained at first attempt if a well-suited method is applied. The necessary data for calculation of T ad, Tad and for some CHNOF explosives are given in Table 1. To express detonation pressure as a function of T ad or Ta d, or square of loading density, we used the experimentally measured P CJ versus various combinations of T ad or Tad and at given loading densities for eleven CHNOF explosives (Abdelazim, 1986; Dobratz and Crawford, 1985; Horning et al., 1970; Mader, 1998) which cover a wide range of explosives in oxygen balance. As shown in Figs. 1 and 2, the following linear relationships are obtained: PCJ = 7.83(Tad)1/202 0.69 PCJ = 9.02(Tad)1/202 1.2 (3) (4)

where Tad and Tad are the approximate detonation temperatures for confined and unconfined conditions respectively, Q ad is the difference between heat of formation of decomposition products and the explosive in gas phase, C v and C p are the sum of the molar heat capacities of detonation products at constant volume and pressure respectively which can be obtained from standard thermochemistry tables at or near the approximate detonation temperature (Stull and Prohet, 1971). Eqs. (1) and (2) show that a large positive explosive heat of formation favors a high approximate detonation temperature. Semi-empirical quantum mechanical methods have been calibrated to typical organic or biological systems and tend to be inaccurate for problems involving hydrogen-bonding, chemical transitions or nitrated compounds (Akutsu and Tahara, 1991; De Paz

where P CJ and PC J are expressed in kilobars, in moles of gas per gram of explosive and 0 in grams per cubic centimeter. As seen from Eqs. (3) and (4), the pressure performance of explosive according to decomposition procedure can be raised by providing higher values of T ad or Tad and at specified initial density. These simple equations show that by using gas phase heat of formation alone, without correction for crystalline effects, is sufficient to determine detonation pressure. IV. THE VALIDITY OF OBTAINED P CJ EQUATIONS WITH RESPECT TO BKW-EOS The BKW-EOS is the most widely used equation of state in predicting equation of state of high

148

Journal of the Chinese Institute of Engineers, Vol. 29, No. 1 (2006)

Table 1 Parameters used in calculations Explosive a FEFO TFNA TFENA DATB HMX TETRYL TNT NG NM PETN RDX Reaction products 2CO + 2N 2 + 2H 2O + 3CO 2 + 2HF 5CO + 2N 2 + H 2O + H 2 + 3HF 2CO + N 2 + 3HF 6CO + 2.5N 2 + 2.5H 2 4CO + 4N 2 + 4H 2O 7CO + 2.5N 2 + 1.5H 2 + H 2O C(s) + 6CO + 1.5N 2 + 2.5H 2 3CO 2 + 1.5N 2 + 2.5H 2O + 0.25O 2 CO + 0.5N 2 + 0.5H 2 + H 2O 3CO 2 + 2CO + 2N 2 + 4H 2O 3CO + 3N 2 + 3H 2O

H f (kJ/mol) b
-864.4 -737.2 -639.7 58.8 188.4 38.8 -18.0 -542.2 -74.8 -728.0 158.4

Gas phase T ad (K) 4285 3068 2989 2891 4404 3524 2645 4857 3325 4163 4448

Tad (K) 3303 2268 2176 2160 3361 2631 2024 3639 2500 3237 3401

c
0.0344 0.0430 0.0417 0.0453 0.0405 0.0418 0.0441 0.0319 0.0492 0.0348 0.0405

a) See appendix A for glossary of compound names and chemical formulas b) Heat of formation calculated by Joback additive group procedure (Reid et al., 1987) c) Number of gaseous products available per unit weight of explosive

500

500

400

400

PCJ/kbar

PCJ/kbar

300

300

200

200

100

100

0 0 10 20 30 ( Tad)1/2 02 40 50

0 0 10 20 30 ( Tad)1/2 02 40 50

Fig. 1 The experimental C-J Detonation Pressure versus (Tad)1/202, where heat of formation calculated by Joback procedure is used to estimate the approximate detonation temperature. The points are: + NG; HMX; TETRYL; TNT; DATB; PETN; * RDX; NM; FEFO; TFNA; TFENA

