Next Article in Journal
Luminescent Ln-Ionic Liquids beyond Europium
Next Article in Special Issue
Cascade Transformations of 1-R-Ethynyl-9,10-anthraquinones with Amidines: Expanding Access to Isoaporphinoid Alkaloids
Previous Article in Journal
Synthesis and Electrochemical Performance of Electrostatic Self-Assembled Nano-Silicon@N-Doped Reduced Graphene Oxide/Carbon Nanofibers Composite as Anode Material for Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conformational Analysis of N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of Diphenylphosphorylacetic Acid: Dipole Moments, IR Spectroscopy, DFT Study

by
Anastasiia Kuznetsova
1,
Denis Chachkov
2,
Oleg Artyushin
3,
Natalia Bondarenko
4 and
Yana Vereshchagina
1,*
1
Department of Physical Chemistry, A.M. Butlerov Institute of Chemistry, Kazan Federal University, Kremlevskaya 18, 420008 Kazan, Russia
2
Kazan Department of Joint Supercomputer Center of Russian Academy of Sciences, Branch of Federal Scientific Center “Scientific Research Institute for System Analysis of the RAS”, Lobachevskogo 2/31, 420111 Kazan, Russia
3
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, Vavilova 28, 119991 Moscow, Russia
4
Institute of Chemical Reagents and High Purity Chemical Substances of NRC “Kurchatov Institute”, Bogorodsky val 3, 107076 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(16), 4832; https://doi.org/10.3390/molecules26164832
Submission received: 6 July 2021 / Revised: 30 July 2021 / Accepted: 4 August 2021 / Published: 10 August 2021

Abstract

:
Experimental and theoretical conformational analysis of N-methyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide, N-butyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide, and N-octyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide was carried out by the methods of dipole moments, IR spectroscopy, and Density Functional Theory (DFT) B3PW91/6-311++G(df,p) calculations. In solution, these N,N-dialkyl substituted bisphosphorylated acetamides exist as a conformational equilibrium of several forms divided into two groups—with Z- or E-configuration of the carbonyl group and alkyl substituent, and syn or anti arrangement of the phosphoryl-containing fragments relative to the amide plane. The substituents at the phosphorus atoms have eclipsed cis- or staggered gauche-orientation relative to the P=O groups, and cis orientation of the substituents is due to the presence of intramolecular H-contacts P=O...H−Cphenyl or p,π conjugation between the phosphoryl group and the phenyl ring. Preferred conformers of acetamides molecules are additionally stabilized by various intramolecular hydrogen contacts with the participation of oxygen atoms of the P=O or C=O groups and hydrogen atoms of the methylene and ethylene bridges, alkyl substituents, and phenyl rings. However, steric factors, such as a flat amide fragment, the bulky phenyl groups, and the configuration of alkyl bridges, make a significant contribution to the realization of preferred conformers.

1. Introduction

Derivatives of carbamoylmethylphosphine oxides (CMPO) are known as effective extractants of rare earth elements and actinides from mineral acid solutions [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17]. The high coordinating ability of these compounds is due to the presence of polar phosphoryl and amide groups. CMPO are used in industrial processes for reprocessing spent nuclear fuel [18,19] and for the processing of radioactive waste and the separation of transplutonium elements [20,21], as well as for the preparation of composite materials for column chromatography [22,23,24]. Carbamoylmethylphosphine oxides modified with a dialkylamidomethyl coordination group in the methylene bridge are potential neuroprotective agents [25,26]. Manganese (II) complexes with CMPO ligands exhibit luminescent properties [27]. Palladium complexes containing phosphoryl-substituted CMPO as ligand demonstrate catalytic activity in Suzuki reaction [28].
Although carbamoylmethylphosphine oxides have found active practical application for a long time, their spatial structure has been poorly studied. The literature contains data on the structure of complexes of CMPO with rare earth elements and actinides [17,29,30], manganese (II) [27], and mercury (II) [31] in the solid state. The crystal structures of (diethylcarbamoyl)methyldiphenylphosphine oxide and its protonated form [32], diphenylmorpholine carbamoylmethyl phosphine oxide [33], diphenyl-N,N-dimethylcarbamoylmethylphosphine oxide [34], N-aryl-substituted CMPO modified with a phosphoryl group, and its complexes with Pd (II) and Re (I) [35] were determined by X-ray diffraction. Information on the structure in solution is available only for P(X)-modified (X = O, S) N-aryl-substituted CMPOs: experimental and theoretical conformational analysis by the methods of dipole moments and quantum chemistry showed that in solution, these compounds exist as an equilibrium of several conformers with intramolecular hydrogen bonds H⋯X [36]. The deficiency of data on the CMPO’s conformational behavior is a problem in describing the coordinating properties of such compounds and explaining the reaction mechanisms with their participation. N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of diphenylphosphorylacetic acid were synthesized recently [37], but their polarities and conformational structures have not been studied.
In the present work, we investigated the spatial structure of N-methyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide 1, N-butyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide 2, and N-octyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide 3 (Scheme 1) by the methods of dipole moments, IR spectroscopy, and quantum chemistry DFT B3PW91/6-311++G(df,p).

2. Results and Discussion

2.1. Methodology

The studied amides of phosphorylacetic acid (phosphorylacetamides) 13 are polyfunctional polar compounds. Two phosphoryl groups with bulky phenyl substituents at the phosphorus atom are linked through alkyl bridges by an amide group; there is also a flexible alkyl substituent at the nitrogen atom. All fragments of a molecule can freely rotate about single bonds, and functional groups can participate in intra- and intermolecular non-covalent interactions, which leads to the emergence of a large number of possible conformers.
The method of dipole moments is a sensitive instrument for the determination of structure and the study of fine features of spatial and electronic structure of polar organic and organoelement compounds in solution. To determine the experimental values of the dipole moments, we used the second Debye method based on the measurement of the dielectric permittivity of the dilute solutions of the polar substance in a nonpolar solvent [38].
To search the possible conformations of isolated molecules of 13, we applied the Density Functional Theory (DFT) with B3PW91 hybrid functional. This method has been successfully used to study the polarity and spatial structure of similar organophosphorus compounds with double bond phosphorus–chalcogen and aryl substituents [36,39]. The choice of the B3PW91 method was also based on the data [40].

