Next Article in Journal
Structure–Function Relationship of Novel Tetrakis (Mercapto-Terphenyl)Benzene Cobalt (II) Phthalocyanines: Synthesis and Computational Evaluation
Previous Article in Journal
Highly Engineered Cr-In/H-SSZ-39 Catalyst for Enhanced Performance in CH4-SCR of NOx
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Question of the Course of the Hetero Diels–Alder Reactions Between N-(2,2,2-Trichloroethylidene)Carboxamides and Dicyclohexylcarbodiimide: A New Case of the Stepwise Zwitterionic Cycloaddition Process

by
Przemysław Woliński
1,
Karolina Zawadzińska-Wrochniak
2,
Ewa Dresler
3 and
Radomir Jasiński
1,*
1
Department of Organic Chemistry and Technology, Cracow University of Technology, Warszawska 24, 31-155 Kraków, Poland
2
Radom Scientific Society, Rynek 15, 26-600 Radom, Poland
3
Łukasiewicz Research Network—Institute of Heavy Organic Synthesis “Blachownia”, Energetyków 9, 47-225 Kędzierzyn-Koźle, Poland
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(13), 2692; https://doi.org/10.3390/molecules30132692 (registering DOI)
Submission received: 25 May 2025 / Revised: 12 June 2025 / Accepted: 20 June 2025 / Published: 21 June 2025

Abstract

:
The regioselectivity and the molecular mechanism of the Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides and dicyclohexylcarbodiimide were explored based on the ωB97xd/6-311G(d) (PCM) calculations. It was found that the reaction course is determined by polar local interactions. It is interesting that the most favored reaction channel is realized not via classical single-step Diels–Alder mechanism, but according to the stepwise scheme with the intervention of the zwitterionic intermediate. The details of the electron density redistribution along the reaction coordinate were explained using the ELF technique.

1. Introduction

Six-membered nitrogen- and oxygen-containing heterocycles play an important role in many areas of modern pharmacy and biotechnology: as respiratory virus inhibitors [1], anti-inflammatories [2] as well as phytochemicals and antioxidant agents [3,4] along with many other uses [5,6,7,8,9,10] (Scheme 1).
For the preparation of this group of cyclic molecular systems many strategies are possible. However, the most universal protocol is the hetero Diels–Alder (HDA) reaction with the participation of unsaturated molecular segments of conjugated nitroalkenes [11], imines [12], vinyl aldehydes [13], nitrosocarbonyl compounds [14] and many other [15]. Many of these types of transformations require the presence of Lewis acids as catalysts [12,16,17,18]. For example, 2-aryl-1-nitroethenes react with cycloalkenes only in the presence of the SnCl2 (Scheme 2) [19]. Under thermal conditions this transformation is not proceeding.
Very recently, Zadorozhnii and coworkers [20] described the interesting case of the HDA reaction between N-(2,2,2-trichloroethylidene)carboxamides (1) and dicyclohexylcarbodiimide (2). This process is realized without the presence of any catalysts and proceeded with full regioselectivity under very mild conditions (Scheme 3).
However, the phenomenon of the high reaction regioselectivity was not explained. The molecular mechanism of the cycloaddition is not known. It should be underlined at this point that due to evidently electrophilic nature of the N-(2,2,2-trichloroethylidene)carboxamides 1 and nucleophilic character of the dicyclohexylcarbodiimide 2, their cycloaddition can be realized via stepwise mechanism with the intervention of the zwitterionic intermediate [21,22]. Alternatively, the stepwise mechanism can compete with the one-step. Two types of stepwise mechanisms, determined by different regiocontrol, are theoretically possible (Scheme 4).
So, in the framework of our study, we decided to shed a light on the question of the regiodirection as well as mechanistic aspects of title processes. For this purpose, the results of the calculations based on the ωB97X-D/6-311G(d) (PCM) DFT were used. To obtain information about which of the substituents influences the reaction course, we decided to modify the base molecule 1a by introducing an electron-donating substituent (EDG) (case 1b), as well as the electron-acceptor substituent (EWG) (case 1c).

2. Results and Discussion

2.1. Electronic Properties and the Nature of Intermolecular Interactions of Cycloaddition Components

We initiated our study from the analysis of key electronic properties of the components of the HDA process. We performed this investigation based on the reactivity descriptors estimated in the framework of the Conceptual Density Functional Theory (CDFT) [23,24,25]. A similar approach was very recently used for the prediction of the local reactivity of components of many different cycloaddition processes [26,27,28,29,30,31]. As can be seen in Table 1, the electronic chemical potential of N-(2,2,2-trichloroethylidene)carboxamides 1a–c is much higher than the electronic potential of dicyclohexylcarbodiimide 2. This indicates that in the studied HDA reactions the charge should be transferred from molecule 2 to a respective reaction partner. So, in Domingo terminology [32] the analyzed cycloadditions should be treated as the Reverse Electron Density Flux (REDF) processes. According to the universal scale of the organic reactivity [33], the classification of considered addends is different. All N-(2,2,2-trichloroethylidene)carboxamides (1a–c) should be defined as strong electrophiles. For comparison, the global electrophilicity of the dicyclohexylcarbodiimide (2) does not exceed 0.5 eV. So, this component should be treated as the nucleophilic agent. Its properties are quantitatively expressed by a great value in the global nucleophilicity index (2.66 eV). It should be underlined that the difference between global electrophilicities within considered reaction pairs confirms the polar nature [34] of title reactions.
It is generally known [35,36,37] that within polar processes local reactivity and regioselectivity is determined by the most attractive interactions between the most activated reaction centers. In the case of considered heteroanalogues of diene, the most local electrophilicity power is localized on the C4 carbon atom of the O=C-N=C < moiety (Table 1, Figure 1). On the other hand, the most nucleophilic power within the molecule 2 is located on the nitrogen atom of the -N=C=N- moiety (Table 1, Figure 2). The mentioned interactions of most active regions of addends should lead to the formation of cycloadducts 3a–c, which excellently correlate with the experimentally observed regioselectivity.