Fig. 2 The experimental C-J Detonation Pressure versus Tad)1/202, where heat of formation calculated by Joback procedure is used to estimate the approximate detonation temperature. The points are: + NG; HMX; TETRYL; TNT; DATB; PETN; * RDX; NM; FEFO; TFNA; TFENA

explosives (Abdualazeem, 1998). In this method, two sets of parameters are needed to fit specimen explosives, namely parameters fitting RDX and parameters fitting TNT. The first parameters, recommended for most explosives, are those that represent the best fit to RDX. For high explosives whose detonation products contain about half or more of the total number of moles of detonation products as solid carbon, the recommended parameters can be obtained by fitting TNT. Comparisons of calculated P CJ to the new

correlations and BKW-EOS with experimental data are listed in Table 2, which may be taken as appropriate validation tests of the obtained correlation for use with CHNOF explosives. As indicated in Table 2, good agreement is obtained between measured and obtained values of the detonation pressure by the correlation as compared to BKW-EOS over the initial explosive density defined by the experiments. The mean absolute deviation in Eqs. (3) and (4) for these explosives, |(measured-predicted)/measured| 100, are 6.13 and 5.80 respectively.

M. H. Keshavarz and H. R. Pouretedal: Calculation of Detonation Pressures of Condensed CHNOF Explosives

149

Table 2 Comparison between the calculated PCJ by the correlation and BKW-EOS computer code (Mader, 1998) with the experimental values C-J detonation pressure (kbar) Explosive DATB HMX TETRYL NG NM PETN
b

0 (g/cm )
3

Eq. (3) 285.8 376.8 274.0 245.7 128.3 21.0 91.7 262.1 50.8 94.2 126.5 174.2 223.3 339.8 239.6 256.8 202.1

Eq. (4) 284.0 378.7 272.2 244.5 127.7 20.9 92.6 265.8 50.7 94.3 126.9 175.0 224.5 341.8 241.9 253.8 198.1

BKW-EOS (RDX Param.) 282.0 395.0 251.0 246.0 132.0 30.3 101.0 280.0 57.7 98.9 127.1 170.4 218.0 347.0 235.0 242.0 162.0

Experiment a 259 393 260 253 125 24 87 300 47.8 95.9 121.6 166.2 210.8 347.0 250 249 174

1.788 1.90 1.70 1.59 1.135 0.48 0.99 1.67 0.70 0.95 1.10 1.29 1.46 1.80 1.59 1.692 1.523

RDX c

FEFO d TFNA
d

TFENA d

a) Measured values of detonation pressure taken from Dobratz and Crawford (1985) except were noted. b) Horning et al. (1970) c) Abdelazim (1986) d) Mader (1998)

A major shortcoming of the BKW-EOS, in addition to its empirical nature, lies in the fact that recalibration is always required to obtain new sets of fitting parameters every time new chemical species are deal with (Abdualazeem, 1998). The calculated BKW-EOS temperature shows much deviation as compared to experimental values (Mader, 1998). Moreover, not only the predicted temperature, by BKW-EOS, changes with loading density, but also the predicted number of moles changes with initial density. One of the advantages of Eqs. (3) and (4) is that the product of simply calculated parameters, namely and T ad or Tad, does not depend upon loading density. V. CONCLUSIONS Using computed heats of formation of explosive in gas phase instead of experimental condensed heat of formation, and the estimated composition of detonation products are two advantages of this method that can help us to predict PCJ of any CHNOF explosives

without any experimental data about the explosive. The calculated P CJ here, the same as the other procedures (computer codes) (Dobratz and Crawford, 1985), is not very sensitive in showing the accurate values of heat of formation. Therefore, there is no necessity to use complicated methods such as quantum mechanical calculations for evaluating the heat of formation of the explosive in gas phase. Accurate and simple group contribution methods, such as Benson (Benson et al., 1969), Yoneda (Reid et al., 1987) and Joback (Reid et al., 1987), can be used for estimation of gas phase heat of formation. Among the methods Jobacks method is much simpler to use and normally provides reliable estimates (Reid et al., 1987). It is reasonable to expect the calculated and experimental P CJ to differ by 10 to 20% because of the non-steady nature of the detonation wave (Mader, 1998). As shown in Table 2, the results of Eqs. (3) and (4) are comparable with the experimental data and the calculated values by complex BKW-EOS computer code (Mader, 1998). Since heat capacities

