Next Article in Journal
Acid Catalyzed N-Alkylation of Pyrazoles with Trichloroacetimidates
Next Article in Special Issue
A Molecular Electron Density Theory Study of the [3+2] Cycloaddition Reaction of Pseudo(mono)radical Azomethine Ylides with Phenyl Vinyl Sulphone
Previous Article in Journal
A Theoretical Study on the Photochemical Isomerization of 2,6-Dimethylpyrazine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diels–Alder Polar Reactions of Azaheterocycles: A Theoretical and Experimental Study

by
Carla M. Ormachea
,
María Nélida Kneeteman
and
Pedro M. E. Mancini
*
Laboratorio Fester–Química Orgánica (FIQ), Group of Organic Synthesis and Materials (GSOM), Instituto de Química Aplicada del Litoral (IQAL) (UNL-CONICET), Universidad Nacional del Litoral, Santa Fe 3000, Argentina
*
Author to whom correspondence should be addressed.
Submission received: 27 February 2022 / Revised: 23 April 2022 / Accepted: 9 May 2022 / Published: 22 May 2022
(This article belongs to the Special Issue Chemistry of Heterocycles)

Abstract

:
A number of azaheterocycles (pyridines, pyrroles and indoles) have been properly functionalized so that they can act as dienophiles in cycloaddition Diels–Alder processes. This work analyzed the reactive behavior of these molecules through mechanistic analysis and the regioselectivity of the process using computational calculation tools. Based on this knowledge, a study was conducted on the influences of reaction variables, in particular solvent, catalyst and microwave irradiation, to achieve favorable changes—shorter reaction times, more acceptable temperatures and better yields. Theoretical calculations allowed the development of predictive approaches, which were later experimentally corroborated. This analysis allowed us to make reasonable assumptions related to reaction mechanisms, which allowed—through the analysis of corresponding transition states—us to consider such reactions at the boundary between pericyclic and polar processes.

1. Introduction

Since Diels and Alder, in 1928, observed the formation of a 1:1 adduct in reactions between a series of dienes and certain alkenes, including cyclopentadiene and p-benzoquinone, this process has become one of the synthetic methodologies most useful for the formation of carbon–carbon, carbon–heteroatom and heteroatom–heteroatom bonds contained in cyclic products.
The D-A reaction is of great interest in synthesis, not only as a consequence of its versatility but also because of its high regioselectivity, endo-stereoselectivity and cis-stereospecificity, with the retention of a configuration normally observed in the diene and in the dienophile and the exclusive, or at least preponderant, formation of one isomer among several possible ones.
Five-membered aromatic heterocycles, as well as their benzofused analogs, have also been traditionally studied as dienes; however, in recent years, interest in them as dienophiles has increased. In the 1980s, E. Wenkert [1] studied the behavior of α- and β-acylsubstituted furans in thermal Diels–Alder reactions with normal electron demand, showing the possibility that an aromatic system substituted by an electron-withdrawing group can act as a dienophile in such a reaction.
The reactive behavior of a series of azaheterocycles (pyridines, pyrroles and indoles) properly substituted was experimentally and computationally studied in recent years by our group [2,3,4,5,6,7,8,9,10]. These heterocycles can act as dienophiles in processes of DA cycloaddition. To explain reactive behaviors, their electrophilic character has been analyzed. However, the regioselectivity of these processes has been determined through a mechanistic analysis. The aim of the present work is to reconcile different aspects of experimental chemistry with the principles of theoretical chemistry. In particular, studying the influences of reaction variables, solvent and microwave irradiation contributes to the progress of knowledge about these reactions.
Theoretical studies allowed us to make reasonable assumptions related to reaction mechanisms, which allowed—via the analysis of corresponding transition states (TSs)—us to conceive these types of reactions at the boundary between pericyclic and polar processes. From the certainty provided by the experimental work, two models of nitroheteroaromatics electrophiles were discussed: pentaheterocycles and hexaheterocycles.

