Abstract
All possible aza analogs of azulene, containing from one to three nitrogen atoms in the five-membered ring or from one to five nitrogen atoms in the seven-membered ring, have been theoretically considered to obtain information about their stabilities and aromaticities. Total electronic energy and nucleus independent chemical shift (NICS) data have been used to evaluate stability and aromaticity, respectively. The stabilities of the structures are strongly affected by the positions of the nitrogen atoms. Calculations of azaazulenes show that stability is decreased with close proximity of the nitrogen atoms. When nitrogen in the five-membered ring is adjacent to a ring junction, aromaticity of the cyclopentadienyl anion is reduced and that of the tropylium cation is increased. The number of nitrogen atoms affects the aromaticity of the system.
Introduction
Aromatic compounds such as benzene, naphthalene, and anthracene play very important roles in both synthetic and computational organic chemistry. Azulene is an organic molecule that is an isomer of naphthalene. The structures of naphthalene and azulene are isoelectronic in terms of π systems. Both structures possess ten π electrons delocalized in two fused rings. Unlike naphthalene, which is formed by the fusion of two six-membered aromatic rings, azulene is composed of one five- and one seven-membered ionic aromatic ring. The symmetric distribution of electrons through naphthalene makes it nonpolar; however, azulene has a high dipole moment. This polarity can be attributed to azulene being composed of the fused aromatic six π electron cyclopentadienyl anion and aromatic six π electron tropylium cation. To achieve the stable aromatic sextet in both rings, one electron pair from the seven-membered ring is transferred to the five-membered ring [1–3].
Azaazulene compounds can modulate protein kinase (PK) activity and/or act as anticancer agents [4]. PKs are enzymes which catalyze the phosphorylation of specific tyrosine, serine, or threonine residues in cellular proteins. PKs mediate cellular signal transduction in regulating cellular function such as proliferation, differentiation, growth, cell cycle, cell metabolism, cell survival, cell apoptosis, DNA damage repair, cell motility, and response to the microenvironment. Dysregulated PK activity is a frequent cause of diseases such as angiogenesis, cancer, tumor growth, tumor metastasis, atherosclerosis, age-related macular degeneration, diabetic retinopathy, inflammatory diseases, and/or parasitical disease [4].
Aromaticity still continues to be a frequently investigated area of chemistry. The basic criterion for aromatic compounds is that they possess cyclic conjugated π systems containing the proper number of π electrons (i.e., the Hückel rule). Although this criterion is sufficiently strong to predict the aromaticity of a host of neutral and/or charged ring systems, it is not always a good indicator of aromaticity for more complex systems as in the present case.
Aromaticity is expressed by a combination of properties in cyclic delocalized systems. In general, aromaticity is discussed in terms of energetic, structural, and magnetic criteria [5–10]. In 1996, Schleyer and coworkers introduced a simple and efficient probe for aromaticity, namely, nucleus independent chemical shift (NICS) [11], which is the computed value of the negative magnetic shielding at some selected point in space, generally, at a ring or cage center. Negative NICS values denote aromaticity (-11.5 ppm for benzene, -11.4 ppm for naphthalene) and positive NICS values denote antiaromaticity (28.8 ppm for cyclobutadiene), whereas small NICS values indicate nonaromaticity (-2.1 ppm for cyclohexane, -1.1 ppm for adamantane) [11]. NICS may be a useful indicator of aromaticity that usually correlates well with other energetic, structural, and magnetic criteria for aromaticity [12–15]. Resonance energies and magnetic susceptibilities are measures of the overall aromaticity of a polycyclic system, but do not provide information about the individual rings. NICS is an effective probe for local aromaticity of individual rings of polycyclic systems.
The aromatic character of several organic compounds including azulene has been investigated theoretically by Schleyer and coworkers [11]. They applied B3LYP/6-31G(d) and B3LYP/6-31+G(d) methods for the computation of NICS data of the compounds and obtained -8.3 (-7.0) ppm and -21.5 (-19.7) ppm for seven-membered and five-membered rings of azulene, respectively. They reported that NICS values are only somewhat sensitive to the basis set and the use of 6-31+G* (where possible) is recommended [11].
