Next Article in Journal
Electroplating Cobalt Films on Silicon Nanostructures for Sensing Molecules
Next Article in Special Issue
Dihydropyrimidone Derivatives as Thymidine Phosphorylase Inhibitors: Inhibition Kinetics, Cytotoxicity, and Molecular Docking
Previous Article in Journal
Azumamides A-E: Isolation, Synthesis, Biological Activity, and Structure–Activity Relationship
Previous Article in Special Issue
Small Heterocyclic Ligands as Anticancer Agents: QSAR with a Model G-Quadruplex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the Regioselectivity and the Molecular Mechanism of [3 + 2] Cycloaddition Reactions between Nitrous Oxide and Conjugated Nitroalkenes: A DFT Computational Study

1
Łukasiewicz Research Network, Institute of Heavy Organic Synthesis “Blachownia”, Energetyków 9, 47-225 Kędzierzyn-Koźle, Poland
2
Centre of Molecular and Macromolecular Studies, Polish Academy of Sciences, Sienkiewicza 112, 90-363 Lodz, Poland
3
Institute of Organic Chemistry and Technology, Cracow University of Technology, Warszawska 24, 31-155 Cracow, Poland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(23), 8441; https://doi.org/10.3390/molecules27238441
Submission received: 17 November 2022 / Revised: 26 November 2022 / Accepted: 29 November 2022 / Published: 2 December 2022

Abstract

:
Regiochemical aspects and the molecular mechanism of the [3 + 2] cycloaddition between nitrous oxide and conjugated nitroalkenes were evaluated on the basis of the wb97xd/6-311 + G(d) (PCM) computational study. It was found that, independently of the nature of the nitroalkene, all considered processes are realized via polar, single-step mechanisms. All attempts at the localization of hypothetical zwitterionic intermediates were unsuccessful. Additionally, the DFT computational study suggested that, in the course of the reaction, the formation of respective Δ2-4-nitro-4-R1-5-R2-1-oxa-2,3-diazolines was preferred from the kinetic point of view.

1. Introduction

Five-membered heterocycles including nitrogen atom(s) are an important part of many molecular systems with a valuable role in the fields of pharmacy and biotechnology [1,2,3]. The introduction of nitro groups for these types of molecules increases the potential activity of the considered molecules [4]. Additionally, the nitro group opens a wide range of possibilities for the further functionalization of amines [5], nitrile N-oxides [6], oximes [7], nitronic esters [8], and many others [9]. Thus, the question of the preparation of nitro-functionalized five-membered heterocycles has attracted the attention of many researchers around the world.
This work consisted of the continuation of our systematic studies regarding the [3 + 2] cycloaddition (32CA) processes involving conjugated nitroalkenes. Previously, we experimentally and theoretically analyzed 32CAs with the participation of nitrones [10,11], azomethine ylides [12], thiocarbonyl ylides [13], diazo compounds [14], nitrile N-oxides [15,16], and azides [17]. In the framework of this paper, we decided to shed light on the potential regioselectivity and molecular mechanism of 32CA with the participation of nitrous oxide (1) as a three-atom component (TAC) [18]. As 2π components within this study, we tested nitroethene (2a) and its analogs with different types and positions of the functionalization (Scheme 1). In particular, we selected different types of methyl- and nitro-substituted analogs of nitroethene (2b–e).
It must be underlined that the mechanistic aspects of this type of reaction cannot be predicted a priori, because in recent years many examples of stepwise 32CAs with biradical or zwitterionic intermediates have been detected [19]. For this, study we applied results derived from DFT calculations. We hope that our research will be helpful for further experimental studies in the mentioned area.

