Home Physical Sciences Electrophilic substitution reactions of thiophene and thieno[2,3-b]thiophene heterocycles: a DFT study
Article Publicly Available

Electrophilic substitution reactions of thiophene and thieno[2,3-b]thiophene heterocycles: a DFT study

  • Nandini Savoo ORCID logo , Chahlaa Mungur , Lydia Rhyman , Ponnadurai Ramasami EMAIL logo and John A. Joule
Published/Copyright: June 2, 2023

Abstract

We determined the most preferred site for the electrophilic aromatic substitution (SEAr) reactions of thiophene 1 and thieno[2,3-b]thiophene 2 using the N-chlorosuccinimide electrophile in the gas phase and in acetic acid. The B3LYP/6-311G(d), B3LYP-D3/6-311G(d) and M06-2X/6-311G(d) methods were employed to investigate the SEAr reaction mechanisms. We found that, compared to the β-carbon atom, the α-carbon atom in both 1 and 2 is preferred for electrophilic attack both kinetically and thermodynamically.

Introduction

Electrophilic aromatic substitution (SEAr) reactions (Scheme 1) occur when an aromatic ring reacts with an electrophile to generate a product in which one hydrogen atom is replaced by the electrophile [1, 2]. The SEAr reaction often occurs in the presence of a catalyst [1, 37]. Aromatic compounds produced by SEAr reactions are employed in agriculture, pharmacy and industry [6, 812].

Scheme 1: 
SEAr reaction mechanism (a) involving a benzene ring and a halogen; (b) at the α- and β-carbon atoms of a five-membered heterocycle and a halogen; (c) at the α- and β-carbon atoms of a 5:5 [2,3-b]-bicyclic heterocycle and a halogen. X and Y are heteroatoms.
Scheme 1:

SEAr reaction mechanism (a) involving a benzene ring and a halogen; (b) at the α- and β-carbon atoms of a five-membered heterocycle and a halogen; (c) at the α- and β-carbon atoms of a 5:5 [2,3-b]-bicyclic heterocycle and a halogen. X and Y are heteroatoms.

The initial step of the SEAr reactions of benzene involves an interaction between the benzene ring and the electrophile (E+) to form a π-complex [π-complex 1 in Scheme 1(a)] [1, 13]. The π-complex 1 reacts to generate a cationic intermediate, also known as the Wheland intermediate, which is an arenium ion; this is the rate-determining step. This cationic intermediate is called a σ-complex because of the formation of a new σ bond with benzene. The σ-complex is stabilised by resonance. Since the benzene ring loses its aromaticity, the σ-complex is less stable than the benzene ring [1]. Deprotonation takes place in the last step in order to restore the aromaticity of the benzene ring [14]. This happens through an attack of the counteranion of the E+ electrophile on the hydrogen atom of the ring [Scheme 1(a)] [1, 14].

The SEAr reactions involving benzene derivatives are well-established. They are influenced by the electron-donating and electron-withdrawing groups attached to the aromatic ring [15]. The electron-donating and electron-withdrawing groups enhance and inhibit the SEAr reaction by activating and deactivating the aromatic ring, respectively [2].

Other aromatic rings, such as five-membered, benzo-fused and 5:5 bicyclic heterocycles, also undergo SEAr reactions. These heterocyclic compounds are important moieties and precursors which are useful mainly in drug discovery and material science [16], [17], [18], [19], [20], [21]. Five-membered heterocycles are described as π excessive [22]. The lone pair of electrons is delocalised; this completes the sextet of electrons which is required for aromaticity [22].

One of the simplest aromatic heterocycles has a five-membered ring with one heteroatom. The two carbon atoms where electrophilic substitution can take place are the α- and β-carbon atoms with respect to the heteroatom [Scheme 1(b)]. There are reports of high reactivities of the α-carbon atom of five-membered heterocycles, for example thiophene, pyrrole and furan rings [10, 17]. The synthesis of β-substituted five-membered heterocycle derivatives requires increasing the electron-withdrawing ability of a substituent at the α-carbon atom through the use of Lewis acid and protic acids [10, 17]. Belen’kii reported that the favourable site of electrophilic attack may be justified in terms of resonance stabilisation of the Lewis structure of the intermediate by the delocalisation of the positive charge [17]. There is greater resonance stabilisation of the σ-complex when substitution occurs at the α-carbon atom, as illustrated in Scheme 1(b). The trends in the SEAr reactions of benzo-fused five-membered rings, such as indole, are predicted in a similar manner using resonance structures; this is shown in Scheme S1 in the Supplementary Information (SI) document.

