Next Article in Journal
Interlaboratory Evaluation of a User-Friendly Benchtop Mass Spectrometer for Multiple-Attribute Monitoring Studies of a Monoclonal Antibody
Next Article in Special Issue
Influence of the Solvent on the Stability of Aminopurine Tautomers and Properties of the Amino Group
Previous Article in Journal
Effect of Electrode Spacing on the Performance of a Membrane-Less Microbial Fuel Cell with Magnetite as an Additive
Previous Article in Special Issue
Excited-State Polarizabilities: A Combined Density Functional Theory and Information-Theoretic Approach Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aromatic Character and Relative Stability of Pyrazoloporphyrin Tautomers and Related Protonated Species: Insights into How Pyrazole Changes the Properties of Carbaporphyrinoid Systems

Department of Chemistry, Illinois State University, Normal, IL 61790-4160, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2854; https://doi.org/10.3390/molecules28062854
Submission received: 22 February 2023 / Revised: 12 March 2023 / Accepted: 17 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Computational and Theoretical Studies on Isomeric Organic Compounds)

Abstract

:
Pyrazoloporphyrins (PzPs), which are porphyrin analogues incorporating a pyrazole subunit, are examples of carbaporphyrin-type structures with a carbon atom within the macrocyclic cavity. DFT calculations were used to assess a series of 17 PzP tautomers, nine monoprotonated species and four related diprotonated PzP dications. The geometries of the structures were optimized using M06-2X/6-311++G(d,p), and the relative stabilities computed with the cc-PVTZ functional. Nucleus independent chemical shifts, both NICS(0) and NICS(1)zz, were calculated, and the anisotropy of the induced current density (AICD) plots were generated for all of the species under investigation. The results for free base PzPs show that fully aromatic PzP tautomers are not significantly more stable than weakly aromatic cross-conjugated species. In addition, strongly aromatic structures with internal CH2′s are much less stable, a feature that is also seen for protonated PzPs. The degree of planarity for the individual macrocycles does not significantly correlate with the stability of these structures. The results allow significant aromatic conjugation pathways to be identified in many cases, and provide insights into the aromatic properties of this poorly studied system. These investigations also complement experimental results for PzPs and emphasize the need for further studies in this area.

1. Introduction

Investigations into porphyrin-type structures has exploded over the last thirty years [1,2,3]. In 1994, two groups independently discovered N-confused porphyrins (NCPs, 1) (Figure 1) as by-products in the acid-catalyzed condensation of pyrrole with aromatic aldehydes [4,5], and an efficient synthesis of this system was subsequently reported by Lindsey and Geier [6]. Immediately after the initial disclosure of NCPs, a number of related porphyrinoids were reported, including carbaporphyrins 2 and 3 [7,8,9], azuliporphyrins 4 [10,11], benziporphyrins 5 [12,13,14] and tropiporphyrins 6 [15,16]. Carbaporphyrinoid systems, porphyrin-like molecules with one or more carbon atoms within the macrocyclic core, exhibit diverse and unusual reactivity and may exist as aromatic, nonaromatic or antiaromatic species [1,17,18]. In addition, carbaporphyrin-type systems react with many late transition metal cations to generate stable organometallic derivatives under mild conditions [19,20,21,22]. N-Confused porphyrins have been particularly well studied with regard to their metalation properties [22], but complementary organometallic chemistry has also been described for related systems [19,20]. N-Confused porphyrins have two major tautomeric forms, 1a and 1b, that have dramatically different spectroscopic properties. Tautomer 1a, which is favored in nonpolar solvents including chloroform, is fully aromatic, and the proton NMR spectra for NCPs exhibit a strong diamagnetic ring current that results in the inner CH being shifted upfield to approximately −4 ppm. Tautomer 1b is approximately 5 kcal/mol higher in energy [23,24,25,26,27], but is favored in polar aprotic solvents such as DMF or DMSO [23]. This is due to favorable hydrogen bonding interactions between the solvent and the external NH. As 1b is cross-conjugated, the global aromatic character for the macrocycle is greatly diminished.
Carbaporphyrin 2 is structurally similar to NCP tautomer 1b and also exhibits a strongly aromatic character [8]. In order to further investigate these systems, porphyrin analogues 7 with pyrazole subunits were targeted for investigation [28,29]. Structures 2, 1 and 7 provide a series of carbaporphyrinoids with an increasing number of nitrogens in the modified subunit. Therefore, this series can provide new insights into the properties of these systems. In addition, pyrazoloporphyrins (PzPs) might exist in a number of tautomeric forms, including tautomers 7a and 7b, that are analogous to the major species for NCP. Although N-confused porphyrins are commonly synthesized using Rothemund-Lindsey reaction conditions [4,5,6], alternative routes to NCPs have been developed, including ‘2 + 2’ [30] and ‘3 + 1’ strategies [31,32]. The ‘3 + 1’ version of the MacDonald condensation has been a particularly useful method for preparing carbaporphyrinoid systems [33,34] (Scheme 1). The reaction of tripyrrane 8 with indene dialdehyde 9 in the presence of trifluoroacetic acid (TFA) afforded carbaporphyrin 10, while trialdehyde 11 gave modest yields of carbaporphyrin aldehydes 12 [7,35]. Methoxymethylene cyclopentene aldehyde 13 similarly afforded carbachlorin 14, and this can be further oxidized to carbaporphyrin 15 [8]. The same methodology was successfully applied to the preparation of N-confused porphyrins 16 and 17 [31,32]. Unfortunately, attempts to condense pyrazole dialdehyde 18 with 8 failed to give the expected PzP 19 (Scheme 2). However, N-substituted dialdehydes 20 reacted with 8 to give phlorin 21 and careful oxidation with aqueous FeCl3 afforded PzPs 22 [28,29] The presence of N-phenyl, methyl or ethyl substituents prevents the system from tautomerizing into a fully aromatic porphyrinoid. The proton NMR spectra for PzPs 22 showed that they possess a weak diatropic ring current, and this was attributed to contributions from dipolar resonance contributors such as 22′. The sequential protonation of 22 with TFA initially gave monocation 22H+ and then dication 22H22+. Oxidation of the phlorin intermediate with silver(I) acetate gave oxopyrazolophlorins 23 and 24. Metalation with nickel(II) or palladium(II) acetate afforded the corresponding organometallic derivatives 25 and 26, respectively [28,29]. The proton NMR spectra for 25 and 26 showed the presence of moderate diatropic ring currents. However, all vestiges of an aromatic character were lost upon protonation. The aromatic character of 25 and 26 can be attributed to dipolar resonances 25H+ and 26H+ that incorporate 18π electron delocalization pathways. However, protonation of the external nitrogens inhibits this interaction because the corresponding canonical forms 25′H+ and 26′H+ would necessitate the placement of positive charges on adjacent nitrogen atoms. The earliest investigations into pyrazole-containing porphyrinoids were directed at the synthesis of non-aromatic hexaphyrin analogues 27 [36], and texaphyrin analogue 28 has also been reported [37]. More significantly, a fully conjugated hexaphyrin analogue designated Siamese-twin porphyrin 29 has been synthesized [38] (Figure 2). This system proved to be nonaromatic but possesses two binding pockets that are similar to the binding cores of porphyrins. Extensive metalation studies on this system have been reported [39]. Furuta also investigated the synthesis of PzPs [40] (Scheme 3). Tripyrrane 8 was condensed with pyrazole dialdehyde 20d in the presence of trifluoroacetic acid, followed by oxidation with FeCl3, but in this case oxopyrazolophlorin 23d was isolated instead of the expected N-benzyl PzP 22d. Cleavage of the benzyl group with AlCl3 afforded a tautomeric mixture of N-unsubstituted oxopyrazolophlorins 23e and 23e′ [40]. The favored tautomer 23e self-assembled as a hydrogen bonded dimer (Scheme 3). Tripyrrane analogues can be reacted with dialcohols to give modified porphyrinoids. For instance, tert-butyl N-confused tripyrrane 30 reacted with pyrrole dicarbinol 31 to give tert-butyl NCP 32 and subsequent treatment with aqueous sulfuric acid at 160 °C afforded unsubstituted N-confused porphyrin 1 [41]. However, the acid-catalyzed condensation of pyrazole dialcohol 33 with diphenyltripyrrane 34, followed by oxidation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), did not give targeted PzP 35, but instead afforded a mixture of triphenylcorrole 36 and a nonaromatic expanded porphyrinoid 37 that incorporated a 6-carbon chain derived from DDQ [40].
Due to the many difficulties encountered in the synthesis of pyrazoloporphyrins, this system is not well studied. N-Unsubstituted PzPs are presently unknown, and it has not been possible to contrast their tautomeric properties with closely related NCPs. In order to gain a better understanding of the pyrazoloporphyrin system, a computational study was conducted. The goals were threefold. Firstly, the conformations and relative stabilities of 17 potential PzP tautomers were considered. Secondly, the aromatic characteristics of these tautomers were assessed using nucleus independent chemical shift (NICS) calculations and the anisotropy of induced current density (AICD) plots. Thirdly, these methods were used to predict the relative stabilities and aromatic characteristics of protonated PzPs.

