Abstract

Tautomerization of 2-(2-hydroxyphenyl)-1-azaazulene (2OHPhAZ) in the gas phase and ethanol has been studied using B3LYP, M06-2X, and ωB97XD density functional theory (DFT) with different basis sets. For more accurate data, energies were refined at CCSD(T)/6-311++G(2d,2p) in the gas phase. Nuclear magnetic resonance (NMR), aromaticity, Fukui functions, acidity, and basicity were also calculated and compared with experimental data. Time-dependent density functional theory (TDDFT)-solvation model based on density (TDDFT-SMD) calculations in acetonitrile have been utilized for the simulation of UV–vis electronic spectra. In addition, electronic structures of the investigated system have been discussed. The results reveal that the enol form (2OHPhAZ) is thermodynamically and kinetically stable relative to the keto tautomer (2OPhAZ) and different rotamers (2OHPhAZ–R1:R3) in the gas phase and ethanol. A comparison with the experiment illustrates a good agreement and supports the computational findings.
1. Introduction
Azulene, an isomeric structure of naphthalene (Scheme 1), has a fused five- and seven-membered ring.1,2 It has a blue color and high dipole moment.3−6 This nonbenzenoid nonalternant structure attracts the interest of theoretical and experimental studies for its unusual photophysical properties.7,8,17,9−16 Azulene derivatives have biological activities as well they can be anti-inflammatory18 or have anti-cancer properties.19
Scheme 1. Structures of Azulene, 1-Azaazulene, and Tautomeric Forms (Enolimine and Ketoenamine Form) of 2-(2-Hydroxyphenyl)-1-azaazulene.
Azaazulenes, an azulene heterocyclic analogue (Scheme 1), has attracted interest owing to their chemical and physical properties and their biological activity20,21 as anticancer agents.22 1-Azaazulenes is the most stable structure among the various structures of azaazulenes. Some 1-azaazulene derivatives have been used as ligands in metal complexes.23−25 For example, 2-chloro-8-(2-pyridyl)-1-azaazulene acts as a bidentate ligand when reacted with Cu, Fe, and Pd ions and forms trigonal bipyramidal metal chelate complexes.23
1-Azaazulene with a phenol functional group at the 2-position has been synthesized by Oda et al.26 This structure is similar to 8-hydroxyquinoline, which is an important chelating agent.27,28 Thus, 2-(2-hydroxyphenyl)-1-azaazulene (2OHPhAZ) has the potential to be used as a chelating ligand through the imino nitrogen atom and the phenolic oxygen atom. 2OHPhAZ has intense absorption in the visible region, and its long-wavelength absorption band shifts in the presence of several metal ions, and its emission upon excitation in the presence of Pb2+, Al3+, and Sc3+ ions is 103 times higher than the quantum yield determined in the absence of ions, which suggests potential sensing of metal ions by this compound. A previous experimental study26 reported the synthesis of 2OHPhAZ, its X-ray crystallographic analysis, acidity, basicity, and UV–vis absorption, and emission spectra.
Tautomerism is a phenomenon of coexistence of two or more isomers of the same molecule. Many types of tautomerization can be found depending on the transferred atom. Our case study of tautomerism includes the transfer of hydrogen atom from one position to another within the same structure. In the system under consideration, such hydrogen transfer gives keto-enol tautomers. The intramolecular hydrogen transfer between two equilibrium structures exhibits extreme importance in pharmacology, organic chemistry, medicinal chemistry, and molecular biology.29,30 Tautomers are formed by the exchange of hydrogen atoms between nitrogen and oxygen atoms of the heterocyclic ring. The proton transfer and hydrogen bonding are important characteristics of the hydrogen atom in chemistry. 2OHPhAZ has hydroxyl substitution and azaazulenyl rings. Therefore, equilibrium between enolimine (phenol-imine, OH-form) and ketoenamine (C=O, NH-form) tautomeric forms is expected (Scheme 1).
Experimental studies on the subject of tautomerism are a challenging problem in chemistry and molecular biology. Most tautomers are not observed experimentally because of their low concentration. Detailed analysis of the structural and energetic parameters caused by the migration of hydrogen atoms would enable understanding different properties of tautomers. Knowledge of the relative stabilities of tautomeric forms and their mutual conversion represent an important issue from a structural chemistry point of view. Understanding physical and chemical properties of 2OHPhAZ tautomers may assist future experimental studies on their potential use in some applications, particularly in metal complexes for analytical and biological applications. Comparing obtained results with experimental data26 might give important information concerning the origin of the observed spectra and other properties. However, there is a lack of theoretical studies on 2OHPhAZ.
In this work, we present a computational study on 2OHPhAZ (Scheme 1) at the B3LYP/6-31G(d,p) level of theory. The relative stabilities and electronic absorption spectra of their tautomers and rotamers have been conducted using B3LYP, M06-2X, and ωB97XD functional using the 6-311++G(2d,2p) basis set at the B3LYP/6-31G(d,p) optimized structures in the gas phase and ethanol. Also, single point energy calculations at CCSD(T)31,32/6-311++G(2d,2p) in the gas phase have been performed. Two aspects here are important and will be addressed, namely, proton transfer and hydrogen bonding. A strong correlation exists between aromaticity and structure/stability and, therefore, aromaticity of tautomers and rotamers has also been examined in the light of nucleus-independent chemical shift (NICS) index33,34 and harmonic oscillator model of aromaticity (HOMA).35−37
2. Computational Methods
Tautomers, rotamers, and transition states (TSs) for their interconversions were fully optimized at the density functional theory (DFT) of Becke’s three-parameter and Lee–Yang–Parr hybrid functional (B3LYP)38−40 in conjunction with the 6-31G(d,p) basis set. Vibrational frequency calculations have been conducted for each stationary point at the same level of optimization to characterize its nature as a minimum or transition state on the potential energy surface of the relevant systems. TSs were confirmed by the presence of one imaginary frequency that is examined using the ChemCraft 1.8 program.41 Minima show real frequencies. The minimum energy path (MEP) was carried out at the level of optimization (B3LYP/6-31G(d,p)) through intrinsic reaction coordinate (IRC)42,43 in order to verify that the located TSs connect the reactants and the desired products. It has been found that the energy of the transition state for tautomerization is higher than that of the keto form (2OPhAZ) because they include the zero-point correction. Energies were refined at B3LYP, meta hybrid generalized gradient approximation (M06-2X),44,45 and long-range-corrected hybrid functional of Becke’s 97 that include dispersion correction (ωB97XD)46−49 functionals using the 6-311++G(2d,2p) basis set at the B3LYP/6-31G(d,p) optimized structures. For species in which a single-reference wavefunction is a good approximation, the CCSD(T)31,32 calculations represent a powerful tool for accurate reaction and activation energies. Therefore, we have conducted single point energy calculations at CCSD(T)/6-311++G(2d,2p) using the B3LYP/6-31G(d,p) geometries. The zero-point-corrected relative energies (ΔE0) reveal that M06-2X/6-311++G(2d,2p) has a slightly better performance than ωB97XD/6-311++G(2d,2p) relative to the more accurate (CCSD(T)/6-311++G(2d,2p) level. Thus, unless noted otherwise, the M06-2X/6-311++G(2d,2p) level is used for the discussion of energetics.
Natural bond orbital (NBO) analysis50,51 has been performed at B3LYP/6-31G(d,p) using NBO program version 3.152 to compute the atomic charges, orbital interactions, and their impact on the structure and stability of the investigated systems. Optimization of all investigated structures has also been conducted in ethanol at B3LYP/6-31G(d,p) using the solvation model based on density (SMD).53 In addition, the single point solvation effect in ethanol has been calculated using M06-2X, B3LYP, and ωB97XD functionals with 6-311++G(2d,2p) basis sets at the B3LYP/6-31G(d,p)-optimized geometries in ethanol. All calculations were performed using the Gaussian 09 W program.54
Nuclear magnetic resonance (NMR) spectroscopy is a helpful tool for studying tautomeric equilibria. NMR shielding was calculated using the gauge-independent atomic orbital (GIAO) method,55−57 relative to the13C and 1H isotropic chemical shielding of tetramethylsilane (TMS), at B3LYP/6-31G(d,p) in chloroform using the optimized gas-phase geometry. These calculations yield the NICS33,34 index that is used to evaluate aromaticity, antiartomaticity, and nonaromaticity of each ring for the structures under consideration and the corresponding TSs for their conversions. A ghost atom/probe (bq’s) has been located at the center of the ring for determining NICS(0) and at 1 Å perpendicularly above the ring center for determining NICS(1).33,34,58,59 Another criterion of aromaticity is the geometry-based index, HOMA,35−37 which utilizes bond lengths according to the procedures given by Krygowski.36
To understand reactivity and stability of the studied structures, global chemical reactivity descriptors60−63 have been determined using energies of highest occupied and lowest unoccupied molecular orbitals (HOMO and LUMO). Accordingly, ionization potential (IP), electron affinity (EA), absolute hardness (η), softness (S), electronegativity ( χ), chemical potential (μ), and electrophilicity index (ω) were computed at B3LYP/6-31G(d,p) in the gas phase. The global chemical reactivity descriptors have been calculated using the following equations:
| 1 |
| 2 |
| 3 |
| 4 |
| 5 |
| 6 |
| 7 |
The Fukui function represents one of the density functional descriptors used to model chemical reactivity and site selectivity.64 UCA-FUKUI software65 has been used to calculate condensed Fukui functions (CFF) f+(r), f–(r), and f0(r) of the studied structures using atomic populations purposed by natural population analysis (NPA). Fukui functions are calculated utilizing the following equations:
| 8 |
| 9 |
| 10 |
where +, −, and 0 denote the nucleophilic, electrophilic, and radical attack, respectively.
Chattaraj et al. proposed the concept of generalized philicity66 containing information about known various global and local reactivity. The dual descriptor Δf(r) is the difference between nucleophilic and electrophilic attack in the Fukui function. If Δf(r) is larger than zero at the specific site, this site will be favored for nucleophilic attack. To understand stability and reactivity of the studied structures, chemical reactivity toward positive and negative charges can be predicted through mapping the ESP.
Electronic absorption spectra for the studied structures were calculated at TD-DFT PBE67,68 (TD-PBE/6-311+G(d,p)) in acetonitrile using the SMD approach at the B3LYP/6-31G(d,p)-optimized gas-phase geometry. The Perdew–Burke–Ernzerhof method (PBE0)69 accurately estimated electron excitations of most organic dyes.70 However, for the molecule under consideration compared to experimental results, it underestimates λmax by 59 nm, whereas the PBE functional slightly overestimates λmax by 11–13 nm depending on basis sets. The GaussSum program71 was used to simulate the ultraviolet–visible (UV–vis) spectra. The natural transition orbitals (NTOs)72 were analyzed for each electron excitation instead of particularly discussing the canonical orbitals. The frontier orbitals and NTOs were drawn using Chemcraft.41
3. Results and Discussion
3.1. Structural Analysis
Five isomers are discussed for 2OHPhAZ. The optimized structures of these forms are collected in Figure 1. Reliable structures are required in order to determine and rationalize the stability order of the investigated system. Gad et al.16 reported that B3LYP/6-31+G(d,p) produces the same order of stability obtained by the ab initio multilevel CBS-QB3 method. Previous studies44,45,73−77 illustrated high performance of M06-2X and ωB97XD functionals in predicting the trend of tautomers and conformer stability. Therefore, the structures will be discussed at B3LYP/6-31G(d,p) and energies at M06-2X/6-311++G(2d,2p), ωB97XD/6-311++G(2d,2p), and CCSD(T)/6-311++G(2d,2p). However, energies at B3LYP/6-31G(d,p) and B3LYP/6-311++G(2d,2p) are presented in the Supporting Information. To demonstrate the reliability of the obtained data, a comparison between the theory and the experiment has to be established. The geometrical parameters of 2OHPhAZ optimized at B3LYP/6-31G(d,p) compared with the X-ray data of 2OHPhAZ(26) are listed in Table 1. To measure geometrical similarities and differences between two structures, we used the root-mean-square deviation (RMSD).78−81 The RMSD is calculated by the “RMS Compare Structures” utility in the ChemCraft program. A good agreement was recorded between the X-ray and the calculated 2OHPhAZ structure in the gas phase and ethanol with RMSD values of 0.10 and 0.11 Å, respectively. Therefore, the B3LYP/6-31G (d,p) level well reproduces the X-ray structure. This supports the reliability of this level for structure optimization.
