Next Article in Journal
Application of the Fluorescent Dye BODIPY in the Study of Lipid Dynamics of the Rice Blast Fungus Magnaporthe oryzae
Next Article in Special Issue
Theoretical Studies on Catalysis Mechanisms of Serum Paraoxonase 1 and Phosphotriesterase Diisopropyl Fluorophosphatase Suggest the Alteration of Substrate Preference from Paraoxonase to DFP
Previous Article in Journal
Correction: El-kalyoubi, S.; et al. Novel 2-Thioxanthine and Dipyrimidopyridine Derivatives: Synthesis and Antimicrobial Activity. Molecules, 2015, 20, 19263–19276
Previous Article in Special Issue
Theoretical Study of the Photolysis Mechanisms of Methylpentaphenyldimetallanes (Ph3MM′Ph2Me; M, M′ = Si and Ge)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Photophysical Deactivation Channel of Laser-Excited TATB Based on Semiclassical Dynamics Simulation and TD-DFT Calculation

1
Chongqing Key Laboratory of Big Data for Bio Intelligence, Chongqing University of Posts and Telecommunications, Chongqing 40065, China
2
Department of Physical Sciences, Nicholls State University, P.O. Box 2022, Thibodaux 70310, LA, USA
*
Authors to whom correspondence should be addressed.
Molecules 2018, 23(7), 1593; https://doi.org/10.3390/molecules23071593
Submission received: 29 May 2018 / Revised: 23 June 2018 / Accepted: 28 June 2018 / Published: 30 June 2018
(This article belongs to the Special Issue Theoretical Investigations of Reaction Mechanisms)

Abstract

:
A deactivation channel for laser-excited 1,3,5-triamino-2,4,6-trinitrobenzene (TATB) was studied by semiclassical dynamics. Results indicate that the excited state resulting from an electronic transition from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular mrbital (LUMO) is deactivated via pyramidalization of the activated N atom in a nitro group, with a lifetime of 2.4 ps. An approximately 0.5-electron transfer from the aromatic ring to the activated nitro group led to a significant increase of the C–NO2 bond length, which suggests that C–NO2 bond breaking could be a trigger for an explosive reaction. The time-dependent density functional theory (TD-DFT) method was used to calculate the energies of the ground and S1 excited states for each configuration in the simulated trajectory. The S1←S0 energy gap at the instance of non-adiabatic decay was found to be 0.096 eV, suggesting that the decay geometry is close to the conical intersection.

1. Introduction

Energetic materials refer to organic substances that can store a large amount of chemical energy. Under the action of external stimuli, such as shockwave and light, energetic materials readily decompose and release energy [1,2,3]. The energy release mechanism is very complicated. In general, irradiation by an ultraviolet or visible laser can be used to electronically excite a molecule, which can then undergo a series of photochemical and photophysical processes, including radiative transitions and nonradiative deactivation. Nonradiative deactivation can involve internal conversion and intersystem crossing; internal conversion is considered to be a crucial first step for an explosive decomposition reaction because the vibrational energy generated by the deactivation can lead to the breaking of chemical bonds, triggering a chain reaction that eventually releases the stored chemical energy. Nitro-explosives are among of the most important energetic materials. The incomplete combustion of nitro-explosives can lead to the release of aromatic nitro compounds into the environment, which may lead to biological mutation and carcinogenesis. Photochemical degradation is the most common natural elimination method [3]. Thus, a study of the decay properties of excited states of nitro compounds has important civil and military applications.
The presence of nitro groups significantly changes the photochemical and photophysical behavior of molecules and affect the stability and sensitivity of energetic molecules [4,5,6]. Recent literature shows [7,8,9,10,11,12,13,14,15,16,17,18] excited energetic molecules deactivated from the S1 excited state to the S0 ground state through a conical intersection (CI). Experimental and theoretical studies by Bernstein and coworkers [6] found that the conical intersection plays an important role in the non-adiabatic decay of the energetic molecule dimethylnitramine (DMNA). They used the complete active space self-consistent field (CASSCF) method to optimize the (S1/S0) CI point structure of the DMNA molecule along the nitro–nitrite isomerization reaction coordinates. This isomerization process was predicted to take place through a “loose transition state-like geometry”. Bernstein’s group studied the non-adiabatic transitions of many energetic molecules [8,9,10,11,12,13,14] and found that their (S1/S0) CI are similar in structure. Soto, Arenas and collaborators [15,16,17] used the complete active space with second-order perturbation theory (CASPT2) to predict NO2–ONO isomerization in nitramide molecules through a (S1/S0) CI; a similar CI structure in the nitromethane molecule has led to the nitro elimination reaction. The aforementioned theoretical studies were static quantum calculations based on single-point optimization or potential energy surface scanning according to a simple reaction coordinate; time-resolved dynamics simulations were not reported. So far, only Ghosh et al. [17] used ab initio dynamics simulations to study small molecule energetic materials, such as nitromethane, DMNA, and methylnitrate; they found the S1 excited states of all the energetic molecules decayed to ground state through an ultrafast internal conversion with a lifetime shorter than 500 fs.
1,3,5-Triamino-2,4,6-trinitrobenzene (TATB) is a photosensitized insensitive explosive that is widely used in defense, mining, construction, and other fields. Photochromism, which is attributed to the photodecomposition of TATB, is observed under light irradiation [19,20,21,22,23,24]. The excited state of the TATB molecule is regarded as the important initial step of photodecomposition. Time-dependent density functional theory (TD-DFT) was applied to optimize the structure of S1 and T1 in TATB by Xiong et al. [23]. It was proposed that intersystem crossing from S1 to T1 is followed by a NO2−ONO isomerization. Chu et al. [23] utilized transient absorption spectroscopy and TD-DFT calculation to investigate the dynamics properties of excited states of TATB. They found structural changes, charge transfer properties, and the lifetimes of the excited states; their work laid a foundation for the study of the TATB excited states. However, the TD-DFT method, which is based on a single reference state, cannot correctly describe the structure of the non-adiabatic CI and the chemical bond breaking. It is difficult to build a complete potential energy surface and obtain information pertaining to the photochemical deactivation mechanism. The CASSCF/CASPT2 method has been widely used to study the state–state interaction and has achieved good precision in single-point optimization and excitation energy calculations. However, the active space contains a limited number of electrons and orbitals and can only be used to study excited states of small energetic molecules [25,26,27,28], from which the properties of larger energetic molecules are inferred. It is very hard to directly deal with larger aromatic systems such as TATB molecules. Even if some electrons and orbitals can be neglected to reduce the active space, it is difficult to select which ones to neglect, because the molecular structure of the non-adiabatic coupling point on the potential energy surface is unknown. Ab initio excited state dynamic studies of large molecules incur tremendously high calculation costs.
Correct understanding of the mechanism of a chemical reaction requires clarification of how each atom moves during the conversion from reactant to product. However, for a molecule with as many atoms as TATB, it is difficult to take into account the overall movements with CASSCF and TD-DFT. The ab initio dynamics methods, which were recently reviewed by Shalashilin and coworkers [29], are good candidates for addressing this problem, since a large number of degrees of freedom could be included. In this study, the semiclassical electronic radiation ion dynamics (SERID) [30,31] model was used to simulate the decay process of TATB excited states and track the movement of each atom, as well as the energy change over time. By comparing the findings with experimental and other theoretical results, we were able to identify a deactivation channel for the TATB S1 state.