2. Results and Discussion

The title reactions can be theoretically realized according to two competitive, regioisomeric cycloaddition channels: A, leading to Δ2-4-nitro-4-R1-5-R2-1-oxa-2,3-diazolines (3a–d); and B, leading to Δ2-5-nitro-4-R1-5-R2-1-oxa-2,3-diazolines (4a–d) (Scheme 2). In the case of reaction with the participation of 1,2-dinitroethene, only one reaction channel (3e = 4e) is possible due to the symmetrical nature of the 2π component.
We started our study from the analysis of the nature of intermolecular interactions between cycloaddition components. For this purpose, conceptual density functional theory (CDFT) methodologies were applied based on the results derived (according to Domingo’s recommendations [20,21,22]) from B3LYP/6-31G(d) calculations. In the framework of CDFT, it is generally known that values of electronic chemical potential μ can estimate the direction of the electron density flux between components of bimolecular reactions. In the case of all considered reactions, the predicted electron flux should be observed from nitrous oxide (1) to the respective nitroalkene. Therefore, in Domingo’s terminology, all considered cycloadditions should be classified as forward electron density flux (FEDF) processes [22]. Next, estimated values of global electrophilicity show without any doubts that all analyzed reactions should be treated as polar, due to their respective Δω indices (Table 1).
For the exploration of the energetic profiles of the aforementioned processes, the wb97xd functional with the 6-311 + G(d) basis set was applied. A similar level of theory was very recently used for the exploration of different types of bimolecular processes, such as [3 + 2] cycloadditions [14,23] and [4 + 2] cycloadditions [24,25].
Our study started from the model process involving nitroethene (2a) in the simulated toluene solution. The wb97xd/6-311 + G(d) (PCM) calculations showed that the nature of the energetic profiles was qualitatively similar in the case of both considered channels. In particular, between the valleys of the initial molecular system (1 + 2a) and the valleys of the respective cycloadducts, only two critical points were located. These points are connected with the existence of (firstly) pre-reaction molecular complexes (MCs) and (in the second stage) transition states (TSs).
Intermolecular interactions between the initial molecules in the early reaction stages led to the formation of molecular complexes (MCs) (Figure 1). This was associated with a reduction in the enthalpy of the reaction system by about 0.2–0.4 kcal/mol (Table 2). However, the change in the entropy stimulated the positive values of the respective Gibbs free energies. Thus, MCs cannot exist in the reaction environment as stable intermediates. At this stage, no new sigma bonds are formed. All key distances exist beyond the area typical of the new sigma bonds in transition states (r > 3.3 Å). Interestingly, on both reaction paths, the addends adopt the same orientation, causing further regio-orientation within the transition state (Figure 2). Thus, optimized structures should be treated as orientation complexes. Next, the GEDT values show without any doubts that the localized structures do not exhibit the nature of charge-transfer complexes (Table 3). Similar types of pre-reaction complexes were identified very recently in different bimolecular processes [26,27,28,29].
The conversion of MCs, independently of the reaction paths, led directly to the area of the respective transition state (TSA and TSB for paths A and B, respectively). This was accompanied by a substantial increase in the enthalpy of the reaction system, by about 30 kcal/mol. However, from the kinetic point of view, cycloaddition path A is favored, so the formation of Δ2-4-nitro-1-oxa-2,3-diazoline (3a) is more probable in the course of the reaction.
Within TSs, key interatomic distances were reduced substantially. It is interesting that the kinetically favored TSA is less synchronic than TSB, as evident in the light of the Δl values. The formation of new sigma bonds was associated with the flux of the electron density. The obtained results show that the direction of this flux is compatible with the earlier CDFT prediction. On the other hand, values of global electron density transfer (GEDT) for the considered TSs were similar (GEDT = 0.18e and 0.17e for TSA and TSB, respectively). This observation confirms the polar nature of the analyzed process. The IRC calculations connect the localized TSs with valleys of the respective MCs from the one side, and with valleys of the respective cycloadducts from the other side. This excludes the possibility of a stepwise mechanism of the cycloaddition. All attempts at the optimization of hypothetical zwitterionic intermediates (Scheme 3) were unsuccessful.
Next, we used a similar approach to analyze other 32CAs with the participation of different type-1 and -2 substituted analogs of nitroethene (2a). It was found that, independently of the position of the substituent and its nature, cycloaddition processes involving all conjugated nitroalkenes were realized via single-step mechanisms. Additionally, all of the considered reactions should be treated as polar. The polar nature of transition states is increased to some extent with the increase in the polarity of the solvent. However, these changes are not sufficient for the enforcement of possible stepwise, zwitterionic mechanisms. In conclusion, in light of our results, the proposed mechanism can be treated as general for the 32CA processes between nitrous oxide and conjugated nitroalkenes.