Figure 1.
Optimized structures of the 2OHPhAZ tautomers and rotamers at B3LYP/6-31G(d,p).
Table 1. X-Ray and Calculated Geometrical Parameters of 2OHPhAZ at B3LYP/6-31G(d,p) in the Gas Phase and Ethanol.
| bond/anglesa,b | exp. | gas phase | ethanol |
|---|---|---|---|
| N1–C2 | 1.365 | 1.369 | 1.370 |
| C2–C3 | 1.401 | 1.415 | 1.409 |
| C3–C3a | 1.390 | 1.395 | 1.400 |
| C3a–C4 | 1.388 | 1.404 | 1.404 |
| C4–C5 | 1.378 | 1.388 | 1.389 |
| C5–C6 | 1.394 | 1.405 | 1.404 |
| C6–C7 | 1.382 | 1.393 | 1.395 |
| C7–C8 | 1.386 | 1.399 | 1.398 |
| C8–C8a | 1.388 | 1.391 | 1.393 |
| C8a–N1 | 1.353 | 1.349 | 1.353 |
| C2–C1′ | 1.456 | 1.455 | 1.461 |
| C1′–C2′ | 1.408 | 1.426 | 1.424 |
| C2′–C3′ | 1.376 | 1.406 | 1.403 |
| C3′–C4′ | 1.385 | 1.386 | 1.390 |
| C4′–C5′ | 1.367 | 1.404 | 1.402 |
| C5′–C6′ | 1.392 | 1.385 | 1.388 |
| C6′–C1′ | 1.408 | 1.411 | 1.410 |
| C2′–O | 1.353 | 1.341 | 1.352 |
| O–N1 | 2.583 | 2.590 | 2.572 |
| O–H | 0.975 | 0.999 | 1.007 |
| N···H | 1.688 | 1.684 | 1.653 |
| O–H···N | 150.740 | 148.563 | 149.579 |
| N1–C2–C1′–C2′ | 0.829 | –0.001 | –0.009 |
| C2–C1′–C2′–O1 | –0.512 | –0.003 | –0.009 |
| C1′–C2′–O1–H1 | –2.175 | –0.002 | –0.008 |
| RMSD | 0.103 | 0.106 |
As displayed in Figure 1, the 2OHPhAZ form shows an intramolecular hydrogen bond (IHB) between the hydrogen atom of the phenolic group and the nitrogen atom of the azaazulenyl moiety, (O–H···N, 1.68 Å); see Scheme 1. This value matches the H···N bond distance of 1.69 Å reported experimentally using 1H NMR.26 Also, the keto tautomer (2OPhAZ) illustrates an HB between the oxygen atom of the carbonyl group and the hydrogen attached to the nitrogen atom of the five-membered ring (N–H···O, 1.55 Å). The calculated donor–acceptor (O1–N1) distance of 2.59 Å in 2OHPhAZ agrees very well with the experimental value (2.58 Å) and 2.49 Å in 2OPhAZ. The HB strength in the tautomeric structures can be predicted from N–H···O, O–H···N, and O···N bond lengths.82,83 As a result, the keto form has stronger HB, which might be attributed to the effect of the resonance-assisted hydrogen bond (RAHB).35,84 The short donor–acceptor atom distance gives an indication about the existence of a low-barrier hydrogen bond (LBHB). The LBHB is biologically important85 and is characterized by a short donor–acceptor distance of approximately between 2.4 and 2.6 Å.83,86−88
Three rotamers have been found for the enol isomer (2OHPhAZ), Figure 1. In the first rotamer, (R1), the OH bond adopts an anti-orientation with respect to the nitrogen atom. The other two rotamers (R2 and R3) are obtained via rotation around the inter-cycle bond (C2–C1′) of 2OHPhAZ and R1, respectively. In the R1 rotamer, there is no hydrogen bond and, therefore, the donor–acceptor distance of 2.82 Å is longer than the corresponding distance of 2.59 Å in 2OHPhAZ. In R2 and R3, there are two stabilizing interactions that are absent in R1. These are attractive interaction between phenyl hydrogen with nitrogen and azulenyl hydrogen with oxygen. Both interactions exist in R3 and the former one only in R2. Therefore, R2 and R3 are slightly more stable than R1 at all levels.
Relative to the enol form, the calculated RMSD values for keto or rotamers in the gas phase are listed in Table S1. The RMSD value is larger for R1 than the keto tautomer because of the different orientation of the hydroxyl group in the calculated 2OHPhAZ structure, whereas R2 and R3 give the largest deviation.
Tautomerization and rotamerization pass through transitions states TST, TSR1, TSR2, and TSR3. The optimized structures of TST, TSR1, TSR2, and TSR3 for 2OHPhAZ at B3LYP/6-31G(d,p) in the gas phase are presented in Figure S1. For R2 and R3 and their corresponding TSs, the dihedral angles of N1–C2–C1′–C2′, C2–C1′–C2′–O1, and C1′–C2′–O1–H1 are listed in Table S2. The breaking O–H bond in TST is stretched by 43.14% and the formed N–H bond is elongated by 33.79% with 0.146 Å of RMSD compared to 2OHPhAZ. With respect to the 2OPhAZ form, the breaking/forming, N–H/O–H bonds for TST are elongated/shortened by 4.01/7.80% with an RMSD equal to 0.027 Å. Therefore, the TST structure is closer to the structure of the keto form than the enol form. According to the Hammond postulate,89 the transformation of the enol to keto form is endothermic. The donor–acceptor distance in the TST is lower than the corresponding values in enol and keto forms (2.44 Å). The O–H bond length in TSR1 and R1 is comparable but shorter than that in 2OHPhAZ by 0.03 Å because of the presence of HB in the latter structure. The H–O–C angle is rotated by 2.09° from R1. The donor–acceptor distance in the TSR1 is stretched to 2.88 Å.
3.2. Energies and Stability
The potential energy diagram for tautomerization and rotamerization of 2OHPhAZ at ωB97XD/6–311++G(2d,2p), M06-2X/6-311++G(2d,2p), and CCSD(T)/6-311++G(2d,2p) levels of theory is illustrated in Figure 2a,b in the gas phase and ethanol, respectively. Energy profiles for tautomerization and rotamerization from IRC calculations at the B3LYP/6-31G(d,p) level are displayed in the Supporting Information (Figures S2 and S3, respectively). Compared to (CCSD(T)/6-311++G(2d,2p)) results, ωB97XD/6-311++G(2d,2p) and M06-2X/6-311++G(2d,2p) exhibit a good agreement in reproducing ΔE0 for 2OHPhAZ tautomers and rotamers, with a slightly better performance of the latter method (see Figure 2). Therefore, we will continue our discussion at the M06-2X/6-311++G(2d,2p) level unless noted otherwise.
Figure 2.

Potential energy diagram (ΔE0, ΔE0‡, kcal/mol) for tautomerization and rotamerization of 2OHPhAZ (a) in the gas phase and (b) in ethanol using the SMD solvation model at ωB97XD/6-311++G(2d,2p) (bold), M06-2X/6-311++G(2d,2p) (italic), and CCSD(T)/6-311++G(2d,2p) (normal) in the gas phase.
As displayed in Figures 2 and S2, ketonization and rotamerization of 2OHPhAZ are endothermic processes that give further support for the experimental26 finding of the 2OHPhAZ enol form in the gas phase and ethanol solution as a thermodynamically most stable form (Figure 2a,b). It is more stable than the keto form by 10.08 (4.89) kcal/mol using the M06-2X/6-311++G(2d,2p) in the gas phase (ethanol). The stability of the 2OHPhAZ is expected from the aromaticity of the phenyl ring recorded in 2OHPhAZ relative to the broken aromaticity in 2OPhAZ (as will be discussed in Section 3.4). 2OHPhAZ is also more stable than its rotamers R1, R2, and R3 by 12.04 (6.99), 7.73 (5.94), and 7.95 (5.39) kcal/mol in the gas phase (ethanol), respectively. The stabilization of 2OHPhAZ and 2OPhAZ relative to the rotamers R can be rationalized in terms of the RAHB effect according to similar previous studies.84,90−92
In accordance with Scheme 2 as explained by Lenain et al.84 for similar systems, the difference of stability 2OHPhAZ/2OPhAZ is attributed to the direction of the π-electron delocalization that favors/hinders the proton transfer from the donor atom to the acceptor atom. This affects the relative stabilization or destabilization of the IHB. Thus, 2OPhAZ is expected to have IHB that causes destabilization of the π-system.
Scheme 2. RAHB in the 2OHPhAZ-2OPhAZ Tautomer.
From the calculated barrier heights for tautomerization, it can be noticed that IHB plays a vital role in the kinetic stability of the TST form in accordance with the results obtained by Garcia-Viloca et al. for the maleate anion.93 The barrier from 2OPhAZ to 2OHPhAZ over TST equals −1.41 kcal/mol in the gas phase at M06-2X/6-311++G(2d,2p). Garcia-Viloca et al.93 reported an energy barrier of −0.85 kcal/mol for the intramolecular proton transfer in the maleic monoanion calculated at MP2/6-31+G(d,p) in the gas phase. TST is considered as a mixture of two RAHB structures with the largest π-electron delocalization within the NCCCO quasi-ring and a higher covalent character of HBs that gives the explanation about the higher stability of TS over 2OPhAZ. In the TST (Figure S3), the N1–H1 bond is formed gradually with the breaking of the O–H bond; the two curves cross each other near the transition state (s = −0.4 amu1/2 bohr). The double bond of C–O and the single bond of N1–C2 are gently formed during the course of the reaction.
Determination of the equilibrium constant depends on Gibbs free energies (ΔG) for the investigated tautomers/rotamers according to K = e( – ΔG/RT) relation, where K is the equilibrium constant, R is the universal gas constant, and T is the temperature in Kelvin. Relative Gibbs free energies (ΔG298, ΔG298#) calculated at the M06-2X/6-311++G(2d,2p) and CCSD(T)/6-311++G(2d,2p) levels of theory are collected in Table 2, and those at the B3LYP/6-31G(d,p), B3LYP/6-311++G(2d,2p), and ωB97XD/6-311++G(2d,2p) levels are listed in Table S3. The values of ΔG298 give nearly the same order of stability for the investigated tautomers/rotamers as relative zero-point-corrected energies. It can be noticed that the energy differences in ethanol are higher than the corresponding results in the gas phase. This observation can be ascribed to the higher difference of dipole moments in ethanol than in the gas phase. The dipole moments of the studied tautomers/rotamers in the gas phase and in ethanol calculated at the M06-2X/6-311++G(2d,2p) levels of theory are also given in Table 2.
Table 2. Relative Gibbs Free Energies (ΔG298) and Gibbs Free Energy Barrier (ΔG298‡), in kcal/mol, and Dipole Moments (μ, in Debye) for the Tautomers and Rotamers of 2OHPhAZ in the Gas Phase and Ethanol at M06-2X/6-311++G(2d,2p) and CCSD(T)/6-311++G(2d,2p).
| compound | gas
phase |
ethanol |
||||
|---|---|---|---|---|---|---|
| M06-Xa | CCSD(T)a | M06-2Xa | ||||
| ΔG298 | μ | ΔG298 | μ | ΔG298 | μ | |
| 2OHPhAZ | 0.00 | 4.69 | 0.00 | 4.98 | 0.00 | 7.74 |
| 2OPhAZ | 10.09 | 6.79 | 10.50 | 7.58 | 4.92 | 13.47 |
| R1 | 11.49 | 1.20 | 11.10 | 1.36 | 6.72 | 2.75 |
| R2 | 7.56 | 3.86 | 7.51 | 4.18 | 5.71 | 6.32 |
| R3 | 7.48 | 2.11 | 7.50 | 2.10 | 4.82 | 3.77 |
| TST | 9.02 | 6.64 | 9.84 | 7.39 | 3.82 | 11.21 |
| TSR1 | 12.83 | 1.94 | 12.23 | 2.17 | 9.23 | 4.20 |
| TSR2 | 11.27 | 4.79 | 10.16 | 5.09 | 8.64 | 7.27 |
| TSR3 | 13.81 | 2.11 | 12.44 | 2.20 | 8.92 | 3.90 |
Method/6-311++G(2d,2p)//B3LYP/6-31G(d,p).
