<<

Canadian Journal of Chemistry

Theoretical study on the acidities of , , and their hydrocarbon analogues in DMSO

Journal: Canadian Journal of Chemistry

Manuscript ID cjc-2018-0032.R1

Manuscript Type: Article

Date Submitted by the Author: 13-May-2018

Complete List of Authors: Ariche, Berkane; Modelisation and computational Méthods Laboratory, Faculty of Sciences, University of Saida, Chemistry RAHMOUNI,Draft Ali; Universite Dr Tahar Moulay de Saida, Chimie Is the invited manuscript for consideration in a Special Not applicable (regular submission) Issue?:

Keyword: pKa calculation, index, Pyrolle, DMSO, DFT

https://mc06.manuscriptcentral.com/cjc-pubs Page 1 of 33 Canadian Journal of Chemistry

Theoretical study on the acidities of pyrrole, indole, carbazole and their

hydrocarbon analogues in DMSO

Ariche Berkane, Rahmouni Ali*

Modeling and Calculation Methods Laboratory, University of Saida, B.P. 138, 20002 Saida, Algeria

*Corresponding author

[email protected]

Abstract

SMD and IEFPCM continuum solvation models have been used, in combination with three

quantum chemistry methods (B3LYP,Draft M062X and CBSQB3) to study the acidities of

pyrrole, indole and carbazole as well as their hydrocarbon analogues in DMSO, following a

direct thermodynamic method. Theoretical parameters such as aromaticity indices (HOMA

and SA), molecular electrostatic potential (MEP) and atomic charges have been calculated

using B3LYP/6311++G(d,p) level of theory. Calculated pKa values indicate that there is

generally good agreement with experimental data, with all deviations being less than the

acceptable error for a directly calculated pKa values. The M062X functional combined to

SMD solvation model provide the more accurate pKa values. MEP surfaces show clearly the

electron density change accompanying the deprotonation process and explain the relative

stability of conjugate bases. The HOMA aromaticity indices seem to be directly related to

acidity strength. The collected data have been used to elucidate the pKa trends for the series of

molecules under consideration.

Key words: pKa Calculation, Aromaticity, Density functional theory, CBSQB3, SMD,

Electrostatic interaction.

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 2 of 33

1. Introduction

Over the past several decades, acidbase equilibrium has remained one of the key concepts in chemistry. Thus, different computational strategies have been proposed to predict accurate pKa values1,2. For efficient computations in solution, thermodynamic cycles are generally combined with high level quantum chemical calculations in the gas phase and convenient calculation level for implicit solvation models3,4.

The use of thermodynamic cycles represents the traditional approach in theoretical pKa

* estimations, whereby solvation free energies (∆G solv) and gas phase deprotonation free

° energy (∆G gas) are combined so as to obtain the free energy of deprotonation in solution

* 5,6 (∆G soln) . The thermodynamic cycles are mainly employed due to the fact that solvation

* models are adjusted to deliver accurate solvation free energies (∆G solv), but the modest levels of theory at which they are parameterizedDraft are not accurate enough to estimate deprotonation

* 1 free energy in solution (∆G soln) . By applying a thermodynamic cycle, one can make use of highlevel ab initio calculations or experimental data in the gas phase to enhance the accuracy

* of the resulting free energy (∆G soln).

Over the past few years, numerous thermodynamic cycles have been proposed in the literature1,7. The most elementary one is called direct or absolute method, which is depicted in

Scheme 1. A number of alternatives to the direct method are widely employed in the literature; such as the proton exchange or isodesmic reaction scheme8, the hybrid cluster– continuum9 and the implicitexplicit models10. It should be noted that some recent proposed

11,12 approaches for pKa calculations avoid gasphase calculations and yield accurate results .

Solvation free energies of molecular and ionic species are generally calculated using continuum models. In these models, species are covered by a molecularshaped cavity embedded in a dielectric continuum13. The interaction between the charge distribution of the solute and the dielectric continuum provides the electrostatic contribution, which is the

https://mc06.manuscriptcentral.com/cjc-pubs Page 3 of 33 Canadian Journal of Chemistry

dominant term of solvation free energies for polar and charged solutes. Nowadays,

the most common implicit solvation models are the PCM family of models (eg. IEFPCM,

CPCM and IPCM1,1418, the Minnesota solvation models, SMx (e.g., x=D19 or x=620) and the

ConductorLike Screening Models of Klamt and coworkers (e.g., COSMO21 or COSMO

RS22). Such models can be applied to a wide variety of . They can estimate solvation

free energies for typical neutral solutes, in either aqueous or nonaqueous solvents, accurate to

within 1 kcal/mol. However, the errors incurred for ionic solutes are considerably larger and

can exceed 4 kcal/mol2325.

Solvation free energies as well as acidity data in organic solvents like dimethyl sulfoxide

(DMSO) are less abundant in comparison with those related to aqueous media2830. Such

information is the principal key for understanding many chemical and biochemical

processes3133, in particular the roleDraft that the specific and nonspecific solventsolute

interactions occupy in the binding between species and their target sites34. Moreover, most of

the compounds screened for biological activities are soluble in DMSO at room temperature35

37. Thus, DMSO is usually used for in vitro experimentation and in vivo administration of a

wide range of hydrophobic chemical species38.