2. Materials and Methods

2.1. Experimental Section

DA cycloaddition reactions of azaheterocycles were carried out using different techniques to analyze the solvent effect (conventional solvent vs. ionic liquids), heating method (thermal or microwave) and reaction time.
Thermal reactions using molecular solvents: Thermal reactions were carried out in closed ampoules. Different temperature ranges and various reaction times were used. These reactions were considered as references. Toluene was used as the solvent.
Thermal reactions using ionic liquids: in order to improve the experimental conditions verified under thermal conditions, different types of ionic liquids (ethylammonium nitrate, 1-methylimidazolium tetrafluoroborate, 1-methylimidazolium hexafluorophosphate, 1-n-butyl-3-methylimidazolium tetrafluoroborate and 1-n-butyl-3-methylimidazolium hexafluorophosphate) were used as reaction media.
Reactions applying the microwave irradiation technique: Reactions were developed in a microwave reactor—Monowave-300—using different temperature (or power) ranges and reaction times. The reactions were developed in the presence of classic organic solvents, using protic ionic liquids, and finally in the absence of a solvent. As such, it was possible to analyze the influence of microwave heating on these types of reactions and the degree of response to the different solvents.

2.1.1. Thermal Reactions

The reactions were performed in 5 mL glass ampoules containing a solution of 1 mmol of dienophile and an excess of diene (the diene/dienophile molar ratios being 12:1 for isoprene (the excess used in the case of isoprene is higher since it tends to polymerize) and 3:1 for the other dienes), using 1 mL of toluene as the reference solvent or, alternatively, the corresponding ionic liquid (IL). The solution was heated at 120 °C for 72 h for molecular solvents and 60 °C for 24 h when using IL. The reactions were monitored using thin-layer chromatography (TLC) carried out with 254 nm UV silica gel plates. Purification of the reaction crude and isolation of the different products were carried out using column chromatography, utilizing neutral alumina as the stationary phase and a hexane-ethyl acetate mixture as the eluent. Prior to purification, the reaction crudes were treated. In the case of the cycloadditions carried out in toluene, the solvent was removed using a rotary evaporator, and in the case of the crude reaction corresponding to the ionic liquids, an extraction was carried out with a molecular solvent.

2.1.2. Microwave Reactions

The reactions were developed in a single-mode-type microwave reactor in 10 mL glass vials, adding 1 mmol of the dienophile and different amounts of diene, using 1 mL of toluene as the solvent or, alternatively, the corresponding IL in the same amount. In this case, reactions were also developed in solvent-free conditions. The reactions were monitored using thin-layer chromatography (TLC) carried out with 254 nm UV silica gel plates. At the end of the reaction time, the reaction crude was separated and purified using column chromatography in the same way as in the thermal reactions. The reactions were performed at 180 °C for 30 min. The diene/dienophile ratio used was the same as that in the thermal reactions.

2.2. Computational Details

All the DFT [11] calculations were made using B3LYP functionals [12], with the standard 6-311+G(d,p) basis set [13,14].
The geometry optimizations were made using the Berny analytical gradient optimization method. Frequency computations were used to characterize the stationary points, verifying that the TSs only had one imaginary frequency. In order to determine the energy profiles, the IRC paths were searched [15] by connecting each TS to the two associated minima (reactants and products) of the proposed mechanism using the second-order González–Schlegel [16] integration method. The solvent effects of toluene were considered using the polarizable continuum model (PCM) from Tomasi’s group in the framework of the self-consistent reaction field (SCRF) [17] utilizing a single point energy calculation of the geometries previously optimized in the gas phase. The stationary point electronic structures were analyzed using the natural bond orbital (NBO) method [18]. All the calculations were carried out with the Gaussian 09 suite of programs [19].
The global electrophilicity index [20], ω, is defined by the expression ω = (µ2/2η), µ being the electronic chemical potential and η the chemical hardness. Both values may be approximate in terms of the one-electron energies of the frontier molecular orbital HOMO (εH) and LUMO (εL) as µ ≈ (εH + εL)/2 and η ≈ (εL − εH) [21]. Later, an empirical (relative) nucleophilicity index [22], N, was introduced, which is based on the HOMO energies calculated within the Kohn–Sham scheme and defined as N = EHOMO (Nu) − EHOMO (TCE). The nucleophilicity index is referred to as tetracyanoethylene (TCE) due to the fact that it presents the lowest HOMO energy of a considerable series of molecules investigated in polar cycloadditions. As such, one can conveniently work with a scale of positive nucleophilic values. Electrophilic Pk+ and nucleophilic Pk Parr functions [23] were obtained with the analysis of the Mulliken atomic spin density (ASD) of the radical anion and radial cation of the reagents.