3. Computational Details

All quantum chemical calculations were performed using the ‘Prometheus’ infrastructure, which was shared by the ACK ‘Cyfronet’ in Cracow. The wb97xd method [30] and the 6-311 + G(d) basis set were implemented in the Gaussian 09 package [31].
All critical structures were optimized using the Berny algorithm and were characterized by frequency calculations. It was found that all addends, molecular complexes (MCs), and products had positive Hessian matrices, whereas all transition states (TSs) had one negative eigenvalue in their Hessian matrices. For all transition states, intrinsic reaction coordinate (IRC) calculations were performed. The solvent effect was implemented using the polarizable continuum model (PCM) [32]. Global electron density transfer (GEDT) between substructures of the transition state [33] was calculated according to the following equation:
GEDT = ΣqA
where qA is the net Mulliken charge, and the sum is taken over all of the atoms of the nitroalkene.
All visualizations were prepared using GaussView [34].
In turn, the σ-bond development (l) indices were designated based on the following formula [10]:
  l X Y = 1 r   X Y T S     r   X Y P r   X Y P
The results of the quantum chemical calculations are shown in Table 1 and Table 2.
CDFT reactivity indices [20,21,22] were calculated at the B3LYP/6-31G(d) computational level because the electrophilicity scale is established at that level. The global electrophilicity index (ω) is given by the expression ω = μ2/2η, in terms of the electronic chemical potential (μ) and chemical hardness (η). Both quantities may be approached in terms of the one-electron energies of the frontier molecular orbitals HOMO and LUMO, as μ ≈ (EHOMO + ELUMO)/2 and η ≈ ELUMO − EHOMO, respectively.

4. Conclusions

Our computational wb97xd/6-311 + G(d) (PCM) study shed light on the probable regioselectivity as well as the molecular mechanisms of [3 + 2] cycloaddition processes with the participation of nitrous oxide as a TAC, along with different types of conjugated nitroalkenes as 2π components. In particular, the obtained results in the framework of the conceptual DFT approach indicate the polar, FEDF-type nature of all of the analyzed reactions. The analysis of the global electron density transfer between substructures of localized transition states led to similar conclusions. All localized transition states exhibited an asynchronous nature. However, the influence of the polarity of the solvent on the synchronicity of TSs was not high. All attempts at the optimization of hypothetical zwitterionic intermediates were unsuccessful. In light of our results, the proposed one-step polar mechanism can be treated as general for the 32CA processes between nitrous oxide and conjugated nitroalkenes. Experimental aspects of the title reactions will be the subject of our further studies.