In this paper, we studied theoretically the acidities of pyrrole (1), indole (2) and carbazole

(3) as well as their hydrocarbon analogues 1,3cyclopentadiene (4), indene (5) and

(6) in DMSO. As shown in Table 1, the acidities in DMSO of 1, 2 and 3 increase in that order,

while the acidities of 4, 5 and 6 are in the reverse order. These acidity rankings can be

attributed primarily to the presence of the rings and to the successive

increase/decrease in the aromaticities of conjugate bases39. In addition, the fact that intrinsic

acidity of acids NH is greater than that of acids CH, due to the greater

electronegativity of nitrogen than carbon, the relatively high acidities of 4 and 5 in

comparison with those of 1 and 2 are unanticipated. The main purpose of the present study is

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 4 of 33

to examine carefully some electronic and structural proprieties such as aromaticity indices, molecular electrostatic potential, dipole moment and Mulliken atomic charges, for each compound, in order to explain acidity ranking of hydrocarbon cycles as well as nitrogen containing heterocycles.

2. Methods

All calculations have been conducted using the GAUSSIAN 09 software package41. Three quantum chemistry methods, namely the B3LYP hybrid functional42,43, the M062X Minnesota functional44 and the CBSQB3 composite method4547, have been used to fully optimize structures either in gas phase or in solution. The 6311++G(d,p) basis set has been adopted for

DFT calculations. The continuum solvation models SMD19 and IEFPCM17 were used to perform solution calculations. The IEFPCMDraft model was applied in combination with the PAULING4849 and BONDI49 cavities to yield the IEFPCM/PAULING and IEF

PCM/BONDI solvation free energies respectively.

It is well known that CBSQB3 method is reputed by its expensive cost and moderate estimation of pKa values. However, such a method provides accurate estimates of gas phase thermochemistry data7,12. The purpose of its use in the present work is, on the one hand, the comparison of his gas phase deprotonation free energies with the values obtained through

DFT functionals, on the other hand, the examination of their estimated solvation free energies for each combination of cavity and continuum model.

Thermodynamic cycle of Scheme 1 shows the chemical reactions for the dissociation of

HA acid in the gas phase and in the solution phase. The directly estimated pKa values may be obtained via thermodynamic cycle through Eqs. 13.

∆∗ = () (1) ()

https://mc06.manuscriptcentral.com/cjc-pubs Page 5 of 33 Canadian Journal of Chemistry

∗ ° ∗ ∗ ∗ ∆() = ∆() + ∆()( ) + ∆()( ) − ∆()() + (2)

° ° ° ° ∆() = ()( ) + ()( ) − ()() (3)

Where ∆G°(gas) is the gas pas deprotonation free energy. The terms ∆G*(solv)(A ) and

∆G*(solv)(HA) correspond to the free energy change related to the solvation process of A and

HA species, respectively. Given that ∆G*(solv) corresponds to the free energy change related to

the solvation process (gas  solution), where the solvated compound is in its equilibrium

geometry in the respective phases1,7,24, all structures were optimized in the gas phase as well

as in solution. The stationary points were confirmed by frequency calculations for each level

of theory. The factor RTlnRT of Eq. 2 accounts for the correction needed to convert between

standard states, since gas phase free energies are calculated with respect to a standard state of

1atm and solution phase free energiesDraft correspond to a standard state of 1 mol/L. R and are

the gas constant in units of J/mol.K and L.atm/mol.K, respectively. In the absolute method, all

of the abovementioned free energies can be estimated using quantum chemistry methods,

+ + except G°(gas)(H ) and ∆G*(solv)(H ), due to the fact that the proton is a particular quantum

+ entity which contains no electrons. The calculation of G°(gas)(H ) using the standard equations

of thermodynamics and the SackurTetrode equation50 yields a value of 26.27 kJ/mol. While

+ solvation free energy of proton in DMSO, ∆G*(solv)(H ), has been determined either

experimentally or using theoretical computations, several values have been proposed, the

latest and more accepted value is 1114.62 kJ/mol, which is proposed by Knapp and

coworkers51.

Assuming that the molecular electrostatic potential, MEP, can indicate the change in the

electronic density accompanying protonation/deprotonation reactions, it can explain the

relative stability of molecules as well as their conjugate bases, which in turn can be related to

acidic strength. GaussView 5 software52 has been utilized to construct the MEP surfaces,

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 6 of 33

which are calculated for SMD/B3LYP/6311++G(d,p) optimized geometries. The red color of surfaces indicates the most electronrich site, while blue represents the most electronpoor site, the area of yellow/green color indicate an intermediate region.

In order to investigate the aromatic character for the considered systems, the Harmonic

Oscillator Model of Aromaticity (HOMA)53,54 and Shannon Aromaticity (SA) indices55 have been evaluated using Multiwfn software56. The input data of Multiwfn are related to optimized geometries obtained using Gaussian 09, according to SMD/B3LYP/6311++G(d,p) combination (see Tables S1 and S2 of Supporting information).