2.2. Critical Points on Reaction Profiles and Critical Structures

We started the exploration of reaction profiles from the model cycloaddition 1a + 2. According to the applied experimental conditions, a carbon tetrachloride solution was used in these calculations. It was found that the nature of the energy profiles for both competitive reaction paths are completely different. In the case of the reaction path A (associated with the formation of the adduct 3a), between the valleys of addends and adducts four critical points were detected and verified (Figure 3 and Figure 4). In contrast, in the case of the reaction path B (associated with the formation of the adduct 4a), between valleys of addends and adducts only two critical points were detected and verified (Figure 4 and Figure 5).
In the framework of path A, the interactions between reaction components lead initially to the formation of the pre-reaction molecular complex MCA (Figure 3 and Figure 4). This process is realized as barrierless and is accompanied by the reduction of the enthalpy of the reaction system by 12.1 kcal/mol (Table 2). At the same time, the degree of order of the molecular system is substantially increased. In consequence, the formation of the MCA is associated with the fundamental reduction of the entropy of the reaction system. In consequence, the Gibbs free energy of the formation of the MCA achieves a positive value (0.9 kcal/mol). In consequence, the MCA cannot exist as a stable intermediate. Within the MCA the distances between reaction sites are beyond the range typical for new C-O and C-N sigma bonds in transition states [38,39,40]. However, reaction sites adopt the orientation, which determines the further regioisomerism of the cycloadduct (Figure 3). Thus, the localized structure can be treated as the orientation complex [41]. Analysis of the distribution of charges on reaction sites shows that the electron transfer between substructures (GEDT) is not realized at this stage. So, the considered intermediate is not the charge transfer complex. It should be mentioned at this point that a similar type of pre-reaction complexes was detected experimentally regarding the cycloaddition between ozone and ethene [42].
The further conversion of the reaction system along the reaction coordinate leads to the critical point connected with the existence of the transition state. This is the TSA transition state. The enthalpy of activation for this transformation is, however, very low. So, this process should be treated as kinetically allowed even at lower temperatures. In contrast to many known transition states typical for HDA reactions, within the localized TS only one new sigma bond is formed (C4-C5). At the same time, the great electron density transfer from substructure 2 to substructure 1a is observed (GEDT = 0.33 e; Table 2). This confirms the polar nature of the transition state, which was predicted based on the reactivity descriptors. The IRC calculations connect this critical point directly with the valley of the MCA and the second valley connected with the existence of the second reaction intermediate (IA) (Figure 3). Within the IA molecule, the bond C4-C5 is already fully completed. Subsequently, the great value of the GEDT index shows clearly its zwitterionic character. These types of zwitterionic intermediates were detected earlier regarding some DA and HDA processes [38,43]. To complete the formation of the target molecule 3a, the formation of the second new sigma bond C6-O1 is necessary. This process is realized in the framework of the next critical points on the reaction profile—the transition state TS2A (Figure 3). Within this structure the key distance C6-O1 is reduced up to 2.127 Å. This is a rather typical distance for the new sigma C-O bond in the transition state [44,45,46,47,48]. It should be noted that the polarity of this TS is rather high, which is confirmed by a great value in the GEDT index. The IRC calculations connect this critical point directly with the intermediate IA and with the target product 3a. From a thermodynamic point of view, the formation of this adduct is beneficial due to a negative value of the Gibbs free energy of the formation (Table 2).
The molecular mechanism is completely different for the reaction of path B. Within the initial phase, the intermolecular interactions between addends direct the reaction system to the area of the pre-reaction molecular complex MCB (Figure 4). This process is associated with the reduction of the enthalpy of the reaction system. This change is substantially smaller than in the case of the analogous transformation leading to the MCA. Within the MCB reaction centers adopt specific orientation, but it is a different orientation than in the case of the MCA intermediate (Table 3, Figure 5). These distances exist beyond the range typical for new sigma bonds in transition states. Analogously to MCA, MCB does not exhibit the nature of the charge transfer complex. The further conversion of the reaction system leads to the area of the transition state TSB. This transition requires very high activation energy (more than 60 kcal/mol (Table 2). So, in comparison to the activation barrier on path A, the process of TSB formation should be treated as forbidden from the kinetic point of view. Structurally, the TSB exhibits a completely different nature than transition states observed on path A. Within this structure, two new sigma bonds are formed. These are the C4-C5 and N6-O1 bonds. The advancement of these bonds is slightly different. Subsequently, the polarity of this transition state is relatively smaller than in the case of transition states detected on path A. The IRC calculations connect the TSB directly with the MCB and 4a structures. However, it should be underlined that the transformation of the reaction system to the adduct 4a is fully unfavored due to thermodynamic factors. An extremely high Gibbs free energy in this reaction excludes the possibility of the formation of 4a along the reaction course, independently of the abovementioned kinetic factors.
In the next research stage, we also explored the influence of the polarity of solvent on the reaction course (Table 2, Table 3 and Table 4). For this purpose, the carbon tetrachloride was replaced with acetonitrile in DFT computations. It was found that in the acetonitrile solution, all activation barriers are higher by about 2 kcal/mol in comparison to the analogous process in carbon tetrachloride. These changes are not enough to reverse the reaction regioselectivity. From a mechanistic point of view, the scheme of the transformation of the molecular system along the reaction coordinate is analogous independently of the polarity of the solvent. The substituent effect on the reaction course was also explored (Table 2, Table 3 and Table 4). We detected that the influence of the nature of substituents at the aryl ring of carboxamides 1a–c on the reaction regioselectivity and kinetic profiles is lower than in the case of the solvent effect. In consequence, the proposed regioselectivity scheme, as well as the molecular mechanism, can be treated as general for some range of HDA reaction components.

