Next Article in Journal
Towards Greener Mechanosynthesis of Functional Calixarenes
Previous Article in Journal
Molecular Docking Studies on Synthetic Therapeutic Agents for COVID-19
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Theoretical Study of the Aza˗Wittig Reaction, Me3P=NR (R = Methyl or Phenyl) with Aldehyde Using the DFT and DFT-D Methods (Dispersion Correction) †

1
LCPM, Chemistry Department, Faculty of Sciences, University of Oran 1, Ahmed Benbella, Es-Senia, Oran 31000, Algeria
2
Research Centre in Analytical Chemistry and Physics (CRAPC), BP 248, Algiers RP, Algiers 16004, Algeria
3
Department of Materials Technology, Faculty of Physics, University of Science and Technology Mohamed Boudiaf, Oran 31000, Algeria
*
Author to whom correspondence should be addressed.
Presented at the 24th International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2020; Available online: https://ecsoc-24.sciforum.net/.
Chem. Proc. 2021, 3(1), 47; https://doi.org/10.3390/ecsoc-24-08349
Published: 14 November 2020

Abstract

:
The Aza-Wittig reaction plays an important role in organic transformations and has seen considerable development in the construction of cyclic and acyclic compounds in neutral solvents in the absence of catalysts with a high yield of products. Today, the use of iminophosphoranes (phosphazenes) has become a powerful tool in organic synthesis strategies directed toward the construction of nitrogenous heterocycles. These can react with carbonyl compounds, an excellent method for the construction of C=N double bonds via intermolecular and intramolecular Aza-Wittig reactions. In this study, we are interested in the theoretical study of the reaction path of the Aza˗Wittig reaction between trimethyl-iminophosphoranes (CH3)3P=NR, R = CH3 or Ph) for the two-substituent methyl and phenyl with acetaldehyde, including continuum solvation. Our calculations were carried out by means of ab initio calculations using DFT (Density Function Theory) with a dispersion correction term using the Grimme method in the program Gaussian09, using the functional B3LYP-GD3BJ with a 6-31G** base. The results obtained allowed us to highlight the mechanisms of the Aza-Wittig reaction in particular after the addition of the term of dispersion correction or van der Waals correction, which provides a new description of this reaction and its chemical path.

1. Introduction

Phosphazenes (iminophosphoranes) of the simple structure R3P=NRʹ were first described by Staudinger and Meyer in 1919 [1], but their applications only began in the 1950s. There are different names to designate this chemical function P=N, such as iminophosphorane, phosphinimine, λ-phosphazene, or phosphinimide.
However, compared to extensive studies of the Wittig reaction [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17], little attention has been paid to the Aza-Wittig reaction. In 1997, Jugo Koketsu and colleagues [18] are the first to investigate the Aza-Wittig reaction of iminopnictoranes (H3NMH, M = P, As, Sb, and Bi) with formaldehyde (H2CO) at the MP2 level. They came to the conclusion that the Aza-Wittig reaction was more favorable for phosphorus for As, Sb, and Bi.
The Aza-Wittig reaction is arguably the most important reaction in the iminophosphorane reactivity panel. To elucidate the mechanism of Aza-Wittig reactions, it is important to isolate their intermediates, although such isolations are generally difficult due to their instability. This reaction consists of a nucleophilic attack of the iminophosphoranes on the carbonyl (or thio-carbonyl) derivatives, followed by the removal of phosphine oxide (or phosphine sulfide). This reactivity very clearly resembles that of phosphorus ylides (Scheme 1), thus giving its name.
The mechanism of the Aza-Wittig reaction is broadly known; we know in particular that it has two stages with fairly similar speeds. On the other hand, little experimental data are available on the intermediates (betaine or oxazaphosphetidine) of the reaction path, even if some special cases have been studied and characterized [12,13]
The greatest advances in the mechanism of the Aza-Wittig reaction have been made by theoretical studies of the reaction path using post Hartree–Fock methods such as Møller–Plesset at orders 2 and 4 and DFT [14,15,16]
As part of this work, we studied the reaction of Aza-Wittig for the two substituents of R, methyl and phenyl (CH3)3=NR with acetaldehyde (C2H4O) (see Figure 1). However, to our knowledge, no theoretical study on the DFT-D calculation (van der Waals dispersion correction) has been published on the Aza-Wittig reaction. The present work can provide theoretical information on this reaction.