Stanger describes major paradigms in the field of aromaticity and emphasizes the contradictions and paradoxes between these paradigms, and between different measures of aromaticity [16]. The effect of centric perturbation of a heteroatom to the aromatic rings of well-known aromatic compounds have always found application in both theoretical and experimental studies [17–21].
The present article reveals the results of the theoretical investigation on the effect of centric mono, di, tri, tetra, and penta aza substitution on the aromaticity of the parent azulene system. Moreover, the effect of the position of the aza substitution was investigated by the applications of NICS and electronic energy calculations.
Method of calculation
The initial geometry optimizations of all structures leading to energy minima were achieved by using the semi-empirical PM3 self-consistent field molecular orbital (SCF MO) method [22, 23] at the restricted level [24]. Then, geometry optimizations were achieved within the framework of density functional theory (DFT, B3LYP) [25, 26] at the level of 6-31+G(d,p) [24] without any symmetry constrains. The exchange term of B3LYP consists of hybrid Hartree-Fock and local spin density (LSD) exchange functions with Becke’s gradient correlation to LSD exchange [27]. The correlation term of B3LYP consists of the Vosko, Wilk, Nusair (VWN3) local correlation function [28] and the Lee, Yang, Parr (LYP) correlation correction function [29]. The predictions by the B3LYP method are often in qualitative agreement with the experiment [30–32].
The normal mode analysis for the structures yielded no imaginary frequencies for the 3N-6 vibrational degrees of freedom, where N is the number of atoms in the system. This indicates that the optimized structure of each molecule corresponds to at least a local minimum on the potential energy surface.
Absolute nuclear magnetic resonance (NMR) shielding values [33] were calculated using the gauge independent atomic orbital (GIAO) method [34], with the restricted closed shell formalism employing a 6-31+G(d,p) basis set over B3LYP/6-31+G(d,p) optimized geometries. NICS data were obtained by the computation of absolute NMR shielding at the ring centers, NICS(0).
The geometry optimizations and NICS calculations of the present systems were performed by using the Gaussian 03 package program [35].
Results and discussion
All possible aza analogs of azulene, containing from one to three nitrogen atoms in the five-membered ring or from one to five nitrogen atoms in the seven-membered ring were theoretically analyzed by the application of the B3LYP/6-31+G(d,p) level of theory in order to evaluate their stabilities and aromaticities. The mono and/or dicentric perturbations on positions of ring junctions were excluded because they would result in nonaromatic systems.
The molecules were labeled according to the positions of the nitrogens on the system (see Figure 1). The numbering scheme is illustrated for 5-N1 and 7-N1. The abbreviation 5-N1 presents azaazulene structure where aza substitution is at the five-membered ring and at position 1; 7-N1,N3,N5 means that substitution is on the seven-membered ring at positions 1, 3, and 5. Some of the monoaza and diaza derivatives of azulenes have been synthesized by several research groups: 5-N1 [36–38], 7-N1 [39, 40], 7-N2 [41], 5-N1,N2 [42], 5-N1,N3 [43].

The structures under consideration.
Energetics
The zero point corrected total electronic energies of the present systems were obtained by using the B3LYP/6-31+G(d,p) method. For each series of azaazulenes, the electronic energies of the molecules were calculated and relative energies were derived with respect to the most stable compound. Figure 2 shows the calculated relative electronic energies of the mono (black), di (blue), tri (red), and tetra (green) aza substituted azulene derivatives. The lowest energy compounds were found to be the 5-N1, 5-N1,N3, 7-N1,N3,N5, and 7-N1,N2,N3,N5 isomers for mono, di, tri, and tetra aza substituted azulene derivatives, respectively, and the graph was drawn accordingly. The graph of the relative energies indicates that the stability of the structures decreases for adjacent nitrogen atoms. The only exception is 5-N1,N2,N3, which might be due to preservation of aromaticity on both rings (see below). The drastic instability of 7-N1,N2, 7-N2,N3, 7-N1,N2,N3, and 7-N2,N3,N4 compared to the others can be attributed to vicinal positioning of the two nitrogen atoms for the diaza substituted derivatives and the sequential positioning of three nitrogen atoms for the triaza substituted derivatives. This result is consistent with a decrease of aromaticity in these derivatives.