2.3. The ELF Study of the Reagents and BET Study of Mechanism of the Model Reaction of 1a and 2

First, to shed some light on the electronic structure of our model reagents, carboxamide 1a and carbodiimide 2, we performed ELF analysis using the TopMod 09 package [49]. The obtained ELF valence basin attractors with ELF-based Lewis-like structures can be seen in Figure 6. The molecule of carboxamide 1 presents two monosynaptic basins, V(O1) and V’(O1), integrating 5.27 e which represents the nonbonding electron density of the oxygen atom of the carbonyl group. Interestingly, the disynaptic basin V(O1,C2) with a population of 2.40 e would represent a single bond in the Lewis-like structure of 1a. On the other hand, the imine region is made up of two disynaptic basins, V(C4,N3) and V’(C4,N3), integrating 1.73 e and 1.33 e, respectively, and a monosynaptic basin V(N3) of nitrogen’s nonbonding electron density. The Natural Population Analysis (NPA) shows a significant positive charge of 0.71 e on the carbon atom of the carbonyl group, and a slightly positive charge of 0.12 e on carbon C4. The imine’s nitrogen N3 and oxygen O1 atoms exhibit a moderately negative charge of −0.48 and −0.57 e, respectively. The symmetric molecule of carbodiimide 2 consists of two disynaptic basins, V(N5,C6) and V’(C5,C6), and monosynaptic basin V(N5), integrating 1.71 e, 1.62 e, and 2.72 e. The mirrored basins V(C6,N7), V’(C6,N7) and V(N7) show identical populations. The NPA revealed a positive charge of 0.65 e on the central carbon C6 and negative charges of −0.55 e on the nitrogen atoms N5 and N7.
In the next step, we studied the mechanism of the model reaction of carboxamide 1a with carbodiimide 2 using bonding evolution theory (BET) analysis [50].
The addition of carbodiimide 2 to carboxamide 1a studied with BET revealed four phases in its reaction path; the critical points of this path are named PA. The BET data is collected in Table 5, and the ELF-based Lewis-like structures are depicted in Scheme 5. The molecular complex shows only minimal changes in electron density compared to separate reagents 1a and 2. The first significant change takes place in PA1 where two disynaptic basins, V(C4,N3) and V’(C4,N3), merge into one V(C4,N3), integrating 2.95 e. The GEDT in PA1 is only 0.01 e, but the transition state TSA located in this phase exhibits a significantly increased GEDT by 0.35 e. The III phase, d(C4-N5) = 1.796 Å, starts with the creation of a new bond C4-N5 by donation of a part of the nonbonding electron density of N5, represented by monosynaptic basin V(N5), to carbon atom C4. The newly created disynaptic basin V(C4,N5) shows a population of 1.83 e, where integration of V(N5) decreases to 0.96 e; the structure PA2 and the directly preceding PA2′ can be seen in Figure 7. The last phase starts at PA3 where a merging of two disynaptic basins, V(N5,C6) and V’(N5,C6), of the carbodiimide region into one V(N5,C6) can be observed. The final structure of this phase, which is the intermediate zwitterion IA, exhibits a slight increase in the nonbonding electron densities of atoms O1, N3 and N5. On the other hand, the disynaptic basins V(O1,C2), V(C4,N3) and V(N5,C6) show a decrease in population, interestingly, as well as V(C4,N5) of the newly created bond C4-N5. These changes are in line with thhe electron density distribution of cycloadduct 3a.
The second part of the reaction of carboxamide 1a with carbodiimide 2 proceeds by cyclization of zwitterion IA in six phases based on the BET analysis; the critical points of this path are named PB. The BET data is collected in Table 6, and the ELF-based Lewis-like structures are depicted in Scheme 6. During the first phase a rearrangement of electron density takes place, in particular the disynaptic basins V(O1,C2) and V(N5,C6) decrease in population and at the same time the disynaptic basin V(N3,C2) and monosynaptic basins V(N5) and V(N7) show an increase. In this phase the transition state TS2A is located and a GEDT between carboxamide and carbodiimide fragments of the structure starts to decrease. Point PB2 begins phase II with the creation of a new monosynaptic basin V(C6) integrating 0.04 e, and the disappearance of V’(O1), with its population transferred to V(O1). At the next phase the basin V(C6) is destroyed, and at the same time the population of basin V(O1) increases to 5.84 e. The GEDT decreases to 0.31 e. Point PB3, d(O1-C6) = 1.703 Å and d(C4-N5) = 1.474 Å, marks the creation of a second new bond O1-C6 by donation of the nonbonding electron density of O1, represented by monosynaptic basin V(O1), to C6. The newly created disynaptic basin V(O1,C6) integrated 0.97 e, while the population of V(O1) decreased to 4.91 e. The point PB3 at which the new bond is established and directly preceding point PB3′ is shown on Figure 8. Phase V starts with a split of disynaptic basin V(N3,C2) into two, V(N3,C2) and V’N3,C2), integrating 1.42 and 1.50, respectively. The V(O1,C6) basin of the new bond increases in population to 1.09 e. The last phase begins with the creation of a second monosynaptic basin V’(N5). In the meantime, the V(O1) basin’s integration decreases to 4.49 e, and the GEDT lowers to 0.14 e. In the product 3a, a further increase in the population of monosynaptic basins V’(N5) and V(N7) was observed, while integrations of V(N5), V(C6,N7) and V’(C6,N7) decreased.

3. Computational Details

The explorations of the reaction profiles were performed based on the quantum-chemical DFT calculations. The ωB97X-D/6-311G(d) level of theory from the Gaussian 09 software package [51] was used. Transition states were localized using the QST2 procedure. At the same time, an alternative approach was applied for localization of transition states on reaction paths leading to adducts 3 and 4. In particular, the molecular systems were scanned with a gradual reduction of key interatomic distances. Using both different ways we obtained the same sequence of critical structures along the reaction path, for the A and B cycloaddition channels. All attempts at searching for the alternative mechanisms of reactions A and B were not successful.
All localized and optimized stationary points were characterized using vibrational analysis. It was found that starting molecules, intermediates and products had positive Hessian matrices. On the other hand, all transition states (TS) showed only one negative eigenvalue in their Hessian matrices. The calculations were performed for T = 298 K.
Intrinsic reaction coordinate (IRC) calculations were performed for the verification of all optimized transition states. The presence of the solvent in the reaction environment (tetrachloromethane, acetonitrile) was included using the IEFPCM algorithm [52].
The global electron density transfer (GEDT) [53] within critical structures (MCs and TSs) was estimated using the following formula:
GEDT = −ΣqA
where qA is the net charge, and the sum is taken over all the atoms of the substructure of the nitroalkene.
The same level of theory was used within the ELF analysis
Global and local electronic properties of the reactants were estimated according to the equations recommended earlier by Parr and Domingo [54,55,56]. According to Domingo’s recommendation, for this purpose the ωB97X-D/6-311G(d) level of theory was used. All molecules were fully optimized. Next, electronic chemical potentials (μ) and chemical hardness (η) were evaluated in terms of one-electron energies of FMO (ΕHOMO andΕLUMO) using the following equations:
μ ≈ (EHOMO + ELUMO)/2
η ≈ ELUMO − EHOMO
Next, the values of µ and η were then used to calculate the global electrophilicity index (ω) according to the following formula:
ω = μ2/2η
Subsequently, global nucleophilicity (N) [57] can be expressed in terms of the equation (EHOMO (tetracyanoethene) (11.4 eV) and was calculated using the ωB97xd/6-311G(d) level of theory):
N = EHOMO − EHOMO (tetracyanoethene)
The local electrophilicity (ωk) condensed to atom k was calculated by projecting the index ω onto any reaction center k in the radical cation of the molecule using Parr functions P+k [58] represented by Mulliken spin densities:
ωk = P+k·ω
The local nucleophilicity (Nk) condensed to atom k in the radical anion of the molecule was calculated using global nucleophilicity N and Parr functions P-k [58] according to the following formula:
Nk = Pk·N

4. Conclusions

This ωB97xd/6-311G(d) (PCM) computational study offers a comprehensive view on factors that determine the global and local reactivity of the components of the Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides and dicyclohexylcarbodiimide. It was found that the considered processes should be treated as polar and determined by the local interaction between the C4 carbon atom of the O=C-N=C< moiety of the heteroanalogue of diene and the nitrogen atom of the -N=C=N- moiety of the dicyclohexylcarbodiimide. The polar nature of the title reactions is strong, which determines the stepwise cycloaddition mechanism with the intervention of the zwitterionic intermediate on a kinetically favored path. It should be underlined, that the competitive reaction path that leads to the regioisomeric adduct should be treated as forbidden from a kinetic point of view and not beneficial from a thermodynamic point of view. The details of the electron density redistribution along the reaction coordinate can be explained using the ELF technique.