2. Computational Methods

All the calculations reported in this paper were performed within density functional theory using the functional B3LYP [19,20] and B3LYP-GD3BJ (empirical dispersion) with a base 6-31G** [21]. Each point of the reaction path was fully optimized and verified to be a transition state (TSX) with imaginary frequency or stable species with real frequencies (i.e., RC “reagents”, “intermediate” INX and PC “products”). Frequency analysis was performed to obtain thermochemical information about the reaction processes at 298 K. All calculations are performed using the Gaussian-09 program, and the resulting surfaces were visualized using GaussView 6. IRC (intrinsic reaction coordinate) calculations were also performed to confirm the nature of the transition states.
Single-point energy calculations were also carried out for all species involved in the reaction at gas phase B3LYP-GD3BJ/6-31G** optimized geometries in THF (tetrahydrofuran) solvent using the integral equation formalism polarizable continuum model (IEFPCM).

3. Results and Discussion

3.1. Reaction of Methylimino(trimethyl)phosphorane with Acetaldehyde

Using DFT calculations at the B3LYP/6-31G** and B3LYP˗GD3BJ/6˗31G** level, it was shown that the reaction of the Aza-Wittig continues through (CH3)3P = NCH3 + O=CH2CH3 → RC → TS1 → IN1 → TSr → IN2 → TS2 → PC → (CH3)3P = O + CH3N=CH2CH3, where RC and PC represents a reactive and product complex, TS represents a transition state, and IN represents an intermediate state Scheme 2.
In our work, we found that the Aza-Wittig reaction between iminophosphorane and aldehyde is strongly exothermic and involves a stable [2 + 2] cycloaddition–cycloreversion sequence, with two intermediates IN1 and IN2 (Figure 2).
The first TS1 transition structure was located at only 10.44 and 8.73 kcal/mol, for B3LYP and B3LYP-D3, above the reactants, which corresponds to a supra–supra geometry (see Table 1). This first cycloaddition [2 + 2] turns out to be quite asynchronous (Table 2), with a calculated synchronicity of 0.93 and 0.94. This observation is consistent with the analysis performed by McEwen for the similarity reaction of Wittig on the binding order using a semi-empirical method, which showed that the new C–C bond is formed at about 40% in the TS, while the P–O binding is still negligible at this stage [22,23]. The addition barrier is high in the B3LYP functional compared to the addition barrier in B3LYP-D3, after incorporation of dispersion corrections.
We also found an IN1 reaction intermediate of oxazaphosphetidine, in which the P=O bond was 1.84. Our results also indicated that IN2 is a fairly stable reaction intermediate.
In the intermediary of IN1, the coordination geometry around phosphorus was approximately a trigonal bipyramid, with oxygen derived from the aldehyde group found at the apical position. This then underwent pseudo-rotation on a low barrier to give an IN2 isomer with the iminophosphorane nitrogen in the apical position (IN2) (see Table 3).
The IRC calculations performed for the three TSs (TS1, TSr, TS2) showed that the localized TSs are well connected to the intermediates and final products. Indeed, optimizing the geometry of the last structure obtained on the IRC curve in the intermediate direction to IN1, IN2, and products gave a structure identical to the intermediate IN1, IN2, and final product.
The cycloreversion step [2 + 2] of the reaction took place from IN2 to TS2 had an activation energy greater than that calculated for the first step in Table 4; the incorporation of dispersion correction in functional B3LYP decreased the energy of activation corresponding to cycloaddition–cycloreversion of oxazaphosphetidine. With the effects on the molecular structure of the products being minor, this difference can be explained by the emergence of a dipole–dipole interaction between the two reagents at the addition. Note that the addition of the dispersion correction in the Hamiltonian more precisely in the exchange and correlation equation, for the functional B3LYP, slightly stabilized the structures of intermediates via cycloaddition. Thus, the addition process is more favored kinetically.
On the other hand, for the energy calculation including dispersion correction, B3LYP-D3/6˗31G** (THF) seems to underestimate the energy barriers by −0.69 kcal/mol compared to that calculated at B3LYP-D3/6-31G**. The energy values for the addition steps with addition of solvent and dispersion correction were estimated to be −13.57 kcal/mol. Lower energy barriers suggest that the inclusion of dispersion correction and solvent had a significant effect on the reaction. For the second stages, the energy values were −15.86 kcal/mol.
Table 5 shows that the negatively charged iminophosphorane nitrogen was attracted to the carbon atom of the positive charge aldehyde in TS1 because there was a wide region of negative electrostatic potential around the nitrogen atom, which confirmed the negative charge in the carbon; this charge decreased in the intermediate oxazaphosphetidine IN1. These potential electrostatic surfaces also support the initial stages of the reaction path in which a nucleophilic attack of nitrogen against the carbon aldehyde occurred.