2. Results

2.1. Aromatic and Nonaromatic Tautomers of Pyrazoloporphyrins

Computational methods can provide valuable insights into the structures and aromatic properties of porphyrinoid systems [42,43,44,45,46,47,48,49,50]. In our earlier work, we have assessed the relative stabilities and aromatic properties of a wide range of tautomeric structures and protonated species for carbaporphyrins [51], benzo- and naphthocarbaporphyrins [52,53], heterocarbaporphyrins [54], carbachlorins [8,55], dicarbaporphyrins [56,57,58], benziporphyrins [59], pyreniporphyrins [60], neo-confused porphyrins [52,61], azuliporphyrins [62], 21-oxyazuliporphyrins [63], tropiporphyrins [62], oxyquinoliziniporphyrins [64] and carbatriphyrins [65]. In this study, a series of 17 free base PzP tautomers (Table 1 and Table 2) were examined. The structures were optimized using M06-2X with the triple-ζ basis set 6-311++G(d,p). Seven tautomers are fully conjugated (Table 1), or contain a cross-conjugated element, while the 10 remaining tautomers have methylene bridges (Table 2) that interrupt macrocyclic conjugation. The relative stability and π-delocalization pathways in these tautomers were investigated. The peripheral atoms for PzP are numbered 1–20, and the inner atoms are designated 21–24, as shown for structure 7a in Figure 1. Two hydrogens are relocated in each of the tautomeric structures, and the numerical positions of these hydrogens are designated in the structure notation. For instance, the most stable tautomer of pyrazoloporphyrin is PzP-22,24-H, where the “mobile” hydrogens are located at positions 22 and 24.
Tautomer PzP-22,24-H is very close in energy to the cross-conjugated form PzP-2,23-H (Table 1). This differs from NCPs, where the equivalent aromatic tautomer is approximately 5 kcal/mol lower in energy than the cross-conjugated version. It is noteworthy that the relative energies and Gibbs free energies are very similar for all of the calculated tautomers, demonstrating that entropic factors are not significant. As is the case for aromatic carbaporphyrins and NCPs, PzP has three internal hydrogens, and the crowded cavity leads to the pyrazole ring being pivoted out of the macrocyclic plane (Table 3). Nevertheless, some distortion to the macrocycle is present in PzP-2,23-H as well. Related tautomers PzP-22,23-H, PzP-2,22-H and PzP-2,24-H are 6.67–7.88 kcal/mol higher in energy than PzP-22,24-H (Table 1). This is due to less effective hydrogen bonding interactions within the core [66,67,68,69], as well as less than optimal placement of the hydrogens leading to unfavorable steric interactions. Two additional fully conjugated forms with internal methylene units, PzP-21,22-H and PzP-21,23-H, are possible, but these are 18.32 and 11.27 kcal/mole higher in energy, respectively, than PzP-22,24-H. The former is less stable due to the core arrangement being less well suited for hydrogen bonding interactions as well as placing the internal hydrogens at adjacent positions. Nevertheless, both tautomers are virtually planar and possess 18π electron conjugation pathways [70]. This result demonstrates that global aromaticity and overall planarity are not in themselves sufficient to determine the stability of these structures. It is noteworthy that only one example of a carbaporphyrinoid system with an internal CH2 is presently known [57]. Tautomers with interrupted conjugation due to the presence of a methylene bridge (Table 2) were all much less stable with relative energies of 22.28–32.76 kcal/mol higher than PzP-22,24-H. The degree of distortion for these structures varies considerably, although PzP-2,5-H, PzP-2,10-H, PzP-2,15-H, PzP-2,20-H and PzP-5,23-H are all essentially planar. This factor does not play a role in determining stability, although the first four are particularly unstable.
The aromatic character of the individual tautomers were probed using nucleus independent chemical shift (NICS) calculations and the anisotropy of induced current density (AICD) plots. NICS calculations [71] were carried out using the GIAO method. NICS calculations assess the chemical shifts that might be observed in proton NMR studies but does not require a nucleus to be present at the site of interest. Large negative values correspond to highly shielded regions that result from aromatic ring currents, while positive values correspond to deshielded regions. The latter could result from paratropic ring currents (antiaromaticity), but will also emerge when a ring is placed just outside of a major aromatic conjugation pathway. Two types of NICS calculations, NICS(0) and NICS(1)zz, were conducted. Standard NICS calculations take into account the effects due to σ and π electrons and do not always reliably assess aromatic properties, but NICSzz primarily measures the effects due to the π system [72,73,74,75]. NICSzz calculations were performed 1 Å above the ring. The two techniques were mostly consistent, but the numerical values obtained for NICS(1)zz are much larger than those obtained using NICS(0). In the discussion, the NICS(1)zz values are emphasized. Favored tautomer PzP-22,24-H gave strongly aromatic NICS(0) and NICS(1)zz values of −13.05 and −30.62 ppm, respectively, while cross-conjugated form PzP-2,23-H gave very weakly diatropic values of −1.76 and −2.88 ppm. Given that these structures have very similar stabilities, the aromatic nature of PzP-22,24-H appears to offer few benefits. PzP-22,23-H has comparable aromatic properties to PzP-22,24-H, while PzP-2,22-H and PzP-2,24-H are similar, albeit slightly more diatropic, to PzP-2,23-H. High energy tautomers PzP-21,23-H and PzP-21,22-H with internal CH2 units both gave strongly aromatic NICS(1)zz values (Table 1). The NICS(0)/NICS(1)zz values for the individual pyrazole and pyrrole rings in these structures were also determined. These rings were designated a, b, c and d, where ring a corresponds to the pyrazole unit (Table 1 and Table 2). For PzP-22,24-H, rings b and d gave NICS(1)zz strongly negative values (ca. −30 ppm) but ring c gave a smaller negative value of −11.73 ppm and ring a was only −2.12 ppm. These results suggest that an 18π electron circuit that passes through the inside of rings a and c and the outside of rings b and d is favored (Figure 3). This type of conjugation pathway is also favored for NCP 1a and carbaporphyrins 2 and 3 (Figure 1). For tautomer PzP-2,23-H, ring a gives a strongly negative value of ca. −20 ppm, indicating that the pyrazole unit is strongly aromatic but primarily isolated from the rest of the macrocycle. This is also seen for tautomers PzP-2,22-H and PzP-2,24-H. Tautomer PzP-22,23-H gives strongly negative values for rings c and d that are consistent with the presence of an 18π electron pathway (Figure 3). Tautomer PzP-21,23-H has strongly negative NICS(1)zz values for ring a and c, while PzP-21,22-H gives similarly negative values for rings a and d. These results show that 18π electron circuits that pass around the outside of ring a are present (see Figure 3). Tautomers with interrupted conjugation (Table 2) give varying values for the individual rings. In PzP-2,5-H, PzP-2,10-H, PzP-2,15-H and PzP-2,20-H, the pyrazole unit gives strongly negative NICSzz(1a) values, but the aromatic properties of this ring are diluted in the remaining tautomers.
In addition to the NICS calculations, all the investigated structures were probed using anisotropy of induced current density (AICD) [76] (see Supporting Information). AICD plots show the induced flow of electrons due to an applied magnetic field and allow the visualization of aromatic ring currents. The plot for PzP-22,24-H confirms the presence of the proposed 18π electron circuit (Figure 4), but a degree of bifurcation is also present. The main aromatic pathways pass through the inside of the pyrazole ring, but the routes are less well delineated in ring c. This observation is also consistent with the NICS calculations. Unfavored aromatic tautomer PzP-21,23-H also shows the presence of an 18π electron pathway, but alternative routes through the pyrrolic subunits are clearly evident. Bond length calculations provide further insights (Figure 5). For PzP-22,24-H, the closed circuit representing the proposed aromatic pathway shows little bond length alternation apart from the carbon-carbon bonds within the individual pyrrole units. C6-C7, C7-C8, C16-C17 and C18-C19, all of which fall within the aromatic pathway, gave values of 1.432–1.433 Å, but the equivalent bonds in ring c, C11-C12 and C13-C14, gave slightly longer values of 1.462 Å. C12-C13 for ring c gave a relatively short bond length of 1.345 Å compared to the equivalent bonds C7-C8 and C17-C18 in rings b and d, which came out as 1.368 Å. These trends are consistent with the proposed conjugation pathway. For PzP-2,23-H, a substantial bond length alternation is evident, as would be expected for a nonaromatic porphyrinoid.