2.2. Polarity of Phosphorylacetamides 13

We have determined previously unknown polarities of compounds 13. The experimental values of the dipole moments were determined in benzene solutions using the second Debye method based on the measurement of the dielectric constant of the dilute solutions of a polar substance in a nonpolar solvent. The experimental dipole moments of 13 are listed in Table 1; their values are sufficiently high (Table 1) and are typical for the polarities of the compounds of tetra-coordinated phosphorus (2.5–5.0 Debye) [41].

2.3. Experimental and Theoretical Conformational Analysis of Phosphorylacetamides 13

Conformers for the compounds were built using the Gauss View 6.0 imaging software. All possible conformations for compounds 13 were built by sequential rotation of the parts of the molecules relative to single bonds. Conformations with overlapping atoms or with a too close arrangement, which does not correspond to the physical meaning, were immediately discarded. Thus, there were about a thousand conformers. First, the calculations were carried out for them using small basis set 6-31G(d). Based on these results, we have discarded degenerate structures. Then, conformers with high relative energies (more than 15 kJ/mol) were discarded as unlikely. At the next stage, the calculations were carried out using the extended basis set 6-311++G(df,p). We considered the preferred conformers with relative energies less than 6 kJ/mol without taking into account the mirror isomers with identical energies and theoretical dipole moments. For all preferred conformers, their relative energies and theoretical polarities were computed, and the dipole moments were calculated according to the vector-additive scheme (Table 2). Conformational diversity of the compounds 13 is due to the presence of a large number of rotation axes and several functional groups in their molecules. The percentage of conformers in the mixture was calculated on the basis of the theoretical values of Gibbs energies (Table 2).
According to the results of theoretical calculations, five energetically preferred conformers were found for N-methylacetamide 1 (Figure 1); their characteristics are listed in Table 2 and Table 3. In conformers 1a1e, the phosphorus atoms have a pyramidal structure, the phenyl substituents have cis- or gauche-orientation, and the methyl or ethyl bridges have gauche-orientation relative to the P=O bonds (α, β, γ, ι, κ, and λ angles in Table 3).
The amide fragment is flat, the carbonyl and methyl groups have Z-orientation relative to the Csp2–N bond in conformers 1a, 1c, 1e, and E-orientation in 1b and 1d; the bulky diphenylphosphoryl fragments are syn-located relative to amide plane in conformers 1a1c and anti-located in 1d and 1e (Figure 1).
The conformers 1a1e differ in the arrangement of the following bonds relative to each other: the P1–Csp3 and C=O bonds have synclinal orientation in 1c and anticlinal orientation in all other conformers; the N–Csp3 and Csp3–P2 bonds are practically coplanar in conformers 1a1d (antiperiplanar orientation), and they have anticlinal orientation in 1e. The Csp3–Csp3 and N–Csp3(methyl) bonds have synclinal orientation in conformers 1a, 1c, 1e, and anticlinal orientation in 1b and 1d; that is, the ethyl fragment is practically perpendicular to the amide plane (Table 3). The N–Csp3 and Csp3–P2 bonds lie in the same plane in 1a1e (θ angle, Table 3).
All conformers 1ae are additionally stabilized by intramolecular hydrogen contacts (Figure 1, Table 4). In conformers 1b and 1d, there is an interaction between the oxygen atom of the phosphoryl group P1=O and the hydrogen atom of the methyl substituent, resulting in the formation of seven-membered intramolecular heterocycles. The bifurcate hydrogen bond between the oxygen atom of the P1=O group and the hydrogen atoms of the ethylene bridge and one of the phenyl rings at the second diphenylphosphoryl group is observed in 1a, 1c, and 1e, resulting in the formation of seven- or eight-membered intramolecular cycles. In addition, in conformer 1c, H-contacts between the oxygen atom of the second phosphoryl group P2=O and the hydrogen atom of one of the phenyl substituents at the P1 atom (C2–H1⋯O2=P 2 2.391 Å, C2–H1⋯O2 156°), as well as between the oxygen atom of the carbonyl group and the hydrogen atom of the other phenyl substituent at the P1 atom (C5–H2⋯O3–C7 2.300 Å, C5–H2⋯O3 141°) are observed. The second bifurcate bond is formed by the oxygen atom of the P2 = O group and hydrogen atoms of the phenyl ring at the first phosphorus atom and the methylene bridge (C2–H1⋯O2 = P2 2.304 Å, C2–H1⋯O2 167° and C6–H3⋯O2=P2 2.283 Å, C6–H3⋯O2 164° respectively) in conformer 1e.
Elongation of the alkyl substituent at the nitrogen atom led to an increase in the number of energetically preferred conformers of N-butyl-N-[(2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide 2 to 6 (Figure 2); their characteristics are listed in Table 2 and Table 3. The common features of conformers 2af are a pyramidal structure of phosphorus atoms and a planar amide fragment, which are also characteristic for conformers 1ae. In conformers 2af, the phenyl substituents are predominantly cis-oriented, and the methyl or ethyl bridges are gauche-orientated relative to the P = O bonds (α, β, γ, ι, κ, and λ angles in Table 3).
The carbonyl group and butyl substituent at the nitrogen atom have mutual Z-orientation in conformers 2a, 2c, 2d, and E-orientation in 2b, 2e, and 2f (ε angle, Table 3); the bulky diphenylphosphoryl fragments are syn-located relative to amide plane in conformers 2ae and anti-located in 2f (Figure 2).
The difference between conformers is due to the different mutual arrangement of the following bonds: P1–Csp3 and Csp2=O bonds have synclinal orientation in 2c and anticlinal orientation in all other forms (δ angle, Table 3). The Csp3–Csp3 and N–Csp3(butyl) bonds have synclinal orientation in 2a and 2d, and anticlinal orientation in 2b, 2c, 2e, and 2f (η angle, Table 3); the N–Csp3 and Csp3–P2 bonds are antiperiplanar in 2a2f (θ angle, Table 3).
An additional variety of conformers 2af compared to compound 1 is due to the presence of a longer alkyl substituent. However, the difference between preferred conformers is only due to the rotation relative to the N–Csp3(R) bond, because conformers with a zigzag alkyl radical possess low energies. The Csp2–N and Csp3(R)–Csp3(R) bonds have synclinal orientation in 2a, 2c, and 2d and anticlinal orientation in 2b, 2e, and 2f (ζ angle, Table 3).
For conformers 2ad and 2f, an additional stabilization effect is observed due to the formation of intramolecular hydrogen contacts (Table 4). In 2b and 2f, the interaction occurs between the oxygen atom of the group P1=O and one of the hydrogen atoms of the butyl substituent, resulting in the formation of seven-membered intramolecular rings. In conformers 2a, 2c, 2d, and 2f, a bifurcate hydrogen bond is observed between the oxygen atom of the group P1=O and the hydrogen atoms of the ethylene bridge and one of the phenyl rings at the P2 atom, resulting in the formation of seven- or eight-membered intramolecular cycles. In addition to the bifurcate bond, a contact between the oxygen atom of the second phosphoryl group and the hydrogen atom of one of the phenyl substituents at the P1 atom is formed in conformer 2c (C2–H1⋯O2=P2 2.257 Å, C2–H1⋯O2 176°).
