CP620, Shock Compression of Condensed Matter - 2001 edited by M. D. Furnish, N. N. Thadhani, and Y. Hone © 2002 American Institute of Physics 0-7354-0068-7/02/$ 19.00 HOW POINT AND LINE DEFECTS AFFECT DETONATION PROPERTIES OF ENERGETIC SOLIDS Maija M. Kuklja Department of Electrical and Computer Engineering, Michigan Technological University, 1400 Towmend Drive, Houghton, MI, 49931-1295, USA Abstract. An ab initio study is performed for the initiation of chemistry in high explosive crystals from a solid-state physics viewpoint. Specifically, we are looking for the relationship between the defect-induced deformation of the electronic structure of solids, electronic excitations, and chemical reactions under shock conditions. Band structure calculations by means of the Hartree-Fock method with correlation corrections were done to model an effect of a strong compression induced by a shock/impact wave on the crystals with and without lattice defects. Based on the results obtained, an excitonic mechanism of the earliest stages for initiation of high explosive solids is discussed with application to cyclotrimethylene trinitramine (also known as RDX) crystal. Experimental verification of the validity of the proposed model is reported for RDX and heavy metal azides. Thus, the key role of electronic excitations facilitated by edge dislocations in explosive solids is established and analyzed. Favorable conditions for RDX explosion as a result of shock compression are discussed. atomic/electronic rearrangement taking place in pico- to femto- second range time scales. We believe that accurate models and simulations connecting the microscopic properties of atoms and molecules to the macroscopic behavior of materials will bring very new perspectives in an initiation of detonation theory at large. A wealth of experimental evidence, indicating that electronic excitations (electrons, holes, excitons) play a key role in the initiation process, exist nowadays. A variety of initiation models has been proposed. Experiment and theory have suggested that the sensitivity of HE to initiation by mechanical perturbations or shocks is a strong function of solid-state properties including crystal structure and lattice defects. Several studies on the electronic structure of defects in unstable solids are available [1,2,3]. Yet, no complete microscopic theory of early stages for initiation currently exists in a widely agreed upon form. Specifically, the problem of how the energy of the shock wave is INTRODUCTION Modeling of materials that encompass a range of length and time scales is crucial to advances in the understanding such complex phenomena as explosive decomposition of materials. In this article we discuss a possibility of a solid state chemical reaction involving electronic excited states. We attempt to link computationally obtained results from simulation of the electronic structure of different lattice defects in organic high explosive (HE) solids, a theoretically predicted model for the excitonic mechanism of initiation in the RDX ([CH2N-NO2]3) crystals and experimentally discovered new optical and electronic properties of HE. We also show how the macro-behavior of solids such as electrical conductivity, optical absorption and luminescence, observed for seconds and minutes, can be explained and consistently interpreted using theoretically developed quantum chemical models for lattice defects and an 454 transferred into the individual molecules, and how the energy released can support a chemical reaction, remains unclear. Also, the relationship between the defect structure, electronic excitations, and chemical processes under shock conditions in the detonation front is not yet established. In this article, aimed at revealing this relationship, we analyze the initiation process from a solid-state physics viewpoint. The theoretical approach is based on band-structure calculations of a perfect and defective material under shock conditions. Specifically, we model an effect of the strong compression induced by a shock/impact wave on the crystals with and without lattice defects such as a molecular vacancy, a vacancy complex, nanocracks, edge dislocations, and a free surface. The relevant experimental data are also discussed in great detail. Energy dependence was investigated by proportionally decreasing all lattice constants and keeping the structure of the RDX molecules in the crystal unaltered. The pressure was calculated using the low-temperature formula P=-dU/dV. To simulate a molecular vacancy in RDX, we remove one of the eight molecules from the orthorhombic RDX unit cell. Use of periodic boundary conditions (supercell model) results in a model crystal with a concentration of defects as high as 12.5%. By doubling the size of the supercell and changing positions of vacancies, we were able to study the effect of defect distribution, in particular the formation of vacancy dimers, on the electronic and optical properties of the solid. In a similar manner we modeled two-dimensional defects, such as fine nano-cracks, placed close