Relative energies of compounds of the four series.
NICS
The most well-known aromatic compound is benzene with an excellent delocalization of six π electrons. Thus, replacement of one carbon atom with a heteroatom decreases the aromaticity of the system to some extent due to the electronegativity difference between carbon and other atoms. The aromaticity of a conjugated system decreases even more with the substitution of a second or third heteroatom as in the present case. The NICS data for the parent azulene and mono, di, tri, tetra, and pentaaza substituted azulene derivatives are given in Table 1.
The NICS (ppm) data of azulene and its aza analogs.
Structure | Ring 5 | Ring 7 |
---|---|---|
Azulene | -18.1 | -5.5 |
5-N1 | -15.0 | -6.9 |
5-N2 | -19.5 | -3.8 |
5-N1,N2 | -15.8 | -6.1 |
5-N1,N3 | -12.4 | -7.7 |
5-N1,N2,N3 | -14.5 | -7.2 |
7-N1 | -17.1 | -5.7 |
7-N2 | -18.1 | -4.1 |
7-N3 | -17.6 | -6.1 |
7-N1,N2 | -16.9 | -4.3 |
7-N1,N3 | -16.2 | -6.2 |
7-N1,N4 | -17.3 | -4.0 |
7-N1,N5 | -15.5 | -5.9 |
7-N2,N3 | -17.6 | -4.1 |
7-N2,N4 | -17.7 | -2.7 |
7-N1,N2,N3 | -17.6 | -4.6 |
7-N1,N2,N4 | -16.0 | -2.0 |
7-N1,N2,N5 | -16.1 | -4.1 |
7-N1,N3,N4 | -15.2 | -3.7 |
7-N1,N3,N5 | -14.1 | -6.3 |
7-N2,N3,N4 | -18.7 | -3.2 |
7-N1,N2,N3,N4 | -14.4 | -1.7 |
7-N1,N2,N3,N5 | -14.3 | -4.4 |
7-N1,N2,N4,N5 | -15.7 | -2.5 |
7-N1,N2,N3,N4,N5 | -6.4 | 2.1 |
Azulene structure is composed of two fused ionic aromatic rings, one of them being an electronically rich cyclopentadienyl anion and the other an electronically deficient tropylium cation. Thus, the NICS data for the five-membered ring is expected to be greater than that of benzene, whereas the seven-membered ring should possess a smaller absolute NICS value. The calculated NICS data for azulene are -18.1 ppm and -5.5 ppm for the five-membered and seven-membered rings, respectively, as expected.
In the present study, the influence of position and number of nitrogen atoms on the aromaticity of the system was investigated. It is obvious that heteroatom substitution to an aromatic structure disturbs electronic delocalization resulting in a net decrease in the aromaticity of the system. Mono aza substitution to the five-membered ring at position 1 (5-N1) decreased the NICS data to -15.0 ppm and increased the NICS value of the seven-membered ring to -6.9 ppm. By contrast, mono aza substitution to the five-membered ring at position 2 (5-N2) increased the NICS value of the five-membered ring to -19.5 ppm and decreased the aromaticity of the seven-membered ring to -3.8 ppm. The results of the calculations show that substitutions that are close to fusion points pull the electrons from the electron-rich part and increase electron density of the electron-deficient part. A similar result is obtained for 5-N1,N2 and 5-N1,N3. The aromaticity of the seven-membered ring is strongly affected by the number of nitrogen substitutions. The aromaticity of the seven-membered ring decreases with increases in the number of aza substitutions, and the ring becomes nonaromatic at penta aza substitution (7-N1,N2,N3,N4,N5). The seven-membered ring possesses the lowest aromaticity value when the substitution is far from the fusion points unlike in the five-membered ring. This can be explained by the presence of electronegative nitrogens that pull electrons from the five-membered ring into the seven-membered ring.