Determining the favourable site for SEAr reaction in 5:5 bicyclic systems using resonance structures is intriguing and complex. Scheme 1(c) shows that the stabilisation of the positive charge involves the disruption of the aromaticity of the adjacent five-membered ring for the α-attack but not for the β-attack. Hence, an electrophilic attack at the β-carbon atom is predicted through the resonance structures. However, experiments performed with 5:5 bicyclic heterocycles reported SEAr reactions at the α- and α′-carbon atoms, instead of at the β-carbon atom. Bugge et al. carried out the bromination of thieno[2,3-b]thiophene and thieno[3,2-b]thiophene by using N-bromosuccinimide in glacial acetic acid and obtained the α-product [23]. N-Chlorosuccinimide (NCS) was reacted with thieno[2,3-b]thiophene in acetic acid to form the α-chlorothieno[2,3-b]thiophene and α,α′-dichlorothieno[2,3-b]thiophene products [24].

Such experimental results imply that resonance structures cannot be employed to determine the site which is the most prone to SEAr reactions in the case of 5:5 bicyclic heterocycles.

In order to gain further insights into the preference of either α- or β-substitution in 5:5 bicyclic heterocycles, we performed computational mechanistic studies using the density functional theory (DFT) method. We investigated the mechanisms of the SEAr reactions of a five-membered ring and a 5:5 bicyclic heterocycle through an analysis of the thermodynamic and kinetic parameters in the gas phase and in solution. NCS was used as the electrophile to determine the most preferred site in thiophene 1 and thieno[2,3-b]thiophene 2 for SEAr reaction. NCS was selected as it was reported that N-halosuccinimide compounds are an effective source of electrophile for the direct halogenation of activated and deactivated aromatic systems [25], [26], [27], [28], [29]. Acetic acid was used as the solvent to mimic the experimental conditions used in previous reports []. The reactions investigated in this manuscript are shown in Scheme 2.

Scheme 2: 
SEAr reactions studied involving (a) thiophene 1 and (b) thieno[2,3-b]thiophene 2; (c) Geometry of NCS.
Scheme 2:

SEAr reactions studied involving (a) thiophene 1 and (b) thieno[2,3-b]thiophene 2; (c) Geometry of NCS.

Methodology

The chlorination reactions of 1 and 2 were investigated with NCS using the B3LYP/6-311G(d) method. The B3LYP [33], [34], [35] functional was employed in several recent reports to study properties and mechanisms of heterocycles [36], [37], [38], [39], [40]. Full geometry optimisations were performed in the gas phase using the Gaussian 16 package [41] available in SEAGrid [42], [43], [44], [45]. The transition state (TS) was identified through vibrational frequency [46] and intrinsic reaction coordinate (IRC) [47] analyses. Ground state structures were determined through the absence of imaginary frequency. The relative electronic energies (ΔE) reported include the zero-point energy (ZPE) correction. The B3LYP/6-311G(d) method was employed based on the polarisable continuum model (PCM) [48] to perform bulk solvation with acetic acid. The GaussView 6.0 [49], Chemcraft 1.8 [50] and CYLview [51] software were used for the visualisation of molecules and reactions.

The computations were repeated by considering empirical dispersion (B3LYP-D3) [52] and by using the M06-2X [53] functional due to hydrogen bonding interactions. The trends observed were in agreement with the B3LYP/6-311G(d) method. Hence, only results obtained with the B3LYP/6-311G(d) method are reported in this paper; the results obtained with the B3LYP-D3/6-311G(d) and M06-2X/6-311G(d) methods are reported in the Supplementary Information.

Results and discussion

SEAr reactions of thiophene

The SEAr reaction of 1 may occur at the α- and β-carbon atoms. Unlike SEAr of benzene [1, 13], we found that no cationic intermediate was formed in the reaction of 1. Two pathways were located for the α-SEAr reaction, namely pathways A and B. Only one pathway was located for the β-SEAr reaction (pathway C). The α- and β-SEAr reaction pathways are initiated with a reactant complex (RC) which is formed by the interaction of 1 with NCS. This leads to a TS which forms a product complex (PC). The PC separates into the products (P). Figure 1 shows the potential energy surfaces (PESs) and the geometry of the TSs obtained for the α- and β-SEAr reaction pathways of 1 with NCS. Table 1 lists the relative energy in terms of ΔE, ΔH and ΔG values of the RCs, TSs, PCs and Ps of the α- and β-SEAr reaction pathways with respect to the separate reactants.