3.3. NMR Analysis
LBHB can be investigated by NMR calculation.94−96 In the NMR spectrum, the appearance of a low-field proton signal (high chemical shifts) is a well-known effect of forming a hydrogen bond with a sign for an LBHB. The calculated 13C and 1H NMR chemical shifts for 2OHPhAZ in CHCl3 show good agreement with the experimental findings in CDCl3. Plotting of the calculated 13C and 1H NMR chemical shifts against the experimental values of 2OHPhAZ is displayed in Figure 3. A high correlation was noticed, which gives confidence on the used computational procedures. Detailed data about NMR of the studied tautomers and rotamers at B3LYP/6-31G(d,p) are summarized in Table S4. An inspection of Table 3 indicates that C2′ in the 2OPhAZ has the highest chemical shift (171.33 ppm) because of the electron-withdrawing effect of the attached oxygen atom of the carbonyl group. Apparently, H1 of the 2OPhAZ has the highest chemical shift (19.40 ppm) that accompanies the formation of strong LBHB (N1–H1···O1), followed by H1 of the 2OHPhAZ (14.76 ppm). On contrary, H1 of the studied rotamers have lower chemical shift that range from 4 to 6 ppm. The calculated 1H NMR chemical shifts for LBHB reveal an agreement, to some extent, with those reported by Hibbert and Emsley97 for a proton chemical shift. Therefore, the calculated 1H NMR chemical shifts help in confirming the presence of IHB and distinguish LBHB in the studied tautomers.
Figure 3.
Plot of the experimental chemical shifts verses the calculated 13C and 1H NMR chemical shifts at the B3LYP/6-31G(d,p) level of theory 2OHPhAZ using the GIAO method in CHCl3.
Table 3. 13C and 1H NMR Chemical Shifts (in ppm) for C2′ and H1 Attached to O/N at B3LYP/6-31G(d,p) for 2OHPhAZ Tautomers and Rotamers Using the GIAO Method in CHCl3.
| C2′ | H1 attached to O/N | |
|---|---|---|
| 2OHPhAZ | 157.64 | 14.76 |
| 2OPhAZ | 171.33 | 19.40 |
| R1 | 150.67 | 4.51 |
| R2 | 150.87 | 5.97 |
| R3 | 151.47 | 4.67 |
3.4. Aromaticity
Aromaticity can be assessed through energy, structural, or magnetic and reactivity criteria. One of the widely used criteria to determine the aromaticity of the studied compound is nucleus-independent chemical shifts (NICS), which is a magnetic-based criterion.34,98,99 NICS is defined as the negative of the absolute magnetic shielding calculated at selected points near the investigated molecule starting from zero to a couple of angstroms, in the ring center, above it or sometimes on grids. Negative NICS values imply diatropric ring current that reflects aromatic character, and positive ones indicate paratropic ring current, which leads to antiaromaticity, while zero NICS means nonaromatic. Although the NICS index is somewhat sensitive to basis set, previous studies34,100 calculated NICS as single point at B3LYP/6-31G(d,p) with a reasonable accuracy. NICS(0) refers to the isotropic shielding at the center of the ring, while NICS(1) is computed at 1 Å above the ring center. However, NICS(0)zz and NICS(1)zz correspond to the out-of-plane component of the magnetic shielding tensor. Among the four NICS indices, the NICS(1)zz will be used in further discussion because it gives a reliable measure of magnetic shielding of the out-of-plane component of NICS99,101−105 and correlates well with π-electron delocalization and the ring current.103−105 Because NICS is not sufficient to predict the antiaromaticity of the heteraromatic system,104−107 another criterion has been used to determine aromaticity such as the geometrical-based HOMA.35−37 The higher the HOMA value, the more aromatic is the ring in question, and hence, more delocalized the π electrons of the calculated ring. It is defined as follows:
| 11 |
where n represents the total number of bonds in the molecule, and αi is a normalization constant (αCC = 257.7 and αCN = 93.52). Ropt is the optimized bond length, for C–C = 1.388 Å and for C–N bond Ropt = 1.334 Å. It is worth noting that the HOMA is a relative index and that the maximum aromaticity is found when HOMA = 1 and it equals zero for a nonaromatic system.108
For the investigated systems, NICSzz and HOMA calculations were performed at the B3LYP/6-31G(d,p) level in the gas and ethanol to evaluate the aromatic character of each ring are presented in Table 4. Apparently, the NICS and HOMA results in Table 4 provide a good match with previous studies on benzene,109 pyrrole,110 C7H7+,111 azulene,34,112,113 and azaazulene114 and also can support the order of stability. Both NICS(1)zz and HOMA indices show good performance with the relative energies of enol and keto tautomers. The phenyl ring of 2OHPhAZ has a high value of HOMA (0.89) and a negative value of NICS(1)zz (−18.66 ppm) than 2OPhAZ (NICS(1)zz/HOMA; −12.01/0.51) in the gas phase as its aromaticity does not affect the delocalization in the RAHB. The lower values of HOMA of the five-membered ring in the enol form 2OHPhAZ compared to keto tautomer 2OPhAZ point out to the participation of its electron delocalization to the IHB ring and, therefore, increase RAHB in the enol form. The seven-membered ring for all the studied structures has a higher HOMA value than the five-membered one, indicating their higher aromatic character. Very negligible changes of magnetic indices of aromaticity are observed upon the presence of the polar solvent, like water or chloroform. π-electron delocalization of the phenyl ring in the studied rotamers has a slightly higher aromaticity than that in the enol and keto forms. This picture is in line with other observations.84,115
Table 4. NICS(1)zz (in ppm) and the HOMA Index of the Studied Tautomers and Rotamers Calculated at B3LYP/6-31G(d,p) in the Gas Phase and Ethanola.
| gas
phase |
ethanol |
||||||
|---|---|---|---|---|---|---|---|
| compound | NICS(0)iso | NICS(1)zz | HOMA | NICS(0)iso | NICS(1)zz | HOMA | |
| benzene | –9.84 (−8.06)b | –29.44 (−29.25)b | 0.98 (0.99)b | –9.51 | –28.59 | 0.98 | |
| pyrrole | –15.97 (−17.3)c | –29.80 | 0.92 (0.93)d | –15.75 | –30.01 | 0.92 | |
| C7H7+ | –6.81 (−6.7)e | –26.34 | 0.97 | –6.87 | –26.36 | 0.98 | |
| azulene | five-membered ring | –19.42 (−21.5)c(−18.1)f | –44.70 | 0.29 (0.42)g | –19.23 | –51.49 | 0.29 |
| seven-membered ring | –6.95 (−8.3)c(−5.5)f | –21.07 | 0.52 (0.60)g | –6.84 | –21.18 | 0.52 | |
| aza-azulene | five-membered ring | –16.89 ( −15.0)f | –45.47 | 0.43 | –16.66 | –44.63 | 0.49 |
| seven-membered ring | –8.03 (−6.9)f | –24.43 | 0.60 | –7.97 | –24.61 | 0.65 | |
| 2OHPhAZ | five-membered ring | –14.22 | –34.83 | 0.47 | –14.34 | –34.83 | 0.53 |
| seven-membered ring | –6.79 | –20.23 | 0.64 | –7.15 | –20.69 | 0.68 | |
| six-membered ring | –8.11 | –18.66 | 0.89 | –8.33 | –19.30 | 0.91 | |
| 2OPhAZ | five-membered ring | –11.85 | –26.67 | 0.53 | –12.90 | –29.20 | 0.61 |
| seven-membered ring | –3.49 | –11.97 | 0.66 | –5.00 | –16.30 | 0.73 | |
| six-membered ring | –3.69 | –12.01 | 0.51 | –4.73 | –14.12 | 0.62 | |
| R1 | five-membered ring | –14.44 | –37.76 | 0.42 | –14.44 | –37.76 | 0.48 |
| seven-membered ring | –7.20 | –22.16 | 0.61 | –7.14 | –22.33 | 0.65 | |
| six-membered ring | –8.92 | –18.36 | 0.94 | –8.84 | –18.45 | 0.93 | |
| R2 | five-membered ring | –14.42 | –36.26 | 0.45 | –14.40 | –35.96 | 0.51 |
| seven-membered ring | –7.16 | –20.37 | 0.64 | –7.15 | –20.69 | 0.68 | |
| six-membered ring | –8.77 | –20.94 | 0.94 | –8.80 | –21.17 | 0.94 | |
| R3 | five-membered ring | –14.90 | –39.58 | 0.40 | –14.66 | –38.76 | 0.49 |
| seven-membered ring | –7.05 | –21.52 | 0.62 | –7.02 | –21.67 | 0.66 | |
| six-membered ring | –8.98 | –22.25 | 0.94 | –8.80 | –22.20 | 0.93 | |
The stability and aromaticity are related concepts, and thus, some of the DFT descriptors can be discussed to analyze aromaticity.116−120Table 5 lists the energies of the HOMO, the LUMO, energy gap (Eg), vertical IP, and EA as well as η, S, χ, μ, and ω. From Table 5, it has been found that the keto form has the lowest HOMO–LUMO gap (2.30 eV) followed by enol forms (3.08 eV). The rotamers have the highest energy gap (3.35–3.47 eV). It is known that the lower the energy gaps, the higher the reactivity of the molecule.121,122 Thus, the keto form is expected to have high chemical reactivity, low hardness, high softness, and highest electrophilicity than enol and their rotamers. This might explain the low aromaticity of the keto form and its high global activity. In accordance with the summation of the calculated NICS and HOMA indices presented in Table 4, the enol and rotamers forms illustrate lower electrophilicity (ω) and higher hardness (η) than the keto form, which reflects high aromatic character of the former structures.
Table 5. Global Chemical Descriptor (eV) of the Studied Structures at B3LYP/6-31G(d,p) in the Gas Phase.
| EHOMO | ELUMO | Eg | IP | EA | χ | Η | S | Μ | ω | |
|---|---|---|---|---|---|---|---|---|---|---|
| 2OHPhAZ | –5.46 | –2.38 | 3.08 | 5.46 | 2.38 | 3.92 | 1.54 | 0.33 | –3.92 | 4.99 |
| 2OPhAZ | –4.85 | –2.55 | 2.30 | 4.85 | 2.55 | 3.70 | 1.15 | 0.43 | –3.70 | 5.94 |
| R1 | –5.51 | –2.04 | 3.47 | 5.51 | 2.04 | 3.77 | 1.74 | 0.29 | –3.77 | 4.10 |
| R2 | –5.69 | –2.34 | 3.35 | 5.69 | 2.34 | 4.01 | 1.68 | 0.30 | –4.01 | 4.81 |
| R3 | –5.48 | –2.06 | 3.43 | 5.48 | 2.06 | 3.77 | 1.71 | 0.29 | –3.77 | 4.14 |
3.5. Charge Distribution, Fukui Functions, and ESP Analysis
The charge distribution and the ESP surface are widely used to determine the reactivity of a given molecule and its expected interaction with other systems. NPA charges of some selected atoms of the investigated structures have been calculated in the gas phase at the B3LYP/6-31G(d,p) level (Table 6). For 2OHPhAZ tautomers and rotamers. The results indicate high negative charges on the N1 and O1 atoms and high positive charge on H1. The highest negative and positive charges on the O (−0.701 e), N (−0.557 e), and H (0.522 e) atoms for the enol form indicate the larger electrostatic interaction through N···H–O IHB. The high negative charge on the O and N atoms of enol sheds light on the potential use of this structure as a bidentate ligand when they come close to metal ions.