Author Contributions

Methodology, R.J.; Formal analysis, P.W. and R.J.; Investigation, P.W., K.Z.-W. and R.J.; Data curation, E.D.; Writing—original draft, P.W. and 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.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Acknowledgments

We gratefully acknowledge Polish high-performance computing infrastructure PLGrid (HPC Center: ACK Cyfronet AGH) for providing computer facilities and support within computational grant no. PLG/2025/018201.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Koubi, Y.; Moukhliss, Y.; Hajji, H.; Abdessadak, O.; Alaqarbeh, M.; Ajana, M.A.; Maghat, H.; Lakhlifi, T.; Bouachrine, M. Computational structure—Biological activity and retrosynthesis investigations of 1,2,3-triazole-quinoline hybrid molecules as potential respiratory virus inhibitors. Chem. Heterocycl. Comp. 2024, 60, 491–504. [Google Scholar] [CrossRef]
  2. Abbasi, M.; Farnia, S.M.F.; Tahghighi, A. The possibility of applying some heteroatom-decorated g-C3N4 heterocyclic nanosheets for delivering 5-aminosalicylic acid anti-inflammatory agent. Chem. Heterocycl. Comp. 2024, 60, 655–662. [Google Scholar] [CrossRef]
  3. Elbouhi, M.; Tabti, K.; Ouabane, M.; Alaqarbeh, M.; Elkamel, K.; Lakhlifi, T.; Sbai, A.; Bouachrine, M. A computational exploration of the antioxidant potential of conjugated quinazolinone Schiff bases. Chem. Heterocycl. Comp. 2024, 60, 627–638. [Google Scholar] [CrossRef]
  4. Gassim, H.B.M.A.; Hassan, M.; Abadi, R.S.M.; Mustafa, Y.A.A. Phytochemical constituents and antioxidant activity of Ricinus communis Linn leaf and seeds extracts. Sci.Rad. 2024, 3, 74–88. [Google Scholar] [CrossRef]
  5. Sadowski, M.; Kula, K. Nitro-functionalized analogues of 1,3-Butadiene: An overview of characteristic, synthesis, chemical transformations and biological activity. Curr. Chem. Lett. 2024, 13, 15–30. [Google Scholar] [CrossRef]
  6. Alobaide, Z.N.; Habel, A.; Awin, T.; Buzgaia, N.; Elamari, A.; Elteera, F.; Elabbar, F. Arum Cyrenaicum: A comprehensive review of Phytochemistry and Bioactivity. Sci.Rad. 2024, 3, 218–227. [Google Scholar] [CrossRef]
  7. Shiri, P.; Roosta, A.; Ramezanpour, S.; Amani, A.M. Editorial: Six-membered heterocycles: Their synthesis and bio applications. Front. Chem. 2023, 11, 1229825. [Google Scholar] [CrossRef]
  8. Jasiński, R. Synthesis of 1,2-oxazine N-oxides vianoncatalyzedhetero-Diels–Alder reactions of nitroalkenes (microreview). Chem. Heterocycl. Comp. 2024, 60, 121–123. [Google Scholar] [CrossRef]
  9. Poonam, N.J.; Dhadda, S.; Jangid, D.K. Recent advances in the synthesis of five- and six-membered heterocycles as bioactive skeleton: A concise overview. Chem. Select 2022, 7, e202103139. [Google Scholar]
  10. Sadowski, M.; Synkiewicz-Musialska, B.; Kula, K. (1E,3E)-1,4-Dinitro-1,3-butadiene—Synthesis, Spectral Characteristics and Computational Study Based on MEDT, ADME and PASS Simulation. Molecules 2024, 29, 542. [Google Scholar] [CrossRef]
  11. Wen, C.; Dechsupa, N.; Yu, Z.; Zhang, X.; Liang, S.; Lei, X.; Xu, T.; Gao, X.; Hu, Q.; Innuan, P.; et al. Pentagalloyl Glucose: A Review of Anticancer Properties, Molecular Targets, Mechanisms of Action, Pharmacokinetics, and Safety Profile. Molecules 2023, 28, 4856. [Google Scholar] [CrossRef] [PubMed]
  12. Hellel, D.; Chafaa, F.; Nacereddine, A.K. Synthesis of tetrahydroquinolines and quinoline derivatives through the Lewis acid catalysed Povarov reaction: A comparative study between multi step and multi-component methods. Sci. Rad. 2023, 2, 295–308. [Google Scholar] [CrossRef]
  13. Ameur, S.; Barhoumi, A.; Abdallaoui, H.E.A.; Syed, A.; Belghiti, M.E.; Elgorban, A.M.; Wong, L.S.; Wang, S.; El Idrissi, M.; Zeroual, A.; et al. Molecular docking, exploring diverse selectivities and mechanistic insights in the cycloaddition reaction between 3-benzoylpyrrolo-[1,2-a]quinoxaline-1,2,4(5H)-triones and butyl vinyl ether. Chem. Heterocycl. Comp. 2024, 60, 584–591. [Google Scholar] [CrossRef]
  14. Bodnar, B.S.; Miller, M.J. The Nitrosocarbonyl Hetero-Diels–Alder Reaction as a Useful Tool for Organic Syntheses. Angew. Chem. Int. Ed. 2011, 50, 5630–5647. [Google Scholar] [CrossRef] [PubMed]
  15. Jørgensen, K. Hetero-Diels−Alder Reactions of Ketones—A Challenge for Chemists. Eur. J. Org. Chem. 2004, 2004, 2093–2102. [Google Scholar] [CrossRef]
  16. Waldmann, H. Asymmetric Hetero Diels-Alder Reactions. Synthesis 1994, 1994, 535–551. [Google Scholar] [CrossRef]
  17. Jørgensen, K. Catalytic Asymmetric Hetero-Diels–Alder Reactions of Carbonyl Compounds and Imines. Angew. Chem. Int. Ed. 2000, 39, 3558–3588. [Google Scholar] [CrossRef]
  18. Vermeeren, P.; Hamlin, T.A.; Fernández, I.; Bickelhaupt, F.M. How Lewis Acids Catalyze Diels–Alder Reactions. Angew. Chem. Int. Ed. 2020, 59, 6201–6206. [Google Scholar] [CrossRef]
  19. Denmark, S.E.; Kesler, B.S.; Moon, K.C. Inter- and intramolecular [4+2] cycloadditions of nitroalkenes with olefins. 2-Nitrostyrenes. J. Org. Chem. 1992, 57, 4912. [Google Scholar] [CrossRef]
  20. Zadorozhnii, P.V.; Pokotylo, I.O.; Kiselev, V.V.; Kharchenko, A.V. Synthesis of (Z)-N,3-dicyclohexyl-6-substituted-4-(trichloromethyl)-3,4-dihydro-2H-1,3,5-oxadiazin-2-imines via [4+2] hetero Diels-Alder reaction: Their spectral characteristics and molecular structure. Chem. Data Coll. 2023, 48, 101093. [Google Scholar] [CrossRef]
  21. Jasiński, R. On the Question of Stepwise [4+2] Cycloaddition Reactions and Their Stereochemical Aspects. Symmetry 2021, 13, 1911. [Google Scholar] [CrossRef]
  22. Kącka-Zych, A. Participation of Phosphorylated Analogues of Nitroethene in Diels–Alder Reactions with Anthracene: A Molecular Electron Density Theory Study and Mechanistic Aspect. Organics 2020, 1, 36–48. [Google Scholar] [CrossRef]
  23. 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]
  24. Hoffmann, G.; Chermette, H.; Morell, C. Revisiting nucleophilicity: An index for chemical reactivity from a CDFT approach. J. Mol. Model. 2024, 30, 232. [Google Scholar] [CrossRef] [PubMed]
  25. Rincón, L.; Rodríguez, W.M.; Mora, J.R.; Zambrano, C.; Seijas, L.E.; Reyes, A.; Torres, F.J. A redefinition of global conceptual density functional theory reactivity indexes by means of the cubic expansions of the energy. Phys. Chem. Chem. Phys. 2025, 27, 8174–8185. [Google Scholar] [CrossRef] [PubMed]