2.2. Mono- and Diprotonated Pyrazoloporphyrins

A series of 9 monoprotonated pyrazoloporphyrin tautomers were considered (Table 4). For this series, only structures that were fully conjugated were investigated. The most stable tautomer, PzP-2,22,24-H+, was protonated onto an external nitrogen. The lowest energy tautomer that had been internally protonated, PzP-22,23,24-H+, was calculated to be nearly 17 kcal/mol higher in energy. The two remaining externally protonated tautomers, PzP-2,22,23-H+ and PzP-2,23,24-H+, are only approximately 4 kcal/mol higher in energy, and the differences can be attributed to a combination of steric and poorer hydrogen bonding interactions. Five of the monocations have internal methylene units. The most stable of these, PzP-2,21,23-H+, is still over 17 kcal/mol higher in energy than PzP-2,22,24-H+. This species is externally protonated, but the most stable version with an inner CH2 that is internally protonated, PzP-21,22,24-H+, is nearly 21 kcal/mol higher in energy. The three most favored tautomers, PzP-2,22,24-H+, PzP-2,22,23-H+ and PzP-2,23,24-H+, are by far the least aromatic forms, with NICS(1)zz values between −4.5 and −8.5 ppm. The AICD plot for PzP-2,22,24-H+ (Figure 6a) does not show clear global aromatic pathways. Furthermore, the calculated bond lengths for PzP-2,22,24-H+ show a considerable amount of bond length alternation (Figure 7). In contrast, the AICD plot for PzP-22,23,24-H+ (Figure 6b) indicates that a 19-atom pathway is favored (Figure 8), and this corresponds to the presence of an 18π electron cationic [19]annulene component. This analysis is supported by the NICS(1)zz data, as these show large negative values for rings b, c and d, but not for the pyrazole ring. PzP-2,21,22-H+, which is externally protonated but with an internal methylene unit, gave an AICD plot (Figure 6c) that was consistent with the presence of an 18-atom 18π electron pathway, while the AICD plot for PzP-21,22,24-H+ (Figure 6d), which also has an internal methylene unit but is protonated internally, favors a 19-atom 18π electron circuit (Figure 8). Again, these conjectures are supported by the NICS(1)zz calculations (Table 5).
Four diprotonated PzP dications were considered. By far the most stable tautomer for this series was [PzP-2,22,23,24-H]2+, but this species is the least aromatic with a NICS(1)zz value of only −7.49 ppm. The remaining dications, [PzP-2,21,22,23,24-H]2+, [PzP-2,22,23,24-H]2+ and [PzP-2,22,23,24-H]2+, have internal CH2 units and NICS(1)zz values between −24.93 and −29.07 ppm. The calculated bond lengths for [PzP-2,22,23,24-H]2+ show extensive bond length alternation (Figure 7), as would be expected for a weakly aromatic species.

3. Computational Methods

All DFT calculations used the Gaussian 16 revision C.01 [77]. Geometry optimizations were performed using M06-2X/6-311++G(d,p) [78,79,80]. Using the same functional and basis set, the corresponding vibrational frequencies were computed to confirm the absence of imaginary frequencies and derive zero-point energy and vibrational entropy corrections from unscaled frequencies. Single point energy calculations were performed on the optimized minima using M06-2X/cc-PVTZ [81]. NICS values were calculated using the GIAO method, and AICD plots were obtained from CGST calculations. NICS(0) was calculated at the mean position of the four heavy atoms at the center of the macrocycle. NICS(a), NICS(b), NICS(c), and NICS(d) were calculated at the mean position of the five atoms that make up each component ring in the macrocycle. Additionally, NICS(1)zz, NICS(1a)zz, NICS(1b)zz, NICS(1c)zz, and NICS(1d)zz were obtained by applying the same calculations 1 Å above each of the corresponding points extracting the zz contribution of the magnetic tensor. The resulting energies, cartesian coordinates, and AICD plots can be found in the Supplementary Materials.