3. Results and Discussion

3.1. Experimental Results

To explore the behaviors of azaheterocycles as electrophilic dienophiles in D-A reactions, isoprene, 1-trimethylsilyloxy-1,3-butadiene and 1-methoxy-3-trimethylsilyloxy-1,3-butadiene (Danishefsky’s diene) were selected as nucleophilic dienes. This selection was chosen to be able to compare the influences of ILs and molecular solvents on the reactivity and the regioselectivity of the P-D-A reactions. Figure 1 shows all the products of the Diels–Alder reactions carried out using the different techniques mentioned above. Table 1 details the yields obtained in each case.
Protic ionic liquids, with their specific properties, high polarizability and ability to form hydrogen bonds, first produce a drastic reduction in time and an increase in reaction yields. This acceleration cannot be explained based on a single theory; however, hydrophobic interactions and Lewis acid catalysis may be useful concepts for understanding the effect of these salts on the D-A reaction [24].
The yields obtained in each case are detailed in Table 1. The products a and b of each reaction were considered as a 50/50 mixture of their isomers (substitution of the methyl group in C5 and C6 for products I and II derived from pyrrole; C2 and C3 for products III and IV derived from indole; and C6 and C7 for product VI derived from pyridine). The yields were considered based on the consumption of dienophiles and were calculated by weighing them. In order to be able to easily compare the different reaction conditions, the percentage corresponding to IL was calculated as an average of the yields obtained with [HMIM] [BF4], [HMIM] [PF6] and NEA.

3.2. Computational Results

3.2.1. Electron Demand

For the reactivity study, the calculation of the global electrophilicity values ω was performed in terms of the electronic chemical potential μ and the chemical hardness η (Table 2). As such, whether the proposed dienophiles behave as electrophiles and the maximum number of electrons that they can accept can be evaluated. The values calculated taking into account the reaction media modify the values but not the tendency of them (see Supplementary Materials). In Table 2, only the gas-phase values are considered.
In the same way, to determine the reactivity of the dienes, their global reactivity indices were calculated based on the energies of the frontier molecular orbitals (Table 3).
When compared with the chemical potential of the dienophiles, the dienes present a larger value, for which the charge transfer is from the diene to the dienophile. These reactions, controlled by the HOMO orbitals of the diene and the LUMO of the dienophile, are located within those deemed to be “with normal electron demand”.