  26. CDFT Sadowski, M.; Dresler, E.; Zawadzińska, K.; Wróblewska, A.; Jasiński, R. Syn-Propanethial S-Oxide as an Available Natural Building Block for the Preparation of Nitro-Functionalized, Sulfur-Containing Five-Membered Heterocycles: An MEDT Study. Molecules 2024, 29, 4892. [Google Scholar] [CrossRef]
  27. Dresler, E.; Wróblewska, A.; Jasiński, R. On the energetic aspects and molecular mechanism of 3-nitro-substituted 2-isoxazolines formation via nitrile N-oxide [3+2] cycloaddition: An MEDT computational study. Molecules 2024, 29, 3042. [Google Scholar] [CrossRef]
  28. Zawadzińska, K.; Gadocha, Z.; Pabian, K.; Wróblewska, A.; Wielgus, E.; Jasiński, R. The First Examples of [3+2] Cycloadditions with the Participation of (E)-3,3,3-Tribromo-1-Nitroprop-1-Ene. Materials 2022, 15, 7584. [Google Scholar] [CrossRef]
  29. Kula, K.; Łapczuk, A.; Sadowski, M.; Kras, J.; Zawadzińska, K.; Demchuk, O.M.; Gaurav, G.K.; Wróblewska, A.; Jasiński, R. On the Question of the Formation of Nitro-Functionalized 2,4-Pyrazole Analogs on the Basis of Nitrylimine Molecular Systems and 3,3,3-Trichloro-1-Nitroprop-1-Ene. Molecules 2022, 27, 8409. [Google Scholar] [CrossRef]
  30. 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. [Google Scholar] [CrossRef]
  31. Dresler, E.; Woliński, P.; Wróblewska, A.; Jasiński, R. On the Question of Zwitterionic Intermediates in the [3+2] Cycloaddition Reactions between Aryl Azides and Ethyl Propiolate. Molecules 2023, 28, 8152. [Google Scholar] [CrossRef] [PubMed]
  32. Domingo, L.R.; Ríos-Gutiérrez, M. A Useful Classification of Organic Reactions Based on the Flux of the Electron Density. Sci. Rad. 2023, 2, 1–24. [Google Scholar] [CrossRef]
  33. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Electrophilicity w and Nucleophilicity N Scales for Cationic and Anionic Species. Sci. Rad. 2025, 4, 1–17. [Google Scholar] [CrossRef]
  34. Domingo, L.R.; Sáez, J.A. Understanding the mechanism of polar Diels–Alder reactions. Org. Biomol. Chem. 2009, 7, 3576–3583. [Google Scholar] [CrossRef] [PubMed]
  35. Sadowski, M.; Kula, K. Unexpected Course of Reaction Between (1E,3E)-1,4-Dinitro-1,3-butadiene and N-Methyl Azomethine Ylide—A Comprehensive Experimental and Quantum-Chemical Study. Molecules 2024, 29, 5066. [Google Scholar] [CrossRef]
  36. Aitouna, A.O.; Syed, A.; Alfagham, A.T.; Mazoir, N.; de Julián-Ortiz, J.V.; Elgorban, A.M.; El idrissi, M.; Wong, L.S.; Zeroual, A. Investigating the chemical reactivity and molecular docking of 2-diazo-3,3,3-trifluoro-1-nitropropane with phenyl methacrylate using computational methods. Chem. Heterocycl. Comp. 2024, 60, 592–599. [Google Scholar] [CrossRef]
  37. Mtiraou, H.; Ghabi, A.; Al-Ghulikah, H.; Habib, M.A.; Hajji, M. Structural and chemical reactivity insights of a benzimidazolidinone-based N-heterocycle: A multiapproach quantum-chemical study. Chem. Heterocycl. Comp. 2024, 60, 611–616. [Google Scholar] [CrossRef]
  38. Sadowski, M.; Dresler, E.; Wróblewska, A.; Jasiński, R. A New Insight into the Molecular Mechanism of the Reaction between 2-Methoxyfuran and Ethyl (Z)-3-phenyl-2-nitroprop-2-enoate: An Molecular Electron Density Theory (MEDT) Computational Study. Molecules 2024, 29, 4876. [Google Scholar] [CrossRef]
  39. Dresler, E.; Wróblewska, A.; Jasiński, R. Understanding the Molecular Mechanism of Thermal and LA-Catalysed Diels–Alder Reactions between Cyclopentadiene and Isopropyl 3-Nitroprop-2-Enate. Molecules 2023, 28, 5289. [Google Scholar] [CrossRef]
  40. Łapczuk, A.; Ríos-Gutiérrez, M. Mechanistic Aspects of [3+2] Cycloaddition Reaction of Trifluoroacetonitrile with Diarylnitrilimines in Light of Molecular Electron Density Theory Quantum Chemical Study. Molecules 2025, 30, 85. [Google Scholar] [CrossRef]
  41. Pintér, B.; De Proft, F.; Veszprémi, T.; Geerlings, P. Theoretical Study of the Orientation Rules in Photonucleophilic Aromatic Substitutions. J. Org. Chem. 2008, 73, 1243–1252. [Google Scholar] [CrossRef] [PubMed]
  42. Gillies, C.W.; Gillies, J.Z.; Suenram, D.R.; Lovas, F.J.; Kraka, E.; Cremer, D. Van der Waals complexes in 1,3-dipolar cycloaddition reactions: Ozone-ethylene. J. Am. Chem. Soc. 1991, 113, 2412–2421. [Google Scholar] [CrossRef]
  43. Jasiński, R. On the question of selective protocol for the preparation of juglone via (4+2) cycloaddition involving 3-hydroxypyridazine: DFT mechanistic study. Chem. Heterocycl. Comp. 2023, 59, 179–182. [Google Scholar] [CrossRef]
  44. Kula, K.; Sadowski, M. Regio- and Stereoselectivity of [3+2] Cycloaddition Reactions between (Z)-1-(Anthracen-9-Yl)-N-Methyl Nitrone and Analogs of Trans-β-Nitrostyrene on the Basis of MEDT Computational Study. Chem. Heterocycl. Comp. 2023, 59, 138–144. [Google Scholar] [CrossRef]
  45. Fryźlewicz, A.; Olszewska, A.; Zawadzińska, K.; Woliński, P.; Kula, K.; Kącka-Zych, A.; Łapczuk-Krygier, A.; Jasiński, R. On the Mechanism of the Synthesis of Nitrofunctionalised Δ2-Pyrazolines via [3+2] Cycloaddition Reactions between α-EWG-Activated Nitroethenes and Nitrylimine TAC Systems. Organics 2022, 3, 59–76. [Google Scholar] [CrossRef]
  46. 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]
  47. Karaś, A.; Łapczuk, A. Computational model of the formation of novel nitronorbornene analogs via Diels–Alder process. Reac. Kinet. Mec. Cat. 2025. [Google Scholar] [CrossRef]
  48. Fałowska, A.; Grzybowski, S.; Kapuściński, D.; Sambora, K.; Łapczuk, A. Modeling of the General Trends of Reactivity and Regioselectivity in Cyclopentadiene–Nitroalkene Diels–Alder Reactions. Molecules 2025, 30, 2467. [Google Scholar] [CrossRef]
  49. Noury, S.; Krokidis, X.; Fuster, F.; Silvi, B. Topmod Package; Universite Pierre et Marie Curie: Paris, France, 1997. [Google Scholar]
  50. Krokidis, K.; Noury, S.; Silvi, B. Characterization of Elementary Chemical Processes by Catastrophe Theory. J. Phys. Chem. A 1997, 101, 7277–7282. [Google Scholar] [CrossRef]
  51. 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]
  52. 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]
  53. 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]
  54. 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]
  55. Parr, R.G.; Szentpály, L.V.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  56. Domingo, L.R. 1999–2024, a Quarter Century of the Parr’s Electrophilicity ω Index. Sci. Rad. 2024, 157–186. [Google Scholar] [CrossRef]