Author Contributions

Conceptualization, R.J.; methodology, R.J.; software, E.D.; validation, E.D. and A.W.; investigation, E.D., A.W. and R.J.; writing—original draft preparation, R.J.; writing—review and editing, R.J.; visualization, E.D., A.W. and R.J.; supervision, R.J.; project administration, R.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Luo, L.; Wang, Q.; Xiang, Y.; Peng, X.; Hu, C. Synthesis and biological evaluation of novel thiazolo[4,5-d]pyrimidin-7(6H)-ones as topoisomerase I inhibitors. Chem. Heterocycl. Compd. 2021, 57, 1220–1229. [Google Scholar] [CrossRef]
  2. Nguyen, D.T.; Ngo, T.H.; Tran, H.T.; Dinh, T.P.; Do, P.T.; Nguyen, H.B.; Tran, L.T.P.; Ta, H.M. Synthesis and Anticancer Activity of 11-azaartemisinin Derivatives Bearing 1,2,3-triazole Moiety. Chem. Heterocycl. Compd. 2021, 57, 1037–1044. [Google Scholar] [CrossRef]
  3. Bhat, S.I.; Kigga, M.; Heravi, M.M. Multicomponent reactions based on in situ generated isocyanides for the construction of heterocycles. Chem. Heterocycl. Compd. 2021, 57, 709–719. [Google Scholar] [CrossRef]
  4. Noriega, S.; Cardoso-Ortiz, J.; López-Luna, A.; Cuevas-Flores, M.D.R.; Flores De La Torre, J.A. The Diverse Biological Activity of Recently Synthesized Nitro Compounds. Pharmaceuticals 2022, 15, 717. [Google Scholar] [CrossRef]
  5. Orlandi, M.; Brenna, D.; Harms, R.; Jost, S.; Benaglia, M. Recent Developments in the Reduction of Aromatic and Aliphatic Nitro Comppounds to Amines. Org. Process Res. Dev. 2018, 22, 430–445. [Google Scholar] [CrossRef]
  6. Kącka-Zych, A.; Jasiński, R. Understanding the molecular mechanism of γ-elimination of nitrous acid in the framework of the molecular electron density theory. J. Comput. Chem. 2021, 42, 1195–1203. [Google Scholar] [CrossRef] [PubMed]
  7. Cai, S.; Zhang, S.; Zhao, Y.; Wang, D.Z. New Approach to Oximes through Reduction of Nitro Compounds Enabled by Visible Light Photoredox Catalysis. Org. Lett. 2013, 15, 2660–2663. [Google Scholar] [CrossRef]
  8. Kornblum, N.; Brown, R.A. The Synthesis and Characterization of Nitronic Esters. J. Am. Chem. Soc. 1964, 86, 2681–2687. [Google Scholar] [CrossRef]
  9. Hao, F.; Nishiwaki, N. Recent Progress in Nitro-Promoted Direct Functionalization of Pyridones and Quinolones. Molecules 2020, 25, 673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Jasiński, R. A stepwise, zwitterionic mechanism for the 1,3-dipolar cycloaddition between (Z)-C-4-methoxyphenyl-N-phenylnitrone and gem-chloronitroethene catalysed by 1-butyl-3-methylimidazolium ionic liquid cations. Tetrahedron Lett. 2015, 56, 532–535. [Google Scholar] [CrossRef]
  11. Fryźlewicz, A.; Łapczuk-Krygier, A.; Kula, K.; Demchuk, O.M.; Dresler, E.; Jasiński, R. Regio- and stereoselective synthesis of nitrofunctionalized 1,2-oxazolidine analogs of nicotine. Chem. Heterocycl. Compd. 2020, 56, 120–122. [Google Scholar] [CrossRef]
  12. Żmigrodzka, M.; Dresler, E.; Hordyjewicz-Baran, Z.; Kulesza, R.; Jasiński, R. A unique example of noncatalyzed [3+2] cycloaddition involving (2E)-3-aryl-2-nitroprop-2-enenitriles. Chem. Heterocycl. Compd. 2017, 53, 1161–1162. [Google Scholar] [CrossRef]
  13. Jasiński, R. In the searching for zwitterionic intermediates on reaction paths of [3 + 2] cycloaddition reactions between 2,2,4,4-tetramethyl-3-thiocyclobutanone S-methylide and polymerizable olefins. RSC Adv. 2015, 5, 101045–101048. [Google Scholar] [CrossRef]
  14. Fryźlewicz, A.; Kącka-Zych, A.; Demchuk, O.M.; Mirosław, B.; Woliński, P.; Jasiński, R. Green synthesis of nitrocyclopropane-type precursors of inhibitors for the maturation of fruits and vegetables via domino reactions of diazoalkanes with 2-nitroprop-1-ene. J. Clean. Prod. 2021, 292, 126079. [Google Scholar] [CrossRef]
