About the Aromaticity of symm-Triaminotrinitrobenzene - The Journal

Feb 27, 2019 - ... dynamics at the Car–Parrinello level. Different criteria of the aromaticity were combined with the study of conformational flexib...
0 downloads 0 Views 316KB Size
Subscriber access provided by WEBSTER UNIV

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

About the Aromaticity of Symm-triaminotrinitrobenzene Iryna V. Omelchenko, Oleg V Shishkin, Przemyslaw Dopieralskie, and Zdzislaw Latajka J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.9b00433 • Publication Date (Web): 27 Feb 2019 Downloaded from http://pubs.acs.org on March 1, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

About the Aromaticity of Symmtriaminotrinitrobenzene Iryna V. Omelchenkoa,b,*, Oleg V. Shishkina,b,†, Przemyslaw Dopieralskic, Zdzislaw Latajkac a Department

of X-ray Diffraction Study and Quantum Chemistry, SSI “Institute for Single

Crystals” NAS of Ukraine, 60 Nauky ave., Kharkiv 61072, Ukraine b

Faculty of Chemistry, V.N.Karazin Kharkiv National University, 4 Svobody sq., Kharkiv

61077, Ukraine c

Faculty of Chemistry, University of Wroclaw, F. Joliot-Curie 14, 50-383 Wroclaw, Poland

*

Corresponding author. E-mail: [email protected]. Phone/Fax: +380573410258



Deceased.

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

Abstract Aromaticity and structural features of the isolated symm-triaminotrinitrobenzene (TATB) were examined using the non-empirical ab initio quantum chemical method and molecular dynamics at the Car-Parrinello level. Different criteria of the aromaticity were combined with the study of conformational flexibility of molecule and analysis of the electron density distribution. It was found that cooperative effect of the resonance-assisted hydrogen bonds results in the ultimate decreasing aromaticity of the benzene ring in TATB. Values of HOMA index indicate that it could be classified as low-aromatic in equilibrium state at zero temperature but completely nonaromatic at the room temperature. An extremely high flexibility of the molecule is also not typical for aromatic rings. The electron delocalization in H-bonded O=N–C=C–N–H quasiaromatic rings was found to be greater than in the benzene ring of TATB.

ACS Paragon Plus Environment

2

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction Symm-triaminotrinitrobenzene (TATB) is a well-studied compound known for its non-linear optical and explosive properties combined with the extraordinary thermal, impact, and shock stability in the solid state1,2. Its’ individual molecular structure overcrowded with both π-donor and π-acceptor substituents is the subject of the particular theoretical interest because of the strong intramolecular push-pull interactions3–5. One of the main questions about the structure concerns the aromatic interactions in the TATB molecule. They are associated with the tendency to the tautomerization and the proton transfer, the properties are believed to be the main factors that affect impact sensibility and explosion dynamics6,7. Also the aromaticity interferes with the charge-transfer effect that is responsible for the non-linear optical (NLO) properties8,9. Although the degree of aromaticity of some well-known conjugated molecules, as well as its quantitative criteria, is still under discussion10,11, all benzene derivatives have always been an islet of consistency in the turbulent sea of the aromatic definitions, undoubtedly treated as highly aromatic substances10,12–14. However, the case of TATB seems to need deeper investigation of the matter. One of the most remarkable features of its structure is the number of strong intramolecular resonance-assisted hydrogen bonds15–17 between amino and nitro substituents18. It is known that influence of such intramolecular H-bonding on the aromaticity and π-electron delocalization could be rather strong19–24, especially for the ortho-substituted rings with strong H-bonding between substituents. In some particular non-aromatic conjugated systems, such as malonaldehyde or malondialdehyde derivatives, the electron delocalization in the hydrogenbonded quasi-aromatic rings25 could be comparable to the delocalization in the aromatic πsystems26–28. In the previous studies of the structure and aromaticity of the series of

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 30

aminonitrobenzenes29 it was found that the degree of benzene ring aromaticity decreases essentially with the increasing of the number of such hydrogen bonds between amino and nitro groups, mutually enhancing each other in the non-additive manner. The substituent effect on aromatic ring could also be strongly affected by involving the substituent into intermolecular hydrogen bonding21. Structural consequences of the intramolecular push-pull effect in aminonitrobenzenes have also been associated with the H-bonding, too29. Also it is known that conformations of overcrowded benzene derivatives with 5 to 6 substituents could be strongly affected direct interactions between substituents, either attractive or repulsive, and it could result in strong deformation of benzene ring from planar geometry30,31. The planarity of the benzene ring is known as a property – and the necessary condition – of the aromaticity of cyclic conjugated system, however it was already shown that small deviation from planarity does not disrupt aromatic conjugation and could be used as a measure of the degree of aromaticity32,33, and the Car-Parrinello molecular dynamics (CMPD) studies revealed that the slightly non-planar conformations are even typical for aromatic molecules including benzene at the room temperature in gas phase34,35. Presence of several substituents could force the major out-of-plane deformation unfavorable for aromatic delocalization31. Thus, the TATB molecule contains six substituents that could affect properties of benzene ring by push-pull effect, by resonance-assisted hydrogen bonding and by sterical hindrance, so that is the question in what extent the direct intramolecular interactions between substituents could disturb the persistent electronic structure of the benzene ring. The particular issue concerns limitations of methods that are reasonable for investigation of aromaticity-related properties of TATB. The experimental structural studies of individual TATB molecule are still complicated due to the explosibility of the compound and its extraordinary