4. Conclusions

Pyrazoloporphyrins (PzPs) represent a logical extension of the structural concepts associated with N-confusion, but to date research on this system has been limited, and N-unsubstituted pyrazoloporphyrins are not presently known. DFT calculations have been used to assess the properties of N-unsubstituted PzPs, and the results provide new and unexpected insights. The aromatic properties were probed using anisotropy of induced current density (AICD) plots and nucleus independent chemical shift (NICS) calculations. The two most stable tautomers of PzP differ in energy by <0.4 kcal/mol, even though one form is strongly aromatic, while the other version is cross-conjugated and only weakly aromatic. Tautomers with internal CH2 units are also very aromatic species, but have much higher calculated energies. Monoprotonation can give rise to a number of monocationic species, nine of which were considered. The favored tautomers are externally protonated with very weak diatropic characteristics. Internal protonation gives a cation that is strongly aromatic, but it is nearly 17 kcal/mol higher in energy than the most favorable externally protonated form. Cations with internal methylenes are also very aromatic species but are relatively unstable. Diprotonated PzP dications were also considered. The most stable dication is weakly aromatic, and fully aromatic cations with inner CH2′s are >15 kcal/mol higher in energy. The results are in accord with the reported properties for 2-substituted PzPs, and highlight the profoundly altered properties of this system compared to NCPs and carbaporphyrins.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28062854/s1, Table S1: AICD plots for pyrazoloporphyrin tautomers and related protonated species; Table S2: Calculated energies (hartrees) for all the optimized structures. Table S3: Cartesian coordinates.