2. Results and Discussion

To obtain an initial geometry of the TATB molecule for dynamics simulation, Gaussian 09 [32] was run using MP2/6-311+g(2d,2p) to generate an optimized geometry. Then, a 1000 fs dynamics run was performed in SERID for this single molecule with constant temperature and pressure. Structures at 10 equal time intervals were saved. From these, 10 trajectories can be simulated by taking each of the structures as the initial geometry, coupled with a laser pulse, for a SERID run. The average orbital energies were found to be −6.4 eV, −6.4 eV, −3.2 eV, and −2.8 eV for HOMO-1, HOMO, LUMO, and LUMO+1, respectively; HOMO-1 and HOMO are degenerate orbitals because of molecular symmetry; this is also observed in the calculation at the b3lyp/6-311++g(d,p) level [24]. Femtosecond laser pulses were applied to the equilibrium structure with a full-width at half maxima (FWHM) of 50 fs and photon energy of 3.2 eV (387 nm), which corresponds to the LUMO–HOMO energy gap. The excitation wavelength in the transient spectroscopy experiment of Chu et al. [23] was 400 nm (3.1 eV). The time step was set to 0.05 fs, and data were printed every 20 steps. To ensure sufficient relaxation time, we set our simulation time to 10 ps.
Out of 201 reaction trajectories, 37 trajectories stayed in ground state due to insufficient fluence, 26 trajectories stayed in excited electronic state (with vibrational relaxation), and 136 trajectories led to molecular fragmentations (such as –NO2, –O, and other carbonitride fragments, etc.) because of very high fluence. In these cases, excited state decay to the S0 state is not observed. The photophysical electronic decays S1←S0 were observed in two trajectories. Most of the trajectories (136/201) lead to molecular fragmentations, suggesting that TATB detonation is a very probable outcome of laser excitation. Detailed description of the photochemical reaction, i.e., photodissociation of TATB, will be described in another paper; this paper focuses on the non-adiabatic deactivation pathway of excited state. One trajectory with a laser frequency of 3.2 eV and fluence of 32.94 J·m−2 is reported here. In this trajectory, a photophysical deactivation channel of the TATB molecule via nitro relaxation is observed.
In Figure 1, four representative snapshots show important changes in molecular structure at different times in the simulation. Atomic labels are also shown in Figure 1a. At 0 fs, TATB molecules are basically planar structures, and there are a large number of intramolecular hydrogen bonds between adjacent –NO2 and –NH2 groups, which is the reason for the high melting point and good stability of TATB. Several key structural parameters are also identified in Figure 1a. The excited state is formed after the laser excitation. The main structural change is that the nitro group linked by the C2 atom rotates about the C2–N8 bond axis. The plane of O13–N8–O14 is nearly perpendicular to the plane of the benzene ring at 500 fs, while the displacements of other atoms are negligible. This suggests vibrational activation of the nitro group (hereafter referred to as the “activated nitro group” or ANG). At 2500 fs, the vibrational energy of ANG reaches a maximum, and a pyramidal structure is formed with N8 at the apex. At this time, the molecule decays to the ground state and the ANG rotates around the C2–N8 bond axis as the molecule returns to its original configuration. Significant structural changes were observed only in the ANG, indicating that the excitation is localized on the ANG. CASSCF level calculations of energetic molecules, such as nitrobenzene (NB) [25,26] and dimethylnitramine (DMNA) [27], show that excitation to the lowest excited states (n→π*) is localized at the nitro group; TD-DFT studies of TATB molecules [23] also yielded similar results. The ANG is not necessarily the nitro linked to C2 atom, it could also be another nitro linked to C4 or C6 atoms in the TATB molecule. For instance, the ANG was the nitro linked to the C6 atom in the other photophysical decay trajectory we obtained.
The time evolution of the orbital energies and electronic populations of the HOMO-1, HOMO, LUMO, and LUMO+1 are summarized in Figure 2. Diagrams of the four frontier orbitals are also listed. The electron density in HOMO-1 is mainly distributed between C4 and the linked nitro group and is mainly distributed among C2, C6, and their linked nitro groups in HOMO. Both orbitals exhibit non-bonding (n) character. The electron in LUMO, which shows π* anti-bonding character, is distributed equally among the three NO2 groups, whereas the electron in LUMO-1 is distributed primarily between NO2 at the C4 and C6 positions and their corresponding NH2 groups. The lowest excited state should result from an n→π* excitation, similar to the smaller nitro-containing energetic molecules [25,26,27,28]. It can be seen from Figure 2a that the LUMO energy decreased rapidly after laser irradiation, and HOMO energy increased significantly at the same time. It means the electronic transition from HOMO to LUMO is the dominant excitation. After 500 fs, the energy of HOMO and LUMO fluctuated between −4.4 and −5.8 eV, respectively. At 2500 fs, the energy gap of LUMO–HOMO reaches a minimum of 0.04 eV, and the orbital energies rapidly return to their original values. Figure 2b shows the variation of frontier orbital electron