Table 6. NPA Charges of Some Selected Atomsa of the Investigated Systems Calculated at the B3LYP/6-31G(d,p) Level of Theory in the Gas Phase.
| N | O | C2 | C1′ | C2′ | H1 | |
|---|---|---|---|---|---|---|
| 2OHPhAZ | –0.557 | –0.701 | 0.229 | –0.169 | 0.386 | 0.522 |
| 2OPhAZ | –0.543 | –0.688 | 0.250 | –0.202 | 0.439 | 0.486 |
| R1 | –0.466 | –0.667 | 0.213 | –0.125 | 0.368 | 0.489 |
| R2 | –0.560 | –0.567 | 0.193 | 0.070 | 0.273 | 0.312 |
| R3 | –0.586 | –0.562 | 0.234 | 0.055 | 0.284 | 0.318 |
The calculated condensed Fukui functions (fk+, fk–, and fk0) evaluated from NPA for the investigated tautomer and rotamers at B3LYP/6-31G(d,p) are given in Tables 7 and S5. Higher values indicate more reactivity.123,124 An inspection of these data reveals that the preferred site for electrophilic attack for 2OHPhAZ and 2OPhAZ tautomers are C6, and for R1, R2 and R3 are C7 and C5 of the highest value of f–. On the other hand, H1 and C6 (2OHPhAZ, 2OPhAZ and R1) are the sites prone to nucleophilic attack as they have the highest f+ values. Table 7 shows that at the DFT level, the most susceptible site to a nucleophilic attack for R2 is located on O1 and C3′. For Rotamer R3, the reactivity descriptors show that H1 and C3′ are more reactive site for nucleophilic attack. For 2OHPhAZ, 2OPhAZ, R1 (H1, C6), R2 (C6, C5′), and R3 (O1, C3′) are the most active site for free radical reactions.
Table 7. Condensed Fukui Functions (fk+, fk–, fk0) of the Reactive Sitesa Evaluated from Natural Population Analysis for the Investigated Tautomer and Rotamers at B3LYP/6-31G(d,p).
| 2OHPhAZ |
2OPhAZ |
R1 |
R2 |
R3 |
|||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| atom | fk– | fk+ | fk0 | fk– | fk+ | fk0 | fk– | fk+ | fk0 | fk– | fk+ | fk0 | fk– | fk+ | fk0 |
| N1 | 0.0374 | –0.0937 | 0.0282 | 0.0239 | –0.0802 | 0.0282 | –0.0528 | –0.0036 | 0.0282 | –0.0345 | 0.0299 | 0.0023 | –0.0237 | 0.0188 | 0.0024 |
| H1 | 0.0031 | 0.5219 | 0.2625 | 0.0388 | 0.4862 | 0.2625 | 0.0362 | 0.4887 | 0.2625 | 0.0010 | 0.0000 | 0.0005 | 0.0205 | –0.0073 | 0.0066 |
| O1 | –0.0194 | –0.0576 | 0.0385 | –0.0319 | –0.0452 | 0.0385 | –0.0534 | –0.0237 | 0.0385 | –0.9348 | 0.9560 | 0.0106 | –0.0333 | 0.9640 | 0.4653 |
| C2 | 0.0222 | 0.0067 | 0.0144 | 0.0013 | 0.0276 | 0.0144 | 0.0388 | –0.0099 | 0.0144 | 0.0396 | –0.0139 | 0.0129 | 0.0451 | –0.0155 | 0.0148 |
| C3 | 0.0258 | 0.0159 | 0.0208 | 0.0188 | 0.0229 | 0.0208 | 0.0340 | 0.0077 | 0.0208 | 0.0440 | –0.0628 | 0.0094 | 0.0292 | –0.0008 | 0.0142 |
| C4 | 0.0327 | 0.0563 | 0.0445 | 0.0461 | 0.0428 | 0.0445 | 0.0298 | 0.0591 | 0.0445 | 0.0309 | 0.0492 | 0.0400 | 0.0314 | 0.0503 | 0.0409 |
| C5 | 0.0348 | 0.0163 | 0.0255 | 0.0207 | 0.0304 | 0.0255 | 0.0442 | 0.0068 | 0.0255 | 0.0441 | 0.0166 | 0.0303 | 0.0458 | 0.0093 | 0.0275 |
| C6 | 0.0406 | 0.0694 | 0.0550 | 0.0595 | 0.0506 | 0.0550 | 0.0369 | 0.0731 | 0.0550 | 0.0373 | 0.0703 | 0.0538 | 0.0375 | 0.0678 | 0.0527 |
| C7 | 0.0333 | 0.0123 | 0.0228 | 0.0118 | 0.0339 | 0.0228 | 0.0454 | 0.0003 | 0.0228 | 0.0466 | 0.0096 | 0.0281 | 0.0479 | 0.0035 | 0.0257 |
| C8 | 0.0077 | 0.0564 | 0.0320 | 0.0499 | 0.0142 | 0.0320 | –0.0046 | 0.0686 | 0.0320 | –0.0069 | 0.0719 | 0.0325 | –0.0058 | 0.0684 | 0.0313 |
| C1′ | 0.0093 | 0.0285 | 0.0189 | 0.0425 | –0.0047 | 0.0189 | –0.0344 | 0.0722 | 0.0189 | 0.0818 | –0.0707 | 0.0056 | 0.0199 | –0.0697 | 0.0249 |
| C2′ | –0.0174 | –0.0567 | 0.0370 | –0.0706 | –0.0035 | 0.0370 | 0.0011 | –0.0751 | 0.0370 | 0.1316 | –0.1869 | 0.0277 | 0.0242 | –0.2072 | 0.0915 |
| C3′ | –0.0099 | 0.0292 | 0.0096 | 0.0002 | 0.0190 | 0.0096 | 0.0161 | 0.0031 | 0.0096 | –0.5130 | 0.5707 | 0.0288 | 0.0217 | 0.5713 | 0.2965 |
| C4′ | 0.0361 | 0.0367 | 0.0364 | 0.0350 | 0.0378 | 0.0364 | 0.0431 | 0.0297 | 0.0364 | 0.0377 | 0.0256 | 0.0316 | 0.0391 | 0.0257 | 0.0324 |
| C5′ | 0.0300 | 0.0516 | 0.0408 | 0.0399 | 0.0417 | 0.0408 | 0.0226 | 0.0590 | 0.0408 | 0.0456 | 0.0378 | 0.0417 | 0.0173 | 0.0076 | 0.0124 |
| C6′ | 0.0220 | 0.0156 | 0.0188 | 0.0264 | 0.0112 | 0.0188 | 0.0286 | 0.0090 | 0.0188 | –0.5548 | 0.5513 | 0.0018 | 0.0271 | 0.5524 | 0.2897 |
The charged regions in the molecule can be depicted by ESP maps. The different colors represent different values of the electrostatic potential. The potential increases in the order red < orange < yellow < green < blue. The red color in ESP maps represents the most negative electrostatic potential while the blue color reflects the most positive electrostatic potential regions. The ESP surfaces of investigated structures obtained using B3LYP/6-31G(d,p) are depicted in Figure 4. ESP of enol/keto forms shows the localization of a significant negative charge on the O atom while blue color appears around some hydrogen atoms of the azaazulene ring. Therefore, the O atom has the ability for electron donation to electron-deficient centers such as metal ions. A slight electron-donating ability of the N atom has been noticed as well. A remarkable blue color on the H1 attached to O atoms in rotamers has been observed where there is no nearby nitrogen (R1 and R3) or C=C bonds (R2). A clear electron-rich area (red color) has been observed in the region between N1 and O1 atoms in R1, which denotes a bidonating ability of this compound when interacting with metal ions as reported experimentally for similar azaazulene derivatives.23−26
Figure 4.
Molecular ESP surfaces of the investigated structure at B3LYP/6-31G(d,p).
3.6. Acidity and Basicity
Our investigated compound has one labile proton which is attached to either a nitrogen or oxygen atom. The reliable and quick estimation of the acidity and basicity of a molecule without synthesis and experimental determination is very important to interpret the structure reactivity and property relationships. Moreover, knowledge of the acidity constants, (pKa), is important for determining equilibrium constant (K) of reactions, especially that involving proton transfers. However, when the determination of pKa experimentally is difficult, computational approaches can be applied to estimate the pKa using the thermodynamic free energies cycle that is shown in Scheme 3. Consequently, DFT calculations were applied to study the possible protonated (cation) and deprotonated (anion) forms of 2-(2-hydroxyphenyl)-1-azaazulene that is depicted in Figure 5. The protonated form, denoted as AH2+, has a net charge of +1, while the corresponding enol/rotamers or keto, AH, is neutral. According to the optimized structure of the protonated form, the hydroxyl group of the phenyl ring has been rotated to give the rotamer structure. Therefore, throughout the acidity constant calculation from the protonated form, we will consider the energy of the rotamer. On the other hand, the deprotonated form, denoted as A–, has a net charge of −1. The equations used for calculating pKa values are given below:
| 12 |
| 13 |
| 14 |
| 15 |
where Gi(g) is the standard free energy of the species “i” in the gas phase, ΔGi(solv) is the solvation free energy of “i”, and Gi(aq) is the free energy change in the aqueous phase. The GH+(g) and ΔGH+(solv) values are −6.28125,126 and – 265.90 kcal/mol,127,128 respectively.
Scheme 3. Thermodynamic Cycle Connecting Gas (g) and Aqueous (s) Phases for pKa Calculation.
Figure 5.
Optimized structures and the relative energy of protonated (AH2+, cation) and deprotonated (A–, anion) forms of 2OHPhAZ at B3LYP/6-31G(d,p).
The experimental acidity constants were taken in our consideration. The correlation between the experimental and the calculated acidity constants revealed that the ωB97XD/6-311++G(2d,2p) level yields the closest pKa value to experiment. As given in Table 8, the experimental pKa of 2OHPhAZ was reported as 12.7,26 and the calculated pKa from the protonated form is found as 15.84, 11.25, 8.23, and 12.47 using B3LYP/6-31G(d,p), B3LYP/6-311++G(2d,2p), M06-2X/6-311++G(2d,2p), and ωB97XD/6-311++G(2d,2p) levels, respectively. By comparing the pKa obtained by deprotonation, we can find that the enol form is a weaker acid than its keto and rotamers. This can be attributed to the strength of the OH and NH bonds, presence or absence of hydrogen bonds, and the stability of the resulting conjugate base upon deprotonation. More negative charges on the nitrogen and oxygen atoms for enol than keto and rotamers forms (Table 6) might reflect the strength of the OH and NH bonds.
Table 8. Calculated Acidity Constant (pKa) for the Protonated And Deprotonated Forms in Ethanol at B3LYP,a M06-2X,a and ωB97XDa.
| B3LYP/6-31G(d,p) | B3LYPa | M06-2Xa | ωB97XDa | |
|---|---|---|---|---|
| AH2+ (R1) → AH (R1) + H+ | 15.84 | 11.25 | 8.23 | 12.47 |
| AH2+ (R2) → AH (R2) + H+ | 8.34 | 5.26 | 1.76 | 6.16 |
| AH2+ (R3) → AH (R3) + H+ | 11.34 | 18.45 | 4.85 | 9.49 |
| AH (enol) → A1– + H+ | 28.90 | 21.30 | 20.44 | 23.17 |
| AH (keto) → A1– + H+ | 26.55 | 19.92 | 16.84 | 21.14 |
| AH (R1) → A1– + H+ | 21.97 | 15.50 | 15.51 | 17.78 |
| AH (R2) → A2– + H+ | 18.98 | 13.07 | 13.06 | 15.58 |
| AH (R3) → A2– + H+ | 19.84 | 3.15 | 13.72 | 15.99 |
Method/6-311++G(2d,2p)// B3LYP/6-31G(d,p).
In this study, we examine the intrinsic basicity in the gas phase that can be given by proton affinity (PA), which is the negative of the protonation reaction of AH.
| 16 |
According to a previous study on the 2-pyridone tautomer and rotamer,73 Michelson et al. found that the stable enol form has lower PA followed by the keto form with difference 5.7 kcal/mol at M06-2X/6-311 + G(2df,2p), and the rotamer has slightly higher PA. Our calculated PA of 2-(2-hydroxyphenyl)-1-azaazulene tautomer differs by 8.6 kcal/mol (241.88 (enol) versus 250.51 (keto), kcal/mol), while the rotamer forms are slightly higher in the calculated PA than the enol form (254.92, 242.81, and 248.13 kcal/mol for R1, R2, and R3, respectively) at ωB97XD/6-311++G(2d,2p). The calculated PA at B3LYP/6-31G(d,p), B3LYP/6-311++G(2d,2p), M06-2X/6-311++G(2d,2p), and ωB97XD/6-311++G(2d,2p) levels has been listed in Table S6.