  57. Domingo, L.R.; Chamorro, E.; Pérez, P. Understanding the Reactivity of Captodative Ethylenes in Polar Cycloaddition Reactions. A Theoretical Study. J. Org. Chem. 2008, 73, 4615–4624. [Google Scholar] [CrossRef]
  58. 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]
Scheme 1. Examples of bioactive six-membered nitrogen- and oxygen-containing heterocycles.
Scheme 1. Examples of bioactive six-membered nitrogen- and oxygen-containing heterocycles.
Molecules 30 02692 sch001
Scheme 2. Experimentally observed course of the hetero Diels–Alder reactions between 2-aryl-1-nitroethenes and cycloalkenes.
Scheme 2. Experimentally observed course of the hetero Diels–Alder reactions between 2-aryl-1-nitroethenes and cycloalkenes.
Molecules 30 02692 sch002
Scheme 3. Experimentally observed course of the hetero Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides (1 and dicyclohexylcarbodiimide (2).
Scheme 3. Experimentally observed course of the hetero Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides (1 and dicyclohexylcarbodiimide (2).
Molecules 30 02692 sch003
Scheme 4. Possible molecular mechanisms for the hetero Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides (1 and dicyclohexylcarbodiimide (2).
Scheme 4. Possible molecular mechanisms for the hetero Diels–Alder reactions between N-(2,2,2-trichloroethylidene)carboxamides (1 and dicyclohexylcarbodiimide (2).
Molecules 30 02692 sch004
Figure 1. The distribution of the local electrophilicity within N-(2,2,2-trichloroethylidene)carboxamide (1ac) molecules.
Figure 1. The distribution of the local electrophilicity within N-(2,2,2-trichloroethylidene)carboxamide (1ac) molecules.
Molecules 30 02692 g001aMolecules 30 02692 g001b
Figure 2. The distribution of the local electrophilicity within dicyclohexylcarbodiimide (2) molecule.
Figure 2. The distribution of the local electrophilicity within dicyclohexylcarbodiimide (2) molecule.
Molecules 30 02692 g002
Figure 3. Views of critical structures of the reaction 1a + 2 → 3a in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Figure 3. Views of critical structures of the reaction 1a + 2 → 3a in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Molecules 30 02692 g003
Figure 4. Enthalpy profiles for the [4+2] cycloaddition between N-(2,2,2-trichloroethylidene)carboxamide (1a) and dicyclohexylcarbodiimide (2) in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Figure 4. Enthalpy profiles for the [4+2] cycloaddition between N-(2,2,2-trichloroethylidene)carboxamide (1a) and dicyclohexylcarbodiimide (2) in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Molecules 30 02692 g004
Figure 5. Views of critical structures of the reaction 1a + 2 → 4a in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Figure 5. Views of critical structures of the reaction 1a + 2 → 4a in the CCl4 environment in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Molecules 30 02692 g005
Figure 6. ωB97X-D/6-311G(d) ELF valence basin attractors with the most relevant valence basin populations and ELF-based Lewis-like structures with natural atomic charges of carboxamide 1a and carbodiimide 2. Negative charges are colored in red and positive in blue.
Figure 6. ωB97X-D/6-311G(d) ELF valence basin attractors with the most relevant valence basin populations and ELF-based Lewis-like structures with natural atomic charges of carboxamide 1a and carbodiimide 2. Negative charges are colored in red and positive in blue.
Molecules 30 02692 g006
Figure 7. ELF valence basin attractors of structures of the most important topological changes, creation of C4-N5 bond, during the reaction of carboxamide 1a and carbodiimide 2 along the reaction path leading to zwitterion IA. Structures with apostrophe represent the first structure before the critical point.
Figure 7. ELF valence basin attractors of structures of the most important topological changes, creation of C4-N5 bond, during the reaction of carboxamide 1a and carbodiimide 2 along the reaction path leading to zwitterion IA. Structures with apostrophe represent the first structure before the critical point.
Molecules 30 02692 g007
Scheme 5. Simplified representation of the molecular mechanism of the addition of carboxamide 1a and carbodiimide 2 by Lewis-like structures based on the topological analysis of the ELF along the reaction path leading to zwitterion IA.
Scheme 5. Simplified representation of the molecular mechanism of the addition of carboxamide 1a and carbodiimide 2 by Lewis-like structures based on the topological analysis of the ELF along the reaction path leading to zwitterion IA.
Molecules 30 02692 sch005
Scheme 6. Simplified representation of the molecular mechanism of cyclization of zwitterion IA by Lewis-like structures based on the topological analysis of the ELF along the reaction path leading to product 3a.
Scheme 6. Simplified representation of the molecular mechanism of cyclization of zwitterion IA by Lewis-like structures based on the topological analysis of the ELF along the reaction path leading to product 3a.
Molecules 30 02692 sch006
Figure 8. ELF valence basin attractors of structures of the most important topological changes, creation of O1-C6 bond, during cyclization of zwitterion IA along the reaction path leading to product 3a. Structures with apostrophe represent the first structure before the critical point.
Figure 8. ELF valence basin attractors of structures of the most important topological changes, creation of O1-C6 bond, during cyclization of zwitterion IA along the reaction path leading to product 3a. Structures with apostrophe represent the first structure before the critical point.
Molecules 30 02692 g008
Table 1. Global and local electronic properties of N-(2,2,2-trichloroethylidene)carboxamides (1ac) and dicyclohexylcarbodiimide (2).
Table 1. Global and local electronic properties of N-(2,2,2-trichloroethylidene)carboxamides (1ac) and dicyclohexylcarbodiimide (2).
Global PropertiesLocal Properties
Molecules 30 02692 i001
μ
[eV]
η
[eV]
ω
[eV]
N
[eV]
PNPCNN
[eV]
NC
[eV]
P+CP+OωC
[eV]
ωO
[eV]
1a−5.068.861.451.91 0.3330.1760.480.25
1b−4.878.611.382.23 0.3350.1750.460.24
1c−5.498.981.681.42 0.2950.1970.490.33
2−3.0511.380.412.660.473−0.1631.25−0.43
Table 2. Kinetic and thermodynamic parameters for the [4+2] cycloadditions between N-(2,2,2-trichloroethylidene)carboxamides (1a–c) and dicyclohexylcarbodiimide (2) in the light of the ωB97X-D/6-311G(d) (PCM) calculations (ΔH, ΔG are in kcal/mol; ΔS are in cal/molK).