Conclusion
Centrically nitrogen substituted azulene derivatives were theoretically considered by the application of the DFT B3LYP/6-31+G(d,p) method to obtain information about their stabilities and aromaticities. The stabilities of the structures are strongly affected by the position of the aza points. The closer the nitrogens to one another, the less aromatic and thus less stable the system is. When the aza substitution is close to the fusion points for the five-membered ring, the aromaticity of the cyclopentadienyl anion part decreases and the aromaticity of the tropylium cation part increases.
References
[1] Badger, G. M. The Structures and Reactions of the Aromatic Compounds; Cambridge University Press: Cambridge, 2011.Search in Google Scholar
[2] Gleiter, R.; Haberhauer, G. Aromaticity and Other Conjugation Effects; Wiley-VCH: Weinheim, 2012.Search in Google Scholar
[3] Hopf, H. Classics in Hydrocarbon Chemistry: Syntheses, Concepts, Perspectives; Wiley-VCH: Weinheim, 2000.Search in Google Scholar
[4] Lee, O. Azaazulene Compounds, Patent Number: 20100280012, 2010.Search in Google Scholar
[5] Schleyer, P. R.; Jiao, H. What is aromaticity? Pure Appl. Chem. 1996, 68, 209–218.Search in Google Scholar
[6] Schleyer, P. R. Introduction: aromaticity. Chem. Rev. 2001, 101, 1115–1118.Search in Google Scholar
[7] Cyranski, M. K.; Krygowski, T. M.; Katritzky, A. R.; Schleyer, P. R. To what extent can aromaticity be defined uniquely? J. Org. Chem.2002, 67, 1333–1338.Search in Google Scholar
[8] Glukhovtsev, M. N. Aromaticity today: energetic and structural criteria. J. Chem. Educ. 1997, 74, 132–136.Search in Google Scholar
[9] Krygowski, T. M.; Cyranski, M. K.; Czarnocki, Z.; Hafelinger, G.; Katritzky, A. R. Aromaticity: a theoretical concept of immense practical importance. Tetrahedron2000, 56, 1783–1796.Search in Google Scholar
[10] Minkin, V. I.; Glukhovtsev, M. N.; Simkin, B. Y. Aromaticity and Antiaromaticity: Electronic and Structural Aspects; Wiley: New York, 1994.Search in Google Scholar
[11] Schleyer, P. R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Hommes, N. J. R. E. Nucleus independent chemical shifts: a simple and efficient aromaticity probe. J. Am. Chem. Soc.1996, 118, 6317–6318.Search in Google Scholar
[12] Jiao, H.; Schleyer, P. R. Aromaticity of pericyclic reaction transition structures: magnetic evidence. J. Phys. Org. Chem. 1998, 111, 655–662.Search in Google Scholar
[13] Schleyer, P. R.; Kiran, B.; Simion, D. V.; Sorensen, T. S. Does Cr(CO)3 complexation reduce the aromaticity of benzene? J. Am. Chem. Soc. 2000, 122, 510–513.Search in Google Scholar
[14] Patchkovskii, S.; Thiel, W. Nucleus-independent chemical shifts from semiempirical calculations. J. Mol. Model. 2002, 6, 67–75.Search in Google Scholar
[15] Quinonero, D.; Garau, C.; Frontera, A.; Ballester, P.; Costa, A.; Deya, P. M. Quantification of aromaticity in oxocarbons: the problem of the fictitious ‘nonaromatic’ reference system. Chem. Eur. J.2002, 8, 433–438.Search in Google Scholar
[16] Stanger, A. What is … aromaticity. A critique of the aromaticity concept – Can it really be defined? Chem. Commun.2009, 1939–1947.10.1039/b816811cSearch in Google Scholar PubMed