3.2.2. Stereochemistry

To analyze the stereochemistry of the reaction, the local indices were calculated. Considering that the most favorable charge transfer is from the most nucleophilic to the most electrophilic center, the bond formed between this to atoms is the one that determines the isomer obtained.
Local indices were determined using the Fukui function. It must be taken into account that the values of the local indices are numerically comparable for the different atoms within the same molecule since they are relative to each system. For the dienophiles (Table 4), the beta carbon to the nitro group turns out to be the most electrophilic in all cases, and it will be the one that reacts with the most nucleophilic center of the diene. Regarding local properties, the nitro group, in addition to providing the molecule with the necessary electrophilic character, becomes responsible for the regiochemistry of the reaction.
With the diene being the nucleophilic counterpart of the process, the local nucleophilicity indices in C1 and C4 and the atoms involved in the new cycloadduct were calculated for it (Table 5). For isoprene, the most nucleophilic center corresponds to C1. However, the values of the nucleophilicity of carbons 1 and 4 do not present a significant difference, which explains why isomers are obtained in comparable proportions in the experimental work. For 1-trimethylsilyloxy-1,3-butadiene, although the difference in local indices is more marked than in the case of isoprene, in the reactions detailed below, it was observed that there is only one product; this is due to the fact that the substituent in position 1 has a tendency to be eliminated in the aromatization process, for which it is not possible to determine whether both isomers or only one of them are formed in the first stage, or in what proportion. The Danishefsky diene would be located as a marginal electrophile, being the most nucleophilic of the series. The greatest difference in nucleophilicity observed is for the Danishefsky diene (0.91 eV), a value sufficient for only one of the isomers to be the product of cycloaddition. In each reactive pair, the favored isomer is the one corresponding to the C4 bond of the Danishefsky diene and C adjacent to the nitro group of each dienophile, which agrees with the experimental results.

3.2.3. Mechanism Analysis

The analysis of the synchronous or asynchronous character of the reaction can be carried out by relating the values of the bond formation distances of the atoms involved in the transition state; that is, Δr = r1 − r2, where r1 and r2 are the distances from C2 and C3 of the dienophile to C1 and C4 of the diene, always taking r1 as the greatest distance. However, the energy of the reaction was evaluated through the difference in energy (ΔE) of the TSs (activation energy) and of the cycloadduct/s (CAs) (enthalpy of reaction) with respect to the sum of the energy of the reagents (see Supplementary Materials).
The activation energy (in the gas phase) associated with the nucleophilic attack of the diene is higher than the one involving the effect of IL. The polar nature of D-A reactions was assessed by analyzing charge transfer (CT) in the TS. In the TSs, the CT flowing from the diene to the dienophile is higher when ILs are considered. These values indicate the zwitterionic character of the TSs and explain why the reactions that involve IL achieve similar yields in shorter times. The polar character of the reaction increases because of the hydrogen bond formed, which improves the CT process and increases the asynchronous character of the reactions.
To study the effects of microwave irradiation on polar Diels–Alder processes, the displacement vectors of the IR frequency vibrations were analyzed. This values shows that there are stretching vibrations with similar atomic motions to the electronic changes implicated in the formation of the TSs.
The stretching vibrations of the dienophiles and the systems dienophile+diene are between 1300 and 1400 cm−1. This indicates that any extra energy between 1700 and 1400 cm−1 can favor the process by improving the electronic changes that are necessary to obtain TS geometry. Experimentally, this is observed as a reduction in the reaction time. It should be noted that the values of the stretching vibrations between the TS and the TS-IL are close for the studied systems, which means that the microwave irradiation effect should be equivalent in all cases. This agrees with the experimental results, considering that microwave irradiation significantly reduces reaction times and improves yields, although no significant difference is observed between the type of solvent chosen.