population with time. After laser irradiation, about one electron is excited from HOMO to LUMO, and a small amount goes to the LUMO+1. The excited state decays to ground state after the “crossing” point, as marked in Figure 2a, which corresponds to a LUMO-to-HOMO transition in Figure 2b. The laser frequency used in simulation (3.2 eV) is close to the excitation wavelength used in the experiment (400 nm, i.e., 3.1 eV), suggesting that the density-functional-based tight binding (DFTB) method with parametric quantification is adequate for the calculation of electron energy. Due to a limitation of the Ehrenfest dynamics theory, the excited states decay to a mixed state rather than a pure state, so there is a small amount of electronic population of other orbitals. The simulation shows that with a 100 fs laser duration, the excited state deactivation takes place at 2500 fs. Therefore, the lifetime of the excited state can be regarded as 2400 fs. The experimental results of femtosecond transient spectra [23] show that the excited state lifetime is 6 ps. The calculated lifetime and result experimentally are of the same order of magnitude. Considering the fact that a single TATB molecule was simulated in vacuo without any solvent effect included, it is fair to say the simulation results are consistent with experimental observations.
Figure 3a shows how the C2–N8 bond length changes with time. The average bond length of C2–N8 in the ground state is 1.40 Å, compared with the optimized ground state molecular structure of 1.431 Å at the MP2/6-311++g(2d,2p) level and 1.438 Å as reported in Reference [23]. After laser irradiation, the bond length of C2–N8 extends to an average of 1.46 Å. TD-DFT calculations [23] of TATB excited states show that C2–N8 increases to 1.474 Å in the S1 state. After 2500 fs, the C2–N8 bond returns to its initial value due to the molecular decay to the ground state. In a SERID simulation, the number of electrons at every atom over time is obtained by projecting the electron wavefunction into the atomic orbital. Initially (at 0 fs), the valence electron numbers at C2 and N8 are 4.1 and 3.8 Å due to O (in the ANG) being highly electronegative. After excitation, it is clear from Figure 3b that the number of C2 valence electrons decreases to an average of 3.6 and the number of N8 atom valence electrons increases to an average of 4.3, implying that about 0.5 electrons are transferred from the C2 to the N8 atom. Chu et al. [23] analyzed the charge distributions of the ground state and the first excited state of TATB molecule and found net 0.339-electron from the benzene ring to the ANG due to excitation. According to the frontier orbital properties in Figure 2, the transfer of electrons from C2 to N8 implies that the number of orbital electrons increases at the C–NO2 site, resulting in a decreased bond order, which explains the increased C2–N8 bond length.
It is worth noting that almost all the excited state intramolecular electron transfers happen between C2 and N8 atoms, as shown in Figure 4a. The color map clearly shows charge transfers between N8 (light blue–dark blue color) and C2 (yellow-red color). Slight charge transfers are observed at N7 and N9. The kinetic energy distribution diagram (Figure 4b) shows a significant rise in the kinetic energy of O13 and O14 at 1500 fs. These O atoms are part of the ANG. This suggests that electronic excitation energy is converted to vibrational energy that is mostly concentrated on the ANG. This could possibly lead to reactions that involve C–NO2 bond breaking, N–O bond cleavage, and ring opening. Indeed, the breaking away of O atom(s) from nitro is a precursor of the generally accepted NO2–ONO isomerization mechanism [16,17]. Nitro elimination and formation of fragments by aromatic ring opening observed in other energetic molecules could also happen in TATB.
In order to investigate how vibrations affect deactivation, we compared the changes in all the degrees of freedom with the excitation energy. We found that the pyramidalization of the N8 atom in the ANG plays an important role is the deactivation of the TATB molecule. Figure 5a displays the variation of the C2–O13–O14–N8 dihedral angle and LUMO–HOMO energy gap with time. Initially, with energy gap being 3.2 eV, the C2–O13–O14–N8 dihedral angle is about 0°, indicating that the ANG is in the TATB molecular plane. After excitation, the energy gap decreases gradually, accompanied by an increase in the C2–O13–O14–N8 dihedral vibration, leading to the pyramidal structure shown in Figure 1c. With the concurrent decrease in the LUMO–HOMO energy gap, we can surmise that the pyramidalization of the C2–O13–O14–N8 dihedral angle is the essential vibrational mode associated with the nonradiative deactivation channel. At 2500 fs, the LUMO–HOMO energy gap reaches a minimum of 0.096 eV and the dihedral angle C2–O13–O14–N8 reaches a minimum of −35.2°. The N8 atom deviates from the planarity to the maximum extent at that time. After deactivation, the ANG still retains a large vibrational amplitude, which suggests the conversion of excited electronic energy to vibrational energy. However, the C2–O13–O14–N8 angle has an average value of 0°, suggesting a return to the planar structure. The pyramidal structure has also been found in the (S1/S0) CI structure of non-adiabatic decay [25,26,27,28] of other small energetic