According to the quantum chemical calculations, for N-octyl-N-[(2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide 3, six energetically preferred conformers were found (Figure 3); their characteristics are listed in Table 2 and Table 3. As for compounds 1 and 2, conformers 3af are characterized by a pyramidal structure of the phosphorus atoms and a flat structure of the nitrogen atom. In conformers 3af, the phenyl substituents are predominantly cis-oriented and the methyl or ethyl bridges are gauche-orientated relative to the P=O groups (α, β, γ, ι, κ, and λ angles in Table 3).
As in the case of butyl substituted amide 2, the carbonyl group and octyl substituent at the nitrogen atom have mutual Z-configuration in conformers 3a, 3c, 3d, and E-configuration in 3b, 3e, and 3f (dihedral angle ε, Table 3); the diphenylphosphoryl fragments are syn-located relative to amide plane in conformers 3ae and anti-located in 3f (Figure 2).
Conformers 3ae differ in the structure of the carbamoylmethyl fragment and the ethylene bridge; the P1–Csp3 and C=O bonds are synclinal in 3c, while in the rest of the conformers, these bonds are anticlinal (dihedral angle δ, Table 3). The Csp3–Csp3 and N–Csp3(R) bonds are synclinal in 3a and 3d, and anticlinal in 3b, 3c, 3e, and 3f (dihedral angle η, Table 3). The N–Csp3 and Csp3–P2 bonds have antiperiplanar mutual arrangement in conformers 3af (dihedral angle θ, Table 3).
An additional difference between conformers is due to the rotation of the octyl substituent relative to the N–Csp3(R) bond. The Csp2–N and Csp3(R)–Csp3(R) bonds have mutual synclinal arrangement in 3a, 3c, and 3d and anticlinal arrangement in 3b, 3e, and 3f (dihedral angle ζ Table 3).
The intramolecular hydrogen contacts are possible in conformers 3a3d and 3f (Table 4). In conformers 3b and 3f, an interaction between the oxygen atom of the P1=O group and one of the hydrogen atoms of the octyl substituent is observed, resulting in the formation of seven-membered intramolecular cycles. The bifurcate hydrogen bonds between the oxygen atom of the first phosphoryl group and the hydrogen atoms of the ethylene bridge and one of the phenyl rings at the P2 atom arise in conformers 3a, 3c, 3d, and 3f. Moreover, a contact between the oxygen atom of the second phosphoryl group and the hydrogen atom of one of the phenyl substituents at the P1 atom is formed in conformer 3c (C2–H1⋯O2 = P2 2.273 Å, C2–H1⋯O2 176°).
It should be noted that in conformers 1a, 2a, and 2d, 3a, and 3d, the phosphoryl group P1=O and one of the phenyl substituents at the P1 atom have a completely eclipsed cis-orientation (α angle, Table 3); their coplanar arrangement promotes p,π-conjugation in the molecules and additionally stabilizes these conformers (Figure 4 and Figure S1). A similar fact was previously described for (arylcarbamoylmethyl)diphenylphosphine oxides and sulfides [38].
Conformers 1c, 1e, 2c, and 3c have the smallest dipole moments, both theoretical and calculated by the vector-additive scheme (exaltation between the experimental and calculated values—Δμ is 3.38, 1.75, 2.85, and 2.49 D, respectively—see Table 1 and Table 2), which can presumably be explained by the largest number of intramolecular hydrogen contacts in these forms, causing a decrease in their dipole moments. However, these low-polarity forms have slightly higher Gibbs energies, and their content in conformational equilibrium is insignificant (Table 2).
We have registered the IR spectra of compounds 13 in the solid state, in the melt, and in solution of the trichloromethane (Figure 5, Figures S2 and S3). The comparison of IR spectra of 13 showed a change in the number of bands in the range of 700–800 cm−1, corresponding to the out-of-plane deformation vibrations of the C–H bonds in the phenyl substituents. For compound 1, three absorption bands are observed in the melt, whereas seven bands are observed in the solid sample: the 717 cm−1 band splits in two (717 cm−1 and 724 cm−1), and new bands at 737 cm−1, 758 cm−1, and 770 cm−1 appear. In the case of compound 2, there are also three absorption bands in the melt and seven bands in the solid state: the 717 cm−1 band splits in two (715 cm−1 and 723 cm−1), the 793 cm−1 band splits in two (790 cm−1 and 795 cm−1), and new bands at 740 cm−1 and 774 cm−1 appear. The IR spectra of 3 also contain three vibrational bands in the melt and six bands in the solid state. New bands at 760 cm−1, 777 cm−1, and 809 cm−1 appear. These IR spectral data indicate the presence of conformational equilibrium in acetamides 13.
It should be noted that there is no noticeable shift and no change in the number of vibration bands of the phosphoryl and carbonyl groups when changing the state from solid to liquid (Table 5). A split peak is observed for the stretching vibrations of the P=O bond in the spectrum of the solid sample of 1 as well as in simulated spectra of conformers 1ae, 2af, and 3af. The theoretical frequencies of the stretching vibrations of the C=O and P=O groups were calculated using the scaling factor 0.96 [42].
A comparison of experimental and calculated according to vector scheme dipole moments, IR data, and theoretical results showed that in solution, compounds 13 exist as a conformational equilibrium of several forms, among which highly polar ones prevail.
Preferred conformers 1ae, 2af, and 3af can be divided into two groups, with Z- or E-configuration of the carbonyl and alkyl groups relative to the Csp2–N bond in the planar amide fragment; in addition, the phosphoryl-containing fragments can be syn- or anti-oriented relative to the amide plane (Table 2, Figure 6).
According to NMR spectroscopy data (1H, 13C{1H}, 31P{1H}, in CDCl3 solutions), diphenylphosphinylacetic acid amides 13 exist as two conformers in a ratio of ≈2.5–3.0:1, depending on the structure of the alkyl substituent at the nitrogen atom [37]. However, it was not possible to establish the structure of these two conformers.
A comparative analysis of the obtained experimental and theoretical results and the data of NMR spectroscopy [37] allowed us to conclude that in solution, compounds 13 exist as two sets of conformers: in the first set, the C=O and alkyl groups have Z-configuration in rotamers 1a, 1c, and 1e; 2a, 2c, and 2d; 3a, 3c, and 3d, while in the second set, these groups have E-configuration in rotamers 1b and 1d; 2b, 2e, and 2d; 3b, 3e, and 3d, and in all forms except 1e, 2f, and 3f, the phosphoryl-containing fragments are syn-located relative to the plane of the amide fragment. We believe that the first set of conformers corresponds to the minor conformer, and the second set corresponds to the majority conformer, found by NMR spectroscopy data, which is confirmed by the total conformer ratio (Table 2). Elongation of the alkyl substituent at the nitrogen atom (methyl–butyl–octyl) in a series of 123 led to an increase in the number of preferred conformers from five for 1 to six for 2 and 3. However, the structures of the corresponding preferred conformers differ slightly, and steric factors—a flat amide fragment, the size of bulky phenyl groups, and the configuration of alkyl bridges—are important.