  15. Zawadzińska, K.; Ríos-Gutiérrez, M.; Kula, K.; Woliński, P.; Mirosław, B.; Krawczyk, T.; Jasiński, R. The Participation of 3,3,3-Trichloro-1-nitroprop-1-ene in the [3+2] Cycloaddition Reaction with Selected Nitrile N-Oxides in the Light of the Experimental and MEDT Quantum Chemical Study. Molecules 2021, 26, 6774. [Google Scholar] [CrossRef] [PubMed]
  16. Woliński, P.; Kącka-Zych, A.; Demchuk, O.M.; Łapczuk-Krygier, A.; Mirosław, B.; Jasiński, R. Clean and molecularly programmable protocol for preparation of bis-heterobiarylic systems via a domino pseudocyclic reaction as a valuable alternative for TM-catalyzed cross-couplings. J. Clean. Prod. 2020, 275, 122086. [Google Scholar] [CrossRef]
  17. Jasiński, R. Nitroacetylene as dipolarophile in [2+3] cycloaddition reactions with allenyl-type three-atom components: DFT computational study. Monatsh. Chem. 2015, 146, 591–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Ríos-Gutiérrez, M.; Domingo, L.R. Unravelling the Mysteries of the [3+2] Cycloaddition Reactions. Eur. J. Org. Chem. 2019, 2019, 267–282. [Google Scholar] [CrossRef]
  19. Jasiński, R.; Dresler, E. On the Question of Zwitterionic Intermediates in the [3+2] Cycloaddition Reactions: A Critical Review. Organics 2020, 1, 5. [Google Scholar] [CrossRef]
  20. Domingo, L.R.; Aurell, M.J.; Pérez, P.; Contreras, R. Quantitative characterization of the global electrophilicity power of common diene/dienophile pairs in Diels–Alder reactions. Tetrahedron 2002, 58, 4417–4423. [Google Scholar] [CrossRef]
  21. Pérez, P.; Domingo, L.R.; Aurell, M.J.; Contreras, R. Quantitative characterization of the global electrophilicity pattern of some reagents involved in 1,3-dipolar cycloaddition reactions. Tetrahedron 2003, 59, 3117–3125. [Google Scholar] [CrossRef]
  22. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Applications of the Conceptual Density Functional Theory Indices to Organic Chemistry Reactivity. Molecules 2016, 21, 748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Kula, K.; Zawadzińska, K. Local nucleophile-electrophile interactions in [3+2] cycloaddition reactions between benzonitrile N-oxide and selected conjugated nitroalkenes in the light of MEDT computational study. Curr. Chem. Lett. 2021, 10, 9–16. [Google Scholar] [CrossRef]
  24. Kula, K.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Jasiński, R. Analysis of the possibility and molecular mechanism of carbon dioxide consumption in the Diels-Alder processes. Pure Appl. Chem. 2021, 93, 427–446. [Google Scholar] [CrossRef]
  25. Kącka-Zych, A.; Pérez, P. Perfluorobicyclo[2.2.0]hex-1(4)-ene as unique partner for Diels–Alder reactions with benzene: A density functional theory study. Theor. Chem. Acc. 2021, 140, 17. [Google Scholar] [CrossRef]
  26. Kącka-Zych, A. Understanding the uniqueness of the stepwise [4+1]cycloaddition reaction between conjugated nitroalkenesand electrophilic carbene systems with a molecular electrondensity theory perspective. Int. J. Quantum Chem. 2021, 121, 26440. [Google Scholar] [CrossRef]
  27. Kacka-Zych, A. Push-pull nitronates in the [3+2] cycloaddition with nitroethylene: Molecular Electron Density Theory study. J Mol. Graph. Model. 2020, 97, 1075492. [Google Scholar] [CrossRef]
  28. Kącka-Zych, A. The Molecular Mechanism of the Formation of Four-Membered Cyclic Nitronates and Their Retro (3+2) Cycloaddition: A DFT Mechanistic Study. Molecules 2021, 26, 4786. [Google Scholar] [CrossRef] [PubMed]