molecules, such as NB, TNA, TNT, and DMNA. Figure 5b displays the variation of N8–O13 and N8–O14 distances with time. Starting at 1.296Å, both N8–O13 and N8–O14 bonds lengthen after laser application; large fluctuations of about 1.37 Å are observed between 300 and 2500 fs. Compared to Figure 3 and Figure 4, it was found that in the C2–O13–O14–N8 dihedral angle, N8–O13, N8–O14, and C2–N8 bonds lengths significantly changed from 300 to 2500 fs. This suggests that activation of the ANG involves electron transfer from C2 to N8 and increased vibrational amplitude in the ANG as intramolecular hydrogen bonds are broken. At 2500 fs, the N8–O13 bond reaches a maximum of 1.442 A. N8–O13 and N8–O14 bonds sharply vibrate when electrons decay but still have an average value of 1.3 Å, suggesting a return to ground state values. The increased vibrational amplitudes cause the N–O bonds in the ANG to lengthen to as much as 1.5 Å, suggesting a propensity for N–O bond breaking during deactivation.
Since the Ehrenfest dynamics method ignores electron spin, the SERID method cannot identify the multiplicity of excited states in the simulation. The CASSCF method can optimize the structure of different electronic states and the conical intersection of the adiabatic potential energy surface, but it is difficult to calculate for TATB if one were to include all the π electrons and orbitals. Even if some electrons and orbitals are neglected to reduce the active space, it would be hard to choose which ones to neglect because of the lack of the conical intersection geometry on the potential energy surface. Although the simulation found that the excited state is basically located in the ANG, the degrees of freedom (including pyramidalization of N atom, bending of O–N–O, stretch of N–O, and rotation of nitro, etc.) were still too complex for confirming reaction coordinates. For this study, the ground state and the lowest singlet excited state potential energy surfaces were calculated along the simulation trajectory at the unrestricted TD-B3LYP/cc-pVDZ level to explore a sample strategy to investigate deactivation pathways. The potential energy curves of S0 and S1 are displayed in Figure 6a. We are interested in the deactivation channels, so only the energies of trajectory points between 2400 and 2600 fs are given in Figure 6b. In the potential energy calculation, the ground state energy of the first conformation (i.e., conformation at 0 fs) was assumed to be zero. However, we found the S0 curve reached a minimum of −0.91 eV at 135 fs due to conformation and energy fluctuations in the dynamics simulation. Thus, the geometry at 135 fs should be regarded as S0 minimum. Therefore, both S0 and S1 potential curves should be shifted 0.91 eV upward based on this geometry. The S1 minimum of 1.146 eV at 700 fs should be increased to 2.056 eV. All the energy values of trajectory points are increased by 0.91 eV in Figure 6b. It can be seen from Figure 6b that at 2500 fs, the energy gap between S1 and S0 is at a minimum (0.096 eV); this can be regarded as corresponding to a CI. It also means that the configuration at 2500 fs is very close to the structure of TATB at the CI. Therefore, the decay in the trajectory can be regarded as the S1→S0 non-adiabatic deactivation via CI.
The CI structure is given in Figure 6c. Some critical geometrical parameters are also included. There are significant differences of structure between the ANG and the meta nitro group. In the ground state TATB molecule, the three nitro groups are coplanar with the benzene ring because of the formation of a p–π conjugated system. At the CI, however, the N8 atom is off the plane, forming a pyramid with the C2, O13, and O14 atoms and the N at the meta-nitro group remained in plane. At CI, the bond lengths of C2–N8, N8–O13, and N8–O14 were 1.474, 1.397, and 1.442 Å, respectively; these are essentially sigma bonds. The bond angles C2–N8–O13, C2–N8–O14, and O13–N8–O14 were 112.1, 107.4, and 110.3°, respectively. They are very similar to the bond angles in nonequivalent sp3 hybridization, as in the NH3 molecule. On the other hand, the bond lengths in meta-nitro are shorter than the ones in the ANG and could be regard as π bonds. The bond angles of the meta-nitro are close to 120°, suggesting sp2 hybridization for the N atom. In order to determine the relaxation path involved in the fast internal conversion of TATB from the S1 state, the minimum energy connecting the Franck–Condon (FC) points of S1 and S0 surfaces of TATB to the minimum energy (S1/S0) CI was determined, as seen in Figure 6d. The TD-DFT-calculated vertical excitation energy (here we defined the vertical excitation energy as the S1–S0 energy gap at S0 minimum) for the lowest lying excited electronic states (S1) of TATB is 3.134 eV. The calculated vertical excitation energy is in good agreement with the available experimental [24] excitation wavelength of 400 nm (3.1 eV). Direct comparison of the excitation energy used in experimental work with the calculated vertical excitation energy reveals that TATB is excited below the Franck–Condon (FC) point of the S1 state at 400 nm. Our calculated results are in close agreement with Xiong et al. [23] at the TD-B3LYP/6-311++G (d, p) level. They found that the vertical excitation energy and one S1 minimum were 350.5 kJ·mol−1 (3.63 eV) and 218 kJ·mol−1 (2.26 eV), respectively.