The HOMA index is a geometric indicator related to the degree of bondlength alternation, which can be calculated via the following formula:

= 1 − ∑ ( − ) (4) Draft

Where n is the number of bonds forming the aromatic cycle, Ri are the bond lengths in the actual geometry of the cycle obtained from quantum chemical calculations and Ropt is the mean bond length for which the energy of the cycle has an optimal value54, α is an empirical constant chosen to give HOMA = 0 for Kekulé benzene structure and 1 for system with all bonds equal to the optimal value Ropt. As normalized, HOMA value is close to 1 for aromatic cycles, close to 0 for nonaromatic systems and negative for antiaromatic ones.

The SA index is an electronicbased aromaticity descriptor, based on the electron delocalization derived from the AIM theory of Bader57. The SA index is defined as:

= ln() − ∑ . ln () (5)

Where n is the number of bonds in the cycle, and pi is the normalized electron density at the critical point of the ith bond. The socalled bond critical point BCP, as defined by Bader, corresponds to a minimum of the electron density in the normal direction of a bond which is a maximum in the two orthogonal directions58. The SA index for a cycle is a measure of the

https://mc06.manuscriptcentral.com/cjc-pubs Page 7 of 33 Canadian Journal of Chemistry

localization of electron density, thus less SA index value indicates that the cycle is more

aromatic and vice versa.

3. Results and Discussion

We will discuss our results in the following order, first the solvation free energies and pKa

values will be presented, then the electronic properties (acidic site charge, dipole moment and

Molecular Electrostatic Potential MEP) will be discussed, and finally the aromaticity indices

(Harmonic Oscillator Model of Aromaticity index HOMA and Shannon Aromaticity index

SA) will be examined.

Solvation free energies of the Ncycles 1, 2 and 3, as well as their conjugate bases 1H, 2

H and 3H are presented in Table 2. For each system, computations were carried out using SMD, IEFPCM/PAULING and IEFPCM/BONDIDraft solvation models in combination with the B3LYP/6311++G(d,p), M062X/6311++G(d,p) levels of theory and CBSQB3 composite

method. As it can be seen in Table 2 IEFPCM/PAULING and IEFPCM/BONDI models

provide close values for neutral molecules, while SMD values are more negative. In addition,

the examination of Fig. 1 reveals that values obtained with SMD decrease from 1 to 3,

whereas, those calculated with IEFPCM increase in that order. Values obtained with all three

models for conjugate bases increase from 1 to 3, but the increase in magnitude is more

important for IEFPCM in comparison with SMD model. Seeing the solvation free energies of

1H and 2H, values obtained with SMD are slightly greater than those related to IEFPCM

models. However, for 3H anion SMD and IEFPCM/BONDI models give large negative

values of solvation free energies. Regarding the quantum chemistry methods, for all the

systems, M062X functional provides the lowest values, while data obtained with B3LYP and

CBSQB3 are nearly of the same magnitude.

Table 3 collects the pKa values of Ncycles 1, 2 and 3, estimated using absolute method

based on thermodynamic cycle described in Methods section. Before examining pKa values, it

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 8 of 33

is first worth noting that we have used absolute method due to its simplicity and low computational cost. It should be noted that Coote and coworkers1 suggested that the acceptable error margin for a directly calculated pKa values, in water solution, should be around 3.5 pKa units. When using the same definition of acceptable error margin, and based on Pliego and Riveros estimation of associated errors for solvation free energies in DMSO28, the acceptable error margin, in that solvent, becomes 4.25 pKa units.

As can be seen, all the calculated pKa values decrease from 1 to 3, so they mimic the general trends of experimental data given in Table 139,40. However, it should be pointed out that the pKa values calculated using the combinations SMD/M062X and IEF

PCM/BONDI/B3LYP are very accurate, but those evaluated using SMD/B3LYP, IEF PCM/PAULING/M062X and IEFPCM/BONDI/M062XDraft combinations are less precise. The mean absolute deviation (MAD), together with the maximum absolute deviation

(ADmax), related to pKa values are provided in table S8 of supporting information and depicted in Fig. 2. According to Table 3 and Fig. 2 the highest MAD is 2.36 pKa units, which is related to the combination IEFPCM/PAULING/M062X. The best agreement with experimental pKa values is provided by the combinations SMD/M062X, IEFPCM/BONDI/ B3LYP and IEF

PCM/PAULING/B3LYP, for which the MAD/ADmax are 0.26/0.47, 0.32/0.64 and 0.62/1.03 pKa units respectively. The composite method CBSQB3 provides modest deviations for all three solvation models, the MAD/ADmax of 0.88/1.29, 0.97/1.75 and 1.23/1.63 pKa units are obtained for IEFPCM/BONDI, IEFPCM/PAULING and SMD model, respectively.