ACS Paragon Plus Environment

4

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

stability in the crystalline phase2. Thus the most precise experiments were carried out for the substance in the solid state5,36. However, the molecular structure is obligatory affected with the intermolecular interactions that are observed in the crystals and solutions. Effects of the crystal packing could introduce essential corrections to the structure of individual molecule5,37,38 . That seems to be important for TATB molecule for two main reasons: (a) the intermolecular hydrogen bonds in the crystal compete with intramolecular5 and may affect their influence, and (b) the overcrowded nitroaromatic molecules are known to be rather flexible31,39 but the TATB planarity may be enforced by the crystal surrounding. Also, it is usually not possible to measure the physico-chemical properties of individual molecule in the crystal directly due to the collective effects. The quantum chemical studies is the perspective way to investigate the TATB properties, both individual and collective, however it is known to be very sensible against the computational methods40. Therefore some of the results obtained in the previous studies of TATB (HOMA and NICS aromaticity values with MP2/cc-pVDZ method, as well as flexibility investigation29) need the careful revision. In the present study, we examined the structure, conformational properties, NICS indices, and intramolecular interactions in the TATB molecule in order to ascertain its actual aromaticity degree and the degree of changes of structural and electronic features associated with the aromaticity. Computational Details Ab initio calculations: the structure of isolated TATB molecule has been optimized using the Møller-Plesset second order perturbation theory (MP2)41 with series of Dunning correlationconsistent basis sets42 and also using the DFT M06-2X method43. The character of stationary point on potential energy surface was checked by calculations of the Hessian matrix at the M062X level of theory only, no negative eigenvalues of the Hessian were found. The deformation

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 30

energy at zero temperature was defined as increasing of the total energy of the molecule upon the relaxed potential energy surface (PES) scan along the endocyclic torsion angles by the given angle value, from 0° to 30°, as described in previous studies14,33. PES scans were performed at the MP2/cc-pVTZ level. Since the molecule is not strictly symmetrical in calculations, different angles revealed slightly different values of deformation energies, but only the smallest values were taken as “deformation energy”. Amino group pyramidality was calculated as the sum of bond angles centered on the corresponding nitrogen atom. Aromaticity was estimated by HOMA aromaticity index44 and values of the z-component of nuclear independent chemical shifts at the point located 1 Å above the ring critical point, NICS(1)zz45, the last calculated at the M062X/cc-pVTZ level. Considering that the molecule is not perfectly planar in that calculation, the mean plane of the non-hydrogen atoms were taken as (xy) plane, and the results for the points 1 Å above and below the plane were averaged out. AIM analysis46 was performed with AIMALL program47 using the wavefunction from MP2/aug-cc-pVTZ calculation, and the NBO analysis48 was performed at the MP2/cc-pvtz (orbital decomposition, Wiberg bond orders) and M06-2X/ccpVTZ (atomic charges and E(2) interaction energies) levels. CPMD calculations: all calculations are based on ab initio molecular dynamics49 using the efficient Car–Parrinello50 propagation scheme as implemented in the CPMD program package51. These pseudopotential calculations have been carried out using the PBE52 exchange–correlation functional within the spin–restricted Kohn–Sham formalism together with a plane-wave basis set at a kinetic energy cutoff of 100 Ry, gamma–point sampling of the Brillouin zone, and Troullier– Martins53 normconserving pseudopotentials. The supercell for all these calculations was a cubic box 16 Å in length. All dynamic simulations were performed in the canonical ensemble at 298 K using Nosé–Hoover chain thermostats54 in order to control the kinetic energy of the nuclei as

ACS Paragon Plus Environment

6

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

well as the fictitious kinetic energy of the orbitals. A molecular dynamics time step of Δt = 3 a.u. (≈ 0.073 fs) was used for the integration of the Car–Parrinello equations of motion using a fictitious mass parameter for the orbitals of 400 a.u. together with the proper atomic masses. After initial equilibration period (ca. 30 000 steps) the Car-Parrinello molecular dynamics simulations were performed and the data were collected over trajectories spanning 300 000 steps (ca. 22 ps). Additional processing of the experimental X-ray diffraction data of the TATB molecule taken from the CSD database55 was accomplished in order to normalize all N–H distances (N–H bonds were fixed at 1.015 Å) except the data sets containing already normalized N–H distances. All experimental N–H…O bond lengths were given for the normalized N–H distances. Benchmark of reproducibility of computations results. It should be noted that the geometrical parameters at the equilibrium point crucially depend on the basis set but much less depend on the method (Table 1). Unfortunately the effect of the basis set extending and improving the method cannot be verified due to the lack of experimental data in the gas phase. One can see the drastic shortening of the endocyclic C–C bonds as well as C-NH2 and C-NO2 bonds with extending the basis set from double-zeta to triple-zeta. Values of the torsion angles are strongly affected by the presence of the diffuse functions in the basis set but that effect is much less pronounced in the triple-zeta basis. It is possibly due to the intramolecular basis set superposition error56, so in this case at least the triple-zeta basis is obligatory for the correct estimation of the geometrical parameters: the smaller sets presumably would give the erroneous results. Also, it is worth mention that in all calculations, the TATB molecule adopted the C1 symmetry. Here and further in Results&Discussion section, we give the range of the values of

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 30

geometrical parameters if they differ for different positions in molecule but merge them in one value if the difference is negligible. Table 1. Selected structural parameters of the TATB molecule (bonds lengths and H-bonds distances, Å, torsion angles, deg, amino groups pyramidalities Ʃ(NH2), deg) estimated using different basis sets and methods at the equilibrium point in vacuum. method/basis C–C

C–NH2

C–NO2

H…O

C-C-C-C

C-C-N-O

Ʃ(NH2)

MP2/ccpvdz

1.443

1.336

1.445

1.702 - 1.705

5.7

7.8 - 14.6

360.0

MP2/aug-cc1.445 pvdz

1.338

1.436

1.704

0.0

10.4

360.0

MP2/cc-pvtz

1.430

1.329

1.431

1.719

2.3 - 2.4

15.4 - 18.1 360.0

MP2/aug-cc1.432 pvtz

1.329

1.428

1.714 - 1.716

0.1 - 0.3

15.2 - 15.8 360.0

M062X/aug-ccpvtz

1.325

1.435

1.760

0.1 - 0.3

19.5

1.431

360.0

Results and discussion Selected geometrical parameters and HOMA aromaticity index of single molecule, obtained at the equilibrium point (MP2/aug-cc-pVTZ results), in the dynamical motion at room temperature (CPMD results), and taken from the single crystal X-ray diffraction study36 are given in the Table 2. For CPMD results, mean HOMA value was calculated as the mean value of HOMA index at every point of trajectory. 1. Molecular geometry, intramolecular hydrogen bonding, and aromaticity. Specific features of molecular geometry are among the most important evidences of the aromatic character of the ring10. The structural properties of TATB molecule indicate its drastic difference