to each other. The modeling of dislocations is more challenging. The crystal is assumed to be composed of stacked parallel layers of one-unit-cell thick. To produce a dislocation with the Burgers vector of [001] one can slide one crystalline plane with respect to another part of the solid, along the z direction. This slip produces a plastic deformation in the material, and the boundary between the slipped and unslipped regions is the edge dislocation. Its position is marked by the termination of an extra vertical half-plane of the RDX molecules crowded into the upper half of the crystal. In our calculations we modeled the dislocation core by a onedimensional (polymer) model. The periodic translation was applied along the dislocation line d [010] and no dislocation-dislocation interactions were included. The free surface of RDX was simulated by means of two different solid state models such as a periodic slab model and a molecular cluster model [6]. The lattice relaxation induced by defects was neglected in this study to reduce computational costs. One can find more details regarding the calculation techniques and approximations used in our studies elsewhere [1-2]. The results obtained allowed us to make two important observations. First, the most studied defects induce a reduction of the band-gap, even at zero pressure. The band-gap is most sensitive to the presence of dislocations due to a strong internal stress. Second, volumetric contraction usually leads to the additional narrowing of the gap. This effect strongly depends on the nature and spatial SUMMARY OF THE RESULTS Our calculations of the electronic structure of explosive solids containing various defects are performed by means of the standard Hartree-Fock (HF) method for a periodic system by the code CRYSTAL95/98 [4]. In this study, we used the 621g split valence set with the modified scaling factor of 1.10 for the outer (most diffuse) Gaussian function. In all calculations, internal geometry parameters of the RDX molecule such as bond lengths, angles, and torsion angles were taken from experimental data [5] and fixed, while the lattice constants were varied to minimize the total energy. Electron correlation corrections based on the second-order many-body perturbation theory (MBPT) are included in the calculations to correct the value of the optical band gap of the system, which is overestimated in the HF limit. Our main objective was to study the effect of defects on the RDX optical band gap under shock compression. Our calculations have been performed for the crystal structures both in equilibrium and under external hydrostatic compression. The band gap as a function of the crystal compression V/Vo was carefully analyzed. A response of the ideal and defective RDX crystals to the shock/impact wave excitation was modeled quantum-mechanically with the rigid molecule approximation. Thus, isotropic compressibility of the crystal in terms of Volume455 orientation of defects in the solid. For instance, in crystals containing nano-cracks the gap weakly depends on the pressure, while in crystals with edge dislocations the gap drops to 1.5 eV at relatively small compression, about 92% of a normal crystal volume Vo. The correlation correction, which is largely due to polarization effects, reduces the computed pure, perfect RDX gap by 8.42 eV to a value of 5.25 eV. Optical absorption experiments [7] yield 3.4 eV. In particular, it was found that strong intramolecular light absorption at wavelengths below 340 nm (3.4 eV) leads to the formation of NO2 radicals. It was also revealed that this weak absorption band in the near ultraviolet is not observed in spectra of the solvated or gas-phase RDX compounds [7]. This fact coupled with a value difference of nearly 2 eV between calculation and experiment, leads us to assume that this absorption is associated with the essentially solid-state of the matter, with lattice defects. In our calculations, we model the perfect crystals while the real solids with some imperfections were investigated in experiments. Following our recent study, the dislocation-related deformation of the electronic structure narrows the band gap of RDX to 3.3 eV, much closer to the experimental value (3.4 eV). This creates a new absorption band, which was not observed for either the perfect RDX crystal or the isolated RDX molecule [1]. The obtained results conclude that the band-gap is most sensitive to the presence of dislocations, which, unlike other defects, produce local energy levels in the forbidden gap [8]. Positions of these dislocation-induced electronic states in the band gap are strongly dependent on shear stress and/or molecular motion in the solid [8]. The dramatic narrowing of the optical band gap, related to the high shear strain, leads to the splitting off of local levels from both the top of the valence band and the bottom of the conduction band. These local states, having bonding and antibonding character, respectively, are attributed to the weakest N-NO2 bond in any RDX molecule located near the dislocation core. From