Fig. 1: 
PESs of the 1 + NCS α- and β-SEAr reactions. ΔE values are in kcal mol−1. Bond distances are in Å. All values resulting from the solvent-phase computations are in parentheses.
Fig. 1:

PESs of the 1 + NCS α- and β-SEAr reactions. ΔE values are in kcal mol−1. Bond distances are in Å. All values resulting from the solvent-phase computations are in parentheses.

Table 1:

ΔE, ΔH and ΔG, in kcal mol−1, of the RCs, TSs, PCs and Ps of the α- and β-SEAr reactions of 1 with NCS. The solvent-phase energy values are in parentheses.

ΔE ΔH ΔG
Pathway A (α-SEAr reaction)

RC 1αA −1.7 (−0.9) −0.8 (+0.5) +5.2 (+5.8)
TS 1αA +58.5 (+57.2) +59.0 (+58.1) +69.5 (+67.3)
PC 1αA −29.4 (−28.4) −28.7 (−27.1) −21.2 (−20.7)
P 1αA −26.9 (−26.9) −27.0 (−26.4) −26.9 (−28.0)

Pathway B (α-SEAr reaction)

RC 1αB −2.1 (−1.1) −1.2 (+0.3) +4.7 (+4.8)
TS 1αB +58.2 (+56.8) +58.6 (+57.8) +69.3 (+67.0)
PC 1αB −10.2 (−9.1) −9.5 (−7.9) −1.9 (−1.1)
P 1αB −6.4 (−7.2) −6.5 (−6.8) −6.2 (−7.8)

Pathway C (β-SEAr reaction)

RC 1βC −1.9 (−0.7) −1.0 (+0.7) +6.0 (+5.1)
TS 1βC +65.5 (+63.8) +66.0 (+64.9) +76.2 (+73.6)
PC 1βC −13.2 (−12.1) −12.6 (−11.0) −3.9 (−3.9)
P 1βC −8.8 (−9.8) −8.9 (−9.4) −8.6 (−10.5)

The gas-phase pathway A of the α-SEAr reaction starts with RC 1αA in which the oxygen atom of NCS interacts with the α-hydrogen atom of the thiophene ring at a distance of 2.407 Å (Figure S1 in Supplementary Information). This O‧‧‧H interaction is not stabilising enough such that the ΔE of RC 1αA (−1.7 kcal mol−1; Table 1) is close to that of the separate reactants. RC 1αA leads to TS 1αA with a ΔE of +58.5 kcal mol−1. The geometry of TS 1αA comprises of a four-atom ring which involves the carbon, chlorine, nitrogen and hydrogen atoms with bond distances as shown in Fig. 1(a). TS 1αA generates PC 1αA (ΔE = −29.4 kcal mol−1) in which the chlorine atom from NCS substitutes the α-hydrogen atom of 1. This hydrogen atom bonds to the nitrogen atom of NCS. Moreover, an interaction between the β-hydrogen and the oxygen atoms occurs at a distance of 2.457 Å in PC 1αA (Figure S1).

RC 1αB in the pathway B of the gas-phase α-SEAr reaction is also formed through an interaction of the β-hydrogen atom of 1 and the oxygen atom of NCS. The O‧‧‧H distance is 2.461 Å (Figure S1) and the ΔE is −2.1 kcal mol−1 (Table 1). RC 1αB overcomes an activation barrier leading to TS 1αB in which a ring is formed as depicted in Fig. 1(b). Compared to TS 1αA , in TS 1αB , the oxygen atom of NCS interacts with the α-hydrogen atom of 1 at a distance of 2.010 Å. TS 1αB has a ΔE value of +58.2 kcal mol−1 (Table 1), which is comparable to that of TS 1αA (+58.5 kcal mol−1; Table 1) and thus, these two pathways are in competition. TS 1αB leads to PC 1αB where an interaction occurs between the nitrogen and β-hydrogen atoms at 2.709 Å. The relative energy of PC 1αB is −10.2 kcal mol−1, which indicates that the formation of the O–H bond is less favoured than the N–H.