[17] Türker, L.; Gümüş, S.; Atalar, T. Structural and molecular orbital properties of some boroxine derivatives – a theoretical study. Bull. Korean Chem. Soc.2009, 30, 2233–2239.Search in Google Scholar
[18] Türker, L.; Gümüş, S. Structural and molecular orbital studies of Si-phenyl silaanthracenes. Polycycl. Arom. Comp. 2010, 30, 61–74.Search in Google Scholar
[19] Gümüş, S. A computational study on substituted diazabenzenes. Turk. J. Chem.2011, 35, 803–808.Search in Google Scholar
[20] Gümüş, S. The aromaticity of substituted diazanaphthalenes. Comput. Theor. Chem.2011, 963, 263–267.Search in Google Scholar
[21] Gümüş, S.; Türker, L. Substituent effect on the aromaticity of 1,3-azole systems. Hetercycl. Commun.2012, 18, 11–16.Search in Google Scholar
[22] Stewart, J. J. P. Optimization of parameters for semi empirical methods I. Method. J. Comput. Chem.1989, 10, 209–220.Search in Google Scholar
[23] Stewart, J. J. P. Optimization of parameters for semi empirical methods II. Applications. J. Comput. Chem.1989, 10, 221–264.Search in Google Scholar
[24] Leach, A. R. Molecular Modeling; Longman: Essex, 1997.Search in Google Scholar
[25] Kohn, W.; Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev.1965, 140, A1133–A1138.Search in Google Scholar
[26] Parr, R. G.; Yang, W. Density Functional Theory of Atoms and Molecules; Oxford University Press: London, 1989.Search in Google Scholar
[27] Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A1988, 38, 3098–3100.10.1103/PhysRevA.38.3098Search in Google Scholar
[28] Vosko, S. H.; Vilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: a critical analysis. Can. J. Phys.1980, 58, 1200–1211.Search in Google Scholar
[29] Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B1988, 37, 785–789.Search in Google Scholar
[30] Scuseria, G. E. Comparison of coupled-cluster results with a hybrid of Hartree-Fock and density functional theory. J. Chem. Phys.1992, 97, 7528–7530.Search in Google Scholar
[31] Sosa, C.; Lee, C. Density functional description of transition structures using nonlocal corrections. Silylene insertion reactions into the hydrogen molecule. J. Chem. Phys.1993, 98, 8004–8011.Search in Google Scholar
[32] Wilson, P. J.; Amos, R. D.; Handy, N. C. Density functional predictions for metal and ligand nuclear shielding constants in diamagnetic closed-shell first-row transition-metal complexes. Phys. Chem. Chem. Phys.2000, 2, 187–194.Search in Google Scholar
[33] Pulay, P.; Hinton, J. F.; Wolinski, K. Nuclear Magnetic Shieldings and Molecular Structure; Tossel, J. A., Ed.; NATO ASI Series C; Kluwer: Netherlands, 1993.Search in Google Scholar
[34] Hehre, W. J.; Radom, L.; Schleyer, P. R.; Pople, J. A. Ab Initio Molecular Orbital Theory; Wiley: New York, 1986.Search in Google Scholar
[35] Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Chesseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C. Gaussian 03; Gaussian Inc.: Wallingford, CT, 2004.Search in Google Scholar
[36] Abe, N.; Ishikawa, N.; Hayashi, T.; Miura, Y. Synthesis and some reactions of 2-(2-aminoanilino)cyclohepta[b]pyrroles: leading to 5H-cyclohepta[1′,2′:4,5]pyrrolo[2,3-b][1,5]benzodiazepine, a novel 20π antiaromatic system, and cyclohepta[1′,2′:4,5] pyrrolo[1,2-a]benzimidazoles. Bull. Chem. Soc. Japan1990, 63, 1617–1622.Search in Google Scholar