3.7. UV–Vis Spectral Analysis
Potential use of any compound as a dye or sensor can be deduced from its photophysics and photochemistry. The first maximum wavelength of absorption using TDDFT calculation of the 2OHPhAZ in the gas phase by means of different solvation models PCM, CPCM, and SMD at different levels of theory such as B3LYP, CAM-B3LYP, PBE,PBE0, ωb97X-D, M06-2X, and CIS with the 6-311+G(d,p) basis sets is collected in Table 9. The results of calculations are compared with the available experimental data. The TDDFT-PBE/SMD model chemistry exhibits a good quantitative agreement regarding the first and second maximum excitation peak (Eex; the discrepancy between the calculated and experimental first maximum and second excitation energy 11 and 12 nm, respectively).26
Table 9. First λmax for 2OHPhAZ Using Different Functionals (Solvation Model, Acetonitrile)/6-311+G(d,p)//B3LYP/6-31G(d,p) with a Variety of Solvation Models (Experimental26 Value Is 495 nm)a.
| model\functional | B3LYP | CAM-B3YP | PBE | PBEb | ωB97X-D | M06-2X | CIS |
|---|---|---|---|---|---|---|---|
| CPCM | 448 (0.2715) | 423 (0.2067) | 513 (0.2005) | 514 (0.2005) | 422 (0.1906) | 422 (0.2048) | 361 (0.5763) |
| SMD | 445 (0.2765) | 420 (0.2166) | 506 (0.2131) | 508 (0.2125)c | 418 (0.2003) | 419 (0.2143) | 360 (0.6115) |
| PCM | 447 (0.2626) | 422 (0.1992) | 512 (0.1934) | 513 (0.1933) | 421 (0.1836) | 421 (0.1973) | 358 (0.5405) |
| gas phase | 459 (0.2169) | 429 (0.1523) | 553 (0.1051) | 552 (0.1054) | 429 (0.137) | 428 (0.1513) | 386 (0.2389) |
Values in parentheses represent oscillator strength.
PBE/6-311+G(2d,2p).
The result of PBE0/6-311+G(2d,2p) using the SMD model is 436 nm (0.2739).
The hydrogen bond is one of the reasons responsible for the strength, broadening, and shift of the absorption peaks. Table 9 presents the values of Eex, oscillator strength f, and transition configurations of the intense peaks for enol, keto, and rotamers. The keto form with the short HB is accompanied by a large redshift (extended to 800 nm) followed by the enol form (extended to 650 nm), and the lower shift has been found with the rotamers (extended to 550 nm), as depicted in Figure 6. As expected from the lower Eg value of the keto, enol, and rotamers (Table 5), the maxima in the electronic absorption spectra of keto in comparison with enol, R1,R2, and R3 are shifted bathochromically by 17, 66, 51, and 49 nm, respectively. The intense peaks in the UV–vis spectra of keto spread over the range of 346–638 nm followed by enol that illustrates spreading over the range 272–506 nm, and the their rotamers (R1,R2, and R3) peaks spread over the range (280–479, 273–483, and 276–471 nm, respectively). The strong electronic absorption of keto is attributed to the HOMO–3 to LUMO and HOMO–2 to LUMO+1 transitions. The maximum absorption peak for enol appears at a lower wavelength than keto, which is attributed to HOMO–2 to LUMO+1 and HOMO–1 to LUMO+1 transitions. However, the strong electronic absorption of the rotamer R1 is mainly due to the HOMO–1 to LUMO+2 transitions. The strong electronic absorption of the studied rotamers has different contribution, as shown in Table 10.
Figure 6.

Simulated UV/Vis absorption spectra for the keto, enol, and rotamers of 2OHPhAZ at TD-PBE/6–311 + G(d,p).
Table 10. Excitation Energies (eV) at (TD-PBE-SMD, Acetonitrile)/6-311 + G(d,p)//B3LYP/6-31G(d,p), Oscillator Strengths (f > 0.15), and their Transition Characters for 2OHPhAZ Tautomer and Rotamers.
| compound | state | Ea | f | assignmentb |
|---|---|---|---|---|
| 2OHPhAZ | 1 | 2.45 (506) | 0.2131 | H → L (88%) |
| 4 | 3.28 (378) | 0.2137 | H–2 → L (55%), H–1 → L+1 (27%), H → L+1 (11%) | |
| 6 | 3.76 (329) | 0.2990 | H–2 → L (10%), H–2 → L+1 (27%), H–1 → L + 1 (40%), H → L+1 (12%) | |
| 8 | 4.18 (297) | 0.2947 | H–4 → L (27%), H–2 → L+1 (45%), H–1 → L + 1 (13%), H → L+2 (10%) | |
| 2OPhAZ | 1 | 1.94 (638) | 0.2023 | H → L (89%) |
| 4 | 2.81 (441) | 0.1868 | H–2 → L (90%) | |
| 7 | 3.59 (346) | 0.4396 | H–3 → L (40%), H–2 → L+1 (34%), H → L+2 (14%) | |
| R1 | 1 | 2.59 (479) | 0.1951 | H → L (88%) |
| 6 | 3.84 (323) | 0.2506 | H–4 → L (15%), H-2 → L+1 (31%), H-1 → L+1 (26%), H → L+1 (17%) | |
| 9 | 4.17 (297) | 0.2348 | H–3 → L + 1 (13%), H–2 → L+1 (16%), H → L+2 (50%) | |
| 10 | 4.43 (280) | 0.3647 | H–1 → L+2 (70%) | |
| R2 | 1 | 2.56 (483) | 0.1759 | H → L (86%), H–2 → L (11%) |
| 8 | 4.20 (295) | 0.3645 | H–3 → L+1 (24%), H → L+2 (53%) | |
| 10 | 4.54 (273) | 0.1509 | H–1 → L+2 (82%) | |
| R3 | 1 | 2.63 (471) | 0.2036 | H–2 → L (11%), H → L (84%) |
| 3 | 3.15 (393) | 0.1745 | H–2 → L (65%), H → L+1 (21%) | |
| 6 | 3.81 (325) | 0.1594 | H–2 → L+1 (60%), H–1 → L+1 (11%), H → L+1 (15%) | |
| 8 | 4.18 (297) | 0.4859 | H–1 → L+1 (11%), H → L+2 (66%) |
Values in parentheses are wave lengths in nm.
Only contributions above 10% are shown. H and L represent HOMO and LUMO, respectively.
The NTOs for high-intensity excited states of the investigated systems are shown in Figure 7 to analyze the nature of absorption. The occupied and unoccupied NTOs are referred to as “hole” and “particles” transition orbitals, respectively. The NTOs generally give a simpler description of the excited state than the canonical orbitals. As displayed in Figure S4, where the canonical orbitals were used, the dominant transitions are π–π* for the excitations with some contribution from n−π* excitation. This makes the analysis of excitations cumbersome. However, as depicted in Figure 7, the hole NTOs contributing to the illustrated band shown in Figure 6 and Table 10 of all studied structures are delocalized over the whole molecular skeleton while the particles NTOs are mainly delocalized over either azaazulene or benzene rings. This suggests π–π* excitation.
Figure 7.
NTOs for the excitation with significant and small but non-negligible oscillation strengths for the studied structures at the PBE/6-311+G(d,p) level with solvent effects of acetonitrile through SMD. The displayed occupied (holes) and unoccupied (electrons) NTO pairs are only that have contribution more than 50% to each excited state (λ is eigenvalues of the pairs).
4. Conclusions
Tautomer and the related rotamers of 2-(2-hydroxyphenyl)-1-azaazulene (2OHPhAZ) were theoretically investigated for analyzing various molecular properties, stability, and aromaticity. Relative energy results reveal that M06-2X/6-311++G(2d,2p) has a good performance compared to CCSD(T)/6-311++G(2d,2p). The structural parameters and vibrational frequencies have been discussed at B3LYP/6-31G(d,p) and compared with the available experimental data. Comparisons with the experiment of 2-(2-hydroxyphenyl)-1-azaazulene show a very good agreement. All the levels of calculations demonstrate thermodynamic and kinetic stability of the enol form in a good agreement with the presence of enol experimentally exclusively as the most stable tautomer. The remarkable chemical shift of a proton at 14.76 and 19.40 ppm in the nuclear magnetic resonance spectrum has been attributed to the existence of LBHBs for the enol and keto forms, respectively. The aromaticity of the enol and rotamers forms is slightly higher than the keto form. Hardness and the electrophilicity trend show a good correlation with the HOMA and NICS aromaticity indices. The sites of electrophilic and nucleophilic were determined using Fukui functions. UV–vis absorption spectra (in the gas phase and ethanol) were examined by TD-DFT using B3LYP, CAM-B3LYP, PBE, PBE0, ωb97X-D, M06-2X, and CIS methods with the 6–311+G(d,p) basis sets for 2OHPhAZ. The TDDFT-PBE/SMD approaches exhibit good agreement regarding the first and second maximum excitation peaks. NTOs are used to indicate the π–π* nature of the transitions.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c00866.
Supplementary data associated with this article can be found in the online version. Figure S1: Optimized structures of interconversions of keto-enol tautomers and rotamers (TST, TSR) of 2-(2-Hydroxyphenyl)-1-azaazulene (2OHPhAZ) at B3LYP/6-31G(d,p). Figure S2: Potential energy profiles from the IRC calculation for tautomerization and rotamerization of 2OHPhAZ obtained at the B3LYP/6-31G (d, p) level. Figure S3: Change of bond lengths (angstroms) along reaction coordinates for tautomarization of 2OHPhAZ at the B3LYP/6-31G(d,p) level. Figure S4: Frontier orbitals of the studied structures at B3LYP/6-31G(d,p). Table S1: Root-mean-square atom positional deviations (RMSD) values, in Å, calculated for the studied system on the calculated structure of enol (2OHPhAZ) in the gas phase. Table S2: dihedral angle in the degree of N1–C2–C1′–C2′, C2–C1′–C2′–O1 and C1′–C2′–O1–H1 for R1, R2, and R3 with their corresponding TSs. Table S3: Relative Gibbs free energies (ΔG298) and Gibbs Free Energy Barrier (ΔG298‡) for the tautomer and rotamers forms of 2OHPhAZ in kcal/mol, in the gas-phase and under the effect of ethanol calculated at the B3LYP/6-31G(d,p), B3LYP/6-311++G(2d,2p) and ωB97XD/6-311++G(2d,2p) levels of theory. Table S4: 13C and 1H NMR chemical shifts (in ppm) calculated at the B3LYP/6-31G(d,p) level of theory for the 2OHPhAZ tautomer and rotamers using the GIAO method in CHCl3. Table S5: Condensed Fukui functions (fk+, fk–, fk0) of the hydrogen: atomsa for the investigated tautomer and rotamers at B3LYP/6-31G(d,p). Table S6: Calculate proton affinity (PA) for 2OHPhAZ tautomers and rotamers in the gas phase at B3LYP, M06-2X, and ωB97XD (PDF)
Author Contributions
A.B.E.: conceptualization, methodology, formal analysis, data curation, validation, visualization, and investigation, and writing-review and editing, S.H.E.: conceptualization, methodology, formal analysis, data curation, validation, visualization, investigation, and writing—review and editing. M.A.A.: conceptualization, methodology, formal analysis, data curation, validation, visualization, investigation, and writing—review and editing, M.A.M.M.: conceptualization, methodology, formal analysis, data curation, validation, visualization, investigation, and writing-review and editing. T.T.: formal analysis, data curation, validation, and writing—review and editing. A.M.E.: formal analysis, data curation, validation, visualization, investigation, supervision, and writing—review and editing. All authors read and approved the final manuscript.
The authors declare no competing financial interest.
Notes
Electronic Supporting Information data file is submitted. All data and materials support the published claims and comply with field standards.