ACS Paragon Plus Environment

8

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

from benzene as well as mono- and disubstituted benzene derivatives having the same substituents (aniline, nitrobenzene, ortho-nitroaniline). Endocyclic C–C bond lengths (Table 2) are quite far from the C–C bond lengths for benzene (ca. 1.396 Å), Cipso–Cα bonds in aniline, nitrobenzene and ortho-nitroaniline (ca. 1.395 – 1.410 Å). These bonds are longer in CPMD simulation, likely due to the dynamical vibration motion of the molecule that is not considered in equilibrium point calculations, but both CMPD and ab initio C–C bond lengths fit the range of the bond lengths observed in the crystal. Presence of the neighboring substituents in benzene ring causes elongation of C–C bonds, however, the magnitude of the elongation is much smaller (Cipso–Cα bond length in ortho-nitroaniline with the same pair of substituents is 1.410 Å). Table 2. Geometrical details and aromaticity of TATB (covalent and non-covalent bonds lengths, Å, torsion angles, deg, amino groups pyramidality Ʃ(NH2), deg, and HOMA aromaticity index44 derived from C–C bond lengths) estimated with different methods method

C-C

C-NH2

C-NO2

H…O

C-C-C-C

C-C-N-O

equilibrium

1.432

1.329

1.428

1.715

0.1 - 0.3

15.2 - 15.8 360.0

0.50

1.437

1.7751.786

8.5 14.3

13.5 - 18.9 357.8

0.06

experiment 1.435- 1.309(crystal)[a]36 1.450 1.319

1.4171.422

1.6931.791

1.575.00

3590.70 - 3.22 360

0.24

experiment (crystal)5

1.4141.420

1.6951.720

2.505.78

1.37 - 3.23 360

0.08

CPMD

1.441

1.335

1.445- 1.3181.449 1.321

Ʃ(NH2)

HOMA

-

[a] Experimental geometry with renormalized N–H distances. At the same time, the values of the exocyclic C–N bonds in TATB (1.309 – 1.335 Å for NH2 group and 1.417 – 1.437 Å for NO2, depending on the method) is dramatically smaller than values of similar bonds in all isolated isomeric nitroanilines (1.370 – 1.390 Å and 1.456 – 1.470 Å, respectively). Amino group is planar or almost planar in isolated molecule as well as

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 30

molecule in the crystal, and the whole molecule is slightly bent in both calculations (C-C-N-O torsion angles are 13.5 – 18.9°) but it is planar in the crystal, perhaps due to the extremely low energy of nitro group rotation39,40 that makes the molecular shape very sensitive to the features of intermolecular interaction, like hydrogen bonds and stacking interactions1,57 . The structural aromaticity index HOMA of the benzene ring of TATB is extremely low. Even the highest value of 0.50 obtained by QC modeling at equilibrium geometry and zero temperature is more typical for semi-aromatic and non-aromatic heterocycles44 rather than for benzene derivatives58. Moreover, the dynamical aromaticity of TATB varies from -1.87 to 0.82 (Fig. 1) with the most probable value of 0.06. Thus one can state that at the room temperature in gas phase, more than 95% of TATB molecules are completely non-aromatic.

Figure 1. The HOMA distribution in TATB from the CPMD simulation To verify these conclusions, also the magnetic NICS(1)zz indices were calculated for the molecule in equilibrium geometry. The values in the projections of ring critical points of the benzene ring and all six H-bonded rings were investigated in the MP2/aug-cc-pVTZ geometry

ACS Paragon Plus Environment

10

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

with M06-2X method. The value for the benzene ring is -7.85 ppm that corresponds to the semiaromatic or even non-aromatic ring (it equals to ca. -35 for benzene and -23 ÷ -35 for the most aromatic structures45). 2. Intramolecular hydrogen bonding. The presence of the resonance-assisted hydrogen bonds is known to have substantial effect on the structure of the conjugated molecules26,27,59. Particularly it could reinforce the electron localization due to the positive non-additive interaction with the charge-transfer effect60 that is known to have notable influence on the equilibrium electronic structure of TATB8. The lengths of the H…O contacts in the intramolecular N–H…O bonds (Table 2) indicate strong H-boding. Interestingly, the range of the H…O distances in the crystal embraces the range derived from ab initio and CPMD calculations. They are notably shorter than the H-bond in ortho-nitroaniline (~1.95 Å29,61). At the time, C– NH2 and C–NO2 bonds are notably shortened too but C–C endocyclic bonds are significantly elongated (Table 2) as compare to mono- or disubstituted benzene derivatives with amino and nitro substituents14,29,61. All these are well known structural features of the resonance-assisted Hbonding62 indicating strong π-conjugation of electron density localized in the hydrogen-bonded ring H–N–C–C–N–O, so that ring could be defined as a quasi-aromatic. The QTAIM study of the properties of hydrogen bond critical points reveals their high density values (~0.048 e/a03) close to experimentally observed (0.044 … 0.047 e/a03)5 and low but negative Laplacian (~ 0.039 e/a05) that slightly differs from experimental results (+0.184 … +0.200 e/a05)5. That indicates the considerable contribution of covalent bonding. Energy of each of these intramolecular H-bond estimated following the Espinosa equation63 is 16.4 – 16.5 kcal/mol that almost perfectly matches results from experimental electron density analysis (15.1 – 15.8

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 30

kcal/mol)5. Therefore the cumulative energy of intramolecular H-bonds is about 98 kcal/mol. That is three times above the known resonance energy in the benzene ring (ca. 36 kcal/mol58). NICS(1)zz values calculated for quasi-aromatic H-bonded rings are about -11.4 that indicates higher π-electron delocalization than in carbocyclic ring, whereas the ring size is approximately the same (the 1…4 N–N distance in the H-bonded ring is 2.81 Å while the 1…4 C–C distance in the carbocyclic ring is 2.86 Å). One can state that π-delocalization in H-bonded rings of TATB is higher than in the carbocyclic ring. Thus the key energetic factor that defines the properties of the TATB molecule is not the benzene aromaticity but the intramolecular H-bonding. 2. Conformational flexibility. It was previously found that the isolated benzene and its derivatives are moderately flexible: the energy of deformation of any torsion angle by 30° is about 7 kcal/mol

14

for benzene, slightly dropping down with the aromaticity decrease14,29,33.