3. Materials and Methods

3.1. Materials

N-methyl-N-[(2-(diphenylphosphoryl)ethyl]diphenylphosphorylamide 1, N-butyl-N-[(2-(diphenylphosphoryl)ethyl] diphenylphosphorylacetamide 2, and N-octyl-N-[(2-(diphenylphosphoryl)ethyl] diphenylphosphorylacetamide 3 were synthesized according to the following procedure [37].

3.2. IR Spectroscopy

The infrared spectra of crystals were collected on a FTIR Bruker Vertex 70 spectrometer (Bruker, Ettlingen, Germany) (600–4000 cm−1) with a single reflection, germanium crystal ATR accessory (MIRacle, PIKE Technologies) purged under dry air to remove atmospheric water vapor. The interferograms were recorded with a resolution of 2 cm−1, 128 scans, and Fourier transformed using a Blackman–Harris apodization function. The thin films of molten compounds were produced by the heating of crystal between KBr plates in home-made electrical oven. Temperature was measured by the PT100 sensor and was kept constant using a PID controller to ensure a standard deviation smaller than 1 K. The solid phase of samples was produced by slow cooling of thin film of molten compound (liquid) between KBr plates. The crystallization of the films was observed visually between crossed polarizers. The comparison spectra of the solid phases after melting and followed crystallization and the crystals of the same compounds showed their identity. Thus, there was no decomposition of the samples studied upon melting. KBr cells were used with a spacer (0.2 mm) to achieve the best signal/noise ratio. Concentrations of compounds were varied from 0.05 to 0.1 mol/L. Chloroform purified using molecular sieves was used as a solvent.