The computed solvation free energies of Ccycles 4, 5 and 6 as well as their conjugate bases are presented in Table 4. Looking at collected data, we can observe that values estimated for neutral species using IEFPCM/PAULING model are positive, this is probably due to the nature of cavity created using Pauling radii59. Examining the sign and the

https://mc06.manuscriptcentral.com/cjc-pubs Page 9 of 33 Canadian Journal of Chemistry

magnitude of the overall solvation data, we see that SMD model gives the lowest values, as

indicated by Fig. 3. However for all models there are proportionalities between solvation free

* * energies of neutral and anionic species, i.e. the differences (∆G (solv)(A ) ∆G (solv)(HA)) are

similar, regardless of the type of combination model/cavity/method utilized. The analysis of

data related to quantum chemistry methods reveals that, like Ncycles results, M062X always

provides the most negative values, while B3LYP and CBSQB3 give nearly similar values.

The pKa values computed for Ccycles 4, 5 and 6 are collected in Table 5. Their

MADs/ADmax are represented in Fig.4 and reported in table S8 of supporting information. As

shown, it is possible to state that, like Ncycles, all calculated pKas mimic the experimental

ones; they increase from 4 to 6. Moreover, the obtained results seem to be almost insensitive

to the utilized solvation model and cavity type, but they depend directly on the calculation Draft method. The B3LYP functional and CBSQB3 composite method give almost the same pKa

values, which are slightly greater than those obtained with M062X functional. Such a result is

due partially to the fact that M062X underestimate the gas phase deprotonation free energies

° (∆G gas), as it can be seen in the table S3 of supporting information. On the other hand, the

evaluated solvation free energies related to M062X are also lower to those related to B3LYP

and CBSQB3. These differences can be attributed to the nature and the parameterization of

Minnesota functional.

The M062X functional combined to SMD model provides the lowest pKa error, the

MAD/ADmax are 1.43/2.68 pKa units, respectively. Such a result can be simply explained by

the fact that SMD model is parameterized using M062X functional to provide accurate

solvation free energies. The combination SMD/CBSQB3 provides the largest pKa error; the

MAD and ADmax are 2.20 and 4.01 pKa units, respectively. These deviations are due to the

evaluated solvation free energies, since the gas phase free energies calculated with CBSQB3

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 10 of 33

are very accurate. As can be seen, despite the magnitude of such MADs, they are still less

1,28 than the acceptable error margin for a directly calculated pKa values .

According to figures 2 and 4, one can notice that unlike Ncycles, the obtained deviations for Ccycles seem to be less sensitive to the type of the combination of solvation model and calculation method, this fact can be explained by examining the solvation free energies and gas phase deprotonation free energies. Solvation free energies calculated with B3LYP and

CBSQB3 for all systems are very close and they are slightly greater than those calculated with M062X functional. However, the calculated gas phase deprotonation free energies of N cycles seem to be more accurate than those related to Ccycles, the errors related to B3LYP and CBSQB3 for Ncycles are around 3 kJ/mol, for Ccycles the error reach 9 kJ/mol for cyclopentadiene. The deviations associatedDraft to the results obtained with M062X are more important, especially for Ccycles. For example, the error related to the deprotonation of cyclopentadiene is 20.77 kJ/mol. The examination of gas phase deprotonation free energies shows that the good estimation of pKa values is due to error cancellation rather than the precise evaluation of gas phase and solution phase free energies.

Figures 5 and 6 show MEP surfaces and dipole moments of studied cycles as well as their corresponding conjugate bases respectively. A deeper look at Fig. 5 reveals that electronpoor regions are located on acidic sites for all molecular systems, which promote the interactions between oxygen atoms of solvent molecules and acidic proton of solutes. Mulliken atomic charges of acidic sites seem to have no relation with acidity strength (see figures S1 and S2 in the Supporting Information), while protons of Ncycles are more charged than those of C cycles. Regarding the dipole moments, it can be seen that their orientations are the same for systems 1 and 4, 2 and 5, as well as 3 and 6. Such orientation is due primarily to geometric similarities and to the presence or absence of benzene rings. The magnitudes of dipole moments are more important for Ncycles due to the presence of nitrogen atoms.

https://mc06.manuscriptcentral.com/cjc-pubs Page 11 of 33 Canadian Journal of Chemistry

Closer inspection of Fig. 6 shows that all conjugate bases have an electronrich region

located inside the cycles, which is due to aromatic effect. Furthermore, Ncycles possess a

narrow region rich in electron density located on nitrogen atoms, leading to a selective

interaction between proton and conjugate bases, but this property is not observed for Ccycles,

and consequently their interaction with protons is not selective. Moreover, 4H has a dipole

moment equal to zero, such a structure is unfavorable to proton recombination and this can

probably explain the relatively high acidity of 4. As can be seen, charges located on nitrogen

atoms are directly related to the acidic strength, thus nitrogen charges decrease in the

following order: 3 (0.283) < 2 (0.350) < 1 (0.435) as pKa (3) < pKa (2) < pKa (1). The

explanation of such relation can be attributed to the nature of NH bound formed when

conjugate base is combined with proton; hence the important nitrogen charge tends to form

highly ionic bond and therefore less stableDraft acid molecule in solution. In the same way, acid 6

is weaker than 4 due to the charges of their basic sites, (0.231 and 0.335 respectively).