4. Conclusions

Throughout this research, a series of D-A reactions were analyzed, both from an experimental point of view and a theoretical point of view.
The theoretical analysis, based on the DFT theory, in combination with a series of complementary experiences, contributes to a better degree of understanding of these particular types of cycloaddition reactions.
Knowing that the theoretical studies are consistent with the experimental results, we were able to develop predictive approaches, reducing the experimental expense of chemical tests. However, these studies made it possible to understand the mechanism of the reactions evaluated, which allowed, via the corresponding proposed transition states, us to conceive this type of reaction at the boundary between pericyclic and polar processes.
It was experimentally feasible to discuss two models of nitroheteroaromatic electrophiles in the function types of the rings—pentaheterocycles (and benzofused-pentaheterocycles) vs. hexaheterocycles—and the consequences of the type of structure on reactivity and regioselectivity.
It was possible to demonstrate that the indices of the chosen reactivity are good indicators of the feasibility of the reaction, and that its derived local indices are very useful in determining the region and stereochemistry of each cycloaddition. Regarding the solvent effect, although the continuous dielectric method is the most used, in the case of PILs, alternative methods of supermolecules should be considered since the reactions are affected by hydrogen bonding interactions and the effect of polarization.
An analysis of the reaction mechanisms made it possible to identify cycloadducts and the corresponding transition states, which allowed us to determine the energies of activation of each process. In all cases, transition states were obtained. Within the mechanistic studies, we observed that the charge transfer was greater when ILs were involved than in solvent-free conditions. The acceleration of reaction times can be understood as an increase in the polar character of the reaction due to the hydrogen bond.
The effect of microwave irradiation seems to be enough in itself to speed up reactive processes, regardless of the solvent used.
It is likely that the main contribution is the theoretical detection of a type of polar Diels–Alder reaction that presents a reaction intermediate, which is significant for this type of process.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/org3020008/s1. File S1: Computational calculations including solvent effect and Mechanistic calculations.

Author Contributions

C.M.O.: formal analysis, data curation and investigation; M.N.K.: conceptualization, resources and project administration; and P.M.E.M.: methodology, writing—review and editing, and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Agencia Nacional de Ciencia y Tecnología (ANCyT) of Argentina, PICT 2014 No. 1587, and by CAI+D 2011, No. 66, 501 201101 00478 LI, at the Universidad Nacional del Litoral, Santa Fe, Argentina. C. Ormachea gratefully acknowledges the Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET) of Argentina for the scholarship.

Data Availability Statement

Not applicable.