Also it was found that the vibrations at room temperature drastically increase the degree of nonplanarity for all aromatic rings34,35,64. The TATB molecule is extremely flexible as compared to benzene. Its deformation energy is 1.8 kcal/mol at zero temperature. Another particular feature of this potential energy surface (PES) is it's anharmonicity (Fig. 2). The dependence between total energy gain and the value of the endocyclic torsion angle for highly aromatic cycles is commonly strictly harmonic (follows the E = kφ2 equation; the corresponding ideal parabolic curve is denoted with a red color on the fig. 2)14,33. One can see that PES of TATB doesn't meet this rule: is has a plateau in the range of ±15° with almost constant value of energy that is not typical for benzene derivatives14 but is rather common for non-aromatic partially saturated rings65,66.

ACS Paragon Plus Environment

12

Page 13 of 30

2.0 1.8 1.6 1.4

Energy (kcal/mol)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1.2 1.0 0.8 0.6 0.4 0.2 0.0 -0.2 -30

-20

-10

0

10

20

30

Torsion angle (deg)

Figure 2. Changes of the energy with the change of the endocyclic torsion angle at equilibrium point. The red curve plots the harmonic energy dependence inherent to highly aromatic rings. CMPD study reveals that at the room temperature TATB is far more flexible than benzene34 (see Fig. 3). The module of each single torsion angle could reach 35°, and the puckering amplitude67 changes from 0 to 0.6. The mean puckering amplitude S is 0.27 here, while it does not exceed 0.16 in benzene and other aromatic rings34,35. That value indicates rather strong puckering: the approximate limit of the near-planar configuration is S~0.1 that corresponds to torsion values ~5°. The typical conformation is far from planar, only 5.6% of the molecules are planar at every moment, puckering parameters67 indicate that 48.9% adopt slightly flattened twist-boat and 27.9% - flattened boat conformation. That agrees well with the extremely low aromaticity in the gas phase and explains a great difference between the values of torsion angles estimated with different methods: the energy of deformation is so small that even the minimal influence or negligible computational error can change the torsion angles up to 20°.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

12

Relative Frequency, %

10

8

6

4

2

0 0

5

10

15

20

25

30

35

Torsion, deg

7 6

Relative Frequency, %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 30

5 4 3 2 1 0 0.0

0.1

0.2

0.3

0.4

0.5

0.6

S (puckering parameter)

Figure 3. Distribution of the torsion angle C(1)–C(2)–C(3)–C(4) (module value) and ring puckering amplitude S at the room temperature according to the CPMD data. 3. Orbital decomposition and intramolecular conjugation: NBO and AIM results. The constructing of the natural bonding orbitals for the TATB molecule leads to unexpected results. The NBO algorithms of the natural localized molecular orbitals (NLMO) threat its molecular structure in a completely different way than the common aromatic benzene derivatives. While the most derivatives retain the orbital structure of the parent benzene ring with both σ and π orbitals at C–C bonds, only the σ orbitals at the C–C bonds were located in the TATB molecule

ACS Paragon Plus Environment

14

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

indicating dramatic weakness of these bonds compared to benzene C–C bonds. However, the substituents effect is not strong enough to form the radialene-like picture of the π orbitals, and all C–N bonds are treated as σ bonds too. This is supported by the values of the Wiberg bond orders (1.14 for C–C bonds, 0.96 for C–NO2 and 1.29 for C–NH2 bonds). In the frame of these limitations, all π electrons of the ring were localized by NBO algorithms as lone pairs on carbon atoms. It rather does not mean the actual presence of the lone pair. In order to find the arguments, we have calculated critical points (CPs) of the L(r) function68, which is defined as minus the Laplacian of the electron density, ∇2ρ46, in the vicinity of all non-hydrogen atoms (see Fig. 4).

Figure 4. Bond critical points (BCPs) of the electron density (small green spheres) and maximums (3;+3 critical points) of L(r) function (large yellow spheres) in the vicinity of the C– NH2, C–NO2 and C–C bonds of TATB. Black lines show bond paths.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 30

The L(r) function is a powerful tool for exploring the electron density map in order to locate the electron pairs68. It provides a quantitative physical basis for the valence-shell electron pair repulsion model69,70. Maximums of the L(r) mapped with (3;+3) CPs represent the local charge concentration areas that could be associated with the spatially localized electron pairs70. The distribution of the (3;+3) CPs of L(r) presented on Fig. 4 is common for a π-conjugated system with sp2-hybridized C, N and O atoms. Each C–C, C–N and N–O bond reveals two L(r) maxima located on the bond path, representing the valence shell charge concentration areas for double bond. It is worth mention that these maxima are located symmetrically or almost symmetrically with respect to the geometrical center of each bond, but the bond critical points are not. BCP position on each of C–N bonds meets the position of (3;+3) CP of L(r) that is closer to the C atom. Thereby the local charge concentration on C–N bonds is strongly shifted to the basin of the N atom that determines highly negative charge of N atom (see below). Two non-bonding (3;+3) CPs of L(r) near O atoms represent positions of the lone pairs, directed at H atoms of neighboring NH2 groups. One can see that there are no non-bonding (3;+3) CPs of L(r) that could be attributed to the lone pair of carbon atoms. Therefore one can suppose that lone pairs on carbon atoms in the NBO interpretation of TATB are no more than erroneous construction due to the extremely delocalized structure with the strong interaction between the ring and substituents. The forced orbital localization is somewhat tricky in the case but it still could provide valuable quantitative estimation of the relative energies of intramolecular conjugation. Analysis of the E(2) energies (Table 3) reveals that the three strongest interactions (264 – 336 kcal/mol) along with six weaker interactions (126 – 139 kcal/mol) correspond to the conjugation of this virtual ‘lone pair’ with the π-system of the substituents. The fourth to sixth ones (158 –