this observation one may conclude that the lowest energy electronic excitation, i.e. promotion of an electron from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) may lead to the breaking of the N-NO2 bond. The nature of the HOMO to LUMO excitation and its relation to the molecular decomposition has been also established in earlier theoretical [9] and experimental [7] studies. Interesting conclusions follow from our recent study on the thermal decomposition of solid RDX [6]. The molecule in the crystal has different energetic barriers for cleavage of the N-NO2 bonds, and a molecule located near the surface can more easily dissociate than a molecule buried in the bulk crystal. Hot spots are known to be associated with lattice defects such as vacancies, voids, pores, cracks, dislocations, and others. It is well established that defective solids, in particular, RDX typically has a high concentration of imperfections even at low temperatures, are much more sensitive to initiation than high quality (perfect) crystals. Our results lead us to the suggestion that one of the reasons for initiation on hot spots is the reduced energetic barrier for the decomposition of molecules placed on defects. Obviously, the surface induced effect must be taken into account studying the initiation mechanisms. Further, we investigated a time-dependent lattice response to the mechanical perturbation, specifically how the vacancy diffusion dynamics is affected by the impact wave progressing across the solid [10]. The developed method based on thermodynamics illustrates that the impact wave interacts with the lattice defects (vacancies) creating the vacancy super-saturation zone, which is moving further with the impact wave. Once the exothermic chemical reaction in the chemically active zone is triggered, this will result in an appearance of the self-sustained regime, and therefore, an impact wave may be transferred to detonation. A good illustrating example is the well examined pore collapse mechanism. The analysis performed may help in a consistent interpretation of the existing experimental data, and in providing useful insights for the understanding of mechanisms for detonation initiation in explosive molecular solids. A MODEL OF INITIATION OF CHEMISTRY In this section we examine the role of electronic excitation in the initiation of chemical reactions for the example of solid RDX. We attempt to understand how such a fast process as electronic excitation, promoted by lattice defect (especially, 456 edge dislocations) and pressure, affects macroproperties of HE and finally leads to an explosion. Let us now consider a way by which the reaction goes through the excited state potential energy surface. We propose the following scenario as illustrated in Figs. 1 and 2. The system is in the ground state (point A) until the impact wave arrives. Volumetric contraction, induced by a shock/impact wave progressing through a solid, leads to an additional narrowing of the optical gap [1-2]. In other words, an impact wave drives the system to the point F. From this point the molecule has a chance to be excited to the triplet state (or an excited singlet state) [7]. The probability for this excitation at point F is higher than at the equilibrium (point A) due to the locally reduced gap and an increased temperature. Our ab initio calculations of the compressibility of the defect-free RDX crystal show that a pressure of 30 GPa causes a gap reduction of about 2 eV. The pressure of -15-20 GPa locally reduces the gap to about 1 eV for RDX containing an edge dislocation. Usually RDX has a rather high concentration of these defects, typically 5*1012 dislocations per cm2 [1]. This corresponds to about 8.5*1016 molecules located near to dislocation cores in a 1mm3 sample. Then, a simple estimate shows that about 3.1*107 such molecules will be excited by the impact front. We would like to note that we use here the energy gap corresponding to the optical (or vertical) electronic transition illustrated by the vertical arrow at point F in Fig. 1. In fact, the thermal excitation of electrons requires even less energy due to nuclear relaxation in the excited state. Furthermore, a dynamic response of the lattice to excitation by a shock/impact wave in terms of probability and population of the electronic excited states, is expected to be much more pronounced. As mentioned above, when the system is excited, an electron is removed from HOMO to LUMO, and the energy is localized on a single N-NO2 stretching mode. If the electronic excitation is long-lived (triplet excited states are known to live much longer than the period of nuclear vibrations), then this excess energy can be released in a form of nuclear motion [11] and, with a high probability, result in the breaking of the N-NO2 bond. In Figs. 1 and 2, this situation is visualized as the system's movement along the path from B to C accompanied by the release of products (R and NO2 radicals) having high kinetic energy. The chemical reaction can go further with an additional energy release as indicated by