Supplementary Material
References
- Anderson A. G.; Nelson J. J. A. Electrophilic Substitution Of Azulene. J. Am. Chem. Soc. 1950, 72, 3824–3825. 10.1021/ja01164a528. [DOI] [Google Scholar]
- Buenker R. J.; Peyerimhoff S. D. Ab Initio SCF Calculations for Azulene and Naphthalene. Chem. Phys. Lett. 1969, 3, 37–42. 10.1016/0009-2614(69)80014-X. [DOI] [Google Scholar]
- Liu R. S. H. Colorful Azulene and Its Equally Colorful Derivatives. J. Chem. Educ. 2002, 79, 183–185. 10.1021/ed079p183. [DOI] [Google Scholar]
- Lemal D. M.; Goldman G. D. Synthesis of Azulene, a Blue Hydrocarbon. J. Chem. Educ. 1988, 65, 923–925. 10.1021/ed065p923. [DOI] [Google Scholar]
- Mikheev Y. A.; Guseva L. N.; Ershov Y. A. The Structure of Dimers and the Nature of Azulene Chromaticity. Russ. J. Phys. Chem. A 2012, 86, 85–92. 10.1134/S0036024412010232. [DOI] [Google Scholar]
- Murakami A.; Kobayashi T.; Goldberg A.; Nakamura S.; Murakami A.; Kobayashi T.; Goldberg A.; Nakamura S. CASSCF and CASPT2 Studies on the Structures , Transition Energies, and Dipole Moments of Ground and Excited States for Azulene. J. Chem. Phys. 2004, 120, 1245–1252. 10.1063/1.1631914. [DOI] [PubMed] [Google Scholar]
- Pariser R. Electronic Spectrum and Structure of Azulene. J. Chem. Phys. 1956, 25, 1112–1116. 10.1063/1.1743159. [DOI] [Google Scholar]
- Birks J. B. The Photophysics of Azulene. Chem. Phys. Lett. 1972, 17, 370–372. 10.1016/0009-2614(72)87099-4. [DOI] [Google Scholar]
- Liu R. S. H.; Asato A. E.. Tuning the Color and Excited State Properties of the Azulenic Chromophore: NIR Absorbing Pigments and Materials. 2003, 4, 179–194, 10.1016/j.jphotochemrev.2003.09.001. [DOI]
- Rekka E.; CHrysselis M.; Siskou I.; Ourounakis A. K. Synthesis of New Azulene Derivatives and Study of Their Effect on Lipid Peroxidation and Lipoxygenase Activity. Chem. Pharm. Bull. 2002, 50, 904–907. 10.1248/cpb.50.904. [DOI] [PubMed] [Google Scholar]
- Wang B.; Lin Y.; Chang J.; Wang P. Theoretical Studies of Azulene and Its Derivatives. Can. J. Chem. 2000, 78, 224–232. 10.1139/v99-241. [DOI] [Google Scholar]
- El-Nahas A. M.; Staykov A.; Yoshizawa K. First-Principles Calculations of Electron Transport through Azulene. J. Phys. Chem. C 2016, 120, 9043–9052. 10.1021/acs.jpcc.6b00767. [DOI] [Google Scholar]
- Haidar E.; Tawfik S. A.; Stampfl C.; Hirao K.; Yoshizawa K.; Nakajima T.; Soliman K. A.; El-Nahas A. M. Attenuation of Redox Switching and Rectification in Azulenequinones/Hydroquinones after B and N Doping: A First-Principles Investigation. Adv. Theory Simul. 2020, 4, 2000203 10.1002/adts.202000203. [DOI] [Google Scholar]
- El-Nahas A. M.; Staykov A.; Yoshizawa K. Electrical Conductivity and Diode-Like Behavior of Substituted Azulene Electrical Conductivity and Diode-Like Behavior of Substituted Azulene. J. Phys. Chem. C 2017, 121, 2504–2511. 10.1021/acs.jpcc.6b10339. [DOI] [Google Scholar]
- Haidar E.-A.; Tawfik S. A.; Stampfl C.; Hirao K.; Yoshizawa K.; El-Demerdash S. H.; Nakajima T.; El-Nahas A. M. Electronic Transport Investigation of Redox-Switching of Azulenequinones/Hydroquinones via First-Principles Studies. Phys. Chem. Chem. Phys. 2019, 21, 17859–17867. 10.1039/C9CP03233A. [DOI] [PubMed] [Google Scholar]
- Gad S. F.; El-Demerdash S. H.; El-Mehasseb I. M.; El-Nahas A. M. Structure, Stability and Conversions of Tautomers and Rotamers of Azulene-Based Uracil Analogue. J. Mol. Struct. 2019, 1182, 271–282. 10.1016/j.molstruc.2019.01.020. [DOI] [Google Scholar]
- Meshhal M. M.; El-Demerdash S. H.; El-Nahas A. M. A Thermochemical Computational Study on Hydroxyquinolines and Their Azulene Analogues. J. Mol. Struct. 2019, 1183, 70–77. 10.1016/j.molstruc.2019.01.047. [DOI] [Google Scholar]
- Ramadan M.; Goeters S.; Watzer B.; Krause E.; Lohmann K.; Bauer R.; Hempel B.; Imming P.; Chemie P.; Philipps-uni V.; Jugendmedizin K.; Philipps-uni V.; Biologie P.; Heinrich-heine-uni V.; Wissenschaften P.; Karl-franzens-uni V.; Gmbh R.; Pharmazie F.; Martin-luther-uni V. Chamazulene Carboxylic Acid and Matricin: A Natural Profen and Its Natural Prodrug, Identified through Similarity to Synthetic Drug Substances. J. Nat. Prod. 2006, 69, 1041–1045. 10.1021/np0601556. [DOI] [PubMed] [Google Scholar]
- Ishihara M.; Wakabayashi H.; Motohashi N.; Sakagami H. Quantitative Structure–Cytotoxicity Relationship of Newly Synthesised Trihaloacetylazulenes Determined by a Semi-Empirical Molecular-Orbital Method (PM5). Anticancer Res. 2011, 31, 515–520. [PubMed] [Google Scholar]
- Abe N.; Gunji T. The Chemistry of Azaazulenes. Heterocycles 2010, 82, 201–248. 10.3987/REV-10-SR(E)1. [DOI] [Google Scholar]
- Kimura M. The Chemistry of Aza-Azulenes. J. Synth. Org. Chem., Jpn. 1981, 39, 690–700. 10.5059/yukigoseikyokaishi.39.690. [DOI] [Google Scholar]
- Turányi T.; Zalotai L.; Dóbé S.; Bérces T. Effect of the Uncertainty of Kinetic and Thermodynamic Data on Methane Flame Simulation Results. Phys. Chem. Chem. Phys. 2002, 4, 2568–2578. 10.1039/b109154a. [DOI] [Google Scholar]
- Abe N.; Hashimoto E.; Fujii H.; Murakami Y.; Tagashira S.; Kakehi A. Synthesis of 2-Chloro-8-(2-Pyridyl)-1-Azaazulene and Its Metal Complexes. Heterocycles 2004, 63, 2341–2348. 10.3987/COM-04-10182. [DOI] [Google Scholar]
- Oda M.; Miyawaki D.; Matsumoto N.; Kuroda S. Palladium-Catalyzed Amination of 2-Chloro-1-Azaazulene with 2-Aminopyridine. Heterocycles 2011, 83, 547–554. 10.3987/COM-10-12109. [DOI] [Google Scholar]
- Oda M.; Ogura K.; Thanh N. C.; Kishi S.; Kuroda S.; Fujimori K.; Noda T.; Abe N. Synthesis and Properties of 2-(2-Pyridyl)-1-Azaazulene. Tetrahedron Lett. 2007, 48, 4471–4475. 10.1016/j.tetlet.2007.05.008. [DOI] [Google Scholar]
- Oda M.; Sugiyama A.; Takeuchi R.; Fujiwara Y.; Miyatake R.; Abe T.; Kuroda S. Synthesis, Molecular Structure, and Properties of 2-(2-Hydroxyphenyl)-1- Azaazulene. Eur. J. Org. Chem. 2012, 6, 2231–2236. 10.1002/ejoc.201101831. [DOI] [Google Scholar]
- Philips J. P. The Reactions of 8-Quinolinol. Chem. Rev. 1956, 56, 271–297. 10.1021/cr50008a003. [DOI] [Google Scholar]
- Casado-Sánchez A.; Uygur M.; González-Muñoz D.; Aguilar-Galindo F.; Nova-Fernández J. L.; Arranz-Plaza J.; Díaz-Tendero S.; Cabrera S.; Mancheño O. G.; Alemán J. 8-Mercaptoquinoline as a Ligand for Enhancing the Photocatalytic Activity of Pt(II) Coordination Complexes: Reactions and Mechanistic Insights. J. Org. Chem. 2019, 2019, 6437–6447. 10.1021/acs.joc.9b00520. [DOI] [PubMed] [Google Scholar]
- Vila J. A.; Arnautova Y. A.; Vorobjev Y.; Scheraga H. A. Assessing the Fractions of Tautomeric Forms of the Imidazole Ring of Histidine in Proteins as a Function of PH. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 5602–5607. 10.1073/pnas.1102373108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minkin V. I.; Olekhnovich L. P.; Zhdanov Y. A. Molecular Design of Tautomeric Compounds. Acc. Chem. Res. 1981, 3, 210–217. 10.1021/ar00067a003. [DOI] [Google Scholar]
- Pople J. A.; Head-Gordon M.; Raghavachari K. Quadratic Configuration Interaction. A General Technique for Determining Electron Correlation Energies. J. Chem. Phys. 1987, 87, 5968–5975. 10.1063/1.453520. [DOI] [Google Scholar]
- Purvis G. D.; Bartlett R. J. A Full Coupled-Cluster Singles and Doubles Model: The Inclusion of Disconnected Triples. J. Chem. Phys. 1982, 76, 1910–1918. 10.1063/1.443164. [DOI] [Google Scholar]
- Buhl M.; van Wiillen C. Computational Evidence for a New C84 Isomer. Chem. Phys. Lett. 1995, 247, 63–68. 10.1016/0009-2614(95)01193-6. [DOI] [Google Scholar]
- von Ragué Schleyer P.; Maerker C.; Dransfeld A.; Jiao H.; van Eikema Hommes N. J. R. Nucleus-Independent Chemical Shifts: A Simple and Efficient Aromaticity Probe. J. Am. Chem. Soc. 1996, 118, 6317–6318. 10.1021/ja960582d. [DOI] [PubMed] [Google Scholar]
- Sobczyk L.; Grabowski S. J.; Krygowski T. M. Interrelation between H-Bond and Pi-Electron Delocalization. Chem. Rev. 2005, 105, 3513–3560. 10.1021/cr030083c. [DOI] [PubMed] [Google Scholar]
- Krygowski T. M. Crystallographic Studies of Inter- and Intramolecular Interactions Reflected in Aromatic Character of Pi-Electron Systems. J. Chem. Inf. Comput. Sci. 1993, 33, 70–78. 10.1021/ci00011a011. [DOI] [Google Scholar]
- Kruszewski J.; Krygows T. M. Definition of Aromaticity Basing on the Harmonic Oscillator Model. Tetrahedron Lett. 1972, 13, 3839–3842. 10.1016/S0040-4039(01)94175-9. [DOI] [Google Scholar]
- Becke A. D. A New Mixing of Hartree-Fock and Local-Density-Functional Theories. J. Chem. Phys. 1993, 98, 1372–1377. 10.1063/1.464304. [DOI] [Google Scholar]
- Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
- Stephens P. J.; Devlin F. J.; Chabalowski C.; Frisch M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. 10.1021/j100096a001. [DOI] [Google Scholar]
- Zhurko G. A.; Zhurko D. A. Chemcraft 1.8. www.chemcraftprog.com.