In pathway C of the β-SEAr reaction in the gas phase, the oxygen atom of NCS interacts with the β-hydrogen atom of 1 in RC 1βC (Figure S1). The ΔE of RC 1βC is −1.9 kcal mol−1 (Table 1). Contrary to the pathway A, the pathway C of the β-SEAr reaction features a TS 1βC with a five-atom ring involving the carbon, chlorine, nitrogen, oxygen and hydrogen atoms [Fig. 1(b)]. A shorter distance (2.060 Å) occurs between the oxygen and hydrogen atoms of TS 1βC , compared to the nitrogen and hydrogen atoms in TS 1αA (2.356 Å; Fig. 1). However, the O‧‧‧H distance in TS 1βC is longer than in TS 1αB (2.010 Å), indicating that the O‧‧‧H interaction in TS 1βC is stronger than in TS 1αA and weaker than in TS 1αB . The ΔE value of TS 1βC is +65.5 kcal mol−1, which is higher than that of the TSs of the α-SEAr reaction. Hence, the α-SEAr reaction is kinetically more favoured than the β-SEAr reaction. TS 1βC leads to PC 1βC (−13.2 kcal mol−1) in which the chlorine atom replaces the β-hydrogen atom of 1. The hydrogen atom bonds to the oxygen atom of NCS as observed in the optimised structure of PC 1αB . This is different from the PC 1αA where the N–H bond is formed. The nitrogen atom in PC 1βC forms a double bond with the neighbouring C atom and interacts with the α′-hydrogen atom (Figure S1).

The pathways of the two SEAr reactions end with different products, P 1αA , P 1αB and P 1βC . The ΔH of P 1αA (−27.0 kcal mol−1) is lower than that of P 1αB (−6.5 kcal mol−1) and P 1βC (−8.9 kcal mol−1); the pathway A is more exothermic than the pathways B and C. P 1αA has a relative Gibbs free energy of −26.9 kcal mol−1, which is lower than that of P 1αB (−6.2 kcal mol−1) and P 1βC (−8.6 kcal mol−1). This difference shows that P 1αA is thermodynamically more stable than P 1αB and P 1βC . Thus, the pathway A of the α-SEAr reaction is both kinetically and thermodynamically more favoured than the β-SEAr reaction. Hence, electrophilic attack occurs at the α-position in 1 through pathway A. This observation is in line with the deduction which was reached through an analysis of the resonance structures in Scheme 1(b).

When the reactions were studied in acetic acid, the two pathways A and B were obtained for the α-SEAr reaction and the pathway C was obtained for the β-SEAr reaction, similar as in the gas phase. The geometries of the RCs and PCs are shown in Figure S2 in the Supplementary Information. On comparing the energy values of the solvent-phase reaction (Table 1), it was observed that the energy values are comparable.

SEAr reactions of thieno[2,3-b]thiophene

The SEAr reaction may occur at the α- and β-carbon atoms of 2. Two pathways (A and B) and one pathway (C) were located for the α- and β-SEAr reactions, respectively. Figure 2 depicts the PESs of the α- and β-SEAr reactions of 2 with NCS and the geometries of TS 2αA , TS 2αB and TS 2βC . Table 2 lists the relative energy in terms of ΔE, ΔH and ΔG of the RCs, TSs, PCs and Ps of the α- and β-SEAr reactions, in kcal mol−1, with respect to the separate reactants.

Fig. 2: 
PESs of the 2 + NCS α- and β-SEAr reactions. ΔE values are in kcal mol−1. Bond distances are in Å. All values resulting from the solvent-phase computations are in parentheses.
Fig. 2:

PESs of the 2 + NCS α- and β-SEAr reactions. ΔE values are in kcal mol−1. Bond distances are in Å. All values resulting from the solvent-phase computations are in parentheses.

Table 2:

ΔE, ΔH and ΔG, in kcal mol−1, of the RCs, TSs, PCs and Ps of the α- and β-SEAr reactions of 2 with NCS. The solvent-phase energy values are in parentheses.

ΔE ΔH ΔG
Pathway A (α-SEAr reaction)

RC 2αA −2.1 (+1.5) −1.1 (+2.9) +4.6 (+8.3)
TS 2αA +54.0 (+54.2) +54.6 (+55.3) +65.2 (+64.2)
PC 2αA −30.2 (−26.2) −29.3 (−24.8) −21.4 (−19.4)
P 2αA −27.1 (−28.1) −27.1 (−27.6) −27.0 (−29.1)

Pathway B (α-SEAr reaction)

RC 2αB −2.4 −1.5 +4.4
TS 2αB +53.9 +54.4 +65.0
PC 2αB −10.9 −10.2 −1.4
P 2αB −6.6 −6.6 −6.3

Pathway C (β-SEAr reaction)

RC −2.4 (+1.1) −1.4 (+2.5) +4.5 (+8.4)
TS +61.0 (+60.6) +61.6 (+61.8) +71.5 (+70.8)
PC −13.1 (−9.7) −12.3 (−8.4) −3.6 (−1.3)
P −8.7 (−11.0) −8.7 (−10.5) −8.4 (−11.5)