[37] Nitta, M.; Iino, Y.; Hara, E.; Kobayashi, T. On the reaction of N-vinyliminophosphoranes. Part 7. A short new 1-azaazulene synthesis. J. Chem. Soc. Perkin Trans. 11989, 51–56.Search in Google Scholar
[38] Wentrup, C.; Becker, J. Synthesis of 1-azaazulene and benz[a]azulene by carbene rearrangement. J. Am. Chem. Soc. 1984, 106, 3705–3706.Search in Google Scholar
[39] Meth-Cohn, O.; Moore, C. A new simple approach to aza- azulene synthesis. J. Chem. Soc. Chem. Commun.1983, 1246–1248.10.1039/c39830001246Search in Google Scholar
[40] Meth-Cohn, O.; Moore, C.; Rooyen, P. H. The synthesis and chemistry of 4-aza-azulene. J Chem. Soc. Perkin Trans. 11985, 1793–1802.Search in Google Scholar
[41] Hafner, K.; Kreuder, M. 5-Aza-azulen. Angewandt. Chem. 1961, 73, 657–657.Search in Google Scholar
[42] Greco, C. V.; Pellegrini, F. C.; Pesce, M. A. An approach to the synthesis of a cycloheptapyrazole (1,2-diaza-azulene). J. Chem. Soc., Perkin Trans.1972, 1, 1623–1625.Search in Google Scholar
[43] Nozoe, T.; Mukai, T.; Murata, I. 1,3-Diazaazulene. J. Am. Chem. Soc. 1954, 76, 3352–3353.10.1021/ja01641a086Search in Google Scholar
©2013 by Walter de Gruyter Berlin Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Articles in the same Issue
- Masthead
- Masthead
- Reviews
- Methods for the synthesis of xanthine-derived polycyclic fused systems
- Heterocyclic synthesis via catalysis of N-heterocyclic carbenes: very classical and very modern chemical species
- Preliminary Communication
- Efficient synthesis of substituted imidazo[4,5-b]pyridines
- Research Articles
- Synthesis of N-(2-pyridyl)imidazolidin-2-ones and 1-(2-pyridyl)-2,3,7,8-tetrahydro-1H-imidazo[2,1-b][1,3,5]triazepin-5(6H)-ones with potential biological activities
- Design, synthesis, antibacterial, and antifungal studies of novel 3-substituted coumarinyl-triazine derivatives
- Synthesis, characterization, and antimicrobial screening of s-triazines linked with piperazine or aniline scaffolds
- Synthesis of 9-(Cn-1F2n-1)-substituted acridine by the reaction of 2-(CnF2n+1)-substituted aniline with ortho-methyl-substituted aromatic Grignard reagent
- Microwave-assisted synthesis of 2-(4,5-dihydro-5-(tetrazolo[1,5-a]quinoline-4-yl)-1H-pyrazol-3-yl)-substituted phenols
- A computational study on azaazulenes
Articles in the same Issue
- Masthead
- Masthead
- Reviews
- Methods for the synthesis of xanthine-derived polycyclic fused systems
- Heterocyclic synthesis via catalysis of N-heterocyclic carbenes: very classical and very modern chemical species
- Preliminary Communication
- Efficient synthesis of substituted imidazo[4,5-b]pyridines
- Research Articles
- Synthesis of N-(2-pyridyl)imidazolidin-2-ones and 1-(2-pyridyl)-2,3,7,8-tetrahydro-1H-imidazo[2,1-b][1,3,5]triazepin-5(6H)-ones with potential biological activities
- Design, synthesis, antibacterial, and antifungal studies of novel 3-substituted coumarinyl-triazine derivatives
- Synthesis, characterization, and antimicrobial screening of s-triazines linked with piperazine or aniline scaffolds
- Synthesis of 9-(Cn-1F2n-1)-substituted acridine by the reaction of 2-(CnF2n+1)-substituted aniline with ortho-methyl-substituted aromatic Grignard reagent
- Microwave-assisted synthesis of 2-(4,5-dihydro-5-(tetrazolo[1,5-a]quinoline-4-yl)-1H-pyrazol-3-yl)-substituted phenols
- A computational study on azaazulenes