3.2. Reaction of Phenylimino(trimethyl)phosphorane with Acetaldehyde

Structure, geometric data and relative energies corresponding to the reaction of phenylimino(trimethyl)phosphorane ((CH3)3P=NPh) with acetaldehyde (CH3C=O) at B3LYP and B3LYP˗D3/6˗31G** are given in Table 6, Table 7 and Table 8.
The first transition state TS had an almost planar structure TS1, in which the dihedral angle ΦPNCO = −7.29 and −8.33 (for B3LYP-D3); for the second transition state, TS2 took the values of ΦPNCO = 1.77 and 8.90. As shown in Figure 3, the second TS barrier was lower compared to TS1, which may explain kinetically why the reaction with phenyl as a substituent in the Aza-Wittig reaction was more stable and why the reaction took place rapidly in the second elimination step, i.e., cyclorevision of imine. Therefore, the Aza-Wittig reaction can be much more kinetically favorable for CH3 in the first cycloaddition step and less favorable in the cycloreversion step, in contrast to the case where the phenyl substituent is bonded in imionphosphrane, with regard to the thermodynamic stability reaction with CH3 which is more favored over phenyl.
The relative energies, enthalpy, and free energy for intermediates and transition states were more stable for functional B3LYP-D3 compared to B3LYP, without dispersion correction. Neglecting dispersion correction with solvent appeared to lead to a huge overestimation of the activation barriers by a single point at B3LYP-D3/6-31G** (THF). Thus, inclusion of dispersion correction significantly lowered the activation barriers for the reaction (see Figure 3).
The activation energies of addition and elimination are reported in Table 9. The addition TS1 have an energy high than the elimination TS2 (the direct barriers to the decomposition of the IN2 are less than the return barrier IN1). This means that the oxaphosphetane formation step must be irreversible.
Table 10 shows that the negatively charged iminophosphorane nitrogen was attracted to the positive charge of the carbon atom of aldehyde in TS1, because there was a wide region of negative electrostatic potential around the nitrogen atom, which confirmed the negative charge in the carbon; this charge decreased in the intermediate oxazaphosphetidine IN1. These potential electrostatic surfaces also support the initial stages of the reaction path in which a nucleophile attack of nitrogen against the carbon aldehyde occurred.

4. Conclusions

The Aza-Wittig reaction of trimethyliminophosphorane (CH3)3P=NR for both methyl and phenyl substituents was investigated by means of ab initio calculations using DFT and DFT˗D (dispersion correction) on the basis of 6-31G**. We conclude in this study that the Aza-Wittig reaction between phosphazenes and aldehydes takes place via a [2 + 2] cycloaddition–cycloreversion mechanism, with oxazaphosphetidines being the reaction intermediates. The two processes [2 + 2] are associated with thermally authorized asynchronous processes.
Methylimino(trimethyl)phosphorane is more reactive than phenylimino(trimethyl)phosphorane and less reactive in the second stage of elimination, i.e., cyclorevision of imine. Intermediate oxazaphosphetidines are four-membered rings in which the nitrogen atom is pyramidal. However, the energy barriers associated with the rotational movement are localized between the two intermediates; therefore, only the second stage leads to the imines and phosphine oxides. The addition of the dispersion correction brings a new description of the reaction and the path of the chemical reaction, because it makes it possible to find a better description of the dispersion interactions in the description of the corrected functions using the DFT method.

Author Contributions

Calculation resources: A.A. and H.S. All authors have read and agreed to the published version of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available in the manuscript.

Acknowledgments

The authors are deeply grateful to the University of Oran (Algeria) for computing resources and computer time used on Haytham at the UCI (Unité de Calcul Intensif), and to the University of Reims Champagne Ardennes (URCA, France) for computing resources and computer time used on Romeo calculator and Al-Farabi Cluster computer of the Ecole Nationale Polytechnique Oran—Maurice Audin.