Author Contributions

All computational studies were performed by D.I.A.; based upon concepts developed by T.D.L. The manuscript was written by T.D.L., and this author was also responsible for project administration and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation under grant number CHE-1855240.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Lash, T.D. Carbaporphyrinoid Systems. Chem. Rev. 2017, 117, 2313–2446. [Google Scholar] [CrossRef] [PubMed]
  2. Lash, T.D. Heteroporphyrins and Carbaporphyrins. In Fundamentals of Porphyrin Chemistry: A 21st Century Approach; Brothers, P.J., Senge, M.O., Eds.; Wiley: Hoboken, NJ, USA, 2022; Volume 1, pp. 385–451. [Google Scholar]
  3. Toganoh, M.; Furuta, H. Creation from Confusion to Fusion in the Porphyrin World—The Last Three Decades of N-Confused Porphyrinoid Chemistry. Chem. Rev. 2022, 122, 8313–8437. [Google Scholar] [CrossRef] [PubMed]
  4. Furuta, H.; Asano, T.; Ogawa, T. N-Confused Porphyrin: A New Isomer of Tetraphenylporphyrin. J. Am. Chem. Soc. 1994, 116, 767–768. [Google Scholar] [CrossRef]
  5. Chmielewski, P.J.; Latos-Grażyński, L.; Rachlewicz, K.; Glowiak, T. Tetra-p-tolylporphyrin with an Inverted Pyrrole Ring: A Novel Isomer of Porphyrin. Angew. Chem. Int. Ed. Engl. 1994, 33, 779–781. [Google Scholar] [CrossRef]
  6. Geier, G.R., III; Haynes, D.M.; Lindsey, J.S. An Efficient One-Flask Synthesis of N-Confused Tetraphenylporphyrin. Org. Lett. 1999, 1, 1455–1458. [Google Scholar] [CrossRef]
  7. Lash, T.D.; Hayes, M.J. Carbaporphyrins. Angew. Chem. Int. Ed. Engl. 1997, 36, 840–842. [Google Scholar] [CrossRef]
  8. Li, D.; Lash, T.D. Synthesis and Reactivity of Carbachlorins and Carbaporphyrins. J. Org. Chem. 2014, 79, 7112–7121. [Google Scholar] [CrossRef]
  9. Garbicz, M.; Latos-Grażyński, L. A meso-Tetraaryl-21-carbaporphyrin: Incorporation of a Cyclopentadiene Unit into a Porphyrin Architecture. Angew. Chem. Int. Ed. 2019, 58, 6089–6093. [Google Scholar] [CrossRef]
  10. Lash, T.D.; Chaney, S.T. Azuliporphyrin, a Case of Borderline Porphyrinoid Aromaticity. Angew. Chem. Int. Ed. Engl. 1997, 36, 839–840. [Google Scholar] [CrossRef]
  11. Lash, T.D. Out of the Blue! Azuliporphyrins and Related Carbaporphyrinoid Systems. Acc. Chem. Res. 2016, 49, 471–482. [Google Scholar] [CrossRef]
  12. Lash, T.D. Oxybenziporphyrin, an Aromatic Semiquinone Porphyrin Analog with Pathways for 18π-Electron Delocalization. Angew. Chem. Int. Ed. Engl. 1995, 34, 2533–2535. [Google Scholar] [CrossRef]
  13. Lash, T.D.; Chaney, S.T.; Richter, D.T. Conjugated Macrocycles Related to the Porphyrins. Part 12. Oxybenzi- and Oxypyriporphyrins: Aromaticity and Conjugation in Highly Modified Porphyrinoid Structures. J. Org. Chem. 1998, 63, 9076–9088. [Google Scholar] [CrossRef]
  14. Lash, T.D. Benziporphyrins, a Unique Platform for Exploring the Aromatic Characteristics of Porphyrinoid Systems. Org. Biomol. Chem. 2015, 13, 7846–7878. [Google Scholar] [CrossRef] [PubMed]
  15. Lash, T.D.; Chaney, S.T. Conjugated Macrocycles Related to the Porphyrins. Part 7. Tropiporphyrin: Tropylium versus Porphyrinoid Aromaticity. Tetrahedron Lett. 1996, 37, 8825–8828. [Google Scholar] [CrossRef]
  16. Bergman, K.M.; Ferrence, G.M.; Lash, T.D. Tropiporphyrins, Cycloheptatrienyl Analogues of the Porphyrins: Synthesis, Spectroscopy, Chemistry and Structural Characterization of a Silver(III) Derivative. J. Org. Chem. 2004, 69, 7888–7897. [Google Scholar] [CrossRef] [PubMed]
  17. Lash, T.D. Carbaporphyrinoids: Taking the Heterocycle Out of Nature’s [18]Annulene. Synlett 2000, 2000, 279–295. [Google Scholar]
  18. Lash, T.D. Recent Advances on the Synthesis and Chemistry of Carbaporphyrins and Related Porphyrinoid Systems. Eur. J. Org. Chem. 2007, 5461–5481. [Google Scholar] [CrossRef]
  19. Lash, T.D. Metal Complexes of Carbaporphyrinoid Systems. Chem. Asian J. 2014, 9, 682–705. [Google Scholar] [CrossRef]
  20. Lash, T.D. Organometallic Chemistry within the Structured Environment Provided by the Macrocyclic Cores of Carbaporphyrins and Related Systems. Molecules 2023, 28, 1496. [Google Scholar] [CrossRef]
  21. Thuita, D.W.; Brückner, C. Metals Complexes of Porphyrinoids Containing Nonpyrrolic Heterocycles. Chem. Rev. 2022, 122, 7990–8052. [Google Scholar] [CrossRef]
  22. Toganoh, M.; Furuta, H. Handbook of Porphyrin Science—With Applications to Chemistry, Physics, Material Science, Engineering, Biology and Medicine; Kadish, K.M., Smith, K.M., Guilard, R., Eds.; World Scientific Publishing: Singapore, 2010; Volume 2, pp. 103–192. [Google Scholar]
  23. Furuta, H.; Ishizuka, T.; Osuka, A.; Dejima, H.; Nakagawa, H.; Ishikawa, Y. NH Tautomerism of N-Confused Porphyrin. J. Am. Chem. Soc. 2001, 123, 6207–6208. [Google Scholar] [CrossRef] [PubMed]
  24. Ghosh, A.; Wondimagegn, T.; Nilsen, H.J. Molecular Structures, Tautomerism, and Carbon Nucleophilicity of Free-Base Inverted Porphyrins and Carbaporphyrins: A Density Functional Theory Study. J. Phys. Chem. B 1998, 102, 10459–10467. [Google Scholar] [CrossRef]
  25. Szterenberg, L.; Latos-Grażyński, L. Structure and Stability of 2-Aza-21-carbaporphyrin Tautomers Prearranged for Coordination. Inorg. Chem. 1997, 36, 6287–6291. [Google Scholar] [CrossRef]
  26. Furuta, H.; Maeda, H.; Osuka, A. Stability and Structure of N-Confused Porphyrins. J. Org. Chem. 2000, 65, 4222–4226. [Google Scholar] [CrossRef]
  27. Furuta, H.; Maeda, H.; Osuka, A. Theoretical Study of Stability, Structures, and Aromaticity of Multiply N-Confused Porphyrins. J. Org. Chem. 2001, 66, 8563–8572. [Google Scholar] [CrossRef] [PubMed]
  28. Lash, T.D.; Young, A.M.; Von Ruden, A.L.; Ferrence, G.M. Adding to the Confusion! Synthesis and Metalation of Pyrazole Analogues of the Porphyrins. Chem. Commun. 2008, 44, 6309–6311. [Google Scholar] [CrossRef] [PubMed]
  29. Young, A.M.; Von Ruden, A.L.; Lash, T.D. Pyrazole Analogues of the Porphyrins and Oxophlorins. Org. Biomol. Chem. 2011, 9, 6293–6305. [Google Scholar] [CrossRef]
  30. Liu, B.Y.; Brückner, C.; Dolphin, D. A meso-Unsubstituted N-Confused Porphyrin Prepared by Rational Synthesis. Chem. Commun. 1996, 32, 2141–2142. [Google Scholar] [CrossRef]
  31. Lash, T.D.; Richter, D.T.; Shiner, C.M. Conjugated Macrocycles Related to the Porphyrins. Part 16. Synthesis of Hexa- and Heptaalkyl Substituted Inverted or N-Confused Porphyrins by the “3 + 1” Methodology. J. Org. Chem. 1999, 64, 7973. [Google Scholar] [CrossRef]