ACS Paragon Plus Environment

16

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

179 kcal/mol) are due to resonance inside nitro group. The remarkable finding is that conjugation interactions between C atoms and N–O bonds are stronger than interactions between neighboring C and N atoms. That is a clear evidence of strong charge transfer from the carbon ring to the oxygen atom and π-system of nitro group that reinforces the intramolecular hydrogen bonds. Such type of conjugation is much more typical for non-aromatic conjugated systems71. At the time no conjugative coupling between the carbon atoms were found among the interactions with highest E(2) energies. That accounts for the low aromaticity of carbon ring. We have compared the present data for TATB with E(2) energies of the ortho-nitroaniline having the same substituents but only one intramolecular H-bond. The strongest interaction of the 186 kcal/mol corresponds to the resonance inside nitro group and two interactions of 120 – 146 kcal/mol correspond to the resonance inside the benzene ring. The similar energy disposition could be found in para-nitroaniline: the strongest interaction here (360 kcal/mol) is due to the conjugation inside the benzene ring, the second one (211 kcal/mol) is due to the resonance inside the nitro group. All other E(2) energies in both these molecules are much weaker (40 kcal/mol and smaller). There are no strong interactions that correspond to the conjugation between the ring and the substituent in nitroanilines, in contrast to TATB. Thus, the overall picture of conjugation energies in TATB is drastically different from the benzene derivatives, and the TATB could not be treated as the benzene derivative from the energetic point of view. Table 3. E(2) orbital interaction energies from the NBO/NLMO study donor

acceptor

energy, kcal/mol

lone pair carbon

antibonding N = O

336

lone pair carbon

antibonding N = O

332

lone pair carbon

antibonding N = O

264

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

lone pair oxygen

antibonding N = O

179

lone pair oxygen

antibonding N = O

165

lone pair oxygen

antibonding N = O

158

lone pair carbon

antibonding C – N

139

lone pair carbon

antibonding C – N

133

lone pair carbon

antibonding C – N

131

lone pair carbon

antibonding C – N

130

lone pair carbon

antibonding C – N

127

lone pair carbon

antibonding C – N

126

Page 18 of 30

Natural charges are equal to -0.37e for each nitro group and +0.18e for each amino group. Carbon atoms connected to the nitro group bear negatively charges (-0.12e), and those connected to amino groups — positively charges (+0.30e). Generally, the structure is strongly charge alternating. The magnitude of the charge separation is also much higher than in isomeric nitroanilines (which carry -0.23 … -0.27e on nitro group and about zero on amino group), and the carbon atoms connected to substituents in nitroanilines are all positively charged (+0.04 +0.16e) in more or less uniform way. The atomic charges estimated from the AIM study are rather different from the NBO charges because of the reciprocal polarization46 but emphasize the redistribution of the electron density from the benzene ring to the substituents. Both nitro and amino groups in the case carry negative charges, and these charges are different, -0.31…-0.71e for nitro groups and -0.07…-0.24e for amino groups (previously reported values are 0.70e/+0.62e 4 and -0.72…-0.75e / -0.16…-0.18e 5). The negative charge on each amino group is mainly due to the amino nitrogen atoms that bear charges of -1.3 - -1.5e (previously reported values are -1.21e

4

and -1.305…-1.326e 5). Carbon atoms bear all extra positive charge but the

charge alternating is observed here as well (+0.28…+0.36e on C atoms connected to nitro groups

ACS Paragon Plus Environment

18

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and +0.63…0.69e on C atoms connected with amino groups, previously reported values are +0.28e/+0.62e 4 and +0.27…+0.28e / +0.63…+0.64e 5). Conclusions Structural features, aromaticity indices, deformation energies and features of intramolecular interactions of the TATB molecule were examined with different quantum chemical methods and compared with known experimental data. Geometrical parameters of the molecule strongly differ from the parameters of benzene and benzene derivatives with smaller number of substituents. Values of aromaticity indices clearly reveal that carbon ring in TATB is non-aromatic, and its extremely high flexibility at zero and room temperatures is also not typical for aromatic rings. The consolidated strength of six intramolecular resonance-assisted N–H…O hydrogen bonds is about 98 kcal/mol that is three times higher than resonance energy of the benzene aromatic πsystem (~37 kcal/mol), and these interactions are the governing factor that determine the molecular structure. According to NICS(1)zz aromaticity indices, the electron delocalization in H-bonded O=N–C=C–N–H quasi-aromatic rings is greater than in the carbon ring. The strongest π-π interactions in TATB molecule correspond to the conjugation between the carbon ring and the substituents, and the values of natural charges reveal highly charge alternating structure that supports the strong hydrogen bonding. ACKNOWLEDGMENT The authors (P.D. and Z.L.) would like to gratefully acknowledge the Wroclaw Centre for Networking and Supercomputing (WCSS) for use of BEM cluster. P.D. gratefully acknowledges financial support from the National Science Centre of Poland (2016/23/B/ST4/01099). I.O. gratefully acknowledges financial support from the National Academy of Sciences of Ukraine (2017-2018, 0117U001685).

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 30

REFERENCES (1)

Boddu, V. M.; Redner, P. Energetic Materials: Thermophysical Properties, Predictions, and Experimental Measurement; CRC Press, 2010.

(2)

Boddu, V. M.; Viswanath, D. S.; Ghosh, T. K.; Damavarapu, R. 2,4,6-Triamino-1,3,5Trinitrobenzene (TATB) and TATB-Based Formulations--a Review. Journal of Hazardous Materials 2010, 181 (1–3), 1–8.

(3)

Davidson, A. J.; Dias, R. P.; Dattelbaum, D. M.; Yoo, C.-S. “Stubborn” Triaminotrinitrobenzene: Unusually High Chemical Stability of a Molecular Solid to 150 GPa. The Journal of Chemical Physics 2011, 135 (17), 174507.