References

  1. Staudinger, H. New organic compounds of phosphorus. III. Phosphine-methylene derivatives and phosphinimines. Helv. Chim. Acta 1919, 2, 635–646. [Google Scholar] [CrossRef]
  2. Naito, T.; Nagase, S.; Yamataka, H. Theoretical study of the structure and reactivity of ylides of N, P, As, Sb, and Bi. J. Am. Chem. Soc. 1994, 116, 10080–10088. [Google Scholar] [CrossRef]
  3. Volatron, F.; Eisenstein, O. Wittig versus Corey-Chaykovsky Reaction. Theoretical study of the reactivity of phosphonium methylide and sulfonium methylide with formaldehyde. J. Am. Chem. Soc. 1987, 109, 1–14. [Google Scholar] [CrossRef]
  4. Streitwieser Jr, A.; Rajca, A.; McDowell, R.S.; Glaser, R. Semipolar phosphorus-oxygen and phosphorus-carbon bonds. A theoretical study of hypophosphite and related methylenephosphoranes. J. Am. Chem. Soc. 1987, 109, 4184–4188. [Google Scholar] [CrossRef]
  5. Dixon, D.A.; Smart, B.E. The structures and energetics of fluorine-substituted phosphonium ylides. J. Am. Chem. Soc. 1986, 108, 7172–7177. [Google Scholar] [CrossRef]
  6. Eades, R.A.; Gassman, P.G.; Dixon, D.A. The conformations and energetics of simple ylides. J. Am. Chem. Soc. 1981, 103, 1066–1068. [Google Scholar] [CrossRef]
  7. Vincent, M.A.; Schaefer III, H.F.; Schier, A.; Schmidbaur, H. Molecular and electronic structure of phosphonium cyclopropylide: A theoretical study. J. Am. Chem. Soc. 1983, 105, 3806–3811. [Google Scholar] [CrossRef]
  8. Hoeller, R.; Lischka, H. A theoretical investigation on the model Wittig reaction PH3CH2+ CH2O. fwdarw. PH3O+ C2H4. J. Am. Chem. Soc. 1980, 102, 4632–4635. [Google Scholar] [CrossRef]
  9. Maryanoff, B.E.; Reitz, A.B. The Wittig olefination reaction and modifications involving phosphoryl-stabilized carbanions. Stereochemistry, mechanism, and selected synthetic aspects. Chem. Rev. 1989, 89, 863–927. [Google Scholar] [CrossRef]
  10. Lischka, H. Electronic structure and proton affinity of methylenephosphorane by ab initio methods including electron correlation. J. Am. Chem. Soc. 1977, 99, 353–360. [Google Scholar] [CrossRef]
  11. Nguyen, M.T.; Hegarty, A.F. An Ab initio study of the diadic prototropic tautomerism H3PX⇄ H2PXH (X= O, NH, CH2). J. Chem. Soc. Perkin Trans. 1987, 2, 47–54. [Google Scholar] [CrossRef]
  12. Bachrach, S.M. Molecular structure of phosphonium ylides. J. Org. Chem. 1992, 57, 4367–4373. [Google Scholar] [CrossRef]
  13. Sasaki, T.; Eguchi, S.; Okano, T. Novel synthesis and reactivity of 4-azahomoadamant-3-ene and 4-aza-4-homobrend-3-enes via intramolecular aza-Wittig reactions. J. Am. Chem. Soc. 1983, 105, 5912–5913. [Google Scholar] [CrossRef]
  14. Sheldrick, W.S.; Schomburg, D.; Schmidpeter, A.; von Criegern, T. Four-And Five-Membered Phosphorus Heterocycles. 39. Structural Changes in The (2+ 2) Cycloaddition of Ketones to Azaphospholes. Chem. Inf. 1980, 11, 304. [Google Scholar] [CrossRef]
  15. Molina, P.; Alajarin, M.; Lopez Leonardo, C.; Claramunt, R.M.; Foces-Foces, M.D.L.C.; Hernandez Cano, F.; Catalan, J.; De Paz, J.L.G.; Elguero, J. Experimental and theoretical study of the R3P+-X-bond. Case of betaines derived from N-iminophosphoranes and alkyl isocyanates. J. Am. Chem. Soc. 1989, 111, 355–363. [Google Scholar] [CrossRef]