In RC 2αA of the pathway A of the gas-phase α-SEAr reaction, the oxygen atom of the NCS molecule interacts with the β- and β′-hydrogen atoms of 2 (Figure S3 in Supplementary Information) at distances of 2.673 and 2.678 Å, respectively. These O‧‧‧H interactions stabilise the RC 2αA with a relative electronic energy of −2.1 kcal mol−1 (Table 2). The pathway A of the α-SEAr reaction features TS 2αA (ΔE = +54.0 kcal mol−1) in which the NCS electrophile interacts with 2 in a similar manner as with 1. The C–Cl–N–H ring is formed with an N‧‧‧H distance of 2.352 Å and a C‧‧‧Cl distance of 2.502 Å. The nitrogen atom completely abstracts the α-hydrogen atom in PC . The oxygen atom interacts with the β- and β′-hydrogen atoms at distances of 2.795 and 2.779 Å, respectively. PC 2αA has a ΔE of −30.2 kcal mol−1.

Contrary to RC 2αA , the oxygen atom of the NCS moiety in RC 2αB (ΔE = −2.4 kcal mol−1; Table 2) interacts with the α-hydrogen atom of 2 at a distance of 2.505 Å (Figure S3) in pathway B in the gas phase. RC 2αB leads to TS 2αB in which the oxygen atom is also involved in the ring in the geometry of TS 2αB . TS 2αB has a ΔE of +53.9 kcal mol−1, which is comparable to TS 2αA . TS 2αB leads to PC 2αB in which an O–H bond is formed and the nitrogen atom interacts with the β- and β′-hydrogen atoms at 2.905 and 2.961 Å, respectively. This PC has a higher ΔE (−10.9 kcal mol−1) than PC 2αA .

The pathway C of the gas-phase β-SEAr reaction starts with RC 2βC (−2.4 kcal mol−1) in which an α′-H‧‧‧O interaction occurs at a distance of 2.467 Å (Figure S2 in Supplementary Information). RC 2βC overcomes an activation energy of +61.0 kcal mol−1 to reach TS 2βC which is structurally comparable to TS 1βC . This activation energy is higher than that of the α-SEAr reaction; the α-SEAr reaction is kinetically more favoured than the β-SEAr reaction. The oxygen atom interacts with the β-hydrogen atom at a distance of 2.116 Å in TS 2βC . The oxygen atom abstracts the β-hydrogen atom in PC 2βC . The nitrogen atom forms a double bond with the neighbouring carbon atom and interacts with the β′-hydrogen atom at a distance of 2.394 Å. PC 2βC (−13.1 kcal mol−1) is higher in ΔE than PC 2αA (−30.2 kcal mol−1) and lower in ΔE than PC 2αB (−10.9 kcal mol−1).

Table 2 shows that the ΔH of P 2αA is lower than that of P 2αB and P 2βC ; therefore, the pathway A of the α-SEAr reaction is more exothermic than the pathways B and C. The pathway A for the α-SEAr reaction of 2 is also thermodynamically more favoured owing to the lower ΔG of P 2αA . The electrophilic attack is expected to occur at the α-carbon atom with NCS as the α-SEAr reaction is both kinetically and thermodynamically more preferred than the attack at the β-carbon atom. The same trend was observed for the reaction of 1 with NCS. However, contrary to the reaction with 1, the preference for the α-SEAr reaction does not correlate with the deduction which was reached through the resonance structures [Scheme 1(c)]. This highlights the importance of mechanistic studies to predict reaction outcomes.

On considering the effect of acetic acid as solvent, the pathway B of the α-SEAr reaction was not located. The geometries of the RCs and PCs for pathways A and C are shown in Figure S4 in the Supplementary Information. Similar to the reactions of 1, the energy values computed in acetic acid are close to those computed in the gas phase. The trends observed also remain unchanged, that is, the reactants 2 and NCS prefer the α-SEAr reaction as shown by the lower activation barrier of the pathway A (+54.2 kcal mol−1) compared to the pathway C (+60.6 kcal mol−1).

Conclusions

The SEAr reactions of thiophene 1 and thieno[2,3-b]thiophene 2 with NCS were investigated to determine whether electrophilic attack occurs at the α- or the β-positions. The investigations were performed in the gas phase and in acetic acid. The Rs, RCs, TSs, PCs and Ps along the PESs of both SEAr reaction mechanisms were optimised using the B3LYP/6-311G(d), B3LYP-D3/6-311G(d) and M06-2X/6-311G(d) methods. The trends observed remain unchanged with the different methods; hence, only the B3LYP/6-311G(d) results are reported. An analysis of the resonance structures shows that electrophilic attack is expected to occur at the α-carbon atom in 1 and at the β-carbon atom in 2. However, on comparing the energetic parameters, the pathway A of the α-SEAr reaction was found to be kinetically and thermodynamically more favoured for both 1 and 2 in the gas phase and in acetic acid. Hence, chlorination with NCS will occur at the α-carbon atom for both 1 and 2, despite what is predicted by the resonance structures. This demonstrates how mechanistic studies help in predicting reaction outcomes. Further mechanistic studies involving the presence of substituents and different electrophiles would aid in deciphering general rules about SEAr reactions of 5:5 bicyclic heterocycles. In the future, we will be applying the methodology used in this paper to examine other 5:5 fused heterocycles: with other heteroatoms, with heteroatoms in other positions, and with different heteroatoms in the two rings. A further extension will deal with 5:5 systems with more than two heteroatoms.