- Gonzalez C.; Sosa C.; Schlegel H. B. Ab Initio Study of the Addition Reaction of the Methyl Radical to Ethylene and Formaldehyde. J. Phys. Chem. 1989, 93, 2435–2440. 10.1021/j100343a042. [DOI] [Google Scholar]
- Gonzalez C.; Schlegel H. B. Reaction Path Following in Mass-Weighted Internal Coordinates. J. Phys. Chem. 1990, 94, 5523–5527. 10.1021/j100377a021. [DOI] [Google Scholar]
- 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 M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215–241. 10.1007/s00214-007-0310-x. [DOI] [Google Scholar]
- Zhao Y.; Schultz N. E.; Truhlar D. G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364–382. 10.1021/ct0502763. [DOI] [PubMed] [Google Scholar]
- Chai J. D.; Head-Gordon M. Systematic Optimization of Long-Range Corrected Hybrid Density Functionals. J. Chem. Phys. 2008, 128, 084106–084114. 10.1063/1.2834918. [DOI] [PubMed] [Google Scholar]
- Mardirossian N.; Head-Gordon M. Ωb97X-V: A 10-Parameter, Range-Separated Hybrid, Generalized Gradient Approximation Density Functional with Nonlocal Correlation, Designed by a Survival-of-the-Fittest Strategy. Phys. Chem. Chem. Phys. 2014, 16, 9904–9924. 10.1039/c3cp54374a. [DOI] [PubMed] [Google Scholar]
- Lin Y. S.; Li G.-D.; Mao S. P.; Chai J. D. Long-Range Corrected Hybrid Density Functionals with Improved Dispersion Corrections. J. Chem. Theory Comput. 2013, 9, 263–272. 10.1021/ct300715s. [DOI] [PubMed] [Google Scholar]
- Chai J. D.; Head-Gordon M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. 10.1039/b810189b. [DOI] [PubMed] [Google Scholar]
- Reed A. E.; Curtiss L. A.; Weinhold F. Intermolecular Interactions from a Natural Bond Orbital, Donor-Acceptor Viewpoint. Chem. Rev. 1988, 88, 899–926. 10.1021/cr00088a005. [DOI] [Google Scholar]
- Reed A. E.; Weinstock R. B.; Weinhold F. Natural Population Analysis. J. Chem. Phys. 1985, 83, 735–746. 10.1063/1.449486. [DOI] [Google Scholar]
- Glendening E. D.; Reed A. E.; Carpenter J. E.; Weinhold F.. NBO Version 3.1; Gaussian Inc.: Pittsburgh, 2003. [Google Scholar]
- Marenich A. V.; Cramer C. J.; Truhlar D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. 10.1021/jp810292n. [DOI] [PubMed] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Had M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09 Revision A.1; Gaussian Inc.: Wallingford, CT, 2009. [Google Scholar]
- Wolinski K.; Hinton J. F.; Pulay P. Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. 10.1021/ja00179a005. [DOI] [Google Scholar]
- Wolff S. K.; Ziegler T. Calculation of DFT-GIAO NMR Shifts with the Inclusion of Spin-Orbit Coupling. J. Chem. Phys. 1998, 109, 895–905. 10.1063/1.476630. [DOI] [Google Scholar]
- Cheeseman J. R.; Trucks G. W.; Keith T. A.; Frisch M. J. A Comparison of Models for Calculating Nuclear Magnetic Resonance Shielding Tensors. J. Chem. Phys. 1996, 104, 5497–5509. 10.1063/1.471789. [DOI] [Google Scholar]
- Stanger A. Nucleus-Independent Chemical Shifts (NICS): Distance Dependence and Revised Criteria for Aromaticity and Antiaromaticity. J. Org. Chem. 2006, 71, 883–893. 10.1021/jo051746o. [DOI] [PubMed] [Google Scholar]
- Fallah-bagher-shaidaei H.; Wannere C. S. Which NICS Aromaticity Index for Planar π Rings Is Best ?. Org. Lett. 2006, 8, 863–866. 10.1021/ol0529546. [DOI] [PubMed] [Google Scholar]
- Lucila L.; Martinez L.; Mitnik G.; Borunda E. O. Theoretical Calculation of Uv-Vis, IR Spectra and Reactivity Properties of Tamoxifen Drug: A Methodology Comparison. MOJ Bioorg. Org. Chem. 2017, 1, 87–95. 10.15406/mojboc.2017.01.00017. [DOI] [Google Scholar]
- Flores M. C.; Márquez E. A.; Mora J. R. Molecular Modeling Studies of Bromopyrrole Alkaloids as Potential Antimalarial Compounds: A DFT Approach. Med. Chem. Res. 2018, 27, 844–856. 10.1007/s00044-017-2107-3. [DOI] [Google Scholar]
- Mishra V. R.; Sekar N. Photostability of Coumarin Laser Dyes - a Mechanistic Study Using Global and Local Reactivity Descriptors. J. Fluoresc. 2017, 27, 1101–1108. 10.1007/s10895-017-2045-y. [DOI] [PubMed] [Google Scholar]
- Das S.; Shedge S. V.; Pal S. Critical Study of the Charge Transfer Parameter for the Calculation of Interaction Energy Using the Local Hard-Soft Acid-Base Principle. J. Phys. Chem. A 2013, 117, 10933–10943. 10.1021/jp407070h. [DOI] [PubMed] [Google Scholar]
- Morell C.; Toro-labbe A. New Dual Descriptor for Chemical Reactivity. J. Phys. Chem. A 2005, 109, 205–212. 10.1021/jp046577a. [DOI] [PubMed] [Google Scholar]
- Sánchez-márquez J.; Zorrilla D.; Sánchez-coronilla A.; Santos D. M. D. L.; Navas J.; Fernández-lorenzo C. Introducing “UCA-FUKUI” Software: Reactivity-Index Calculations. J. Mol. Model. 2014, 20, 1–13. 10.1007/s00894-014-2492-1. [DOI] [PubMed] [Google Scholar]
- Chattaraj P. K.; Maiti B.; Sarkar U. Philicity: A Unified Treatment of Chemical Reactivity and Selectivity Pratim. J. Phys. Chem. A 2003, 107, 4973–4975. 10.1021/jp034707u. [DOI] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 3, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Ernzerhof M.; Perdew J. P. Generalized Gradient Approximation to the Angle- and System-Averaged Exchange Hole. J. Chem. Phys. 1998, 109, 3313–3320. 10.1063/1.476928. [DOI] [Google Scholar]
- Adamo C.; Barone V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model. J. Chem. Phys. 1999, 110, 6158–6170. 10.1063/1.478522. [DOI] [Google Scholar]
- Jacquemin D.; Perpe E. A.; Scuseria G. E.; Ciofini I.; Adamo C. TD-DFT Performance for the Visible Absorption Spectra of Organic Dyes: Conventional versus Long-Range Hybrids. J. Chem. Theory Comput. 2008, 4, 123–135. 10.1021/ct700187z. [DOI] [PubMed] [Google Scholar]
- O’Boyle N. M.; Tenderholt A. L.; Langner K. M. Cclib: A Library for Package-Independent Computational Chemistry Algorithms. J. Comput. Chem. 2008, 29, 839–845. 10.1002/jcc.20823. [DOI] [PubMed] [Google Scholar]
- Martin R. L. Natural Transition Orbitals. J. Chem. Phys. 2003, 118, 4775–4777. 10.1063/1.1558471. [DOI] [Google Scholar]
- Michelson A. Z.; Petronico A.; Lee J. K. 2-Pyridone and Derivatives: Gas-Phase Acidity, Proton Affinity, Tautomer Preference, and Leaving Group Ability. J. Org. Chem. 2012, 77, 1623–1631. 10.1021/jo201991y. [DOI] [PubMed] [Google Scholar]
- Walker M.; Harvey A. J. A.; Sen A.; Dessent C. E. H. Performance of M06, M06-2X, and M06-HF Density Functionals for Conformationally Flexible Anionic Clusters: M06 Functionals Perform Better than B3LYP for a Model System with Dispersion and Ionic Hydrogen-Bonding Interactions. J. Phys. Chem. A 2013, 117, 12590–12600. 10.1021/jp408166m. [DOI] [PubMed] [Google Scholar]
- Hejazi S. A.; Osman O. I.; Alyoubi A. O.; Aziz S. G.; Hilal R. H. The Thermodynamic and Kinetic Properties of 2-Hydroxypyridine/2-Pyridone Tautomerization: A Theoretical and Computational Revisit. Int. J. Mol. Sci. 2016, 17, 1893. 10.3390/ijms17111893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaczorowska M. A.; Kaczmarek-Kędziera A.; Ośmiałowski B. Tautomeric Equilibrium, Proton Affinity and Mass Spectrometry Fragmentation of Flexible Hydrogen-Bonded Precursors and Rigid N → BF 2 Fluorescent Dyes. Sci. Rep. 2021, 11, 87–100. 10.1038/s41598-021-94978-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berenbeim J. A.; Wong N. G. K.; Cockett M. C. R.; Berden G.; Oomens J.; Rijs A. M.; Dessent C. E. H. Unravelling the Keto–Enol Tautomer Dependent Photochemistry and Degradation Pathways of the Protonated UVA Filter Avobenzone. J. Phys. Chem. A 2020, 124, 2919–2930. 10.1021/acs.jpca.0c01295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koehl P. Protein Structure Similarities. Curr. Opin. Struct. Biol. 2001, 11, 348–353. 10.1016/S0959-440X(00)00214-1. [DOI] [PubMed] [Google Scholar]
- Shibuya T. Efficient Substructure RMSD Query Algorithms. J. Comput. Biol. 2007, 14, 1201–1207. 10.1089/cmb.2007.0079. [DOI] [PubMed] [Google Scholar]
- Kabsch W. A Solution for the Best Rotation to Relate Two Sets of Vectors. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 922–923. 10.1107/S0567739476001873. [DOI] [Google Scholar]
- Shibuya T. Searching Protein Three-Dimensional Structures in Faster Than Linear Time. J. Comput. Biol. 2010, 17, 593–602. 10.1089/cmb.2009.0217. [DOI] [PubMed] [Google Scholar]
- Smallwood C. J.; Mcallister M. A. Characterization of Low-Barrier Hydrogen Bonds . 7. Relationship between Strength and Geometry of Short-Strong Hydrogen Bonds. The Formic Acid-Formate Anion Model System. An Ab Initio and DFT Investigation. J. Am. Chem. Soc. 1997, 119, 11277–11281. 10.1021/ja972517p. [DOI] [Google Scholar]
- Mckenzie R. H.; Bekker C.; Athokpam B.; Ramesh S. G.; Mckenzie R. H.; Bekker C.; Athokpam B.; Ramesh S. G. Effect of Quantum Nuclear Motion on Hydrogen Bonding. J. Chem. Phys. 2014, 140, 174508. 10.1063/1.4873352. [DOI] [PubMed] [Google Scholar]
- Lenain P.; Mandado M.; Mosquera R. A.; Bultinck P. Interplay between Hydrogen-Bond Formation and Multicenter π -Electron Delocalization. J. Phys. Chem. A 2008, 112, 10689–10696. 10.1021/jp805084n. [DOI] [PubMed] [Google Scholar]
- Cleland W. W.; Frey P. A.; Gerlt J. A. The Low Barrier Hydrogen Bond in Enzymatic Catalysis. J. Biol. Chem. 1998, 273, 25529–25532. 10.1074/jbc.273.40.25529. [DOI] [PubMed] [Google Scholar]
- Kong X.; Brinkmann A.; Terskikh V.; Wasylishen R. E.; Bernard G. M.; Duan Z.; Wu Q.; Wu G. Proton Probability Distribution in the O ··· H ··· O Low-Barrier Hydrogen Bond: A Combined Solid-State NMR and Quantum Chemical Computational Study of Dibenzoylmethane and Curcumin. J. Phys. Chem. B 2016, 120, 11692–11704. 10.1021/acs.jpcb.6b08091. [DOI] [PubMed] [Google Scholar]