However, acid 5 has two similar basic sites (0.662), which are directly linked to benzene

ring, such atomic charges explain probably the intermediate acidity of 5 in comparison with 4

and 6.

Thermodynamically speaking, the strength of an acid is directly related to the relative

stability of its neutral and ionic forms. Regarding unsaturated cyclic compounds, like those

we are interested in, the aromaticity can cause a specific stabilization either to the acid or to

its conjugate base. Even though aromaticity is not the only parameter governing the acidity of

studied cycles39, it is important to acknowledge its impact on acid strength. We explore herein

the aromaticity of cycles as well as their conjugate bases using HOMA and SA aromaticity

indices.

The calculated HOMA and SA indices are depicted in Figures 7 and 8 respectively. As

expected, Fig. 7 indicates that the aromaticities of 1, 2 and 3 as well as their conjugate bases

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 12 of 33

are very close, due to the presence of nitrogen pairs of electron. Moreover, the HOMA values decrease slightly from 1H to 3H (0.88682, 0.88516 and 0.87405 respectively), indicating that the aromaticity and, consequently, the relative stability of conjugate bases decreases in that order. This explains the acidity trends observed for Ncycles. On the other hand, HOMA values related to 4 and 5 are negative (0.84103 and 0.0799, respectively). That of cyclopentadiene is very negative indicating that it is clearly an antaromatic compound. Thus, it is instable in the neutral form compared to its ionic form. The HOMA value of the indene is close to zero (0.0799), so it is almost a nonaromatic compound. While, HOMA index of fluorene is small and positive (0.23187) indicating that it has a low aromatic character. When exploring the HOMA indices of conjugate bases 4H, 5H and 6H, it can be observed that all values are large and positive (0.78513, 0.83384 and 0.8299, respectively). Therefore, after deprotonation all Ccycles become aromaticDraft systems. Although their aromaticities are nearly similar, the difference between HOMA indices of molecules and their conjugate bases decreases from 4 to 6 (HOMA(iH)HOMA(i) = 1.63 ; 0.91 ; 0.60 for i = 4, 5 and 6 respectively.). Such results explain clearly the acid strength of Ccycles. It is worth noting that all calculated HOMA values of Ncycles are greater than those of Ccycles, revealing that N cycles are more aromatic than Ccycles thanks to nitrogen atom role.

Concerning the SA indices, it is well known that the smaller the SA index, the more aromatic is the system. The scale proposed for these indices indicates that molecule are aromatic when their SA are less than 0.003, they are non aromatic when SA are between

0.003 and 0.005, and antiaromatic for SA greater than 0.00555. The calculated SA are plotted in Fig. 8. These values are unexpected for several considerations. For instance, the obtained values for 3 and 3H compounds are very large (0.059490 and 0.06215, respectively), indicating that they are antiaromatic compounds, which is unlikely. In the same way, the low value obtained for cyclopentadiene indicated that it is almost an aromatic compound. Such

https://mc06.manuscriptcentral.com/cjc-pubs Page 13 of 33 Canadian Journal of Chemistry

obtained values are questionable; it leads us to say that this approach is less robust than the

previous one in term of prediction of aromatic character.

4. Conclusion

In this study, the direct thermodynamic method has been applied to calculate pKa values of

unsaturated N and Ccycles, using B3LYP, M062X and CBSQB3 quantum chemistry

methods in combination with IEFPCM and SMD continuum models. The obtained data are

generally in good agreement with experiment. The M062X functional combined with the

SMD model provides the lowest pKa errors. The CBSQB3 composite method and the B3LYP

functional provide close values for all solvation models. MEP surfaces of molecular systems

reveal that electronpoor regions are located on acidic sites, promoting the interactions between oxygen atoms of solvent moleculesDraft and acidic proton of solutes. However, conjugate bases have an electronrich region located essentially inside the cycles, due to the aromatic

effect. Calculated HOMA aromaticity indices explain clearly that the acid strength is directly

related to the aromaticities of compounds, particularly those of conjugate bases.

Acknowledgements

This work has been supported by the ministry of higher education and scientific research,

under the CNEPRU projects (Approval No. E03620140002).

References

(1) Ho, J. and Coote, M.L. Theor. Chem. Acc. 2010, 125, 3.

(2) Kelly, C.P., Cramer, C.J. and Truhlar, D.G. J. Phys. Chem. B. 2006, 110, 16066.

(3) Ho, J. and Coote, M.L. Wires. Comput. Mol. Sci. 2011, 1, 649.

(4) Ho, J. and Ertem, M.Z. J. Phys. Chem. B. 2016, 120, 1319.

(5) Alongi, K.S. and Shields, G.C. Annu. Rep. Comput. Chem. 2010, 6, 113.

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 14 of 33

(6) Arslan, T. and Aygan, Ö.Ö.T. Comput. Theor. Chem. 2014, 1032, 65.

(7) Ho, J. Aust. J. Chem. 2014, 67, 1441.