Table 2. Kinetic and thermodynamic parameters for the [4+2] cycloadditions between N-(2,2,2-trichloroethylidene)carboxamides (1a–c) and dicyclohexylcarbodiimide (2) in the light of the ωB97X-D/6-311G(d) (PCM) calculations (ΔH, ΔG are in kcal/mol; ΔS are in cal/molK).
SolventPath TransitionΔHΔSΔG
CCl4A1a + 2 → MCA−12.1−43.80.9
1a + 2 → TSA1.5−54.017.6
1a + 2 → IA−1.6−55.214.8
1a + 2 → TS2A−0.9−61.817.5
1a + 2 → 3a−20.4−63.0−1.6
B1a + 2 → MCB−4.9−43.58.0
1a + 2 → TSB62.6−58.479.9
1a + 2 → 4a15.4−61.433.7
MeCNA1a + 2 → MCA−10.9−45.12.5
1a + 2 → TSA2.7−53.818.7
1a + 2 → IA−2.0−54.214.2
1a + 2 → TS2A−0.1−62.718.6
1a + 2 → 3a−18.2−63.90.8
B1b + 2 → MCB−4.6−47.09.4
1b + 2 → TSB64.1−59.581.8
1b + 2 → 4a17.2−61.835.6
CCl4A1b + 2 → MCA−12.3−46.11.4
1b + 2 → TSA0.4−52.215.9
1b + 2 → IA−1.5−62.417.1
1b + 2 → TS2A−0.5−59.017.1
1b + 2 → 3b−20.3−61.5−2.0
B1b + 2 → MCB−5.4−47.58.8
1b + 2 → TSB62.7−58.580.2
1b + 2 → 4b15.6−60.433.6
CCl4A1c + 2 → MCA−13.2−46.50.7
1c + 2 → TSA−2.2−62.116.4
1c + 2 → IA−3.9−55.812.8
1c + 2 → TS2A−2.7−61.015.4
1c + 2 → 3c−20.8−62.3−2.2
B1c + 2 → MCB−5.7−49.29.0
1c + 2 → TSB61.4−55.277.9
1c + 2 → 4c14.4−69.535.2
Table 3. Key parameters for critical structures of the reactions 1a−c + 2 → 3a−c in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Table 3. Key parameters for critical structures of the reactions 1a−c + 2 → 3a−c in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
SolventReactionStructureInteratomic Distances [Å]GEDT
[e]
O1-C2C2-N3N3-C4C4-N5N5-C6C6-O1
CCl41a + 21a 1.217
21.2061.4201.253
MCA1.2071.4101.2543.2071.2213.481
TSA1.2331.3651.3181.8671.2442.837−0.33
IA1.2541.3341.3851.5481.2692.625−0.58
TS2A1.2741.3161.4041.5091.2942.127−0.51
3a1.3571.2621.4411.4681.3761.389
MeCN1a + 21a 1.218
21.2081.4131.252
MCA1.2091.4101.2543.2091.2193.503
TSA1.2311.3701.3141.9081.2422.910−0.32
IA1.2521.3341.3921.5321.2712.765−0.63
TS2A1.2781.3141.4111.4991.3012.065−0.52
3a1.3561.2631.4421.4691.3761.389
CCl41b + 21b 1.217
MCA1.2081.4111.2543.2361.2213.519
TSA1.2331.3711.3171.8601.2442.938−0.34
IA1.2541.3351.3841.5491.2692.622−0.57
TS2A1.2741.3171.4031.5101.2942.134−0.51
3b1.3591.2621.4411.4681.3761.389
CCl41c + 21c 1.217
MCA1.2061.4061.2543.2211.2213.526
TSA1.2291.3711.3131.9051.2422.985−0.32
IA1.2521.3321.3881.5421.2692.649−0.59
TS2A1.2751.3121.4081.5041.2982.094−0.51
3c1.3541.2601.4421.4681.3771.391
Table 4. Key parameters for critical structures of the reactions 1a−c + 2 → 4a−c in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
Table 4. Key parameters for critical structures of the reactions 1a−c + 2 → 4a−c in the light of the ωB97X-D/6-311G(d) (PCM) calculations.
SolventReactionStructureInteratomic Distances [Å]GEDT
[e]
O1-C2C2-N3N3-C4C4-C5C5-N6N6-O1
CCl41a + 2MCB1.2101.4011.2513.7851.2195.086
TSB1.2511.3411.3381.8801.2882.128−0.22
4a1.3491.2701.4441.5311.3971.413
MeCN1a + 2MCB1.2121.3981.2513.8191.2195.215
TSB1.2501.3421.3401.8781.2892.138−0.23
4a1.3481.2711.4461.5311.3971.413
CCl41b + 2MCB1.2101.4021.2513.7911.2195.083
TSB1.2511.3421.3381.8791.2882.127−0.21
4b1.3501.2711.4441.5311.3961.412
CCl41c + 2MCB1.2081.3971.2523.7801.2195.062
TSB1.2491.3391.3401.8771.2892.138−0.24
4c1.3471.2691.4451.5311.3991.414
Table 5. ELF valence basin populations of the IRC points, MCA—zwitterion IA defining the four different phases characterizing the reaction of the carboxamide 1a and carbodiimide 2. Stationary points 1a, 2, and TSA are also included. Distances are given in angstroms, Å, electron populations in an average number of electrons, [e], relative energies in kcal·mol−1, GEDT values in an average number of electrons, [e].
Table 5. ELF valence basin populations of the IRC points, MCA—zwitterion IA defining the four different phases characterizing the reaction of the carboxamide 1a and carbodiimide 2. Stationary points 1a, 2, and TSA are also included. Distances are given in angstroms, Å, electron populations in an average number of electrons, [e], relative energies in kcal·mol−1, GEDT values in an average number of electrons, [e].
Structures1a2MCAPA1TSAPA2PA3IA
Phases--III IIIIV-
d(C4-N5)--3.2072.8721.8671.7961.7381.548
GEDT--0.010.010.350.410.450.56
dE--−13.62−8.430.420.27−0.09−3.96
V(O1)2.63-2.672.632.702.712.732.74
V’(O1)2.64-2.612.652.772.792.822.90
V(O1,C2)2.40-2.392.402.202.162.132.01
V(N3,C2)1.98-2.002.002.232.272.322.44
V(N3)2.53-2.522.643.123.193.233.33
V(C4,N3)1.73-1.662.952.362.242.151.93
V’(C4,N3)1.33-1.39-----
V(N5,C6)-1.711.681.681.761.803.182.99
V’(N5,C6)-1.621.611.631.391.33--
V(N5)-2.722.782.742.740.961.081.35
V(C6,N7)-1.621.601.541.731.751.781.90
V’(C6,N7)-1.711.731.781.881.891.911.93
V(N7)-2.722.732.762.442.412.382.22
V(C4,N5)-----1.831.661.60
Table 6. ELF valence basin populations of the IRC points, zwitterion IA—product 3a, defining the six different phases characterizing the cyclization reaction of the zwitterion IA. Stationary point TS2A is also included. Distances are given in angstroms, Å, electron populations in an average number of electrons, [e], relative energies in kcal·mol−1, GEDT values in an average number of electrons, [e].
Table 6. ELF valence basin populations of the IRC points, zwitterion IA—product 3a, defining the six different phases characterizing the cyclization reaction of the zwitterion IA. Stationary point TS2A is also included. Distances are given in angstroms, Å, electron populations in an average number of electrons, [e], relative energies in kcal·mol−1, GEDT values in an average number of electrons, [e].
StructuresIATS2APB1PB2PB3PB4PB53a
PhasesI IIIIIIVVVI-
d1(C4-N5)1.5481.5091.4861.4781.4741.4691.4551.468
d2(O1-C6)2.6252.1271.8651.7561.7031.6311.4061.389
GEDT0.560.490.370.310.280.250.140.11
dE0.001.2−0.5−2.7−4.1−6.3−16.7−19.7
V(O1)2.742.495.625.844.914.834.494.46
V’(O1)2.903.15------
V(O1,C2)2.011.891.771.701.681.641.591.58
V(N3,C2)2.442.602.772.842.881.421.521.54
V’(N3,C2)-----1.501.501.52
V(N3)3.333.223.093.033.002.952.892.85
V(C4,N3)1.931.881.851.841.841.831.791.79
V(N5,C6)2.992.672.402.222.162.112.042.07
V(N5)1.351.641.851.921.962.001.981.58
V(C6,N7)1.901.931.881.831.811.781.711.68
V’(C6,N7)1.931.811.781.731.701.691.611.53
V(N7)2.222.352.502.562.592.632.752.86
V(C4,N5)1.601.671.701.711.721.721.751.65
V(O1,C6)----0.971.091.501.55
V’(N5)------0.270.73
V(C6)--0.04-----
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