150

Journal of the Chinese Institute of Engineers, Vol. 29, No. 1 (2006)

at constant pressure are given in JANAF Thermochemical Tables (Stull and Prohet, 1971) and are referred to by many books, Eq. (4) can be used more readily as compared to Eq. (3). The advantages of the new correlation are: 1) Easily calculated PCJ as compared to the other complicated computer code, 2) There is no need to know the solid state heat of the formation of the explosive. Jobacks method can be easily used to calculate heat of formation of the explosive in gas phase, 3) Estimation of PCJ is possible for a wide range of loading densities, 4) There is also no need to know the accurate composition of detonation products. ACKNOWLEDGMENTS We are indebted to the research committee of Malek-ashtar University of Technology (MUT) for supporting this work. This research was supported in part by Institute of Chemical and Science Technology-Tehran-Iran Research Council (Grant No. ICST-8I03-2125). REFERENCES Abdelazim, M. S., 1986, Dense Fluid Detonation Performance as Calculated by the Lennard-Jones Equation of State, Dynamics of Explosions, J. R. Bowen et al., eds., American Institute of Aeronautics and Astronautics, Inc., USA. Abdualazeem, M. S., 1998, Condensed Media Shock Waves and Detonations: Equation of State and Performance, High Temperature-High Pressures, Vol. 30, pp. 387-422. Akhavan, J., 1998, The Chemistry of Explosives, The Royal Society of Chemistry, UK, pp. 78, 85-86. Akutsu, Y., and Tahara, S. Y., 1991, Calculations of Heats of Formation for Nitrocompounds by Semi-empirical MO Methods and Molecular Mechanics, Journal of Energetic Materials, Vol. 9, pp. 161-172. Bailey, A., and Murray, S. G., 1989, Explosives, Propellants and Pyrotechnics, Brasseys Ltd., UK. Benson, S. W., Cruickshank, F. R., Golden, D. M., Haugen, G. R., ONeal, H. E., Rodgers, A. S., Shaw, R., and Walsh, R., 1969, Additivity Rules for the Estimation of Thermochemical Properties, Chemical Reviews, Vol. 69, No. 3, 279-324. Chirat R., and Pittion-Rossillon, G., 1981, A New Equation of State for Detonation Products, Journal of Chemical Physics, Vol. 74, No. 8, pp. 23-35. Cook, D. B., 1998, Handbook of Computational Quantum Chemistry, Oxford University Press, UK. Cook, M. A., 1963, The Science of High Explosives,

Reinhold, New York, USA. Cowperthwaite, M., and Zwisler, W. H., 1976, The JCZ Equations of State for Detonation Products and Their Incorporation into the TIGER code, Proceedings of the 6th Symposium (International) on Detonaton, Coronads CA, Office of the Chief of Naval Operations, Washington, DC, USA, pp. 162-172. De Paz J. L., and Ciller, J., 1993, On the Use of AM1 and PM3 Methods on Energetic Compounds, Propellants, Explosives, Pyrotechnics, Vol. 18, pp. 33-40. Dobratz, B. M., and Crawford, P. C., 1985, LLNL Explosives Handbook, Properties of Chemical Explosives and Explosives Simulants, UCRL-52997 Change 2, Lawrence Livermore National Laboratory, University of California, USA. Horninig, H. C., Lee, E. L., Finger, M., and Kurrie, J. E., 1970, Equation of State of Detonation Products, Proceedings of the 5th Symposium (International) on Detonation, Office of Naval Research, ACR-184, Washington, DC, USA, pp. 422-429. Johansson, C. J., and Persson, P. A., 1970, Detonics of High Explosives, Academic Press, New York, USA, p. 34. Kamlet, M. J., and Ablard, J. E., 1968, Chemistry of Detonations. II. Buffered Equlibria, Journal of Chemical Physics, Vol. 48, No. 1, pp. 36-42. Kamlet, M. J., and Dikinson, C., 1968, Chemistry of Detonations. III. Evaluation of the Simplified Calculational Method for Chapman-Jouguet Detonation Pressures on the Basis of Available Experimental Information, Journal of Chemical Physics, Vol. 48, No. 1, pp. 43-50. Kamlet, M. J., and Hurwitz, H., 1968, Chemistry of Detonations. IV. Evaluation of a Simple Predictional Method for Detonation Velocities of C-H-N-O Explosives, Journal of Chemical Physics, Vol. 48, No. 8, pp. 3685-3692. Kamlet, M. J., and Jacobs, S. J., 1968, Chemistry of Detonations. I. A Simple Method for Calculating Detonation Properties of C-H-N-O Explosives, Journal of Chemical Physics, Vol. 48, No. 1, pp. 23-35. Keshavarz, M. H. and Oftadeh, M., 2002, A New Correlation for Predicting the Chapman-Jouguet Detonation Pressure of CHNO Explosives, High Temperature-High Pressures, Vol. 34, pp. 495497. Keshavarz, M. H. and Oftadeh, M., 2003a, Two New Correlations for Predicting Detonating Power of CHNO Explosives, Bulletin of Korean Chemical Society, Vol. 24, No. 1, pp. 19-22. Keshavarz, M. H. and Oftadeh, M., 2003b, Simple Method for Predicting Detonation Pressure of CHNO