3. Materials and Methods

In this study, the SERID model was used to carry out the dynamics simulation of TATB followed by laser excitation. In the simulation, the electronic motion was calculated by quantum mechanics, while the nuclear motion and radiation field were treated using the classical method. The one-electron states were obtained at each time step by solving the time-dependent Schrödinger equation in a nonorthogonal basis,
i Ψ j t = S 1 · H · Ψ j
where Ψ is the electronic wavefunction, S is the overlap matrix for the atomic orbitals, H is Hamiltonian operator. The electronic motion was calculated using the density functional-based tight binding (DFTB) method. The laser pulse was characterized by a vector potential A, which is coupled to the Hamiltonian via time-dependent Peierls substitution.
H a b ( X X ) = H a b 0 ( X X ) exp ( i q c A · ( X X ) )
Here, H a b ( X X ) is the Hamiltonian matrix element for basis functions a and b on atoms at X and X , respectively, and q = e is the charge of the electron.
The vector potential A comprehensively describes the laser pulse (pulse width, photon frequency, pulse intensity).
A = A ( x , t ) = A 0 cos ( π ( t t 0 2 ) t 0 ) cos ( ω t )
Here, t0 is width of pulse, ω is angular frequency, A0 is a constant.
In this model, the nuclear motion is solved by the Ehrenfest equation of motion,
M l d 2 X l α d t 2 = 1 2 j Ψ j + · ( H X l α i 1 2 S X l α · t ) · Ψ j U r e p / X l α
where U r e p is the effective nuclear–nuclear repulsive potential and X l α = X l a is the expectation value of the time-dependent Heisenberg operator for the α coordinate of the nucleus labeled by l (with α = x, y, z). Equation (4) is obtained by neglecting the terms of second and higher order in the quantum fluctuations X ^ X ^ l α in the exact Ehrenfest theorem.
The Hamiltonian matrix, overlap matrix, and ion–ion repulsive potential are determined in a density-functional-based tight binding (DFTB) method [33,34], which is derived from local density approximation (LDA) functional and does have essentially the same strengths and limitations as TD-DFT. In particular, the bonding is well described but the excited state energies are typically too low. For this reason, as is conventional, we matched the effective central photon energy of the laser pulse to the relevant density-functional (rather than experimental) excitation energy; this should not have a significant effect on the interpretation of the results.
The basis functions used in the present simulations are the 1s atomic orbital of H and the valence s and p orbitals of C. (Spin-up and spin-down states are not distinguished.) The Hamiltonian matrix elements and overlap matrix elements are calculated for a dimer for interatomic distances on a relevant scale and the results from the calculations can then be tabulated and employed in time-dependent simulations.
The weakness of the Ehrenfest method is that it amounts to averaging over all the terms in the Born–Oppenheimer expansion,
Ψ t o t a l ( X n , x e , t ) = i Ψ i n ( X n , t ) Ψ i e ( x e , X n )
rather than following the time evolution of a single term—i.e., a single potential energy surface—which is approximately decoupled from all the others. (Here, X n and x e represent the sets of nuclear and electronic coordinates, respectively, and the Ψ i e values are eigenstates of the electronic Hamiltonian at fixed X n .) The strengths of the present approach include the retention of all of the 3N nuclear degrees of freedom (instead of only the 2 or 3 that are typically considered in a potential energy surface calculation) and the incorporation of both the excitation due to a laser pulse and the subsequent de-excitation at an avoided crossing near a conical intersection.
Because the SERID model is based on a mean field theory, it is not clear which electron is excited to a specific electronic excited state. Therefore, it is necessary to assume a reasonable quantum chemistry model. In this study, a TD-DFT method was used to calculate the energy of ground state and excited state every 100 fs in a simulated trajectory (at 10 fs intervals near the deactivation point) to obtain the potential energy surface. This calculation helps to elucidate the non-adiabatic transition of TATB molecules.
The Gaussian 09 quantum chemical package (Gaussian, Inc., Wallingford, CT, 2009) was used for geometry optimizations and TD-DFT calculations. The VMD [35] open source program was used to visualize the molecular structure.