(8) Casasnovas, R., Ortega‐Castro, J., Frau, J., Donoso, J. and Munoz, F. Int. J. Quantum

Chem. 2014, 114, 1350.

(9) Pliego, J.R. and Riveros, J.M. J. Phys. Chem. A. 2001, 105, 7241.

(10) Kelly, C.P., Cramer, C.J. and Truhlar, D.G. J. Phys. Chem. A. 2006, 110, 2493.

(11) Ho, J. Phys. Chem. Chem. Phys. 2015, 17, 2859.

(12) Casasnovas, R., Fernández, D., OrtegaCastro, J., Frau, J., Donoso, J. and Munoz,

F. Theor. Chem. Acc. 2011, 130, 1.

(13) Mennucci, B. and Cammi, R. (eds.), Continuum solvation models in chemical physics: from theory to applications., Chichester, UK, John Wiley & Sons, 2007.

(14) Mennucci, B. and Tomasi, J. J. Chem.Draft Phys. 1997, 106, 5151.

(15) Tomasi, J., Mennucci, B. and Cammi, R. Chem. Rev. 2005, 105, 2999.

(16) Cossi, M., Rega, N., Scalmani, G. and Barone, V. J. Comput. Chem. 2003, 24, 669.

(17) Tomasi, J., Mennucci, B. and Cances, E. J. Mol. Struct. 1999, 464, 211.

(18) Mennucci, B., Cances, E. and Tomasi, J. J. Phys. Chem. B. 1997, 101, 10506.

(19) Marenich, A.V., Cramer, C.J. and Truhlar, D.G. J. Phys. Chem. B. 2009, 113, 6378.

(20) Kelly, C.P., Cramer, C.J. and Truhlar, D.G. J. Chem. Theory Comput. 2005, 1, 1133.

(21) Klamt, A. and Schüürmann, G.J.G.J. Perkin Trans. 1993, 2, 799.

(22) Klamt, A., Jonas, V., Bürger, T. and Lohrenz, J.C. J. Phys. Chem. A. 1998, 102, 5074.

(23) Marenich, A.V., Olson, R.M., Kelly, C.P., Cramer, C.J. and Truhlar, D.G. J. Chem.

Theory Comput. 2007, 3, 2011.

(24) Takano, Y. and Houk, K.N. J. Chem. Theory. Comput. 2005, 1, 70.

(25) Klamt, A., Mennucci, B., Tomasi, J., Barone, V., Curutchet, C., Orozco, M. and Luque,

F.J. Acc. Chem. Res. 2009, 42, 489.

https://mc06.manuscriptcentral.com/cjc-pubs Page 15 of 33 Canadian Journal of Chemistry

(26) Cramer, C.J. and Truhlar, D.G. Acc. Chem. Res. 2008, 41, 760.

(27) GhalamiChoobar, B. and GhiamiShomami, A. Comput. Theor. Chem. 2013, 1018, 66.

(28) Pliego, J.R. and Riveros, J.M. Phys. Chem. Chem. Phys. 2002, 4, 1622.

(29) Topphoff, M., Gruber, D. and Gmehling, J. J. Chem. Eng. Data. 1999, 44, 1355.

(30) Lai, J.T., Lau, F.W., Robb, D., Westh, P., Nielsen, G., Trandum, C., Hvidt, A. and Koga,

Y. J. Solution. Chem. 1995, 24, 89.

(31) Farajtabar, A. and Gharib, F. Monatsh. Chem. 2010, 141, 381.

(32) Catalan, J., Diaz, C. and GarciaBlanco, F. J. Org. Chem. 2001, 66, 5846.

(33) Luzar, A. and Stefan, J. J. Mol. Liq. 1990, 46, 221.

(34) Wang, H. and BenNaim, A. J. Med. Chem. 1996, 39, 1531.

(35) Alvarado, Y.J., ÁlvarezMon, M., Baricelli, J., CalderaLuzardo, J., Cubillán, N., Ferrer

Amado, G., Hassanhi, M., MarreroPonce,Draft Y., Mancilla, V., Rocafull, M.A. and San Antonio

Sánchez, M.E. J. Solution. Chem. 2010, 39, 1099.

(36) Rollas, S. and Küçükgüzel, S.G. Molecules. 2007, 12, 1910.

(37) Morgan, L.R., Thangaraj, K., LeBlanc, B., Rodgers, A., Wolford, L.T., Hooper, C.L.,

Fan, D. and Jursic, B.S. J. Med. Chem. 2003, 46, 4552.

(38) Santos, N.C., FigueiraCoelho, J., MartinsSilva, J. and Saldanha, C. Biochem.

Pharmacol. 2003, 65, 1035.

(39) Bordwell, F.G., Drucker, G.E. and Fried, H.E. J. Org. Chem. 1981, 46, 632.

(40) Bordwell, F.G. and Drucker, G.E. J. Org. Chem. 1980, 45, 3325.

(41) 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.; Hada, 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.;

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 16 of 33

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, N. J.; 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, Ö.; Foresman, J. B.; Ortiz, J. V.;

Cioslowski, J.; Fox, D. J. Gaussian 09, Revision C.01; Gaussian, Inc., Wallingford, CT, 2009.