  16. Lu, W.C.; Liu, C.B.; Sun, C.C. Theoretical study of the H3PNH+ H2CO reaction mechanism via five reaction channels. J. Phys. Chem. A 1999, 103, 1078–1083. [Google Scholar] [CrossRef]
  17. Lu, W.C.; Sun, C.C.; Zang, Q.J.; Liu, C.B. Theoretical study of the aza-Wittig reaction X3P = NH + O = CHCOOH → X3P = O+ HN= CHCOOH for X = Cl, H and CH3. Chem. Phys. Lett. 1999, 311, 491–498. [Google Scholar] [CrossRef]
  18. Koketsu, J.; Ninomiya, Y.; Suzuki, Y.; Koga, N. Theoretical study on the structures of iminopnictoranes and their reactions with formaldehyde. Inorganic Chemistry 1997, 36, 694–702. [Google Scholar] [CrossRef]
  19. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098. [Google Scholar] [CrossRef]
  20. 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 1988, 37, 785. [Google Scholar] [CrossRef]
  21. Hariharan, P.C.; Pople, J.A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  22. Marí, F.; Lahti, P.M.; McEwen, W.E. Molecular modeling of the Wittig olefination reaction: Part 2: A molecular orbital approach at the MNDO-PM3 level. Heteroat. Chem. 1991, 2, 265–276. [Google Scholar] [CrossRef]
  23. Mari, F.; Lahti, P.M.; McEwen, W.E. Molecular modeling of the Wittig reaction. 3. A theoretical study of the Wittig olefination reaction: MNDO-PM3 treatment of the Wittig half-reaction of unstabilized ylides with aldehydes. J. Am. Chem. Soc. 1992, 114, 813–821. [Google Scholar] [CrossRef]
Scheme 1. Scheme showing two possible paths of the Aza-Wittig reaction mechanism.
Scheme 1. Scheme showing two possible paths of the Aza-Wittig reaction mechanism.
Chemproc 03 00047 sch001
Figure 1. Optimized geometries of the reagents used in the Aza-Wittig reaction mechanism with (A)–(C) representing phosphazene with phenyl, phosphazene with methyl, and acetaldehyde, respectively.
Figure 1. Optimized geometries of the reagents used in the Aza-Wittig reaction mechanism with (A)–(C) representing phosphazene with phenyl, phosphazene with methyl, and acetaldehyde, respectively.
Chemproc 03 00047 g001
Scheme 2. Mechanism of Aza-Wittig reaction (methylimino(trimethyl)phosphorane with acetaldehyde).
Scheme 2. Mechanism of Aza-Wittig reaction (methylimino(trimethyl)phosphorane with acetaldehyde).
Chemproc 03 00047 sch002
Figure 2. Potential energy diagram of the reaction of Aza-Wittig of methylimino(trimethyl)phosphorane with acetaldehyde in gas phase B3LYP, B3LYP-D3 and with solvent B3LYP˗D3 (tetrahydrofuran (THF)).
Figure 2. Potential energy diagram of the reaction of Aza-Wittig of methylimino(trimethyl)phosphorane with acetaldehyde in gas phase B3LYP, B3LYP-D3 and with solvent B3LYP˗D3 (tetrahydrofuran (THF)).
Chemproc 03 00047 g002
Figure 3. Potential energy diagram of the reaction of Aza-Wittig of phenylimino(trimethyl)phosphorane with acetaldehyde in gas phase B3LYP, B3LYP-D3 and with solvent B3LYP˗D3 (THF).
Figure 3. Potential energy diagram of the reaction of Aza-Wittig of phenylimino(trimethyl)phosphorane with acetaldehyde in gas phase B3LYP, B3LYP-D3 and with solvent B3LYP˗D3 (THF).
Chemproc 03 00047 g003
Table 1. Relative energies, enthalpy, and Gibbs free energy (kcal/mol) calculated using DFT, at B3LYP and B3LYP-D3 with a base of 6-31G** (D: dispersion corrections of Grimme).