Woliński, P.; Zawadzińska-Wrochniak, K.; Dresler, E.; Jasiński, R. On the Question of the Course of the Hetero Diels–Alder Reactions Between N-(2,2,2-Trichloroethylidene)Carboxamides and Dicyclohexylcarbodiimide: A New Case of the Stepwise Zwitterionic Cycloaddition Process. Molecules 2025, 30, 2692. https://doi.org/10.3390/molecules30132692

AMA Style

Woliński P, Zawadzińska-Wrochniak K, Dresler E, Jasiński R. On the Question of the Course of the Hetero Diels–Alder Reactions Between N-(2,2,2-Trichloroethylidene)Carboxamides and Dicyclohexylcarbodiimide: A New Case of the Stepwise Zwitterionic Cycloaddition Process. Molecules. 2025; 30(13):2692. https://doi.org/10.3390/molecules30132692

Chicago/Turabian Style

Woliński, Przemysław, Karolina Zawadzińska-Wrochniak, Ewa Dresler, and Radomir Jasiński. 2025. "On the Question of the Course of the Hetero Diels–Alder Reactions Between N-(2,2,2-Trichloroethylidene)Carboxamides and Dicyclohexylcarbodiimide: A New Case of the Stepwise Zwitterionic Cycloaddition Process" Molecules 30, no. 13: 2692. https://doi.org/10.3390/molecules30132692

APA Style

Woliński, P., Zawadzińska-Wrochniak, K., Dresler, E., & Jasiński, R. (2025). On the Question of the Course of the Hetero Diels–Alder Reactions Between N-(2,2,2-Trichloroethylidene)Carboxamides and Dicyclohexylcarbodiimide: A New Case of the Stepwise Zwitterionic Cycloaddition Process. Molecules, 30(13), 2692. https://doi.org/10.3390/molecules30132692

Article Metrics

Back to TopTop