4. Conclusions

In this paper, a photophysical deactivation pathway of laser-excited TATB molecules is reported. The vibration and pyramidalization of the ANG is the key factor of deactivation. Electron transfer from the benzene ring to the ANG was observed. The charge transfer leads to the obvious increase of the C–NO2 bond and formation of the C–N σ bond. It is the reason why the ANG can rotate freely in the excited state. The energies of the S0 state and the S1 excited state of each configuration in the trajectory were calculated by the TD-DFT method. It was found that the energy gap of the nonradiative deactivation in the simulation was 0.096 eV, close to the conical intersection. The vertical excitation energy and S1 minimum were also close to experimental values and other theoretical results. The S0–S1 excitation is localized at the ANG, and the electronic excitation energy is converted to vibrational energy that is mostly concentrated on the ANG. Drastic vibrations of C–NO2 and N–O bonds in the ANG suggest that nitro elimination and N–O bond cleavage could happen in TATB subjected to laser irradiation.

Author Contributions

Methodology, S.Y. and Y.D.; Software, S.Y. and Y.D.; Investigation, W.Z., J.S., J.C. and S.G.; Writing-Original Draft Preparation, W.Z. and S.Y.; Writing-Review & Editing, Y.D. and G.L.

Funding

This research was funded by [Pre-research Fund (6140656020206) and Natural Science Foundation of Chongqing, China (cstc2015jcyjA10037, cstc2016jcyjA0588) and Chongqing Municipal Education Commission Foundation (KJ1500432, KJ160404).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Field, J.E. Hot spot ignition mechanisms for explosives. Acc. Chem. Res. 1992, 25, 489–496. [Google Scholar] [CrossRef]
  2. Zhang, S.Q.; Wang, Y.Q.; Zhen, X.M. Density functional theory investigation of the photoisomerization reaction of nitroalkanes and nirroaromatic compounds. Acta Phys.-Chim. Sin. 2006, 22, 1489–1494. [Google Scholar]
  3. Bhattacharya, A.; Guo, Y.Q.; Bernstein, E.R. Nonadiabatic reaction of energetic molecules. Acc. Chem. Res. 2010, 43, 1476–1485. [Google Scholar] [CrossRef] [PubMed]
  4. Yu, H. Environmental carcinogenic polycyclic aromatic hydrocarbons: Photochemistry and phototoxicity. J. Environ. Sci. Health C 2002, 120, 149–183. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, C. Review of the establishment of nitro group charge method and its applications. J. Hazard. Mater. 2009, 161, 21–28. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, C.; Ma, Y.; Jiang, D. Charge transfer in TATB and HMX under extreme conditions. J. Mol. Model. 2012, 18, 4831–4841. [Google Scholar] [CrossRef] [PubMed]
  7. Bhattacharya, A.; Guo, Y.Q.; Bernstein, E.R. Experimental and theoretical exploration of the initial steps in the decomposition of a model nitramine energetic material: Dimethylnitramine. J. Phys. Chem. A 2009, 113, 811–823. [Google Scholar] [CrossRef] [PubMed]
  8. Bhattacharya, A.; Guo, Y.Q.; Bernstein, E.R. Unimolecular decomposition of tetrazine-N-oxide based high nitrogen content energetic materials from excited electronic states. J. Chem. Phys. 2009, 131, 194304. [Google Scholar] [CrossRef] [PubMed]
  9. Bhattacharya, A.; Guo, Y.Q.; Bernstein, E.R. Decomposition of excited electronic state s-tetrazine and its energetic derivatives. J. Chem. Phys. 2011, 134, 024318. [Google Scholar]
  10. Yuan, B.; Bernstein, E.R. Initial mechanisms for the unimolecular decomposition of electronically excited bisfuroxan based energetic materials. J. Chem. Phys. 2017, 146, 014301. [Google Scholar] [CrossRef] [PubMed]
  11. Yuan, B.; Bernstein, E.R. Initial mechanisms for the unimolecular decomposition of electronically excited nitrogen-rich energetic salts with tetrazole rings: (NH4)2BT and TAGzT. J. Chem. Phys. 2016, 145, 064306. [Google Scholar] [CrossRef]
  12. Yuan, B.; Yu, Z.; Bernstein, E.R. Initial mechanisms for the decomposition of electronically excited energetic materials: 1,5′-BT, 5,5′-BT, and AzTT. J. Chem. Phys. 2015, 142, 124315. [Google Scholar] [CrossRef] [PubMed]
  13. Yuan, B.; Yu, Z.; Bernstein, E.R. Initial decomposition mechanism for the energy release from electronically excited energetic materials: FOX-7 (1,1-diamino-2,2-dinitroethene, C2H4N4O4). J. Chem. Phys. 2014, 140, 074708. [Google Scholar] [CrossRef] [PubMed]