  32. Lash, T.D.; Von Ruden, A.L. Synthesis and Reactivity of N-methyl and N-phenyl meso-Unsubstituted N-Confused Porphyrins. J. Org. Chem. 2008, 73, 9417–9425. [Google Scholar] [CrossRef]
  33. Lash, T.D. Porphyrin Synthesis by the “3 + 1” Approach: New Applications for an Old Methodology. Chem. Eur. J. 1996, 2, 1197–1200. [Google Scholar] [CrossRef]
  34. Lash, T.D. What’s in a Name? The MacDonald Condensation. J. Porphyr. Phthalocyanines 2016, 20, 855–888. [Google Scholar] [CrossRef]
  35. Lash, T.D.; Hayes, M.J.; Spence, J.D.; Muckey, M.A.; Ferrence, G.M.; Szczepura, L.F. Conjugated Macrocycles Related to the Porphyrins. Part 21. Synthesis, Spectroscopy, Electrochemistry and Structural Characterization of Carbaporphyrins. J. Org. Chem. 2002, 67, 4860–4874. [Google Scholar] [CrossRef] [PubMed]
  36. Lind, L.; LeGoff, E. Unpublished work cited in Sessler, J.L.; Weghorn, S. Expanded, Contracted & Isomeric Porphyrins. Pergamon: New York, NY, USA, 1997; pp. 364–366. [Google Scholar]
  37. Katiaouni, S.; Dechert, S.; Brückner, C.; Meyer, F. A Versatile Building Block for Pyrazole-Pyrrole Hybrid Macrocycles. Chem. Commun. 2007, 43, 951–953. [Google Scholar] [CrossRef]
  38. Katiaouni, S.; Dechert, S.; Briñas, R.P.; Brückner, C.; Meyer, F. Siamese-Twin Porphyrin: A Pyrazole-Based Expanded Porphyrin Providing a Bimetallic Cavity. Angew. Chem. Int. Ed. 2011, 50, 1420–1424. [Google Scholar]
  39. Dorazio, S.J.; Vogel, A.; Dechert, S.; Nevonen, D.E.; Nemykin, V.N.; Brückner, C.; Meyer, F. Siamese-Twin Porphyrin Goes Platinum: Group 10 Momnometallic, Homobimetallic, and Heterobimetallic Complexes. Inorg. Chem. 2020, 59, 7290–7305. [Google Scholar] [CrossRef]
  40. Ishida, M.; Fujimoto, H.; Morimoto, T.; Mori, S.; Togonoh, M.; Shimizu, S.; Furuta, H. Supramolecular Dimeric Structures of Pyrazole-Containing meso-Oxo Carbaphlorin Analogues. Supramol. Chem. 2017, 29, 8–16. [Google Scholar] [CrossRef]
  41. Morimoto, T.; Taniguchi, S.; Osuka, A.; Furuta, H. N-Confused Porphine. Eur. J. Org. Chem. 2005, 2005, 3887–3890. [Google Scholar] [CrossRef]
  42. Wu, J.I.; Fernańdez, I.; Schleyer, P.v.R. Descriptionof Aromaticity in Porphyrinoids. J. Am. Chem. Soc. 2013, 135, 315–321. [Google Scholar] [CrossRef]
  43. Ghosh, A. First-Principles Quantum Chemical Studies of Porphyrins. Acc. Chem. Res. 1998, 31, 189–198. [Google Scholar] [CrossRef]
  44. Ghosh, A.; Larsen, S.; Conradie, J.; Foroutan-Nejad, C. Local versus Global Aromaticity in Azuliporphyrin and Benziporphyrin Derivatives. Org. Biomol. Chem. 2018, 16, 7964–7970. [Google Scholar] [CrossRef] [PubMed]
  45. Nakagami, Y.; Sekine, R.; Aihara, J.-i. The Origin of Global and Macrocyclic Aromaticity in Porphyrinoids. Org. Biomol. Chem. 2012, 10, 5219–5229. [Google Scholar] [CrossRef] [PubMed]
  46. Aihara, J.-i.; Nakagami, Y.; Sekine, R.; Makino, M. Validity and Limitations of the Bridged Annulene Model for Porphyrins. J. Phys. Chem. A 2012, 116, 11718–11730. [Google Scholar] [CrossRef]
  47. Valiev, R.R.; Fleigl, H.; Sundholm, D. Predicting the Degree of Aromaticity of Novel Carbaporphyrinoids. Phys. Chem. Chem. Phys. 2015, 17, 14215–14222. [Google Scholar] [CrossRef] [Green Version]
  48. Valiev, R.R.; Fliegl, H.; Sundholm, D. Aromatic Pathways in Carbathiaporphyrins. J. Phys. Chem. A 2015, 119, 1201–1207. [Google Scholar] [CrossRef] [PubMed]
  49. Alonso, M.; Geerlings, P.; De Proft, F. Exploring the Structure-Aromaticity Relationship in Hückel and Möbius N-Fused Pentaphyrins Using DFT. Phys. Chem. Chem. Phys. 2014, 16, 14396–14407. [Google Scholar] [CrossRef] [PubMed]
  50. Alonso, M.; Geerlings, P.; De Proft, F. Topology Switching in [32]Heptaphyrins Controlled by Solvent, Protonation, and meso Substituents. Chem. Eur. J. 2013, 19, 1617–1628. [Google Scholar] [CrossRef]
  51. AbuSalim, D.I.; Lash, T.D. Relative Stability and Diatropic Character of Carbaporphyrin, Dicarbaporphyrin, Tricarbaporphyrin and Quatyrin Tautomers. J. Org. Chem. 2013, 78, 11535–11548. [Google Scholar] [CrossRef]
  52. AbuSalim, D.I.; Lash, T.D. Aromatic Character and Stability of Neo-Confused Porphyrin Tautomers and Related Compounds. Org. Biomol. Chem. 2013, 11, 8306–8323. [Google Scholar] [CrossRef]
  53. Grabowski, E.Y.; AbuSalim, D.I.; Lash, T.D. Naphtho [2,3-b]carbaporphyrins. J. Org. Chem. 2018, 83, 11825–11838. [Google Scholar] [CrossRef]
  54. Lash, T.D.; AbuSalim, D.I.; Ferrence, G.M. Telluracarbaporphyrins and Related Palladium(II) Complex: Evidence for Hypervalent Interactions. Inorg. Chem. 2021, 60, 9833–9847. [Google Scholar] [CrossRef] [PubMed]
  55. Noboa, M.A.; AbuSalim, D.I.; Lash, T.D. Azulichlorins and Benzocarbachlorins Derived Therefrom. J. Org. Chem. 2019, 84, 11649–11664. [Google Scholar] [CrossRef] [PubMed]
  56. AbuSalim, D.I.; Ferrence, G.M.; Lash, T.D. Synthesis of an adj-Dicarbaporphyrin and the Formation of an Unprecedented Tripalladium Sandwich Complex. J. Am. Chem. Soc. 2014, 136, 6763–6772. [Google Scholar] [CrossRef] [PubMed]
  57. Lash, T.D.; AbuSalim, D.I.; Ferrence, G.M. adj-Dicarbachlorin, the First Example of a Free Base Carbaporphyrinoid System with an Internal Methylene Unit. Chem. Commun. 2015, 51, 15952–15955. [Google Scholar] [CrossRef]
  58. Jain, P.; AbuSalim, D.I.; Lash, T.D. adj-Dicarbaporphyrinoid Systems: Synthesis, Spectroscopic Characterization, and Reactivity of 23-Carbabenziporphyrins. J. Org. Chem. 2019, 84, 10237–10256. [Google Scholar] [CrossRef]
  59. AbuSalim, D.I.; Lash, T.D. Relative Stability of Benziporphyrin and Naphthiporphyrin Tautomers and the Emergence of Macrocyclic Diatropicity. Org. Biomol. Chem. 2014, 12, 8719–8736. [Google Scholar] [CrossRef]
  60. Gao, R.; AbuSalim, D.I.; Lash, T.D. Pyreniporphyrins, Porphyrin Analogues that Incorporate a Polycyclic Aromatic Hydrocarbon Subunit within the Macrocyclic Framework. J. Org. Chem. 2017, 82, 6680–6688. [Google Scholar] [CrossRef]
  61. AbuSalim, D.I.; Lash, T.D. In Pursuit of Novel Porphyrin Isomers. Aromatic Character and Relative Stability of Conjugated Tetrapyrroles with Two Neo-Confused Rings or with Mixed Neo- Confused and N-Confused Subunits. J. Phys. Chem. A 2015, 119, 11440–11453. [Google Scholar] [CrossRef]
  62. AbuSalim, D.I.; Lash, T.D. Tropylium and Porphyrinoid Character in Carbaporphyrinoid Systems. Relative Stability and Aromatic Characteristics of Azuliporphyrin and Tropiporphyrin Tautomers, Protonated Species, and Related Structures. J. Phys. Chem. A 2019, 123, 230–246. [Google Scholar] [CrossRef]