  29. Jasiński, R. Stepwise, zwitterionic course of hetero-Diels–Alder reaction between 1,2,4-triazine molecular systems and 2-cyclopropylidene-1,3-dimethylimidazoline. Chem. Heterocycl. Comp. 2022, 58, 260–262. [Google Scholar] [CrossRef]
  30. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  31. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  32. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef] [PubMed]
  33. Domingo, L.R. A new C–C bond formation model based on the quantum chemical topology of electron density. RSC Adv. 2014, 4, 32415–32428. [Google Scholar] [CrossRef]
  34. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6.0; Semichem Inc.: Shawnee Mission, KS, USA, 2016. [Google Scholar]
Scheme 1. Components for the model of [3 + 2] cycloaddition reactions.
Scheme 1. Components for the model of [3 + 2] cycloaddition reactions.
Molecules 27 08441 sch001
Scheme 2. Theoretically possible regioisomeric channels of the [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e).
Scheme 2. Theoretically possible regioisomeric channels of the [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e).
Molecules 27 08441 sch002
Figure 1. Enthalpy profile for the [3 + 2] cycloaddition between nitrous oxide (1) and nitroethene (2a) in the toluene solution according to the wb97xd/6-311 + G(d) (PCM) calculations.
Figure 1. Enthalpy profile for the [3 + 2] cycloaddition between nitrous oxide (1) and nitroethene (2a) in the toluene solution according to the wb97xd/6-311 + G(d) (PCM) calculations.
Molecules 27 08441 g001
Figure 2. Views of key structures for [3 + 2] cycloaddition between nitrous oxide (1) and nitroethene (2a) in the toluene solution according to the wb97xd/6-311 + G(d) (PCM) calculations.
Figure 2. Views of key structures for [3 + 2] cycloaddition between nitrous oxide (1) and nitroethene (2a) in the toluene solution according to the wb97xd/6-311 + G(d) (PCM) calculations.
Molecules 27 08441 g002
Scheme 3. Hypothetical zwitterionic intermediates in the reaction between nitrous oxide (1) and nitroethene (2a).
Scheme 3. Hypothetical zwitterionic intermediates in the reaction between nitrous oxide (1) and nitroethene (2a).
Molecules 27 08441 sch003
Table 1. Global electron properties for nitrous oxide (1) and conjugated nitroalkenes (2a–e).
Table 1. Global electron properties for nitrous oxide (1) and conjugated nitroalkenes (2a–e).
μ (eV)η (eV)ω (eV)Δω (eV)
1−4.928.791.37
2a−5.335.452.611.23
2b−5.165.482.431.05
2c−5.985.033.562.18
2d−5.085.482.350.98
2e−6.494.664.523.15
Table 2. Kinetic and thermodynamic description of the considered paths of [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e) according to the wb97xd/6-311 + G(d) (PCM) calculations.
Table 2. Kinetic and thermodynamic description of the considered paths of [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e) according to the wb97xd/6-311 + G(d) (PCM) calculations.
SolventTransitionΔHΔSΔG
Toluene1 + 2a→ MCA−0.2−14.14.0
1 + 2a→ TSA30.3−30.339.3
1 + 2a→ 3a−2.8−31.06.4
1 + 2a→ MCB−0.4−16.04.4
1 + 2a→ TSB31.2−30.440.2
1 + 2a→ 4a−8.1−30.81.1
Acetone1 + 2a→ MCA0.0−11.53.5
1 + 2a→ TSA30.5−31.239.8
1 + 2a→ 3a−3.4−32.46.2
1 + 2a→ MCB−0.7−18.74.9
1 + 2a→ TSB31.5−31.440.9
1 + 2a→ 4a−9.0−32.10.6
Nitromethane1 + 2a→ MCA0.1−11.23.4
1 + 2a→ TSA30.5−31.239.8
1 + 2a→ 3a−3.5−32.46.2
1 + 2a→ MCB−0.7−18.64.9
1 + 2a→ TSB31.6−31.440.9
1 + 2a→ 4a−9.0−32.20.5
Water1 + 2a→ MCA0.1−11.23.4
1 + 2a→ TSA30.5−31.239.8
1 + 2a→ 3a−3.5−32.46.1
1 + 2a→ MCB−0.6−18.74.9
1 + 2a→ TSB31.6−31.440.9
1 + 2a→ 4a−9.1−32.20.5
Toluene1 + 2b→ MCA−0.6−15.03.9
1 + 2b→ TSA31.1−31.840.6
1 + 2b→ 3b−3.8−33.16.1
1 + 2b→ MCB−0.9−16.54.0
1 + 2b→ TSB30.5−31.840.0
1 + 2b→ 4b−8.7−33.41.3
Toluene1 + 2c→ MCA−0.4−15.54.2