  14. Yu, Z.; Bernstein, E.R. On the Decomposition Mechanisms of New Imidazole-Based Energetic Materials. J. Phys. Chem. A 2013, 117, 1756–1764. [Google Scholar] [CrossRef] [PubMed]
  15. Yu, Z.; Bernstein, E.R. Experimental and theoretical studies of the decomposition of new imidazole based energetic materials: Model systems. J. Chem. Phys. 2012, 137, 114303. [Google Scholar] [CrossRef] [PubMed]
  16. Soto, J.; Arenas, J.F.; Otero, J.C.; Peláez, D. Effect of an S1/S0 conical intersection on the chemistry of nitramide in its ground state. A comparative CASPT2 study of the nitro-nitrite isomerization reactions in nitramide and nitromethane. J. Phys. Chem. A 2006, 110, 8221–8226. [Google Scholar] [CrossRef] [PubMed]
  17. Arenas, J.F.; Otero, J.C.; Peláez, D.; Soto, J. CASPT2 Study of the Decomposition of Nitrosomethane and Its Tautomerization Reactions in the Ground and Low-Lying Excited States. J. Org. Chem. 2006, 71, 983–991. [Google Scholar] [CrossRef] [PubMed]
  18. Ghosh, J.; Gajapathy, H.; Konar, A.; Narasimhaiah, G.M.; Bhattacharya, A. Sub-500 fs electronically nonadiabatic chemical dynamics of energetic molecules from the S1 excited state: Ab initio multiple spawning study. J. Chem. Phys. 2017, 147, 204302. [Google Scholar] [CrossRef] [PubMed]
  19. Britt, A.D.; Moniz, W.B.; Chingas, G.C. Free Radicales of TATB. Propellants Explos. Pyrotech. 1981, 6, 94–95. [Google Scholar] [CrossRef]
  20. Williams, D.L.; Timmons, J.C.; Woodyard, J.D. UV-Induced Degradation Rates of 1,3,5-Triamino-2,4,6-Trinitrobenzene (TATB). J. Phys. Chem. A 2003, 107, 9491–9494. [Google Scholar] [CrossRef]
  21. Ao, Y.Y.; Chen, J.; Song, H.T.; Shi, J.M.; Liu, Y.; Chen, H.B.; Li, J.Q.; Peng, J. Mechanism of Radiation-induced Colour Change in 1,3,5-Triamino-2,4,6-trinitrobenzene under γ-Rays. Chin. J. Energy Mater. 2017, 25, 540–545. [Google Scholar]
  22. Sun, Y. Photo-Induced Decomposition of HNS. Ph.D. Dissertation, Nanjing University of Science and Technology, Nanjing, China, 2013. [Google Scholar]
  23. Xiong, Y.; Liu, J.; Zhong, F.C.; Xu, T.; Cheng, K.M. Identification of the Free Radical Produced in the Photolysis of TATB. J. Phys. Chem. A 2014, 118, 6858–6863. [Google Scholar] [CrossRef] [PubMed]
  24. Chu, G.; Lu, F.; Xin, J. Excited-state dynamics and electron transfer process of 1,3,5-triamino-2,4,6-trinitrobenzene. RSC Adv. 2016, 6, 55560–55567. [Google Scholar] [CrossRef]
  25. Quenneville, J.; Greenfield, M.; Moore, D.S.; McGrane, S.D.; Scharff, R.J. Quantum chemistry studies of electronically excited nitrobenzene, TNA, and TNT. J. Phys. Chem. A 2011, 115, 12286–12297. [Google Scholar] [CrossRef] [PubMed]
  26. Giussani, A.; Worth, G.A. Insights into the complex photophysics and photochemistry of the simplest nitroaromatic compound: A CASPT2//CASSCF study on nitrobenzene. J. Chem. Theory Comput. 2017, 13, 2777–2788. [Google Scholar] [CrossRef] [PubMed]
  27. Mewes, J.M.; Jovanovic, V.; Marian, C.M.; Dreuw, A. On the molecular mechanism of non-radiative decay of nitrobenzene and the unforeseen challenges this simple molecule holds for electronic structure theory. Phys. Chem. Chem. Phys. 2014, 16, 12393–12406. [Google Scholar] [CrossRef] [PubMed]
  28. Guo, Y.; Bhattacharya, A.; Bernstein, E.R. Ultrafast S1 to S0 internal conversion dynamics for dimethylnitramine through a conical intersection. J. Phys. Chem. A 2011, 115, 9349–9353. [Google Scholar] [CrossRef] [PubMed]
  29. Makhov, D.V.; Symonds, C.; Fernandez-Alberti, S.; Shalashilin, D.V. Ab initio quantum direct dynamics simulations of ultrafast photochemistry with Multiconfigurational Ehrenfest approach. Chem. Phys. 2017, 493, 200–218. [Google Scholar] [CrossRef]
  30. Dou, Y.S.; Torralva, B.Z.; Allen, R. Semiclassical electron-radiation-ion dynamics (SERID) and cis–trans photoisomerization of butadiene. J. Mod. Opt. 2003, 50, 2615–2643. [Google Scholar] [CrossRef]