  63. Lash, T.D.; Colby, D.A.; El-Beck, J.A.; AbuSalim, D.I.; Ferrence, G.M. Preparation, Structural Characterization, Assessment of Potential Antiaromaticity and Metalation of 21-Oxyazuliporphyrins. Inorg. Chem. 2015, 54, 9174–9187. [Google Scholar] [CrossRef]
  64. Cramer, E.K.; AbuSalim, D.I.; Lash, T.D. Oxyquinoliziniporphyrins: Introduction of a Heterocyclic Dimension to Carbaporphyrinoid Systems. Org. Lett. 2022, 24, 5402–5406. [Google Scholar] [CrossRef] [PubMed]
  65. Lash, T.D.; Stateman, L.M.; AbuSalim, D.I. Synthesis of Azulitriphyrins [1.2.1] and Related Benzocarbatriphyrins. J. Org. Chem. 2019, 84, 14733–14744. [Google Scholar] [CrossRef] [PubMed]
  66. Singh, A.; Sahoo, D.K.; Sethi, S.K.; Jena, S.; Biswal, H.S. Nature and Strength of the Inner-Core H···H Interactions in Porphyrinoids. ChemPhysChem 2017, 18, 3625–3633. [Google Scholar] [CrossRef] [PubMed]
  67. Toganoh, M.; Furuta, H. Theoretical Study on Conformation and Electronic State of Hückel-Aromatic Multiply N-Confused [26]Hexaphyrins. J. Org. Chem. 2010, 75, 8213–8223. [Google Scholar] [CrossRef] [PubMed]
  68. Alonso, M.; Geerlings, P.; de Proft, F. Viability of Mobius Topologies in [26]- and [28]Hexaphyrins. Chem. Eur. J. 2012, 18, 10916–10928. [Google Scholar] [CrossRef]
  69. Toganoh, M.; Furuta, H. Theoretical Study on the Conformation and Aromaticity of Regular and Singly N-Confused [28]Hexaphyrins. J. Org. Chem. 2013, 78, 9317–9327. [Google Scholar] [CrossRef]
  70. Lash, T.D. Origin of Aromatic Character in Porphyrinoid Systems. J. Porphyr. Phthalocyanines 2011, 15, 1093–1115. [Google Scholar] [CrossRef]
  71. Schleyer, P.v.R.; 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. [Google Scholar] [CrossRef]
  72. Fallah-Bagher-Shaidaei, H.; Wannere, C.S.; Corminboeuf, C.; Puchta, R.; Schleyer, P.v.R. Which NICS Aromaticity Index for Planar π Rings Is Best? Org. Lett. 2006, 8, 863–866. [Google Scholar] [CrossRef]
  73. Corminboeuf, C.; Heine, T.; Seifert, G.; Schleyer, P.v.R.; Weber, J. Induced Magnetic Fields in Aromatic [n]-Annulenes Interpretation of NICS Tensor Components. Phys. Chem. Chem. Phys. 2004, 6, 273–276. [Google Scholar] [CrossRef]
  74. Mills, N.S.; Llagostera, K.B. Summation of Nucleus Independent Chemical Shifts as a Measure of Aromaticity. J. Org. Chem. 2007, 72, 9163–9169. [Google Scholar] [CrossRef] [PubMed]
  75. Sola, M.; Feixas, F.; Jimenéz-Halla, J.O.C.; Matito, E.; Poate, J. A Critical Assessment of the Performance of Magnetic and Electronic Indices of Aromaticity. Symmetry 2010, 2, 1156–1179. [Google Scholar] [CrossRef]
  76. Geuenich, D.; Hess, K.; Köhler, F.; Herges, R. Anisotropy of Induced Current Density (ACID), a General Method to Quantify and Visualize Electronic Delocalization. Chem. Rev. 2005, 105, 3758–3772. [Google Scholar] [CrossRef] [PubMed]
  77. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision, C.01; Gaussian, Inc.: Wallingford, CT, USA, 2019. [Google Scholar]
  78. 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 functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  79. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Schleyer, P.V.R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  80. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  81. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
Figure 1. Selected carbaporphyrinoid systems.
Figure 1. Selected carbaporphyrinoid systems.
Molecules 28 02854 g001
Figure 2. Expanded pyrazole-containing porphyrinoids.
Figure 2. Expanded pyrazole-containing porphyrinoids.
Molecules 28 02854 g002
Scheme 1. MacDonald ‘3 + 1’ syntheses of carbaporphyrins and N-confused porphyrins.
Scheme 1. MacDonald ‘3 + 1’ syntheses of carbaporphyrins and N-confused porphyrins.
Molecules 28 02854 sch001
Scheme 2. Synthesis and reactivity of N-substituted pyrazoloporphyrins.
Scheme 2. Synthesis and reactivity of N-substituted pyrazoloporphyrins.
Molecules 28 02854 sch002
Scheme 3. Attempted syntheses of N-unsubstituted pyrazoloporphyrins.
Scheme 3. Attempted syntheses of N-unsubstituted pyrazoloporphyrins.
Molecules 28 02854 sch003
Figure 3. Proposed aromatic circuits in selected pyrazoloporphyrin tautomers.
Figure 3. Proposed aromatic circuits in selected pyrazoloporphyrin tautomers.
Molecules 28 02854 g003
Figure 4. AICD plots (isovalues 0.05) for PzP-22,24-H (a) and PzP-21,23-H (b).
Figure 4. AICD plots (isovalues 0.05) for PzP-22,24-H (a) and PzP-21,23-H (b).
Molecules 28 02854 g004
Figure 5. Calculated bond lengths and interatomic distances (Å) for the core atoms in pyrazoloporphyrin tautomers PzP-22,24-H (a) and PzP-2,23-H (b).
Figure 5. Calculated bond lengths and interatomic distances (Å) for the core atoms in pyrazoloporphyrin tautomers PzP-22,24-H (a) and PzP-2,23-H (b).
Molecules 28 02854 g005
Figure 6. AICD plots (isovalues 0.05) for PzP-2,22,24-H+ (a), PzP-22,23,24-H+ (b), PzP-2,21,23-H+ (c), and PzP-21,22,24-H+ (d).
Figure 6. AICD plots (isovalues 0.05) for PzP-2,22,24-H+ (a), PzP-22,23,24-H+ (b), PzP-2,21,23-H+ (c), and PzP-21,22,24-H+ (d).
Molecules 28 02854 g006
Figure 7. Calculated bond lengths and interatomic distances for the core atoms for pyrazoloporphyrin monocation PzP-2,22,24-H+ (a) and dication [PzP-2,22,23,24-H]2+ (b).
Figure 7. Calculated bond lengths and interatomic distances for the core atoms for pyrazoloporphyrin monocation PzP-2,22,24-H+ (a) and dication [PzP-2,22,23,24-H]2+ (b).
Molecules 28 02854 g007
Figure 8. Proposed aromatic circuits in selected pyrazoloporphyrin cations.
Figure 8. Proposed aromatic circuits in selected pyrazoloporphyrin cations.
Molecules 28 02854 g008
Table 1. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Continuously Conjugated Free Base Unsubstituted Pyrazoloporphyrin Tautomers.
Table 1. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Continuously Conjugated Free Base Unsubstituted Pyrazoloporphyrin Tautomers.
Molecules 28 02854 i001
PzP-22,24-H
Molecules 28 02854 i002
PzP-2,23-H
Molecules 28 02854 i003
PzP-22,23-H
Molecules 28 02854 i004
PzP-2,22-H
Molecules 28 02854 i005
PzP-2,24-H
Molecules 28 02854 i006
PzP-21,22-H
Molecules 28 02854 i007
PzP-21,23-H
ΔG298 (kcal/mol)
ΔE (kcal/mol)
0.00−0.256.207.246.4216.219.17
0.000.376.677.887.0818.3211.27
NICS(0)
NICS(01)zz
−13.05−1.76−11.88−2.49−2.20−11.96−11.08
−30.62−2.88−28.46−5.85−4.87−29.02−27.24
NICS(a)
NICS(a1)zz
+1.26−8.96+03.59−9.87−9.69−15.12−13.57
+2.12−19.84+04.64−23.46−22.70−49.64−46.92