3.3. Dipole Moments

The experimental values of the dipole moments were determined using the second Debye method [38]. Physical parameters of 13 were measured for series consisting of 4–6 solutions in benzene at 25 °C. The dielectric permittivities of solutions of 13 were determined on a BI-870 instrument (Brookhaven Instruments Corporation, New York, USA); the accuracy is ±0.01. The refractive indices of solutions were measured on a RA-500 refractometer (Kyoto Electronics, Kyoto, Japan); the accuracy is ±0.0001.
The experimental dipole moments were calculated by the Debye Equation (1) [38]:
μ = 0.01283 P or . T .
The orientation polarizabilities Por. were calculated by the Guggenheim-Smith Equation (2) [43,44]:
P or . = M d [ 3 α ( ε 0 + 2 ) 2 3 γ ( n 0 2 + 2 ) 2 ]
where M is the molecular weight of a substance, d is the solvent density, α and γ are slopes of the εiwi and n i 2 wi plots; εi, ni, and wi are the dielectric constant, refractive index, and mass fraction of the solute of the ith solution, respectively. Equations for α and γ and the εi–wi and n i 2 wi plots (Figures S4–S6) are given in Supplementary Materials.
In the calculations of dipole moments according to the vector-additive scheme, we used the theoretical geometry parameters and following dipole moments of bonds and groups: m(CPh→P) 1.09 D, calculated from μexp (C6H5)3P [41]; m(C=>O) 1.94 D [45]; m(P=>O) 2.94 D, calculated from μexp C6H5P=O [41]; m(Csp3→P) 0.83 D [41]; m(Csp3→Csp2) 0.75 D, calculated from μexp C6H5CH3 [38]; m(N→Csp2) 0.94 D, calculated from μexp CH3C(O)NH2 [38]; m(H→Csp3) 0.28 D [46].

3.4. Quantum Chemical Calculations

Quantum chemical calculations with full geometry optimization were performed using the density functional theory method B3PW91 [47,48] and the 6-311++G(df,p) [49] extended basis set (calculations of the molecules in vacuum) using a GAUSSIAN 09 software package [50]. In all cases, the geometric parameters of the molecules were fully optimized. The solvent has not been taken into account. Correspondence of the found stationary points to the energy minimums was confirmed by calculation of the second derivatives of energy with respect to the atom coordinates. All structures identified as energy minima were characterized by Hessians containing only positive frequencies.

4. Conclusions

We have determined the polarities and carried out experimental and theoretical conformational analysis of N-methyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide, N-butyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide, and N-octyl-N-[2-(diphenylphosphoryl)ethyl]diphenylphosphorylacetamide by the methods of dipole moments, IR spectroscopy, and DFT calculations.
In solution, compounds 13 exist as a conformational equilibrium of several forms with a predominance of highly polar rotamers. Preferred conformers 1ae, 2af, and 3af are divided into two groups in which the carbonyl group and alkyl substituent have Z- or E-configuration, and the phosphoryl-containing fragments are syn- or anti-located relative to the amide plane. The substituents at the phosphorus atoms have eclipsed cis- or staggered gauche-orientation relative to the phosphoryl groups, and the eclipsed cis-orientation of the substituents is due to the presence of intramolecular hydrogen bonds P = O⋯H–Cphenyl or p,π-conjugation between the phosphoryl group and the phenyl ring. Additional stabilization of preferred conformers of 13 is provided by various intramolecular hydrogen contacts with the participation of oxygen atoms of the phosphoryl or carbonyl groups and hydrogen atoms of the methylene and ethylene bridges, alkyl substituents, and phenyl rings. However, steric factors, such as a flat amide fragment, the size of bulky phenyl groups, and the configuration of alkyl bridges, make a significant contribution to the realization of preferred conformers.
The studied tridental phosphorylated acetamides are prospective extractants of various metal ions and ligands for the preparation of metal complexes with catalytic and biological activity. The results of the present study can be useful for the prediction and rationalization of properties and reactivity of modified carbamoylphosphine oxides.

Supplementary Materials

Figure S1: IR spectra of compound 2 in different aggregate states, Figure S2: IR spectra of compound 3 in different aggregate states, Figure S3. Visualization of the molecular orbitals for 1e (the interaction between the P = O bond and phenyl substituents is absent). The positive and negative potentials are shown as blue and red areas, respectively, Equations for α and γ calculations (Guggenheim–Smith equation) and the εiwi and n i 2 wi plots, Figure S4. The εiwi and n i 2 wi plots for compound 1. Figure S5. The εiwi and n i 2 wi plots for compound 2. Figure S6. The εi–wi and n i 2 wi plots for compound 3.

Author Contributions

Conceptualization, N.B. and Y.V.; investigation, A.K., D.C. and Y.V.; synthesis, N.B. and O.A.; methodology, Y.V.; software, D.C.; project administration, Y.V.; supervision, Y.V.; visualization, A.K. and Y.V.; writing—original draft, A.K., D.C. and Y.V.; writing—review and editing, D.C. and Y.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was performed under financial support by the Russian Foundation for Basic Research (grant 20-03-00119).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 13 are not available from the authors.