Corresponding author: Ponnadurai Ramasami, Computational Chemistry Group, Department of Chemistry, Faculty of Science, University of Mauritius, Réduit 80837, Mauritius; and Centre of Natural Product Research, Department of Chemical Sciences, University of Johannesburg, Doornfontein Campus, Johannesburg 2028, South Africa, e-mail:

Article note: A collection of invited papers based on presentations at the Virtual Conference on Chemistry and its Applications 2022 (VCCA-2022) held on-line, 8-12 August 2022.


Acknowledgements

The authors are grateful to SEAGrid and South African CHPC for providing computing facilities. NS is also grateful to the Higher Education Commission (HEC) of Mauritius.

References

[1] R. J. Ouellette, J. D. Rawn. Electrophilic aromatic substitution. In Organic Chemistry, pp. 375–407, Academic Press, United Kingdom, 2nd ed. (2018).10.1016/B978-0-12-812838-1.50013-XSearch in Google Scholar

[2] W. Wang, X. Yang, R. Dai, Z. Yan, J. Wei, X. Dou, X. Qiu, H. Zhang, C. Wang, Y. Liu, S. Song, N. Jiao. J. Am. Chem. Soc. 144, 13415 (2022), https://doi.org/10.1021/jacs.2c06440.Search in Google Scholar PubMed

[3] A. K. Srivastava, S. K. Pandey, N. Misra. Theor. Chem. Acc. 135, 158 (2016), https://doi.org/10.1007/s00214-016-1918-5.Search in Google Scholar

[4] R. R. Naredla, D. A. Klumpp. Tetrahedron Lett. 53, 4779 (2012), https://doi.org/10.1016/j.tetlet.2012.06.135.Search in Google Scholar PubMed PubMed Central

[5] J. S. Yadav, B. V. S. Reddy, S. Shubashree, K. Sadashiv. Tetrahedron Lett. 45, 2951 (2004), https://doi.org/10.1016/j.tetlet.2004.02.073.Search in Google Scholar

[6] K. Smith, A. G. El-Hiti. Curr. Org. Synth. 1, 253 (2004), https://doi.org/10.2174/1570179043366747.Search in Google Scholar

[7] G. W. Weaver. Electrophilic aromatic substitution. In Organic Reaction Mechanisms Series, pp. 213–255, John Wiley & Sons, Ltd, United Kingdom (2023).10.1002/9781119608288.ch5bSearch in Google Scholar

[8] W. Mao, M. Ning, Z. Liu, Q. Zhu, Y. Leng, A. Zhang. Bioorg. Med. Chem. 20, 2982 (2012), https://doi.org/10.1016/j.bmc.2012.03.008.Search in Google Scholar PubMed

[9] L. Bréthous, N. Garcia-Delgado, J. Schwartz, S. Bertrand, D. Bertrand, J.-L. Reymond. J. Med. Chem. 55, 4605 (2012), https://doi.org/10.1021/jm300030r.Search in Google Scholar PubMed

[10] A. Ghomri, S. M. Mekelleche. J. Theor. Comput. Chem. 10, 435 (2011), https://doi.org/10.1142/s0219633611006566.Search in Google Scholar

[11] L. C. D. Rezende, S. M. G. Melo, S. Boodts, B. Verbelen, F. S. Emery, W. Dehaen. Dye. Pigment. 154, 155 (2018), https://doi.org/10.1016/j.dyepig.2018.01.043.Search in Google Scholar

[12] H. C. Kiliclar, E. Gencosman, Y. Yagci. Macromol. Rapid Commun. 43, 2100584 (2022), https://doi.org/10.1002/marc.202100584.Search in Google Scholar PubMed

[13] M. Liljenberg, J. H. Stenlid, T. Brinck. J. Mol. Model. 24, 15 (2017), https://doi.org/10.1007/s00894-017-3561-z.Search in Google Scholar PubMed PubMed Central