(4)

David Stephen, A.; Srinivasan, P.; Kumaradhas, P. Bond Charge Depletion, Bond Strength and the Impact Sensitivity of High Energetic 1,3,5-Triamino 2,4,6Trinitrobenzene

(TATB)

Molecule:

A

Theoretical

Charge

Density

Analysis.

Computational and Theoretical Chemistry 2011, 967 (2–3), 250–256. (5)

Chua, Z.; Gianopoulos, C. G.; Zarychta, B.; Zhurova, E. A.; Zhurov, V. V.; Pinkerton, A. A. Inter- and Intramolecular Bonding in 1,3,5-Triamino-2,4,6-Trinitrobenzene: An Experimental and Theoretical Quantum Theory of Atoms in Molecules (QTAIM) Analysis. Crystal Growth & Design 2017, acs.cgd.7b00674.

(6)

Bharatama, P. V; Lammertsma, K. Nitro ⇒ Aci-Nitro Tautomerism in High-Energetic Nitro Compounds. Theoretical and Computational Chemistry 2003, 12, 61–89.

(7)

Wu, C. J.; Fried, L. E. Ring Closure Mediated by Intramolecular Hydrogen Transfer in the Decomposition of a Push−Pull Nitroaromatic: TATB. The Journal of Physical Chemistry

ACS Paragon Plus Environment

20

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

A 2000, 104 (27), 6447–6452. (8)

Kakar, S.; Nelson, A. J.; Treusch, R.; Heske, C.; van Buuren, T.; Jiménez, I.; Pagoria, P.; Terminello, L. J. Electronic Structure of the Energetic Material 1,3,5-Triamino-2,4,6Trinitrobenzene. Physical Review B 2000, 62 (23), 15666–15672.

(9)

Balakina, M. Y.; Fominykh, O. D. The Quantum-Chemical Study of Organic Octupolar Chromophore Triaminotrinitrobenzene and the Dimer: Topological Analysis and Nonlinear Optical Characteristics. International Journal of Quantum Chemistry 2011, 111 (11), 2677–2687.

(10)

Minkin, V. I.; Glukhovtsev, M. N.; Simkin, B. Y. Aromaticity and Antiaromaticity: Electronic and Structural Aspects; Wiley: New York, NY, 1994.

(11)

Balaban, A. T.; Oniciu, D. C.; Katritzky, A. R. Aromaticity as a Cornerstone of Heterocyclic Chemistry. Chemical Reviews 2004, 104 (5), 2777–2812.

(12)

Fowler, P. W.; Soncini, A. Aromaticity , Polarisability and Ring Current. Chemical Physics Letters 2004, 383, 507–511.

(13)

Krygowski, T. M.; Stepien, B. T. Changes in Aromaticity in the Ring of Monosubstituted Benzene Derivatives. Polish Journal of Chemistry 2004, 78 (11–12), 2213–2217.

(14)

Shishkin, O. V; Omelchenko, I. V; Krasovska, M. V; Zubatyuk, R. I.; Gorb, L.; Leszczynski, J. Aromaticity of Monosubstituted Derivatives of Benzene. The Application of out-of-Plane Ring Deformation Energy for a Quantitative Description of Aromaticity. Journal of Molecular Structure 2006, 791 (1–3), 158–164.

(15)

Gilli, P.; Bertolasi, V.; Pretto, L.; Ferretti, V.; Gilli, G. Covalent versus Electrostatic

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 30

Nature of the Strong Hydrogen Bond: Discrimination among Single, Double, and Asymmetric

Single-Well

Hydrogen

Bonds

by

Variable-Temperature

X-Ray

Crystallographic Methods in Beta-Diketone Enol RAHB Systems. Journal of the American Chemical Society 2004, 126 (12), 3845–3855. (16)

Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G. Evidence for Intramolecular N−H···O Resonance-Assisted Hydrogen Bonding in β-Enaminones and Related Heterodienes. A Combined Crystal-Structural, IR and NMR Spectroscopic, and Quantum-Mechanical Investigation. Journal of the American Chemical Society 2000, 122 (42), 10405–10417.

(17)

Góra, R. W.; Maj, M.; Grabowski, S. J. Resonance-Assisted Hydrogen Bonds Revisited. Resonance Stabilization vs. Charge Delocalization. Physical Chemistry Chemical Physics : PCCP 2013, 15 (7), 2514–2522.

(18)

Zhang, G.; Musgrave, C. B. Comparison of DFT Methods for Molecular Orbital Eigenvalue Calculations. The Journal of Physical Chemistry A 2007, 111 (8), 1554–1561.

(19)

Karabıyık, H.; Sevinçek, R.; Petek, H.; Aygün, M. Aromaticity Balance, π-Electron Cooperativity and H-Bonding Properties in Tautomerism of Salicylideneaniline: The Quantum Theory of Atoms in Molecules (QTAIM) Approach. Journal of Molecular Modeling 2011, 17 (6), 1295–1309.

(20)

Jezierska-Mazzarello, A.; Panek, J. J.; Szatyłowicz, H.; Krygowski, T. M. Hydrogen Bonding as a Modulator of Aromaticity and Electronic Structure of Selected OrthoHydroxybenzaldehyde Derivatives. The Journal of Physical Chemistry A 2012, 116 (1), 460–475.

ACS Paragon Plus Environment

22

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(21)

Krygowski, T. M.; Zachara-Horeglad, J. E.; Palusiak, M. H-Bonding-Assisted Substituent Effect. The Journal of Organic Chemistry 2010, 75 (15), 4944–4949.

(22)

Sobczyk, L.; Grabowski, S. J.; Krygowski, T. M. Interrelation between H-Bond and PiElectron Delocalization. Chemical Reviews 2005, 105 (10), 3513–3560.