Table 1. Relative energies, enthalpy, and Gibbs free energy (kcal/mol) calculated using DFT, at B3LYP and B3LYP-D3 with a base of 6-31G** (D: dispersion corrections of Grimme).
Structures(B3LYP) (B3LYP-D3)
∆E (kcal/mol)∆G (kcal/mol)∆H (kcal/mol)∆E (kcal/mol)∆G (kcal/mol)∆H (kcal/mol)
Reactants (RC)0.000000.000.000.000.00
TS110.4414.7710.288.7312.178.52
IN 1−6.63−0.05−5.01−8.23−2.46−6.62
TSr−0.716.44−0.98−3.632.75−2.85
IN2−6.63−0.05−5.01−8.23−2.46−6.62
TS212.7517.9712.8011.2016.0611.32
Products (PC)−30.27−29.31−29.53−27.94−26.90−27.15
Table 2. Bond lengths (in Å) and dihedral angles (in degrees) of the Aza-Wittig reaction of Me3P=NCH3. RC, reactive complex; TS, transition state; IN, intermediate; PC, product complext.
Table 2. Bond lengths (in Å) and dihedral angles (in degrees) of the Aza-Wittig reaction of Me3P=NCH3. RC, reactive complex; TS, transition state; IN, intermediate; PC, product complext.
B3LYP/6-31G**
B3LYP-D3/6-31G**
dP–N
Ångström
dC–O
Ångström
dP–O
Ångström
dN–C
Ångström
ΦPNCO
Reactants (RC)1.58
1.58
1.22
1.22
3.63
3.45
3.07
2.84
−2.42
−0.50
TS11.64
1.63
1.28
1.28
2.69
2.70
1.76
1.76
−7.30
−6.75
IN11.72
1.71
1.40
1.40
1.84
1.84
1.47
1.46
−6.73
−6.95
TSr 1.81
1.81
1.42
1.43
1.73
1.73
1.45
1.45
−11.76
−11.66
IN21.72
1.71
1.40
1.40
1.84
1.84
1.47
1.46
−6.75
−6.95
TS22.42
2.41
1.69
1.69
1.59
1.59
1.36
1.36
8.58
10.20
Products (PC)4.13
3.89
3.76
3.53
1.51
1.51
1.27
1.27
−2.11
−2.01
Table 3. Optimized geometries of structures involved in the reaction pathway.
Table 3. Optimized geometries of structures involved in the reaction pathway.
Chemproc 03 00047 i001 Chemproc 03 00047 i002
Reactants (RC)TS1
Chemproc 03 00047 i003 Chemproc 03 00047 i004
IN1TSr
Chemproc 03 00047 i005 Chemproc 03 00047 i006
IN2TS2
Chemproc 03 00047 i007
Products (PC)
Table 4. Activation energy corresponds to cycloaddition–cycloreversion of oxazaphosphetidine for B3LYP, B3LYP-D3.
Table 4. Activation energy corresponds to cycloaddition–cycloreversion of oxazaphosphetidine for B3LYP, B3LYP-D3.
Δ G B 3 L Y P Δ G B 3 L Y P D
RC → TS114.7712.17
IN2 → TS218.0118.52
Table 5. Maps of total electron density and calculated electrostatic potential for TS1 and intermediates IN1; regions of highest electron density distribution are displayed in intense red (−16.31 kcal/mol) and regions of density distribution lower electronics in intense blue (+16.31 kcal/mol).
Table 5. Maps of total electron density and calculated electrostatic potential for TS1 and intermediates IN1; regions of highest electron density distribution are displayed in intense red (−16.31 kcal/mol) and regions of density distribution lower electronics in intense blue (+16.31 kcal/mol).
TS1 Chemproc 03 00047 i008
IN1 Chemproc 03 00047 i009
Table 6. Bond lengths (in Å) and dihedral angles (in degrees) of the Aza-Wittig reaction of methylimino(trimethyl)phosphorane with acetaldehyde (CH3)3P=NPh.
Table 6. Bond lengths (in Å) and dihedral angles (in degrees) of the Aza-Wittig reaction of methylimino(trimethyl)phosphorane with acetaldehyde (CH3)3P=NPh.
B3LYP/6-31G**
B3LYP-D3/6-31G**
dP–N
Ångström
dC–O
Ångström
dP–O
Ångström
dN–C
Ångström
ΦPNCO
Reactants (RC)1.59
1.59
1.22
1.22
3.60
3.46