[14] Y. Osamura, K. Terada, Y. Kobayashi, R. Okazaki, Y. Ishiyama. J. Mol. Struct. Theochem 461–462, 399 (1999), https://doi.org/10.1016/s0166-1280(98)00452-7.Search in Google Scholar

[15] S. Liu. J. Phys. Chem. A 119, 3107 (2015), https://doi.org/10.1021/acs.jpca.5b00443.Search in Google Scholar PubMed

[16] M. Kruszyk, M. Jessing, J. L. Kristensen, M. Jørgensen. J. Org. Chem. 81, 5128 (2016), https://doi.org/10.1021/acs.joc.6b00584.Search in Google Scholar PubMed

[17] L. I. Belen’kii, N. D. Chuvylkin, I. D. Nesterov. Chem. Heterocycl. Compd. 48, 241 (2012), https://doi.org/10.1007/s10593-012-0985-3.Search in Google Scholar

[18] S. Sainas, A. C. Pippione, D. Boschi, M. L. B. T.-A. in H. C. Lolli, Academic Press (2021), Available from: https://link.springer.com/article/10.1007/s10593-012-0985-3#citeas.Search in Google Scholar

[19] N. Meanwell, R. Sistla. A survey of applications of tetrahydropyrrolo-3,4-azoles and tetrahydropyrrolo-2,3-azoles in medicinal chemistry. In Advances in Heterocyclic Chemistry, pp. 31–100, Academic Press, United States (2021).10.1016/bs.aihch.2020.10.004Search in Google Scholar

[20] A. C. Pippione, S. Sainas, D. Boschi, M. L. Lolli. Hydroxyazoles as acid isosteres and their drug design applications—Part 2: bicyclic systems. In Advances in Heterocyclic Chemistry, pp. 273–311, Academic Press, United States (2021).10.1016/bs.aihch.2020.12.002Search in Google Scholar

[21] B. Parrino, S. Cascioferro, D. Carbone, C. Pecoraro, G. Cirrincione, P. Diana. Eight-membered heterocycles with two heteroatoms in a 1,3-relationship of interest in medicinal chemistry. In Advances in Heterocyclic Chemistry, pp. 1–57, Academic Press, United States (2022).10.1016/bs.aihch.2020.11.001Search in Google Scholar

[22] DR. Naghammahmoodaljamali. Int. J. Curr. Res. Chem. Pharm. Sci. 1, 88 (2014).Search in Google Scholar

[23] A. Bugge. Acta Chem. Scand. 23, 2704 (1969), https://doi.org/10.3891/acta.chem.scand.23-2704.Search in Google Scholar

[24] V. Ryabova, L. Ignatovich. Topics in heterocyclic chemistry. In Thiophenes, J. A. Joule (Ed.), Vol. 39, Springer-Verlag, Berlin Heidelberg (2014).Search in Google Scholar

[25] G. K. S. Prakash, T. Mathew, D. Hoole, P. M. Esteves, Q. Wang, G. Rasul, G. A. Olah. J. Am. Chem. Soc. 126, 15770 (2004), https://doi.org/10.1021/ja0465247.Search in Google Scholar PubMed

[26] P. Wang, M. Ji, F. Sha. J. Chem. Res. 38, 622 (2014), https://doi.org/10.3184/174751914x14116517685637.Search in Google Scholar

[27] Z. Gan, B. Hu, Q. Song, Y. Xu. Synthesis 44, 1074 (2012), https://doi.org/10.1055/s-0031-1289732.Search in Google Scholar

[28] S. K. Sharma. Res. J. Chem. Sci. 5, 54 (2015).10.1021/ie504501aSearch in Google Scholar

[29] W. M. Gołebiewski, M. Gucma. Synthesis 23, 3599 (2007), https://doi.org/10.1055/s-2007-990871.Search in Google Scholar

[30] Z. Puterová, T. Bobula, D. Végh. J. Heterocycl. Chem. 45, 201 (2008), https://doi.org/10.1002/jhet.5570450124.Search in Google Scholar

[31] J. Skramstad, A. Lunde, H. Hope, V. Bjørnstad, P. Frøyen. J. Chem. Soc., Perkin Trans. 2, 1453 (2000), https://doi.org/10.1039/b001570i.Search in Google Scholar

[32] S. Egalahewa, M. Albayer, A. Aprile, J. L. Dutton. Inorg. Chem. 56, 1282 (2017), https://doi.org/10.1021/acs.inorgchem.6b02386.Search in Google Scholar PubMed

[33] A. D. Becke. Phys. Rev. A 38, 3098 (1988), https://doi.org/10.1103/physreva.38.3098.Search in Google Scholar PubMed