Acknowledgments

This research was supported by Argentine Agency of Science and Technology (ANCyT)-PICT 2008 Nº 1214 and by CAI+D’s 2009-12/Q271 and 12/Q272 at Universidad Nacional del Litoral, Santa Fe, Argentina.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wenkert, E.; Moeller, P.D.; Piettre, S.R. Five-membered aromatic heterocycles as dienophiles in Diels-Alder reactions. Furan, pyrrole, and indole. J. Am. Chem. Soc. 1988, 110, 7188–7194. [Google Scholar] [CrossRef]
  2. Della Rosa, C.; Kneeteman, M.; Mancini, P.M.E. Comparison of the reactivity between 2- and 3-nitropyrroles in cycloaddition reactions. A simple indole synthesis. Tetrahedron Lett. 2007, 48, 1435–1438. [Google Scholar] [CrossRef]
  3. Biolatto, B.; Kneeteman, M.; Paredes, E.; Mancini, P.M.E. Reactions of 1-Tosyl-3-substituted Indoles with Conjugated Dienes under Thermal and/or hyperbaric conditions. J. Org. Chem. 2001, 66, 3906–3912. [Google Scholar] [CrossRef] [PubMed]
  4. Mancini, P.M.E.; Kneeteman, M.N.; Cainelli, M.; Ormachea, C.M. Cycloaddition Reactions Assisted by Microwave Irradiation: Protic Ionic liquids vs Solvent-free Conditions. Curr. Microw. Chem. 2017, 4, 219–228. [Google Scholar] [CrossRef]
  5. Mancini, P.M.E.; Kneeteman, M.N.; Cainelli, M.; Ormachea, C.M. Nitropyrroles, Diels-Alder reactions assisted by microwave irradiation and solvent effect. An experimental and theoretical study. J. Mol. Struct. 2017, 1147, 155–160. [Google Scholar] [CrossRef]
  6. Ormachea, C.M.; Lopez Baena, A.F.; Della Rosa, C.; Kneeteman, M.N.; Mancini, P.M.E. N-tosyl-nitropyrroles as Dienophiles in Polar Cycloaddition Reactions Developed in Protic Ionic Liquids. J. Chem. Chem. Eng. 2012, 6, 661–667. [Google Scholar]
  7. Mancini, P.M.E.; Ormachea, C.M.; Della Rosa, C.; Kneeteman, M.N.; Suarez, A.G.; Domingo, L.R. Ionic liquids and microwave irradiation as synergistic combination for polar Diels–Alder reactions using properly substituted heterocycles as dienophiles. A DFT study related. Tetrahedron Lett. 2012, 53, 6508–6511. [Google Scholar] [CrossRef]
  8. Mancini, P.M.E.; Kneeteman, M.N.; Lopez Baena, A.F.; Della Rosa, C.D.; Gamboa, G. Microwave Irradiation: Polar Diels-Alder Reactions Using Nitropyrrole and Nitroindole Derivates as Electrophiles. In Proceedings of the 18th International Electronic Conference on Synthetic Organic Chemistry, Santa Fe, Argentina, 1–30 November 2014; Volume 18. [Google Scholar]
  9. Ormachea, C.M. Indolic and Pyridine Derivatives in Polar Diels-Alder Reactions. A Theoretical-Experimental Study. Ph.D. Thesis, Universidad Nacional del Litoral, Santa Fe, Argentina, 2018. Available online: https://hdl.handle.net/11185/1096 (accessed on 10 November 2021).
  10. Della Rosa, C.; Ormachea, C.M.; Sonzogni, A.S.; Kneeteman, M.N.; Domingo, L.R.; Mancini, P.M.E. Polar Diels-Alder reactions developed in a Protic Ionic Liquid: 3-Nitroindole as dienophile. Theoretical Studies Using Dft Methods. Lett. Org. Chem. 2012, 9, 691–695. [Google Scholar] [CrossRef]
  11. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef] [Green Version]
  12. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B Condens. Matter Mater. Phys. 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  14. 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]
  15. Fukui, K. Formulation of the reaction coordinate. J. Phys. Chem. 1970, 74, 4161–4163. [Google Scholar] [CrossRef]
  16. González, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  17. Cances, E.; Mennucci, B.; Tomasi, J. A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics. J. Chem. Phys. 1997, 107, 3032–3041. [Google Scholar] [CrossRef]
  18. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural population analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  19. Frisch, 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 09, Revision A.02; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  20. Parr, R.G.; Szentpaly, V.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  21. Parr, R.G.; Yang, W. Density Functional Theory of Atoms and Molecules; Oxford University Press: New York, NY, USA, 1989. [Google Scholar]