3.12
2.89
−3.38
0.72
TS11.65
1.64
1.29
1.29
2.55
2.57
1.74
1.73
−7.29
−8.33
IN 11.73
1.72
1.40
1.40
1.84
1.84
1.47
1.47
−5.42
−5.35
TSr1.93
1.91
1.45
1.45
1.68
1.68
1.44
1.43
−4.88
−5.37
IN21.92
1.90
1.44
1.44
1.68
1.69
1.44
1.44
−3.55
−3.71
TS22.60
2.50
1.62
1.58
1.59
1.60
1.38
1.39
1.77
8.90
Products (PC)4.05
3.75
3.21
2.97
1.51
1.51
1.28
1.28
9.05
11.62
Table 7. Optimized geometries of structures involved in the reaction pathway.
Table 7. Optimized geometries of structures involved in the reaction pathway.
Chemproc 03 00047 i010 Chemproc 03 00047 i011
Reactants (RC)TS1
Chemproc 03 00047 i012 Chemproc 03 00047 i013
IN1TSr
Chemproc 03 00047 i014 Chemproc 03 00047 i015
IN2TS2
Chemproc 03 00047 i016
Products (PC)
Table 8. Relative energies, enthalpy, and Gibbs free energy (kcal/mol) calculated using DFT, at the theoretical level B3LYP and B3LYP-D3 with a base of 6-31G**.
Table 8. Relative energies, enthalpy, and Gibbs free energy (kcal/mol) calculated using DFT, at the theoretical level B3LYP and B3LYP-D3 with a base of 6-31G**.
Structures(B3LYP) (B3LYP-D3)
∆E (kcal/mol)∆G (kcal/mol)∆H (kcal/mol)∆E (kcal/mol)∆G (kcal/mol)∆H (kcal/mol)
Reactants (RC)0000.000.000.00
TS115.8819.0515.5913.7616.3013.40
IN 11.697.363.03−0.113.691.15
TSr2.189.213.000.236.080.99
IN20.667.512.15−1.304.370.14
TS210.2314.8710.248.8513.318.98
Products (PC)−24.35−23.29−23.67−22.55−21.48−21.89
Table 9. Activation energy corresponding to cycloaddition–cycloreversion of oxazaphosphetidine for functional B3LYP, B3LYP-D3.
Table 9. Activation energy corresponding to cycloaddition–cycloreversion of oxazaphosphetidine for functional B3LYP, B3LYP-D3.
Δ G B 3 L Y P Δ G B 3 L Y P D
RC → TS119.0516.30
IN2 → TS27.358.93
Table 10. Maps of total electron density and calculated electrostatic potential for TS1 and intermediates IN1; regions of highest electron density distribution are displayed in intense red (−16.31 kcal/mol) and regions of density distribution lower electronics in intense blue (+16.31 kcal/mol).
Table 10. Maps of total electron density and calculated electrostatic potential for TS1 and intermediates IN1; regions of highest electron density distribution are displayed in intense red (−16.31 kcal/mol) and regions of density distribution lower electronics in intense blue (+16.31 kcal/mol).
TS1 Chemproc 03 00047 i017
IN1 Chemproc 03 00047 i018
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Adda, A.; Sediki, H. Theoretical Study of the Aza˗Wittig Reaction, Me3P=NR (R = Methyl or Phenyl) with Aldehyde Using the DFT and DFT-D Methods (Dispersion Correction). Chem. Proc. 2021, 3, 47. https://doi.org/10.3390/ecsoc-24-08349

AMA Style

Adda A, Sediki H. Theoretical Study of the Aza˗Wittig Reaction, Me3P=NR (R = Methyl or Phenyl) with Aldehyde Using the DFT and DFT-D Methods (Dispersion Correction). Chemistry Proceedings. 2021; 3(1):47. https://doi.org/10.3390/ecsoc-24-08349

Chicago/Turabian Style

Adda, Abdelghani, and Hayat Sediki. 2021. "Theoretical Study of the Aza˗Wittig Reaction, Me3P=NR (R = Methyl or Phenyl) with Aldehyde Using the DFT and DFT-D Methods (Dispersion Correction)" Chemistry Proceedings 3, no. 1: 47. https://doi.org/10.3390/ecsoc-24-08349

Article Metrics

Back to TopTop