(23)

Krygowski, T. M.; Zachara, J. E.; Szatylowicz, H. Molecular Geometry as a Source of Chemical Information. 3. How H-Bonding Affects Aromaticity of the Ring in the Case of Phenol and p-Nitrophenol Complexes: A B3LYP/6-311+G Study. The Journal of Organic Chemistry 2004, 69 (21), 7038–7043.

(24)

Wu, J. I.; Jackson, J. E.; Schleyer, P. V. R. Reciprocal Hydrogen Bonding-Aromaticity Relationships. Journal of the American Chemical Society 2014, 136 (39), 13526–13529.

(25)

Krygowski, T. M.; Bankiewicz, B.; Czarnocki, Z.; Palusiak, M. Quasi-Aromaticity—what Does It Mean? Tetrahedron 2015, 71 (30), 4895–4908.

(26)

Martyniak, A.; Majerz, I.; Filarowski, A. Peculiarities of Quasi-Aromatic Hydrogen Bonding. RSC Advances 2012, 2 (21), 8135.

(27)

Nowroozi, A.; Nakhaei, E.; Masumian, E. Exploring the Correlation between the πElectron Delocalization and Intramolecular Hydrogen Bond in Malonaldehyde Derivatives; a Quantum Chemical Study. Structural Chemistry 2014, 25 (5), 1415–1422.

(28)

Romero-Fernández, M. P.; Ávalos, M.; Babiano, R.; Cintas, P.; Jiménez, J. L.; Palacios, J. C. Rethinking Aromaticity in H-Bonded Systems. Caveats for Transition Structures Involving Hydrogen Transfer and π-Delocalization. The Journal of Physical Chemistry A 2015, 119 (3), 525–534.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(29)

Page 24 of 30

Omelchenko, I. V.; Shishkin, O. V.; Gorb, L.; Hill, F. C.; Leszczynski, J. Properties, Aromaticity, and Substituents Effects in Poly Nitro- and Amino-Substituted Benzenes. Structural Chemistry 2012, 23 (5), 1585–1597.

(30)

Majerz, I.; Dziembowska, T. Geometric Aspects of Aromaticity: Interrelations between Intramolecular Hydrogen Bonds, Steric Effects and π-Electron Delocalisation in Nitroanilines. European Journal of Organic Chemistry 2011, 2011 (2), 280–286.

(31)

Majerz, I.; Dziembowska, T. Aromaticity of Overcrowded Nitroanilines. The Journal of Physical Chemistry A 2012, 116 (23), 5629–5636.

(32)

Shishkin, O. V; Pichugin, K. Y.; Gorb, L.; Leszczynski, J. Structural Non-Rigidity of SixMembered Aromatic Rings. Journal of Molecular Structure 2002, 616 (1–3), 159–166.

(33)

Omelchenko, I. V; Shishkin, O. V; Gorb, L.; Leszczynski, J.; Fias, S.; Bultinck, P. Aromaticity in Heterocyclic Analogues of Benzene: Comprehensive Analysis of Structural Aspects, Electron Delocalization and Magnetic Characteristics. Physical Chemistry Chemical Physics : PCCP 2011, 13 (46), 20536–20548.

(34)

Shishkin, O. V; Dopieralski, P.; Omelchenko, I. V; Gorb, L.; Latajka, Z.; Leszczynski, J. Dynamical Nonplanarity of Benzene. Evidences from the Car–Parrinello Molecular Dynamics Study. The Journal of Physical Chemistry Letters 2011, 2 (22), 2881–2884.

(35)

Shishkin, O. V; Dopieralski, P.; Omelchenko, I. V; Gorb, L.; Latajka, Z.; Leszczynski, J. Entropy versus Aromaticity in the Conformational Dynamics of Aromatic Rings. Journal of Molecular Modeling 2013, 19 (10), 4073–4077.

(36)

Cady, H. H.; Larson, A. C. The Crystal Structure of 1,3,5-Triamino-2,4,6-Trinitrobenzene.

ACS Paragon Plus Environment

24

Page 25 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Acta Crystallographica 1965, 18 (3), 485–496. (37)

Liu, H.; Zhao, J.; Ji, G.; Wei, D.; Gong, Z. Vibrational Properties of Molecule and Crystal of TATB : A Comparative Density Functional Study. Physics Letters A 2006, 358, 63–69.

(38)

Huang, Z.; Chen, B.; Gao, G. IR Vibrational Assignments for TATB from the Density Functional B3LYP/6-31G(d) Method. Journal of Molecular Structure 2005, 752 (1–3), 87–92.

(39)

Manaa, M. R.; Fried, L. E. Internal Rotation in Energetic Systems: TATB. The Journal of Physical Chemistry A 2001, 105 (27), 6765–6768.

(40)

Manaa, M. R.; Gee, R. H.; Fried, L. E. Internal Rotation of Amino and Nitro Groups in TATB : MP2 Versus DFT ( B3LYP ). Journal of Physical Chemistry A 2002, 106 (37), 8806–8810.

(41)

Møller, C.; Plesset, M. S. Note on an Approximation Treatment for Many-Electron Systems. Physical Review 1934, 46 (7), 618–622.

(42)

Kendall, R. A.; Dunning, T. H.; Harrison, R. J. Electron Affinities of the First-Row Atoms Revisited. Systematic Basis Sets and Wave Functions. The Journal of Chemical Physics 1992, 96 (9), 6796.

(43)

Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06Class Functionals and 12 Other Function. Theoretical Chemistry Accounts 2007, 120 (1– 3), 215–241.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(44)

Page 26 of 30

Cyrański, M. K.; Krygowski, T. M. Two Sources of the Decrease of Aromaticity: Bond Length Alternation and Bond Elongation. Part I. An Analysis Based on Benzene Ring Deformations. Tetrahedron 1999, 55 (19), 6205–6210.

(45)

Chen, Z.; Wannere, C. S.; Corminboeuf, C.; Puchta, R.; Schleyer, P. V. R. NucleusIndependent Chemical Shifts (NICS) as an Aromaticity Criterion. Chemical Reviews 2005, 105 (10), 3842–3888.

(46)

Bader, R. F. W. Atoms In Molecules: A Quantum Theory; Clarendon Press: Oxford, 1994.