  31. Dou, Y.S.; Torralva, B.; Allen, R.E. Interplay of electronic and nuclear degrees of freedom in a femtosecond-scale photochemical reaction. Chem. Phys. Lett. 2004, 392, 352–357. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  33. Porezag, D.; Frauenheim, T.; Köhler, T.; Seifert, G.; Kaschner, R. Construction of tight-binding-like potentials on the basis of density-functional theory: Application to carbon. Phys. Rev. B 1995, 51, 12947–12957. [Google Scholar] [CrossRef]
  34. Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, T.; Suhai, S.; Seifert, G. Self-consistent-charge density-functional tight-binding method for simulations of complex materials properties. Phys. Rev. B 1998, 58, 7260–7268. [Google Scholar] [CrossRef]
  35. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
Figure 1. Snapshots for the deactivation of the excited state of the 1,3,5-triamino-2,4,6-trinitrobenzene (TATB) molecule following irradiation with a laser pulse of 3.2 eV photon energy and fluence of 32.94 J·m−2. Cyan: C, blue: N, red: O, white: H.
Figure 1. Snapshots for the deactivation of the excited state of the 1,3,5-triamino-2,4,6-trinitrobenzene (TATB) molecule following irradiation with a laser pulse of 3.2 eV photon energy and fluence of 32.94 J·m−2. Cyan: C, blue: N, red: O, white: H.
Molecules 23 01593 g001
Figure 2. Variations with time of (a) the orbital energy and (b) electron populations of the HOMO-1, HOMO, LUMO, and LUMO+1 of TATB molecule.
Figure 2. Variations with time of (a) the orbital energy and (b) electron populations of the HOMO-1, HOMO, LUMO, and LUMO+1 of TATB molecule.
Molecules 23 01593 g002
Figure 3. Variations with time of (a) C2–N8 bond length and (b) valence electrons of C2 and N8 atoms in the activated nitro group.
Figure 3. Variations with time of (a) C2–N8 bond length and (b) valence electrons of C2 and N8 atoms in the activated nitro group.
Molecules 23 01593 g003
Figure 4. Distribution contours of (a) charge and (b) kinetic energy of each atom of TATB at current excited state. Atomic labeling displayed in the x-axis is the same as in Figure 1a. The hydrogen atoms were ignored.
Figure 4. Distribution contours of (a) charge and (b) kinetic energy of each atom of TATB at current excited state. Atomic labeling displayed in the x-axis is the same as in Figure 1a. The hydrogen atoms were ignored.
Molecules 23 01593 g004
Figure 5. (a) Comparison between the energy gap of LUMO–HOMO and N8 atom pyramidalization; (b) variations with time of N8–O13 and N8–O14 bond lengths.
Figure 5. (a) Comparison between the energy gap of LUMO–HOMO and N8 atom pyramidalization; (b) variations with time of N8–O13 and N8–O14 bond lengths.
Molecules 23 01593 g005
Figure 6. (a) S0 and S1 potential energy surface along with simulation trajectory from TD-B3LYP/cc-pVDZ calculation; (b) the detailed variations of S0 and S1 potential energy surface near the deactivation point (relative to the S0 minimum (−0.91 eV) shown in (a); (c) the structure at the conical intersection (CI); (d) the minimum energy pathway from the Franck–Condon (FC) point to the CI.
Figure 6. (a) S0 and S1 potential energy surface along with simulation trajectory from TD-B3LYP/cc-pVDZ calculation; (b) the detailed variations of S0 and S1 potential energy surface near the deactivation point (relative to the S0 minimum (−0.91 eV) shown in (a); (c) the structure at the conical intersection (CI); (d) the minimum energy pathway from the Franck–Condon (FC) point to the CI.
Molecules 23 01593 g006

Share and Cite

MDPI and ACS Style

Zhang, W.; Sang, J.; Cheng, J.; Ge, S.; Yuan, S.; Lo, G.V.; Dou, Y. A Photophysical Deactivation Channel of Laser-Excited TATB Based on Semiclassical Dynamics Simulation and TD-DFT Calculation. Molecules 2018, 23, 1593. https://doi.org/10.3390/molecules23071593

AMA Style

Zhang W, Sang J, Cheng J, Ge S, Yuan S, Lo GV, Dou Y. A Photophysical Deactivation Channel of Laser-Excited TATB Based on Semiclassical Dynamics Simulation and TD-DFT Calculation. Molecules. 2018; 23(7):1593. https://doi.org/10.3390/molecules23071593

Chicago/Turabian Style

Zhang, Wenying, Jian Sang, Jie Cheng, Siyu Ge, Shuai Yuan, Glenn V. Lo, and Yusheng Dou. 2018. "A Photophysical Deactivation Channel of Laser-Excited TATB Based on Semiclassical Dynamics Simulation and TD-DFT Calculation" Molecules 23, no. 7: 1593. https://doi.org/10.3390/molecules23071593

Article Metrics

Back to TopTop