References

  1. Turanov, A.N.; Karandashev, V.K.; Artyushin, O.I.; Peregudov, A.S.; Khvostikov, V.A.; Bondarenko, N.A. Extraction Properties of Diphenyl{[N-(2-diphenylphosphinylethyl)-N-alkyl]carbamoylmethyl}phosphine Oxides in Nitric Acid Solutions. Russ. J. Inorg. Chem. 2020, 65, 905–913. [Google Scholar] [CrossRef]
  2. Turanov, A.N.; Karandashev, V.K.; Kharlamov, A.V.; Bondarenko, N.A.; Khvostikov, V.A. Extraction of lanthanides(III) from Perchlorate Solutions with Carbamoyl- and Phosphorylmethoxymethylphosphine Oxides and Tetrabutyldiglycolamide. Solvent Extr. Ion Exch. 2019, 37, 65–80. [Google Scholar] [CrossRef]
  3. Turanov, A.N.; Karandashev, V.K.; Yarkevich, A.N.; Khvostikov, V.A. Extraction of REE(III), U(VI), and Th(IV) from Nitric Acid Solutions with Diphenyl(dibutylcarbamoylmethyl)phosphine Oxide in the Presence of Quaternary Ammonium Bis[(trifluoromethyl)sulfonyl]imides. Radiochemistry 2019, 61, 694–699. [Google Scholar] [CrossRef]
  4. Turanov, A.N.; Karandashev, V.K.; Yarkevich, A.N.; Khvostikov, V.A. Extraction of Rare Earth Elements from Hydrochloric Acid Solutions with Carbamoylmethylphosphine Oxides in the Presence of Quaternary Ammonium Dinonylnaphthalenesulfonates. Russ. J. Inorg. Chem. 2019, 64, 822–827. [Google Scholar] [CrossRef]
  5. Werner, E.J.; Biros, S.M. Supramolecular ligands for the extraction of lanthanide and actinide ions. Org. Chem. Front. 2019, 6, 2067–2094. [Google Scholar] [CrossRef]
  6. Turanov, A.N.; Karandashev, V.K.; Yarkevich, A.N. Extraction of REE(III), U(VI), and Th(IV) from Nitric Acid Solutions with Carbamoylmethylphosphine Oxides in the Presence of Bis[(trifluoromethyl)sulfonyl]imide Ions. Radiochemistry 2018, 60, 170–176. [Google Scholar] [CrossRef]
  7. Turanov, A.N.; Karandashev, V.K.; Yarkevich, A.N. Extraction of Rare-Earth Elements from Hydrochloric Acid by Carbamoyl Methyl Phosphine Oxides in the Presence of Ionic Liquids. Russ. J. Inorg. Chem. 2018, 63, 406–413. [Google Scholar] [CrossRef]
  8. Leoncini, A.; Huskens, J.; Verboom, W. Ligands for f-element extraction used in the nuclear fuel cycle. Chem. Soc. Rev. 2017, 46, 7229–7273. [Google Scholar] [CrossRef]
  9. Turanov, A.N.; Karandashev, V.K.; Vinogradova, N.M.; Sharova, E.V.; Artyushin, O.I. Extraction Properties of P(O)-Modified N-Aryl-Carbamoylmethylphosphine Oxides in Nitric Acid Media. Solvent Extr. Ion Exch. 2014, 32, 408–423. [Google Scholar] [CrossRef]
  10. Odinets, I.L.; Artyushin, O.I.; Sharova, E.V.; Matveeva, E.V.; Turanov, A.N.; Karandashev, V.K. Extraction Properties of Oligodentate Carbamoylmethylphosphine Oxides and β-Aminophosphine Oxides Toward Lanthanides and Rhenium(VII). Phosphorus Sulfur Silicon Relat. Elem. 2013, 188, 146–149. [Google Scholar] [CrossRef]
  11. Turanov, A.N.; Karandashev, V.K.; Yarkevich, A.N. Extraction of REEs(III), U(VI), and Th(IV) from nitric acid solutions with carbamoylmethylphosphine oxides in the presence of an ionic liquid. Radiochemistry 2013, 55, 382–387. [Google Scholar] [CrossRef]
  12. Pitchaiah, K.C.; Sujatha, K.; Rao, C.V.S.B.; Subramaniam, S.; Sivaraman, N.; Rao, P.R.V. Supercritical fluid extraction of uranium and thorium from nitric acid medium using organophosphorous compounds. Radiochim. Acta 2015, 103, 245–255. [Google Scholar] [CrossRef]
  13. Sujatha, K.; Pitchaiah, K.; Sivaraman, N.; Nagarajan, K.; Srinivasan, T.; Rao, P.V. Recovery of plutonium from polymeric waste matrices using supercritical fluid extraction. Desalin. Water Treat. 2013, 52, 470–475. [Google Scholar] [CrossRef]
  14. Sasaki, Y.; Kitatsuji, Y.; Sugo, Y.; Tsubata, Y.; Kimura, T.; Morita, Y. Actinides Extractability Trends for Multidentate Diamides and Phosphine Oxides. Solvent Extr. Res. Dev. Jpn. 2012, 19, 51–61. [Google Scholar] [CrossRef]
  15. Lumetta, G.J.; Gelis, A.V.; Vandegrift, G.F. Review: Solvent Systems Combining Neutral and Acidic Extractants for Separating Trivalent Lanthanides from the Transuranic Elements. Solvent Extr. Ion Exch. 2010, 28, 287–312. [Google Scholar] [CrossRef]
  16. Alyapyshev, M.Y.; Babain, V.; Ustynyuk, Y.A. Recovery of minor actinides from high-level wastes: Modern trends. Russ. Chem. Rev. 2016, 85, 943–961. [Google Scholar] [CrossRef]
  17. Sharova, E.V.; Artyushin, O.I.; Nelyubina, Y.V.; Lyssenko, K.; Passechnik, M.P.; Odinets, I.L. Complexation of N-alkyl(aryl)- and N,N-dialkylcarbamoylmethylphosphine oxides with the f-elements. Russ. Chem. Bull. 2008, 57, 1890–1896. [Google Scholar] [CrossRef]
  18. Horwitz, E.P.; Schulz, W.W. The Truex Process: A Vital Tool for Disposal of U.S. Defense Nuclear Waste. In New Separation Chemistry Techniques for Radioactive Waste and Other Specific Applications; Springer: Dordrecht, The Netherlands, 1991; pp. 21–29. [Google Scholar]
  19. Sharova, E.V.; IArtyushin, O.; Odinets, I.L. Synthetic routes to carbamoylmethylphosphoryl compounds—Extractants for the processing of spent nuclear fuels. Russ. Chem. Rev. 2014, 83, 95–119. [Google Scholar] [CrossRef]