(42) Lee, C., Yang, W. and Parr, R.G. Phys. Rev. B. 1988, 37, 785.

(43) Becke, A.D. J. Chem. Phys. 1993, 98, 5648.

(44) Zhao, Y. and Truhlar, D.G. Theor. Chem. Acc. 2008, 120, 215.

(45) Ochterski, J.W., Petersson, G.A. and Montgomery Jr, J.A. J. Chem. Phys. 1999, 110,

2822. Draft

(46) Montgomery Jr, J.A., Frisch, M.J., Ochterski, J.W. and Petersson, G.A. J. Chem. Phys.

1999, 110, 2822.

(47) Montgomery Jr, J.A., Frisch, M.J., Ochterski, J.W. and Petersson, G.A. J. Chem. Phys.

2000, 112, 6532.

(48) Pauling, L.; The nature of the chemical bond and the structure of molecules and crystals: an introduction to modern structural chemistry; 3rd ed.; Cornell university press, 1960.

(49) Bondi, A. J. Chem. Phys. 1964, 68, 441.

(50) McQuarrie, D.A.; Statistical Mechanics; Harper and Row, New York, 1976.

(51) Rossini, E., Netz, R.R. and Knapp, E.W. J. Chem. Theory. Comput 2016, 12, 3360.

(52) Dennington, R., Keith, T. and Millam, J. GaussView, Version 5, Semichem, Inc.

Shawnee Mission, KS, 2009.

(53) Kruszewski, J. and Krygowski, T.M. Tetrahedron. Lett. 1972, 13, 3839.

(54) Krygowski, T.M. J. Chem. Inf. Comput. Sci. 1993, 33, 70.

https://mc06.manuscriptcentral.com/cjc-pubs Page 17 of 33 Canadian Journal of Chemistry

(55) Noorizadeh, S. and Shakerzadeh, E. Phys. Chem. Chem. Phys. 2010, 12, 4742.

(56) Lu, T. and Chen, F. J. Comput. Chem. 2012, 33, 580.

(57) Bader, R. F.; Atoms in Molecules. International Series of Monographs in Chemistry;

Oxford; Oxford University Press, 1990.

(58) Alkorta, I., Barrios, L., Rozas, I. and Elguero, J. J. Mol. Struct. THEOCHEM. 2000, 496,

131.

(59) Lee, T.B. and McKee, M.L. Phys. Chem. Chem. Phys. 2011, 13, 10258.

Draft

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 18 of 33

Table 1 Abbreviations, names, structures, and pKa values in DMSO of studied cycles.

39,40 N° Name Structure pKa

H N 1 Pyrrole 23.05

H N 2 Indole 20.90

H N 3 Carbazole

19.90

H2 C 4 1,3Cyclopentadiene

18.00

H2 C 5 Indene Draft 20.10

H2 C 6 Fluorene 22.60

https://mc06.manuscriptcentral.com/cjc-pubs Page 19 of 33 Canadian Journal of Chemistry

Table 2 Solvation free energies of 1, 2 and 3 as well as their conjugate bases 1H, 2H and 3

H (kJ/mol).

Models Methods ∆G*solv

1 1H 2 2H 3 3H

SMD B3LYP 20.9 245.2 23.3 218.0 25.9 203.6

M062X 24.1 251.5 27.9 226.3 31.2 213.6

CBSQB3 20.2 246.4 22.5 219.4 24.9 205.6

IEFPCM/ B3LYP 16.7 255.5 14.6 221.3 11.2 198.6

PAULING M062X 19.6 262.9 18.5 230.3 15.5 209.3

CBSQB3 15.9 256.9 13.5 222.6 9.6 200.3 IEFPCM/ B3LYP 17.8Draft 254.4 17.8 224.0 14.5 204.4 BONDI M062X 20.5 260.8 21.7 232.2 19.1 214.4

CBSQB3 17.1 255.5 17.1 225.4 13.7 206.4

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 20 of 33

Table 3 Predicted pKa values of 1, 2 and 3.

Models Methods pKa

1 2 3

B3LYP 24.60 22.97 21.53

SMD M062X 22.76 20.86 20.37

CBSQB3 23.94 22.08 21.53

IEFPCM/ B3LYP 22.02 20.82 20.67

PAULING M062X 19.93 18.47 18.35

CBSQB3 21.30 19.90 19.73

IEFPCM/ B3LYP 22.41 20.91 20.23 BONDI DraftM062X 20.47 18.70 18.08 CBSQB3 21.76 20.05 19.38

https://mc06.manuscriptcentral.com/cjc-pubs Page 21 of 33 Canadian Journal of Chemistry

Table 4 Solvation free energies of 4, 5 and 6 as well as their conjugate bases 4H, 5H and 6

H (kJ/mol).