- Perrin C. L.; Nielson J. B. “Strong” Hydrogen Bonds in Chemistry and Biology. Annu. Rev. Phys. Chem. 1997, 48, 511–544. 10.1146/annurev.physchem.48.1.511. [DOI] [PubMed] [Google Scholar]
- Gilli P.; Bertolasi V.; Pretto L.; Gilli G. The Nature of Solid-State N-H...O/O-H...N Tautomeric Competition in Resonant Systems. Intramolecular Proton Transfer in Low-Barrier Hydrogen Bonds Formed by the ...O=C-C=N-NH... ...HO-C=C-N=N... Ketohydrazone-Azoenol System.A Variable-Temperature X-Ray Cr. J. Am. Chem. Soc. 2002, 124, 13554–13567. 10.1021/ja020589x. [DOI] [PubMed] [Google Scholar]
- Hammond G. S. A Correlation of Reaction Rates. J. Am. Chem. Soc. 1955, 77, 334–338. 10.1021/ja01607a027. [DOI] [Google Scholar]
- Gilli G.; Bellucci F.; Ferretti V.; Bertolasi V. Evidence for Resonance-Assisted Hydrogen Bonding from Crystal-Structure Correlations on the Enol Form of the,& Diketone Fragment. J. Am. Chem. Soc. 1989, 111, 1023–1028. 10.1021/ja00185a035. [DOI] [Google Scholar]
- Palusiak M.; Simon S.; Sola M. Interplay between Intramolecular Resonance-Assisted Hydrogen Bonding and Aromaticity in o-Hydroxyaryl Aldehydes. J. Org. Chem. 2006, 71, 5241–5248. 10.1021/jo060591x. [DOI] [PubMed] [Google Scholar]
- Krygowski T. M.; Zachara J. E.; Moszyn R. Theoretical Study of Changes in π-Electron Delocalization in the Analogues of an Ortho-Hydroxy Schiff Base When the Proton Is Replaced with Li+ or BeH+. J. Chem. Inf. Model. 2005, 45, 1837–1841. 10.1021/ci0502911. [DOI] [PubMed] [Google Scholar]
- Garcia-Viloca M.; Gonzalez-Lafont A.; Lluch M. Theoretical Study of the Low-Barrier Hydrogen Bond in the Hydrogen Maleate Anion in the Gas Phase. Comparison with Normal Hydrogen Bonds. J. Am. Chem. Soc. 1997, 119, 1081–1086. 10.1021/ja962662n. [DOI] [Google Scholar]
- Tolstoy P. M.; Guo J.; Koeppe B.; Golubev N. S.; Denisov G. S.; Smirnov S. N.; Limbach H. Geometries and Tautomerism of OHN Hydrogen Bonds in Aprotic Solution Probed by H/D Isotope Effects on 13 C NMR Chemical Shifts. J. Phys. Chem. A 2010, 114, 10775–10782. 10.1021/jp1027146. [DOI] [PubMed] [Google Scholar]
- Kumar G. A.; Mcallister M. A. Theoretical Investigation of the Relationship between Proton NMR Chemical Shift and Hydrogen Bond Strength. J. Org. Chem. 1998, 63, 6968–6972. 10.1021/jo980759h. [DOI] [PubMed] [Google Scholar]
- Garcia-Viloca M.; Gelabert R.; Gonzalez-Lafont A.; Moreno M.; Lluch J. M. Is an Extremely Low-Field Proton Signal in the NMR Spectrum Conclusive Evidence for a Low-Barrier Hydrogen Bond?. J. Phys. Chem. A 1997, 101, 8727–8733. 10.1021/jp972335h. [DOI] [Google Scholar]
- Hibbert F.; Emsley J. Hydrogen Bonding and Chemical Reactivity. J. Adv. Phys. Org. Chem. 1990, 26, 255–379. 10.1016/S0065-3160(08)60047-7. [DOI] [Google Scholar]
- von Ragué Schleyer P.; Manoharan M.; Wang Z.; Kiran B.; Jiao H.; Puchta R.; van Eikema Hommes N. J. R. Dissected Nucleus-Independent Chemical Shift Analysis of π -Aromaticity and Antiaromaticity. Org. Lett. 2001, 3, 2465–2468. 10.1021/ol016217v. [DOI] [PubMed] [Google Scholar]
- Corminboeuf C.; Heine T.; Seifert G.; von Ragué Schleyer P.; Webe J. Induced Magnetic Fields in Aromatic [n]-Annulenes—Interpretation of NICS Tensor Components. Phys. Chem. Chem. Phys. 2004, 6, 273–276. 10.1039/B313383B. [DOI] [Google Scholar]
- Sasaki Y.; Takase M.; Okujima T.; Mori S.; Uno H. Synthesis and Redox Properties of Pyrrole- and Azulene-Fused Azacoronene. Org. Lett. 2019, 21, 1900–1903. 10.1021/acs.orglett.9b00515. [DOI] [PubMed] [Google Scholar]
- Lazzeretti P. Assessment of Aromaticity via Molecular Response Properties Y. Phys. Chem. Chem. Phys. 2004, 6, 217–223. 10.1039/B311178D. [DOI] [Google Scholar]
- Krygowski T. M.; Ejsmont K.; Stepien B. T.; Cyran M. K.; Poater J.; Sola M. Relation between the Substituent Effect and Aromaticity. J. Org. Chem. 2004, 69, 6634–6640. 10.1021/jo0492113. [DOI] [PubMed] [Google Scholar]
- Krygowski T. M.; Stepien B. T.; Cyranski M. K.; Ejsmont K. Relation between Resonance Energy and Substituentresonance Effect in P-Phenols. Phys. Org. Chem. 2005, 18, 886–891. 10.1002/poc.960. [DOI] [Google Scholar]
- Inostroza D.; Garcıa V.; Yanez O.; Torres-vega J. J.; Vasquez-Espinal A.; Pino-rios R.; Báez-Grez R.; Tiznado W. On the NICS Limitations to Predict Local and Global Current Pathways in Polycyclic Systems. New J. Chem. 2021, 45, 8345–8351. 10.1039/D1NJ01510A. [DOI] [Google Scholar]
- Báez-Grez R.; Rabanal-León W. A.; Alvarez-Thon L.; Ruiz L.; Tiznado W.; Pino-Rios R. Aromaticity in Heterocyclic Analogues of Benzene: Dissected NICS and Current Density Analysis. J. Phys. Org. Chem. 2019, 32, 1–7. 10.1002/poc.3823. [DOI] [Google Scholar]
- Seal P.; Chakrabarti S. Is Nucleus-Independent Chemical Shift Scan a Reliable Aromaticity Index for Planar Heteroatomic Ring Systems?. J. Phys. Chem. A 2007, 111, 9988–9994. 10.1021/jp074440p. [DOI] [PubMed] [Google Scholar]
- Foroutan-Nejad C. Is NICS a Reliable Aromaticity Index for Transition Metal Clusters ?. Theor. Chem. Acc. 2015, 134, 1–9. 10.1007/s00214-015-1617-7. [DOI] [Google Scholar]
- Mrozek A.; Karolak-Wojciechowska J.; Amiel P.; Barbe J. Five-Membered Heterocycles. Part I. Application of the HOMA Index to 1, 2,4-Trizoles. J. Mol. Struct. 2000, 524, 151–157. 10.1016/S0022-2860(99)00459-7. [DOI] [Google Scholar]
- Dominikowska J.; Palusiak M. EL: The New Aromaticity Measure Based on One-Electron Density Function. Struct. Chem. 2012, 23, 1173–1183. 10.1007/s11224-011-9941-6. [DOI] [Google Scholar]
- Frizzo C. P.; Martins M. A. P. Aromaticity in Heterocycles : New HOMA Index Parametrization. Struct. Chem. 2012, 23, 375–380. 10.1007/s11224-011-9883-z. [DOI] [Google Scholar]
- Corminboeuf C.; Heine T.; Webera J. Evaluation of Aromaticity: A New Dissected NICS Model Based on Canonical Orbitals. Phys. Chem. Chem. Phys. 2003, 5, 246–251. 10.1039/b209674a. [DOI] [Google Scholar]
- Meshhal M. M.; Shibl M. F.; El-Demerdash S. H.; El-Nahas A. M. A Computational Study on Molecular Structure and Stability of Tautomers of Dipyrrole-Based Phenanthroline Analogue. Comput. Theor. Chem. 2018, 1145, 6–14. 10.1016/j.comptc.2018.10.003. [DOI] [Google Scholar]
- Novak I. Distorted Naphthalenes and Azulenes. Comput. Theor. Chem. 2017, 1117, 251–257. 10.1016/j.comptc.2017.08.007. [DOI] [Google Scholar]
- Gümüş S. A Computational Study on Azaazulenes. Heterocycl. Commun. 2013, 19, 369–373. 10.1515/hc-2013-0100. [DOI] [Google Scholar]
- Cyranski M. K.; Krygowski T. M.; Wisiorowski M.; van Eikema Hommes N. J. R.; von Ragué Schleyer P. Global and Local Aromaticity in Porphyrins: An Analysis Based on Molecular Geometries and Nucleus Independent Chemical Shift. Angew. Chem., Int. Ed. 1998, 37, 177–180. 10.1002/(SICI)1521-3773(19980202)37:1/23.0.CO;2-H. [DOI] [Google Scholar]
- Chamizo J. A.; Morgado J.; Sosa P. Organometallic Aromaticity. Organometallics 1993, 12, 5005–5007. 10.1021/om00036a047. [DOI] [Google Scholar]
- Roy D. R.; Bultinck P.; Subramanian V.; Chattaraj P. K. Bonding , Reactivity and Aromaticity in the Light of the Multicenter Indices. J. Mol. Struct. THEOCHEMM 2008, 854, 35–39. 10.1016/j.theochem.2007.12.042. [DOI] [Google Scholar]
- De Proft F.; Geerlings P. Relative Hardness as a Measure of Aromaticity. Phys. Chem. Chem. Phys. 2004, 6, 242–248. 10.1039/B312566C. [DOI] [Google Scholar]
- Pausescu I.; Medeleanu M.; Stefanescu M.; Peter F.; Pop R. A DFT Study on the Stability and Aromaticity of Heterobenzenes Containing Group 15 Elements. Heteroat. Chem. 2015, 36, 206–214. 10.1002/hc.21250. [DOI] [Google Scholar]
- De Proft F.; Geerlings P. Conceptual and Computational DFT in the Study of Aromaticity. Chem. Rev. 2001, 2001, 1451–1464. 10.1021/cr9903205. [DOI] [PubMed] [Google Scholar]
- Cai T.; Xu L.; Anderson M. R.; Ge Z.; Zuo T.; Wang X.; Olmstead M. M.; Balch A. L.; Gibson H. W.; Dorn H. C. Structure and Enhanced Reactivity Rates of the D5h Sc3N@C80 and Lu3N@C80 Metallofullerene Isomers : The Importance of the Pyracylene Motif. J. Am. Chem. Soc. 2006, 128, 8581–8589. 10.1021/ja0615573. [DOI] [PubMed] [Google Scholar]
- Chermette H. Chemical Reactivity Indexes in Density Functional Theory. J. Comput. Chem. 1999, 20, 129–154. . [DOI] [Google Scholar]
- Feng L.; Yang H.; Wang F. Experimental and Theoretical Studies for Corrosion Inhibition of Carbon Steel by Imidazoline Derivative in 5% NaCl Saturated Ca (OH)2 Solution. Electrochim. Acta 2011, 58, 427–436. 10.1016/j.electacta.2011.09.063. [DOI] [Google Scholar]
- Rameshkumar S.; Danaee I.; Rashvandavei M.; Vijayan M. Quantum Chemical and Experimental Investigations on Equipotent Effects of (+)R and (−)S Enantiomers of Racemic Amisulpride as Eco-Friendly Corrosion Inhibitors for Mild Steel in Acidic Solution. J. Mol. Liq. 2015, 212, 168–186. 10.1016/j.molliq.2015.09.001. [DOI] [Google Scholar]
- Topol I. A.; Tawa G. J.; Burt S. K.; Rashin A. A. On the Structure and Thermodynamics of Solvated Monoatomic Ions Using a Hybrid Solvation Model. J. Chem. Phys. 1999, 111, 10998–11014. 10.1063/1.480486. [DOI] [Google Scholar]
- Jang Y. H.; Sowers L. C.; Çaǧin T.; Goddard W. A. First Principles Calculation of PKa Values for 5-Substituted Uracils. J. Phys. Chem. A 2001, 2001, 274–280. 10.1021/jp994432b. [DOI] [Google Scholar]
- Tissandier M. D.; Cowen K. A.; Feng W. Y.; Gundlach E.; Cohen M. H.; Earhart A. D.; Coe J. V.; Tuttle T. R. The Proton’s Absolute Aqueous Enthalpy and Gibbs Free Energy of Solvation from Cluster-Ion Solvation Data. J. Phys. Chem. A 1998, 102, 7787–7794. 10.1021/jp982638r. [DOI] [Google Scholar]
- Camaioni D. M.; Schwerdtfeger C. A. Comment on “Accurate Experimental Values for the Free Energies of Hydration of H+, OH−, and H3O+”. J. Phys. Chem. A 2005, 109, 10795–10797. 10.1021/jp054088k. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