M. H. Keshavarz and H. R. Pouretedal: Calculation of Detonation Pressures of Condensed CHNOF Explosives

151

Mixed Explosives, Indian Journal of Engineering & Material Sciences, Vol. 10, pp. 236-238. Keshavarz, M. H. and Oftadeh, M., 2003/2004, New Method for Estimating the Heat of Formation of CHNO Explosives in Crystalline State, High Temperature-High Pressures, Vol. 35/35, pp. 499-504. Keshavarz, M. H., and Pouretedal, H. R., 2004, An Empirical Method for Predicting Detonation Pressure of CHNOFCl Explosives, Thermochimica Acta, Vol. 414, pp. 203-208. Kubota, N., 2002, Propellants and Explosives, WILEY-VCH, Weinheim, Germany, pp. 24-28. Levine, I. N., 1983, Quantum Chemistry, Allyn and Bacon, Inc., Boston, MA, USA. Mader, C. L., 1963, Detonation Properties of Condensed Explosives Computed Using the BeckerKistiakosky-Wilson Equation of State, Los Alamos Scientific Laboratory Report LA-2900, Los Alamos, NM, USA. Mader, C. L., 1998, Numerical Modeling of Explosives and Propellants, 2nd ed., CRC Press, Boca Raton, FL, USA. Meyer, R., Khler, J., and Homburg, A., 2002, Explosives, 5th ed., Wiley-VCH Verlag GmbH, Weinheim, Germany. Politzer, P., Murray, J. S., Grice, M. E., and Sjoberg, P., 1991, Computer-Aided Design of Monopropellants, Chemistry of Energetic Materials, G. A., Olah, and D. R. Squire, ed., Academic Press Inc., San Diego, CA, USA. Reid, R. C., Prausnitz, J. M., and Poling, B. E., 1987, The Properties of Gases and Liquids, 4th ed., McGraw-Hill, New York, USA. Stewart, J. J. P., 1989, Optimization of Parameters for Semiempirical Methods II. Applications, Journal of Computational Chemistry, Vol. 10, 221-264.

Stull, D. R. and Prohet, R., 1971, JANAF Thermochemical Tables, 2nd ed., National Burea of Standard, NSRDS-NBS 37, Washington DC, USA. Tanaka, K., 1985, Detonation Properties of High Explosives Calculated by Revised Kihara-Hikita Equation of State, Proceedings of 8th Symposium (International) on Detonation, Albuquerque, NM, USA, pp. 548-557. Manuscript Received: May 06, 2004 Revision Received: Jan. 24, 2005 and Accepted: Mar. 01, 2005 APPENDIX A: GLOSSARY OF COMPOUND NAMES 1. D A T B : 1 , 3 - d i a m i n o - 2 , 4 , 6 - t r i n i t r o b e n z e n e (C 6H 5N 5O 6) 2. HMX: 1,3,5,7-tetranitro-1,3,5,7-tetraazacyclooctane (C 4H 8N 8O 8) 3. TETRYL: N-methyl-N-nitro-2,4,6-trinitroaniline (C 7H 5N 5O 8) 4. TNT: 2,4,6-trinitrotoluene (C 7H 5N 3O 6) 5. NG: propane-1,2,3-triol trinitrate (C 3H 5N 3O 9) 6. NM: nitromethane (CH 3NO 2) 7. PETN: 2,2-bis[(nitroxy)methyl]-1,3-propanediol dinitrate (C 5H 8N 4O 12) 8. RDX: 1,3,5-trinitro-1,3,5-triazacyclohexane (C 3H 6N 6O 6) 9. FEFO: 1,1'-methylenedioxy bis(2-fluoro-2,2dinitroethane) (C 5H 6N 4O 10F 2) 10. T F N A : N - 2 , 2 - d i n i t r o p r o p y l - N - 2 , 2 , 2 trifuoroethylenenitroamine (C 5H 7N 4O 6F 3) 11. T F E N A : 2 , 2 , 2 - t r i f u o r o - 1 - n i t o a m i n o e t h a n e (C 2H 3N 2O 2F 3)

Anda mungkin juga menyukai