NICS(b) NICS(b1)zz−11.73−0.04−2.83+0.65−4.05+1.24−0.76
−31.37−5.45−13.35−6.40−10.74−5.94−10.11
NICS(c) NICS(c1)zz−2.03−3.97−10.56+1.02+0.40−2.24−12.94
−11.73−10.70−25.72−5.05−6.39−13.27−33.60
NICS(d) NICS(d1)zz−11.72−0.82−13.19−4.33−0.43−12.70−0.76
−31.34−7.07−37.48−11.71−8.75−33.32−10.11
Table 2. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Pyrazoloporphyrin Tautomers with Methylene Bridges.
Table 2. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Pyrazoloporphyrin Tautomers with Methylene Bridges.
Molecules 28 02854 i008
PzP-2,20-H
Molecules 28 02854 i009
PzP-2,15-H
Molecules 28 02854 i010
PzP-2,10-H
Molecules 28 02854 i011
PzP-2,5-H
Molecules 28 02854 i012
PzP-5,22-H
ΔG298 (kcal/mol)
ΔE (kcal/mol)
28.1427.3827.5628.4327.27
32.1531.4931.6732.7630.13
NICS(0)
NICS(01)zz
+0.72+1.09+1.09+0.53−0.55
+5.25+5.76+5.72+4.52+1.87
NICS(a)
NICS(a1)zz
−10.05−9.11−9.01−9.72+1.94
−24.53−22.74−22.69−24.66−7.33
NICS(b)
NICS(b1)zz
+1.50+0.34−0.48−1.01−1.04
−4.15−5.95−7.19−8.46−9.17
NICS(c)
NICS(c1)zz
−0.06+0.97+1.41+0.50−3.85
−7.02−5.31−4.36−5.85−15.13
NICS(d)
NICS(d1)zz
−1.73−1.18−0.36+2.40−9.23
−10.05−8.76−7.55−2.36−23.78
Molecules 28 02854 i013
PzP-5,23-H
Molecules 28 02854 i014
PzP-5,24-H
Molecules 28 02854 i015
PzP-10,22-H
Molecules 28 02854 i016
PzP-10,23-H
Molecules 28 02854 i017
PzP-10,24-H
ΔG298 (kcal/mol)
ΔE (kcal/mol)
19.9323.2324.7821.0622.96
22.2825.7528.0324.2125.85
NICS(0)
NICS(01)zz
−0.42+0.16−0.08+0.36−0.09
+0.64+2.89+0.74+1.84+1.92
NICS(a)
NICS(a1)zz
+3.14+1.24−0.06+2.34+1.14
−4.64−8.70−11.70−7.05−8.67
NICS(b)
NICS(b1)zz
−2.24−9.41−1.89−3.85−9.65
−11.55−22.39−11.35−15.25−22.94
NICS(c)
NICS(c1)zz
−9.64−1.76−1.15−9.31−0.72
−23.17−10.12−9.94−21.91−8.50
NICS(d)
NICS(d1)zz
−0.47+01.05−9.71−1.63−0.55
−8.40−5.19−23.77−11.44−7.54
Table 3. Calculated dihedral angles for the individual pyrrole and pyrazole rings within a series of pyrazoloporphyrin tautomers and related protonated species.
Table 3. Calculated dihedral angles for the individual pyrrole and pyrazole rings within a series of pyrazoloporphyrin tautomers and related protonated species.
MoleculeabbccddaAverage
PzP-22,24-H13.43−1.411.39−13.417.41
PzP-22,23-H−10.36−0.04−6.4816.138.25
PzP-2,23-H9.940.95−1.54−9.735.54
PzP-2,22-H−7.75−1.53−0.6710.155.03
PzP-2,24-H10.93−0.08−2.37−8.095.37
PzP-2,20-H0.000.000.000.000.00
PzP-2,15-H0.010.000.000.000.00
PzP-2,10-H0.000.000.010.000.00
PzP-2,5-H−0.010.01−0.010.130.04
PzP-5,22-H−12.701.67−4.7120.639.93
PzP-5,23-H0.030.00−0.01−0.030.02
PzP-5,24-H11.87−1.58−2.74−10.126.58
PzP-10,22-H−7.460.16−2.488.854.74
PzP-10,23-H−10.59−3.183.7712.837.59
PzP-10,24-H−9.691.950.878.905.35
PzP-21,22-H0.000.000.000.000.00
PzP-21,23-H0.000.01−0.01−0.010.01
[PzP-22,23,24-H]+19.03−13.8113.83−19.0516.43
[PzP-2,22,24-H]+−18.702.12−1.6918.1510.17
[PzP-2,22,23-H]+15.98−0.817.48−20.9911.31
[PzP-2,22,24-H]+21.07−7.640.32−15.3711.10
[PzP-21,22,24-H]+0.02−0.020.19−0.210.11
[PzP-21,22,23-H]+−1.79−2.6910.82−8.065.84
[PzP-2,21,22-H]+0.000.000.000.000.00
[PzP-2,21,23-H]+0.01−0.020.03−0.020.02
[PzP-2,21,24-H]+0.000.000.000.000.00
[PzP-2,22,23,24-H]2+−23.2515.19−14.9622.8519.06
[PzP-2,21,22,24-H]2+−0.010.01−0.010.010.01
[PzP-2,21,23,24-H]2+−7.1012.00−4.20−0.936.06
[PzP-2,21,22,23-H]2+1.563.13−11.826.835.84
Table 4. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Monoprotonated Pyrazoloporphyrin Cations.
Table 4. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Monoprotonated Pyrazoloporphyrin Cations.
Molecules 28 02854 i018
[PzP-2,22,24-H]+
Molecules 28 02854 i019
[PzP-22,23,24-H]+
Molecules 28 02854 i020
[PzP-2,22,23-H]+
Molecules 28 02854 i021
[PzP-2,23,24-H]+
Molecules 28 02854 i022
[PzP-21,22,24-H]+
ΔG298 (kcal/mol)
ΔE (kcal/mol)
0.0016.703.464.07−19.01
0.0016.973.654.2920.89
NICS(0)
NICS(01)zz
−4.50−12.54−2.72−2.99−15.01
−8.55−29.17−4.52−6.91−36.42
NICS(a)
NICS(a1)zz
−11.38+3.39−9.79−9.40−19.92
−19.15+6.46−18.49−20.79−60.83
NICS(b)
NICS(b1)zz
−5.09−12.81−0.22−5.64−11.30
−14.15−39.27−5.89−13.45−31.17
NICS(c)
NICS(c1)zz
−2.43−12.02−4.43−4.99−0.13
−0.95−23.71−11.56−14.65−13.29
NICS(d)
NICS(d1)zz
−5.97−12.80−6.15−1.30−11.30
−16.03−39.26−17.63−10.77−31.16
Molecules 28 02854 i023
[PzP-21,22,23-H]+
Molecules 28 02854 i024
[PzP-2,21,22-H]+
Molecules 28 02854 i025
[PzP-2,21,23-H]+
Molecules 28 02854 i026
[PzP-2,21,24-H]+
ΔG298 (kcal/mol)
ΔE(kcal/mol)
23.9423.9614.9022.71
25.8426.5217.2924.87
NICS(0)
NICS(01)zz
−12.55−12.40−10.92−8.67
−29.99−30.19−26.83−19.98
NICS(a)
NICS(a1)zz
−16.02−21.54−19.10−15.79
−51.77−52.96−47.98−39.29
NICS(b)
NICS(b1)zz
+1.65+6.16+4.60−6.04
+0.93+5.20+2.01−17.53
NICS(c)
NICS(c1)zz
−12.60+3.49−7.56+0.90
−35.25−0.98−22.10−6.01
NICS(d)
NICS(d1)zz
−13.27−8.50+1.51+2.61
−25.94−24.18−4.54−1.77
Table 5. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Diprotonated Pyrazoloporphyrin Dications.
Table 5. Calculated Relative Energies (kcal/mol) and NICS Values (ppm) for Diprotonated Pyrazoloporphyrin Dications.
Molecules 28 02854 i027
[PzP-2,22,23,24-H]2+
Molecules 28 02854 i028
[PzP-2,21,22,24-H]2+
Molecules 28 02854 i029
[PzP-2,21,23,24-H]2+
Molecules 28 02854 i030
[PzP-2,21,22,23-H]2+
ΔG298 (kcal/mol)
ΔE (kcal/mol)
0.0013.2216.5316.04
0.0015.1318.5318.16
NICS(0)
NICS(01)zz
−4.14−11.63−10.59−12.24
−7.49−27.10−24.93−29.07
NICS(a)
NICS(a1)zz
−9.29−20.24−18.67−21.04
−16.86−49.56−46.46−51.64
NICS(b)
NICS(b1)zz
−6.32−6.31−7.68+6.36
−20.75−19.11−14.35+6.24
NICS(c)
NICS(c1)zz
−4.96+2.65−8.50−7.72
−10.04−2.31−27.00−25.27
NICS(d)
NICS(d1)zz
−7.20−7.85+2.19−9.67
−22.55−22.77−2.23−18.57
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

AbuSalim, D.I.; Lash, T.D. Aromatic Character and Relative Stability of Pyrazoloporphyrin Tautomers and Related Protonated Species: Insights into How Pyrazole Changes the Properties of Carbaporphyrinoid Systems. Molecules 2023, 28, 2854. https://doi.org/10.3390/molecules28062854

AMA Style

AbuSalim DI, Lash TD. Aromatic Character and Relative Stability of Pyrazoloporphyrin Tautomers and Related Protonated Species: Insights into How Pyrazole Changes the Properties of Carbaporphyrinoid Systems. Molecules. 2023; 28(6):2854. https://doi.org/10.3390/molecules28062854

Chicago/Turabian Style

AbuSalim, Deyaa I., and Timothy D. Lash. 2023. "Aromatic Character and Relative Stability of Pyrazoloporphyrin Tautomers and Related Protonated Species: Insights into How Pyrazole Changes the Properties of Carbaporphyrinoid Systems" Molecules 28, no. 6: 2854. https://doi.org/10.3390/molecules28062854

Article Metrics

Back to TopTop