  22. Domingo, L.R.; Pérez, P. The nucleophilicity N index in organic chemistry. Org. Biomol. Chem. 2011, 9, 7168–7175. [Google Scholar] [CrossRef] [PubMed]
  23. Domingo, L.R.; Pérez, P.; Sáez, J.A. Understanding the local reactivity in polar organic reactions through electrophilic and nucleophilic Parr functions. RSC Adv. 2013, 3, 1486–1494. [Google Scholar] [CrossRef]
  24. Domingo, L.R.; Aurell, M.; Kneeteman, M.N.; Mancini, P.M.E. Mechanistic details of the domino reaction of nitronaphthalenes with the electron-rich dienes. A DFT study. Theochem 2008, 853, 68–76. [Google Scholar] [CrossRef]
Figure 1. Reaction products of DA cycloadditions between azaheterocycles dienophiles and the different dienes.
Figure 1. Reaction products of DA cycloadditions between azaheterocycles dienophiles and the different dienes.
Organics 03 00008 g001
Table 1. Yields obtained from the reaction products of the DA cycloadditions.
Table 1. Yields obtained from the reaction products of the DA cycloadditions.
ReactionCondition
Thermal/MSThermal/ILMW/MSMW/ILMW/SF
I50% a + 10% b45% a + 10% b30% a55% a60% a
II50% a + 10% b50% a + 10% b24% a56% a59% a
III36% a + 8% b50% a + 2% b50% a60% a62% a
IV32% a + 8% b50% a + 2% b53% a60% a63% a
V43%56%57%60%58%
VI15%25%30%35%35%
VII45%55%51%67%68%
VIII45%54%42%66%63%
IX45%56%61%75%73%
X40%65%61%75%72%
XI-----
XII20%25%23%25%30%
XIII48%58%63%77%76%
XIV48%58%55%72%72%
XV44%74%68%78%77%
XVI44%72%72%80%78%
XVII26%30%32%31%32%
XVIII41%50%52%50%54%
MSMS: molecular solvent, IL: ionic liquid, SF: solvent free.
Table 2. Global properties calculated using B3LYP method and 6-311+G(d,p) level of theory.
Table 2. Global properties calculated using B3LYP method and 6-311+G(d,p) level of theory.
DienophileεHOMOεLUMOμηωN
N-tosylpyrrol−6.63−1.96−4.304.681.972.49
N-tosyl-2-nitropyrrol−7.27−2.81−5.044.472.842.22
N-tosyl-3-nitropyrrol−7.32−2.66−4.994.662.682.17
N-tosylindol−6.08−1.79−3.934.281.813.05
N-tosyl-2-nitroindol−6.74−2.69−4.714.052.742.39
N-tosyl-3-nitroindol−6.67−2.52−4.814.122.812.26
3-nitropyridine−7.68−2.76−5.224.912.771.44
4-nitropiridine-N-oxide−7.25−3.53−5.393.713.912.24
All the values are expressed in eV and correspond to the gas phase. Bolding is because both columns are calculated using the values of the previous ones.
Table 3. Global properties calculated using B3LYP method and 6-311+G(d,p) level of theory.
Table 3. Global properties calculated using B3LYP method and 6-311+G(d,p) level of theory.
DieneεHOMOεLUMOμηωN
Isoprene−6.57−0.98−3.775.591.272.92
1-trimetylsililoxy-1,3-butadiene−5.90−0.72−3.315.181.063.58
Danishefsky’s diene−5.91−0.52−3.225.390.963.57
All the values are expressed in eV and correspond to the gas phase. Bolding is because both columns are calculated using the values of the previous ones
Table 4. Local electrophilicity indices for dienophiles.
Table 4. Local electrophilicity indices for dienophiles.
Dienophileωk (eV)
C2C3
N-tosyl-2-nitropyrrol0.030.19
N-tosyl-3-nitropyrrol0.120.04
N-tosyl-2-nitroindol0.050.36
N-tosyl-3-nitroindol0.480.07
C3C4
3-nitropyridine0.130.35
4-nitropiridine-N-oxide0.350.17
Table 5. Local nucleophilicity indices for dienophiles.
Table 5. Local nucleophilicity indices for dienophiles.
DieneNk (eV)
C1C4
Isoprene1.110.80
1-trimetylsililoxy-1,3-butadiene0.800.97
Danishefsky’s diene1.470.56
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ormachea, C.M.; Kneeteman, M.N.; Mancini, P.M.E. Diels–Alder Polar Reactions of Azaheterocycles: A Theoretical and Experimental Study. Organics 2022, 3, 102-110. https://0-doi-org.brum.beds.ac.uk/10.3390/org3020008

AMA Style

Ormachea CM, Kneeteman MN, Mancini PME. Diels–Alder Polar Reactions of Azaheterocycles: A Theoretical and Experimental Study. Organics. 2022; 3(2):102-110. https://0-doi-org.brum.beds.ac.uk/10.3390/org3020008

Chicago/Turabian Style

Ormachea, Carla M., María Nélida Kneeteman, and Pedro M. E. Mancini. 2022. "Diels–Alder Polar Reactions of Azaheterocycles: A Theoretical and Experimental Study" Organics 3, no. 2: 102-110. https://0-doi-org.brum.beds.ac.uk/10.3390/org3020008

Article Metrics

Back to TopTop