[34] C. Lee, W. Yang, R. G. Parr. Phys. Rev. B 37, 785 (1988), https://doi.org/10.1103/physrevb.37.785.Search in Google Scholar PubMed

[35] B. Miehlich, A. Savin, H. Stoll, H. Preuss. Chem. Phys. Lett. 157, 200 (1989), https://doi.org/10.1016/0009-2614(89)87234-3.Search in Google Scholar

[36] N. O. C. Winter, N. K. Graf, S. Leutwyler, C. Hättig. Phys. Chem. Chem. Phys. 15, 6623 (2013), https://doi.org/10.1039/c2cp42694c.Search in Google Scholar PubMed

[37] M. A. Solomos, J. A. Bertke, J. A. Swift. New J. Chem. 42, 7125 (2018), https://doi.org/10.1039/c8nj00629f.Search in Google Scholar

[38] A. Ashraf, M. Khalid, M. N. Tahir, M. Yaqub, M. M. Naseer, G. M. Kamal, B. Saifullah, A. A. C. Braga, Z. Shafiq, W. Rauf. RSC Adv. 9, 34567 (2019), https://doi.org/10.1039/c9ra05415d.Search in Google Scholar PubMed PubMed Central

[39] L. Feng, X. Ren, Y. Feng, B. Tan, S. Zhang, W. Li, J. Liu. Phys. Chem. Chem. Phys. 22, 4592 (2020), https://doi.org/10.1039/c9cp06910k.Search in Google Scholar PubMed

[40] V. A. Dixit, W. R. F. Goundry, S. Tomasi. New J. Chem. 41, 3619 (2017), https://doi.org/10.1039/c7nj00950j.Search in Google Scholar

[41] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. MontgomeryJr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, D. J. Fox. Gaussian 16, Revision B.01, Gaussian, Inc., Wallingford CT (2016).Search in Google Scholar

[42] S. Pamidighantam, S. Nakandala, E. Abeysinghe, C. Wimalasena, S. Rathnayakae, S. Marru, M. Pierce. Proc. Comput. Sci. 80, 1927 (2016), https://doi.org/10.1016/j.procs.2016.05.535.Search in Google Scholar

[43] N. Shen, Y. Fan, S. Pamidighantam. J. Comput. Sci. 5, 576 (2014), https://doi.org/10.1016/j.jocs.2014.01.005.Search in Google Scholar

[44] R. Dooley, K. Milfeld, C. Guiang, S. Pamidighantam, G. Allen. J. Grid Comput. 4, 195 (2006), https://doi.org/10.1007/s10723-006-9043-7.Search in Google Scholar

[45] K. Milfeld, C. Guiang, S. Pamidighantam, J. Giuliani. Proceedings of the 2005 Linux Clusters: The HPC Revolution, Chapel Hill, North Carolina (2005).Search in Google Scholar

[46] L. Fan, L. Versluis, T. Ziegler, E. J. Baerends, W. Ravenek. Int. J. Quantum Chem. 34, 173 (1988), https://doi.org/10.1002/qua.560340821.Search in Google Scholar

[47] M. Page, C. Doubleday, J. W. McIver. J. Chem. Phys. 93, 5634 (1990), https://doi.org/10.1063/1.459634.Search in Google Scholar

[48] J. Tomasi, B. Mennucci, R. Cammi. Chem. Rev. 105, 2999 (2005), https://doi.org/10.1021/cr9904009.Search in Google Scholar PubMed

[49] R. Dennington, T. Keith, J. Millam. GaussView, Semichem Inc., Shawnee Mission, KS (2009).Search in Google Scholar

[50] G. A. Zhurko, D. A. Zhurko. ChemCraft, Version 1.7 (2013), Available from: http://www.chemcraftprog.com.Search in Google Scholar

[51] C. Y. Legault. CYLview, Version 1.0b, Université de Sherbrooke (2009), Available from: http://www.cylview.org.Search in Google Scholar

[52] S. Grimme, J. Antony, S. Ehrlich, H. Krieg. J. Chem. Phys. 132, 154104 (2010), https://doi.org/10.1063/1.3382344.Search in Google Scholar PubMed

[53] Y. Zhao, D. G. Truhlar. Theor. Chem. Acc. 120, 215 (2008), https://doi.org/10.1007/s00214-007-0310-x.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/pac-2022-1104).


Received: 2022-11-13
Accepted: 2023-04-21
Published Online: 2023-06-02
Published in Print: 2023-07-26

© 2023 IUPAC & De Gruyter

Downloaded on 25.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/pac-2022-1104/html
Scroll to top button