1 + 2c→ TSA31.0−33.440.9
1 + 2c→ 3c−2.3−34.27.9
1 + 2c→ MCB−0.51169.24.8
1 + 2c→ TSB32.8−33.542.8
1 + 2c→ 4c−9.7−33.80.4
Toluene1 + 2d→ MCA−0.6−13.83.5
1 + 2d→ TSA30.8−30.640.0
1 + 2d→ 3d−2.5−32.27.1
1 + 2d→ MCB−0.6−16.34.3
1 + 2d→ TSB33.4−30.842.6
1 + 2d→ 4d−6.8−32.12.8
Toluene1 + 2e→ MC−0.3−20.55.8
1 + 2e→ TS31.1−35.341.7
1 + 2e→ 3e−4.2−35.46.3
Table 3. Key parameters for critical structures on the considered paths of [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e), according to the wb97xd/6-311 + G(d) (PCM) calculations.
Table 3. Key parameters for critical structures on the considered paths of [3 + 2] cycloaddition with the participation of nitrous oxide (1) and conjugated nitroalkenes (2a–e), according to the wb97xd/6-311 + G(d) (PCM) calculations.
SolventReactionStructureInteratomic Distances r (Å)lC3C4lC5O1ΔlGEDT
(e)
O1-N2N2-N3N3-C4C4-C5C5-O1
Toluene1 + 2a11.1791.119
2a 1.320
MCA1.1791.1183.5701.3213.326
TSA1.2251.1592.0581.3831.9710.6010.6300.030.19
3a1.3641.2261.4711.5171.438
MCB1.1791.1183.5581.3213.315
TSB1.2171.1651.9191.3822.0790.6900.4880.200.19
4a1.4561.2061.4641.5291.375
Acetone1 + 2a11.1791.118
2a 1.321
MCA1.1781.1183.4071.3213.294
TSA1.2241.1582.0681.3831.9700.5940.6330.040.18
3a1.3611.2271.4711.5161.441
MCB1.1791.1183.6121.3213.308
TSB1.2171.1651.9131.3822.0910.6930.4780.210.17
4a1.4581.2061.4641.5281.375
Nitromethane1 + 2a11.1791.118
2a 1.321
MCA1.1781.1183.4061.3223.294
TSA1.2241.1582.0691.3831.9700.5940.6340.040.18
3a1.3611.2271.4711.5161.441
MCB1.1791.1183.6121.3213.308
TSB1.2171.1651.9131.3822.0920.6940.4780.220.17
4a1.4581.2061.4641.5281.375
Water1 + 2a11.1791.118
2a 1.321
MCA1.1781.1183.3981.3223.296
TSA1.2241.1582.0691.3831.9700.5940.6340.040.18
3a1.3601.2271.4711.5161.442
MCB1.1791.1183.6121.3213.309
TSB1.2171.1651.9131.3822.0930.6940.4770.220.17
4a1.4591.2061.4641.5281.375
Toluene1 + 2b2b 1.325
MCA1.1781.1193.3051.3253.329
TSA1.2251.1582.0711.3881.9850.5980.6170.020.21
3b1.3691.2231.4771.5261.435
MCB1.1791.1183.5121.3253.153
TSB1.2161.1661.9261.3862.1220.6830.4660.220.19
4b1.4421.2111.4631.5271.383
Toluene1 + 2c2c 1.317
MCA1.1811.1183.9811.3173.150
TSA1.2271.1522.0981.3891.8860.5560.6860.130.34
3c1.3531.2321.4521.5121.436
MCB1.1791.1183.6751.3173.113
TSB1.2121.1661.8661.3862.0570.7140.4630.250.34
4c1.5281.1931.4511.5361.338
Toluene1 + 2d2d 1.326
MCA1.1801.1183.4681.3273.174
TSA1.2241.1612.0411.3892.0150.6070.6080.000.20
3d1.3611.2281.4651.5251.447
MCB1.1781.1193.4631.3263.362
TSB1.2201.1661.9281.3882.0700.6900.4910.200.19
4d1.4601.2061.4721.5361.371
Toluene1 + 2e2e 1.318
MC1.1801.1183.4931.3193.038
TS1.2211.1601.9821.3861.9680.6470.5740.070.37
3e1.4121.2171.4651.5231.380
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dresler, E.; Wróblewska, A.; Jasiński, R. Understanding the Regioselectivity and the Molecular Mechanism of [3 + 2] Cycloaddition Reactions between Nitrous Oxide and Conjugated Nitroalkenes: A DFT Computational Study. Molecules 2022, 27, 8441. https://doi.org/10.3390/molecules27238441

AMA Style

Dresler E, Wróblewska A, Jasiński R. Understanding the Regioselectivity and the Molecular Mechanism of [3 + 2] Cycloaddition Reactions between Nitrous Oxide and Conjugated Nitroalkenes: A DFT Computational Study. Molecules. 2022; 27(23):8441. https://doi.org/10.3390/molecules27238441

Chicago/Turabian Style

Dresler, Ewa, Aneta Wróblewska, and Radomir Jasiński. 2022. "Understanding the Regioselectivity and the Molecular Mechanism of [3 + 2] Cycloaddition Reactions between Nitrous Oxide and Conjugated Nitroalkenes: A DFT Computational Study" Molecules 27, no. 23: 8441. https://doi.org/10.3390/molecules27238441

Article Metrics

Back to TopTop