(47)

Keith, T. A. AIMAll. TK Gristmill Software, Overland Park KS, USA 2014.

(48)

Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Bohmann, J. A.; Morales, C. M.; Weinhold, F. NBO 5.0. Theoretical Chemistry Institute, University of Wisconsin, Madison 2001.

(49)

Marx, D.; Hutter, J. Ab Initio Molecular Dynamics: Basic Theory and Advanced Methods; Cambridge University Press: Cambridge, 2009.

(50)

Car, R.; Parrinello, M. Unified Approach for Molecular Dynamics and Density-Functional Theory. Physical Review Letters 1985, 55 (22), 2471–2474.

(51)

Hutter, J.; Parrinello, M.; Marx, D.; Focher, P.; Tuckerman, M.; Andreoni, W.; Curioni, A.; Boero, M.; Fois, E.; Roetlisberger, U.; et al. CPMD Program Package, IBM Corp 1990-2004, MPI Für Festkörperforschung Stuttgart 1997- 2001. CPMD Program Package, IBM Corp 1990-2004, MPI für Festkörperforschung Stuttgart 1997- 2001. p see http://www.cpmd.org.

(52)

Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made

ACS Paragon Plus Environment

26

Page 27 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Simple. Physical Review Letters 1996, 77 (18), 3865–3868. (53)

Troullier, N.; Martins, J. L. Efficient Pseudopotentials for Plane-Wave Calculations. Physical Review B 1991, 43 (3), 1993–2006.

(54)

Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé–Hoover Chains: The Canonical Ensemble via Continuous Dynamics. The Journal of Chemical Physics 1992, 97 (4), 2635–2643.

(55)

Groom, C. R.; Allen, F. H. The Cambridge Structural Database in Retrospect and Prospect. Angewandte Chemie International Edition 2014, 53 (3), 662–671.

(56)

Asturiol, D.; Duran, M.; Salvador, P. Intramolecular Basis Set Superposition Error Effects on the Planarity of Benzene and Other Aromatic Molecules: A Solution to the Problem. The Journal of Chemical Physics 2008, 128 (14), 144108.

(57)

Zhang, C.; Kang, B.; Cao, X.; Xiang, B. Why Is the Crystal Shape of TATB Is so Similar to Its Molecular Shape? Understanding by Only Its Root Molecule. Journal of Molecular Modeling 2012, 18 (5), 2247–2256.

(58)

Krygowski, T. M.; Cyrański, M. K. Structural Aspects of Aromaticity. Chemical Reviews 2001, 101 (5), 1385–1420.

(59)

Nowroozi, A.; Raissi, H.; Hajiabadi, H.; Jahani, P. M. Reinvestigation of Intramolecular Hydrogen Bond in Malonaldehyde Derivatives: An Ab Initio, AIM and NBO Study. International Journal of Quantum Chemistry 2011, 111 (12), 3040–3047.

(60)

Guevara-Vela, J. M.; Romero-Montalvo, E.; Costales, A.; Pendás, Á. M.; Rocha-Rinza, T. The Nature of Resonance-Assisted Hydrogen Bonds: A Quantum Chemical Topology

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 30

Perspective. Physical Chemistry Chemical Physics 2016, 18 (38), 26383–26390. (61)

Omelchenko, I. V.; Shishkin, O. V. Basis Set Effects on the Structure of Isomeric Nitroanilines: The Role of Basis Set Expansion, Additional Diffuse and Polarization Functions within the Frame of DFT and MP2 Approaches. Functional Materials 2017, 24 (2), 005-277.

(62)

Gilli, G.; Gilli, P. The Nature of the Hydrogen Bond; Oxford University Press, 2009.

(63)

Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen Bond Strengths Revealed by Topological Analyses of Experimentally Observed Electron Densities. Chemical Physics Letters 1998, 285 (3–4), 170–173.

(64)

Furmanchuk, A.; Shishkin, O. V; Isayev, O.; Gorb, L.; Leszczynski, J. New Insight on Structural Properties of Hydrated Nucleic Acid Bases from Ab Initio Molecular Dynamics. Physical Chemistry Chemical Physics : PCCP 2010, 12 (33), 9945–9954.

(65)

Shishkin, O. V; Leszczynski, J. The Contribution of Ring Conformational Flexibility to the Relative Stability of Isomeric Cyclohexadienes. Chemical Physics Letters 1999, 302 (3–4), 262–266.

(66)

Shishkina, S. V.; Shishkin, O. V.; Leszczynski, J. Three-Stage Character of Ring Inversion in Cyclohexene. Chemical Physics Letters 2002, 354 (5–6), 428–434.

(67)

Zefirov, N. S.; Palyulin, V. A.; Dashevskaya, E. E. Stereochemical Studies. XXXIV. Quantitative Description of Ring Puckering via Torsional Angles. The Case of SixMembered Rings. Journal of Physical Organic Chemistry 1990, 3 (3), 147–158.

(68)

Popelier, P. On the Full Topology of the Laplacian of the Electron Density. Coordination

ACS Paragon Plus Environment

28

Page 29 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Chemistry Reviews 2000, 197 (1), 169–189. (69)

Bader, R. F. W.; MacDougall, P. J.; Lau, C. D. H. Bonded and Nonbonded Charge Concentrations and Their Relation to Molecular Geometry and Reactivity. Journal of the American Chemical Society 1984, 106 (6), 1594–1605.

(70)

MacDougall, P. J.; Hall, M. B.; Bader, R. F. W.; Cheeseman, J. R. Extending the VSEPR Model through the Properties of the Laplacian of the Charge Density. Canadian Journal of Chemistry 1989, 67 (11), 1842–1846.

(71)

Shishkina, S. V.; Shishkin, O. V.; Desenko, S. M.; Leszczynski, J. Conjugation and Hyperconjugation in Conformational Analysis of Cyclohexene Derivatives Containing an Exocyclic Double Bond. Journal of Physical Chemistry A 2008, 112 (30), 7080–7089.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 30

TOC Graphic

ACS Paragon Plus Environment

30