Models Methods ∆G*solv

4 4H 5 5H 6 6H

SMD B3LYP 5.8 239.2 8.7 210.2 13.1 193.1

M062X 7.7 248.0 12.2 220.5 18.0 205.1

CBSQB3 4.7 240.6 7.6 211.9 11.6 194.8

IEFPCM/ B3LYP 2.4 239.5 3.2 203.0 3.4 178.2

PAULING M062X 0.7 246.5 0.5 213.8 0.0 190.4

CBSQB3 3.3 241.6 4.4 205.0 5.4 179.8

IEFPCM/ B3LYP 1.2 239.1 3.0 207.1 1.6 186.0

BONDI M062X Draft 2.8 247.3 6.0 216.7 5.6 197.2

CBSQB3 0.3 240.4 2.1 208.7 0.3 187.6

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 22 of 33

Table 5 Predicted pKa values of 4, 5and 6.

Models Methods pKa

4 5 6

B3LYP 18.94 22.85 25.17

SMD M062X 15.32 19.74 23.861

CBSQB3 18.21 22.50 26.61

IEFPCM/ B3LYP 17.45 22.02 26.17

PAULING M062X 14.11 18.69 23.28

CBSQB3 16.63 21.60 26.26

IEFPCM/ B3LYP 18.15 22.39 25.68

BONDI M062X 14.58 19.32 23.07

CBSDraftQB3 17.47 22.10 25.89

https://mc06.manuscriptcentral.com/cjc-pubs Page 23 of 33 Canadian Journal of Chemistry

Figure captions

Scheme 1. The thermodynamic cycle employed for calculating pKa values.

Fig.1 Solvation free energies of Ncycles. Notations refer to: S:SMD, IP:IEFPCM

PAULING, IB: IEFPCM/BONDI.

Fig.2 MAD and ADmax for pKa of Ncycles, notations refer to: S:SMD, IP:IEFPCM

PAULING, IB: IEFPCM/BONDI, B:B3LYP, M:M062X, C:CBSQB3.

Fig.3 Solvation free energies of Ccycles. Notations refer to: S:SMD, IP:IEFPCM

PAULING, IB: IEFPCM/BONDI.

Fig.4 MAD and ADmax for pKa of Ccycles, notations refer to: S:SMD, IP:IEFPCM

PAULING, IB: IEFPCM/BONDI, B:B3LYP, M:M062X, C:CBSQB3. Fig.5 Molecular electrostatic potentialDraft maps MEPs of systems 1 to 6. Red color indicates most electronrich site and blue color most electronpoor site. Isovalue = 0.002 a.u and the

maximum values of electrostatic potential (Vmax) are: 1(0.071), 2(0.073), 3(0.073), 4(0.022),

5(0.033), 6(0.0316).

Fig.6 Molecular electrostatic potential maps MEPs of systems 1H to 6H. Red color indicates

most electronrich site and yellow color most electronpoor site. Isovalue = 0.002 a.u and the

maximum values of electrostatic potential (Vmax) are: 1H(0.214), 2H(0.197), 3H(0.190), 4

H(0.220), 5H(0.202), 6H(0.188).

Fig.7 Obtained HOMA indices for acids and their conjugate bases (HOMA(H)).

Fig.8 Obtained SA indices for acids and their conjugate bases (SA(H)).

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 24 of 33

Draft

https://mc06.manuscriptcentral.com/cjc-pubs Page 25 of 33 Canadian Journal of Chemistry

Draft

Fig.1 Solvation free energies of N-cycles. Notations refer to: S:SMD, IP:IEF-PCM-PAULING, IB: IEF- PCM/BONDI.

291x203mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 26 of 33

Draft

Fig.2 MAD and ADmax for pKa of N-cycles, notations refer to: S:SMD, IP:IEF-PCM-PAULING, IB: IEF- PCM/BONDI, B:B3LYP, M:M062X, C:CBS-QB3.

296x209mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Page 27 of 33 Canadian Journal of Chemistry

Draft

Fig.3 Solvation free energies of C-cycles. Notations refer to: S:SMD, IP:IEF-PCM-PAULING, IB: IEF- PCM/BONDI.

291x203mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 28 of 33

Draft

Fig.4 MAD and ADmax for pKa of C-cycles, notations refer to: S:SMD, IP:IEF-PCM-PAULING, IB: IEF- PCM/BONDI, B:B3LYP, M:M062X, C:CBS-QB3.

291x203mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Page 29 of 33 Canadian Journal of Chemistry

Vmin Vmax

Draft

1 2 3

4 5 6

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 30 of 33

Vmin Vmax

Draft

1-H 2-H 3-H

4-H 5-H 6-H

https://mc06.manuscriptcentral.com/cjc-pubs Page 31 of 33 Canadian Journal of Chemistry

Draft

Fig.7 Obtained HOMA indices for acids and their conjugate bases (HOMA(-H)).

291x203mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry Page 32 of 33

Draft

Fig.8 Obtained SA indices for acids and their conjugate bases (SA(-H)).

291x203mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/cjc-pubs Page 33 of 33 Canadian Journal of Chemistry

1,5

1,0

0,5

0,0 Draft 1 2 3 4 5 6

HOMA -0,5 Acids HOMA(-H)

-1,0 Harmonic OscillatorModel of Aromaticity -1,5

https://mc06.manuscriptcentral.com/cjc-pubs