the arrow in Fig. 2. This increases the local pressure and temperature around the defect and drives the point F for neighboring molecules even higher in energy (F1). It is also seen from Fig. 1 that pressure approaching the metallization pressure (point D) is too high compared to the experimentally observed values. Therefore, metallization of the sample (insulator to metal phase transition), as was suggested earlier, is not the necessary pre-condition for the initiation of chemistry leading to the chain reaction and explosion. Even a moderate reduction of the band gap can trigger initiation. E(eV) 5.25 2.0 -- 0,0 -- •"1M.NO2 R FIGURE 2. Sketch of the RDX molecular energy as a function of the N-NO2 distance R. FIGURE 1. Schematic plot of the crystal energy per unit cell as a function of crystal compression, V/Vo. The energy terms for the the ground state (S) and excited state (T) of the system are shown. 457 Our proposed model of initiation triggering by electronic excitations associated with the dislocation presence in the crystal was illustrated on RDX high explosive molecular crystal. Consistent results were recently obtained for PETN [C(CH2ONO2)4] crystal [8], which represents an intermediate case between primary and secondary high explosives with respect to sensitivity to detonation. A number of experimental and theoretical data described in the literature lead us to believe that the suggested model of initiation may be general for all solid molecular explosives. In the next section we show some evidence in favor of this proposal for some HE. content will ignite at a shock pressure of 2 GPa or less [16]. A dislocation-induced effect on early stages of slow decomposition in the electric field has been experimentally studied for AgN3 whiskers [18]. It was observed that gas (a product of chemical reaction) evolved mostly from dislocation etch pits. Next, mobility of dislocations, caused by their magnetic moment, was used for a significant reduction of the dislocation density in the crystal, which is, in fact, electrical cleaning of the sample. Further, electric current was applied to the cleaned samples for 20 hours, and neither etch pits nor gas emission were observed during all that time. Then, the initial properties of the sample were restored and gas emission regions near etch pits were observed again. The obtained results suggest the association of the regions of the increased chemical reactivity, responsible for the initial stages of decomposition, with dislocations. Furthermore, a series of experiments were performed to study kinetics of the pre-explosion conductivity of AgN3 crystals using YAG:Nd3+ laser pulse initiation (1064 nm, lOmJ, 30 ps) the methodology of which is described elsewhere [17]. Two groups of samples were probed: one set of crystals with the dislocation density -2-103 cm"2, and another set of "pure" crystals with the dislocation density less than 102 cm"2. A comparison of kinetics of pre-explosion conductivity for these two groups of crystals show that chemical decomposition in crystals with a low density of dislocations develops noticeably slower than in the highly defective samples. Experimental studies revealed the chainreaction nature of explosive decomposition of heavy metal azides such as AgN3, T1N3, PbN6 [17-18]. Preexplosion conductivity, optical absorption, and luminescence observed lead us to believe that not only triggering of initiation is related to electronic excitations but also the development of the process. A model based on formation of hot holes (characteristic time of the process is T~10"9 s) and reproduction of their quasi-local hole states (the lifetime of holes in quasi-local state has been found not to exceed 10"14 s) in the band gap was developed to explain explosive behavior of solids. The appearance of such short-lived states creates conditions for the recurrence of the chain of the processes needed for the shock propagation through the solid and finally an explosion. EXPERIMENTAL EVIDENCE The theoretically developed model was buttressed experimentally very recently [12]. The initiation of RDX by laser excitation under modest pressure conditions (up to 5.0 GPa) suggests that sample initiation occurred in a discrete region of the crystals and not uniformly across the sample. From this it was concluded that absorption must have occurred at point or line defects. In that study, theory favored absorption at a dislocation or a vacancy associated with a dislocation core. The gap of the RDX crystal containing this complex defect is very sensitive to pressure and significantly reduced (less than 2.25 eV with the pressure in the GPa range). Thus, it is demonstrated that the energy associated with this defect is consistent with the absorption of the green excitation wavelength observed experimentally, whereas normally RDX is transparent to green light [12,13]. Initiation on voids (the pore collapse mechanism) does not provide a satisfactory explanation here because vacancies, pores, voids, and free surfaces do not produce local electronic states in the band