  20. Myasoedov, B.F.; Kalmykov, S.N.; Kulyako, Y.M.; Vinokurov, S.E. Nuclear fuel cycle and its impact on the environment. Geochem. Int. 2016, 54, 1156–1167. [Google Scholar] [CrossRef]
  21. Ustynyuk, Y.A.; Alyapyshev, M.Y.; Babain, V.; Ustynyuk, N.A. Quantum chemical modelling of extraction separation of minor actinides and lanthanides: The state of the art. Russ. Chem. Rev. 2016, 85, 917–942. [Google Scholar] [CrossRef]
  22. Kameník, J.; Šebesta, F.; John, J.; Böhmer, V.; Rudzevich, V.; Grüner, B. Polyacrylonitrile based composite materials with extracting agents containing chemically bonded CMPO groups for separation of actinoids. J. Radioanal. Nucl. Chem. 2015, 304, 313–319. [Google Scholar] [CrossRef]
  23. Kameník, J.; Šebesta, F. Study of europium and selected actinides uptake on composite material CMPO-PAN. Czechoslov. J. Phys. 2006, 56, D493–D500. [Google Scholar] [CrossRef]
  24. Annam, S.; Rao, C.B.; Sivaraman, N.; Sivaramakrishna, A.; Vijayakrishna, K. Carbamoylmethylphosphine oxide functionalised porous crosslinked polymers towards sequential separation of uranium (VI) and thorium (IV). React. Funct. Polym. 2018, 131, 203–210. [Google Scholar] [CrossRef]
  25. Yarkevich, A.N.; Safronova, Z.V.; Petrova, L.N.; Gabrelyan, A.V.; Zamoyski, V.L.; Grigor’Ev, V.V.; Bachurin, S.O.; Zefirov, N.S. Physiologically active bis(dialkylamides) of phosphoryl-substituted α,ω-dicarboxylic acids. Russ. J. Gen. Chem. 2013, 83, 41–45. [Google Scholar] [CrossRef]
  26. Yarkevich, A.N.; Brel, V.; Makhaeva, G.F.; Serebryakova, O.; Boltneva, N.; Kovaleva, N. Synthesis and investigation of biological activity of phosphorylated amines and amides. Russ. J. Gen. Chem. 2015, 85, 1644–1649. [Google Scholar] [CrossRef]
  27. Artem’Ev, A.V.; Berezin, A.S.; Brel, V.K.; Morgalyuk, V.P.; Samsonenko, D.G. Deep-red phosphorescent organic–inorganic hybrid Mn(II) complexes based on 2-(diphenylphosphoryl)-N,N-diethylacetamide ligand. Polyhedron 2018, 148, 184–188. [Google Scholar] [CrossRef]
  28. Vasil’Ev, A.A.; Aleksenko, V.Y.; Aleksanyan, D.V.; Kozlov, V.A. Catalytic activity of k3-X,N,Y-palladium pincer complexes (X, Y = O, S) with (thio)phosphoryl-substituted carbamoylmethylphosphine oxide and sulfide ligands in the Suzuki cross-coupling. Mendeleev Commun. 2013, 23, 344–346. [Google Scholar] [CrossRef]
  29. Caudle, L.; Duesler, E.; Paine, R. Formation and crystal and molecular structures of dioxouranium(VI) complexes containing bidentate carbamoylmethylphosphinate and phosphine oxide ligands. Inorg. Chim. Acta 1985, 110, 91–100. [Google Scholar] [CrossRef]
  30. Caudle, L.J.; Duesler, E.N.; Paine, R.T. Preparation and structure of a neodymium complex containing bidentate (carbamoylmethyl)phosphine oxide ligands. Inorg. Chem. 1985, 24, 4441–4444. [Google Scholar] [CrossRef]
  31. Rosario-Amorin, D.; Dehaudt, J.; Caudle, L.J.; Dickie, D.; Paine, R.T. Synthesis and molecular structures of mercury(II) complexes of carbamoylmethylphosphoryl ligands. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 520–526. [Google Scholar] [CrossRef]
  32. Antipin, M.Y.; Struchkov, Y.T.; Matrosov, E.I.; Kabachnik, M.I. Structure of organophosphorus compounds. Part XXX. X-ray structural investigation of diphenyl(diethylcarbamoylmethyl) phosphine oxide Ph2P(O)CH2C(O)NEt2 and its molecular complex with perchloric acid Ph2P(O)CH2C(O_NEt2·HClO4. J. Struct. Chem. 1985, 26, 441–446. [Google Scholar] [CrossRef]
  33. Das, D.; Sivaramakrishna, A.; Gopakumar, G.; Rao, C.B.; Sivaraman, N.; Vijayakrishna, K. Diphenylmorpholine CMPO: Synthesis, coordination behavior and extraction studies of actinides. Polyhedron 2018, 141, 215–222. [Google Scholar] [CrossRef]
  34. Yaftian, M.R.; Hassanzadeh, L.; Eshraghi, M.E.; Matt, D. Solvent extraction of thorium (IV) and europium (III) ions by diphenyl-N,N-dimethylcarbamoylmethylphosphine oxide from aqueous nitrate media. Sep. Purif. Technol. 2003, 31, 261–268. [Google Scholar] [CrossRef]
  35. Aleksenko, V.Y.; Sharova, E.V.; Artyushin, O.I.; Aleksanyan, D.V.; Klemenkova, Z.S.; Nelyubina, Y.V.; Petrovskii, P.V.; Kozlov, V.A.; Odinets, I.L. Coordination of P(X)-modified (X=O, S) N-aryl-carbamoylmethylphosphine oxides and sulfides with Pd(II) and Re(I) ions: Facile formation of 6,6-membered pincer complexes featuring atropisomerism. Polyhedron 2013, 51, 168–179. [Google Scholar] [CrossRef]
  36. Vereshchagina, Y.A.; Khanafieva, R.R.; Chachkov, D.V.; Artyushin, O.I.; Sharova, E.V.; Ishmaeva, E.A. Polarity and structure of P(X)-modified (X = O, S) arylcarbamoylmethylphosphine oxides and sulfides. Russ. J. Org. Chem. 2016, 52, 1413–1418. [Google Scholar] [CrossRef]
  37. Bondarenko, N.A.; Belus’, S.K.; Artyushin, O.I.; Peregudov, A.S. New Tridentate Carbamoylmethylphosphine Oxides: Synthesis and NMR Spectra. Russ. J. Gen. Chem. 2020, 90, 2273–2280. [Google Scholar] [CrossRef]
  38. Osipov, O.A.; Minkin, V.I.; Garnovskii, A. Spravochnik po dipol’nym momentam. In Handbook of Dipole Moments; Publishing House Vysshaya Shkola: Moscow, Russia, 1971; p. 416. [Google Scholar]