gap. Therefore, new absorption bands are not expected due to these defects. It has also been reported that eliminating or severely reducing the number of moving dislocations during rapid deformation will reduce the number and intensity of shock induced hot spots within the crystal. Carefully prepared crystals of the explosives RDX, TNT, HMX, and PETN, with limited defect content and few dislocations, are nearly impossible to ignite at shock pressures often in excess of 40 GPa [14-15]. Similar crystals of the same materials but with high defect/dislocation 458 DISCUSSION AND CONCLUSIONS ACKNOWLEDGEMENTS We have studied the initiation of chemistry in high explosive crystals from a solid-state physics viewpoint. In particular, we were attempting to reveal the interplay between the defect-induced local deformation of the electronic structure of solids, fast electronic excitations, and chemical reactions under shock conditions. In other words, we analyzed how the experimentally observed macro-behavior of HE solids can be consistently interpreted using the theoretically developed model for a fast electronic excitation facilitated by a local micro-deformation, and how this can lead to a macro chain reaction propagating through the solid and result in an explosion. Band structure calculations by means of the Hartree-Fock method with correlation corrections were performed to model an effect of the shock/impact wave on the crystals with and without lattice defects. Based on the results obtained, an excitonic mechanism of the earliest stages for initiation of high explosive solids is discussed with application to the RDX crystal. Experimental evidence for feasibility of this mechanism is analyzed. We would like to draw attention to some fundamental aspects of the problem considered here, which are far beyond the only elucidation of the mechanism of HE decomposition. We believe that this article convincingly demonstrates a very interesting possibility for realization of the chemical reaction in the solid state. The necessary condition for the chemical reaction to occur in liquids and gases is known to be the real migration of reactants towards each other, their collisions, and interactions with virgin molecules. The nature of chain reactions in solids is more complex. A completely different situation is possible [17-18]. Here, electronic excitations migrate across the crystal. Localization of these excitations at definite sites of the crystalline lattice (structure or impurity defect) leads to the appearance (birth!) of actual radicals at the necessary site. Thus, a sufficiently slow process of migration of real heavy particles (usually, diffusion process, which is inconsistent with the experimental data on explosive decomposition of solids) is replaced by much faster migration of electronic excitations (quasi-particles). The author is very grateful to C.M. Tarver and E.D. Aluker for their encouragement and stimulating discussions. Research has been supported in part by the Materials Research Institute at the Lawrence Livermore National Laboratory under contract with DOE. REFERENCES 1. Kuklja, M. M., and Kunz, A.B., /. Phys. Chem. Solids 61, 35 (2000); J. Phys. Chem. B 103, 8427 (1999); J.Appl Phys. 87, 2215 (2000); J.Appl. Phys. 86,4428 (1999). 2. Kuklja, M.M., Kunz, A.B. in Multiscale Modelling of Materials, ed. by V.V.Bulatov, T.D.Rubia, R.Pjillips, E.Kaxiras, N.Ghoniem, (Mat. Res. Soc. Proc. 538, Pittsburgh, PA, 1999) 347-352. 3. Reed, E.J., Joannopoulos, J.D., Fried, L.E., Phys. Rev. B, 62, 16500-165009 (2000). 4. Dovesi, R., Saunders, V.R., Roetti, C, Causa, M., Harrison, N.M, Orlando, R., and Apra, E. in CRYSTAL95 User's Manual (University of Torino, Torino, 1996). 5. Choi, C.S., Prince, E. Acta Crystallogr. B, 28, 2857 (1972). 6. Kuklja, M.M., J. Phys. Chem. fl, (in press) (2001). 7. Marinkas, P. L.,J. of Luminescence 15, 57 (1977). 8. Kuklja, M.M., Kunz, A.B., J.Appl. Phys. 89,4962 (2001). 9. Orloff, M. K., Mullen, P. A., and Rauch, F. C., J.Phys. Chem. 74, 2189(1970). 10. Skryl, Yu., and.Kuklja, M.M, these proceedings. 11. Tarver, C.M., Fried, L.E., Ruggiero, A.J., Calef, D.F., in 10th International Detonation Symposium, Boston, MA, pp. 3-10 (1993). 12.Kunz, A.B., Kuklja, M.M., Botcher, T.R., Russell, T.P., Thermochim. Acta (submitted) 2000. 13. Botcher, T.R , Landouceur, H.D., and Russel, T.P., in Shock Compression of Condensed Matter -1997, APS Proceedings 429, ed. S.C.Schmidt, D.P.Dandekar, and J.W.Forbes, 989 (1998). 14. Coffey, C.S., in Structure and Properties of Energetic Materials, ed. by D.H. Liedenberg, R.W. Armstrong, J.J. Oilman, (Mat. Res. Soc. Proc. 296, Pittsburgh, PA, 1993) 63-73. 15. Dick, J. J., A/?/?/. Phys. Lett. 44, 859 (1984). 16. Liddiard, T.P., Forbes, J.W., Price, D. in Proc. of the 9th Symposium on Detonation, 1989, 1235. 17. Aduev, B.P., Aluker, E.D., Belokurov, G.M., Zakharov, Yu.A, Krechetov, A.G., JETP 89, 906 (1999). 18. Kuklja, M.M., Aduev, B.P., Aluker, E.D., Krasheninin, V.I., Krechetov, A.G., Mitrofanov, A.Yu., J. A/7/7/. Phys., 89,4156 (2001). 459
© Copyright 2025 Paperzz