  39. Kuznetsova, A.A.; Ismagilova, R.R.; Chachkov, D.V.; Belogorlova, N.A.; Malysheva, S.F.; Vereshchagina, Y.A. Experimental and Theoretical Conformational Analysis of Tris(4-methylphenyl)phosphine and Its Chalcogenides. Russ. J. Org. Chem. 2020, 56, 2098–2103. [Google Scholar] [CrossRef]
  40. Medvedev, M.G.; Bushmarinov, I.S.; Sun, J.; Perdew, J.P.; Lyssenko, K.A. Density functional theory is straying from the path toward the exact functional. Science 2017, 355, 49–52. [Google Scholar] [CrossRef]
  41. Ishmaeva, E.A.; Timosheva, A.P.; Timosheva, N.V.; Vereshchagina, Y.A. Spravochnik po dipol’nym momentam fosfororganicheskih soedinenii. In Handbook of Dipole Moments of Organophosphorus Compounds; Publishing House of Kazan University: Kazan, Russia, 1998; p. 120. [Google Scholar]
  42. Johnson, R. Computational Chemistry Comparison and Benchmark Database (CCCBDB). Available online: https://cccbdb.nist.gov/ (accessed on 4 August 2021).
  43. Guggenheim, E.A. A proposed simplification in the procedure for computing electric dipole moments. Trans. Faraday Soc. 1949, 45, 714–720. [Google Scholar] [CrossRef]
  44. Smith, J.W. Some developments of Guggenheim’s simplified procedure for computing electric dipole moments. Trans. Faraday Soc. 1950, 46, 394–399. [Google Scholar] [CrossRef]
  45. Zimin, M.G.; Galeeva, R.M.; Pudovik, A.N. Dipole moments of organophosphorus compounds. Russ. Chem. Bull. 1971, 20, 473–477. [Google Scholar] [CrossRef]
  46. Gribov, L.A.; Popov, E.M. Valence-optical method and theoretical investigation of intensities and polarizations of fundamentals of polyatomic molecules in absorption spectra. Dokl. Akad. Nauk. SSSR 1962, 145, 761–763. [Google Scholar]
  47. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  48. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef] [PubMed]
  49. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  50. 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]
Scheme 1. Structures of compounds 13, selected atom numbers are given.
Scheme 1. Structures of compounds 13, selected atom numbers are given.
Molecules 26 04832 sch001
Figure 1. Preferred conformers of 1 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Figure 1. Preferred conformers of 1 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Molecules 26 04832 g001aMolecules 26 04832 g001b
Figure 2. Preferred conformers of 2 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Figure 2. Preferred conformers of 2 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Molecules 26 04832 g002aMolecules 26 04832 g002b
Figure 3. Preferred conformers of 3 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Figure 3. Preferred conformers of 3 according to DFT calculations (some hydrogen atoms are omitted). Distances are given in Å.
Molecules 26 04832 g003aMolecules 26 04832 g003b
Figure 4. Visualization of the molecular orbitals № 102 for 1a, № 112 for 2a, № 112 for 2d, № 124 for 3a, and № 123 for 3d corresponding to the interaction of the P=O bond and phenyl substituents. The positive and negative wave functions are shown as blue and red areas, respectively. The isocontour value is 0.02 a.u. The frontier orbitals are not given, since we do not consider the reactivity of the compounds in this context.
Figure 4. Visualization of the molecular orbitals № 102 for 1a, № 112 for 2a, № 112 for 2d, № 124 for 3a, and № 123 for 3d corresponding to the interaction of the P=O bond and phenyl substituents. The positive and negative wave functions are shown as blue and red areas, respectively. The isocontour value is 0.02 a.u. The frontier orbitals are not given, since we do not consider the reactivity of the compounds in this context.
Molecules 26 04832 g004
Figure 5. IR spectra of compound 1 in different aggregate states. D is the absorbance.
Figure 5. IR spectra of compound 1 in different aggregate states. D is the absorbance.
Molecules 26 04832 g005
Figure 6. Structures of preferred conformers of 13.
Figure 6. Structures of preferred conformers of 13.
Molecules 26 04832 g006
Table 1. Coefficients of calculated equations (α, γ), orientation polarizabilities (Por.), and experimental dipole moments (μ) of compounds 13 in benzene.
Table 1. Coefficients of calculated equations (α, γ), orientation polarizabilities (Por.), and experimental dipole moments (μ) of compounds 13 in benzene.
CompoundαγPor., cm3μ, D
16.1060.303547.825.15
25.0700.214496.844.90
34.0910.259432.324.77
Table 2. Relative energies (ΔE), Gibbs free energies (ΔG, 298.15 K) (kJ/mol), theoretical (μtheor) and calculated by the vector-additive scheme (μcalc)) dipole moments (D), percentage (n) of preferred conformers of 13 according to DFT B3PW91/6-311++G(df,p).
Table 2. Relative energies (ΔE), Gibbs free energies (ΔG, 298.15 K) (kJ/mol), theoretical (μtheor) and calculated by the vector-additive scheme (μcalc)) dipole moments (D), percentage (n) of preferred conformers of 13 according to DFT B3PW91/6-311++G(df,p).
ConformerΔEΔGμtheorμcalcn, %
1aZ, syn0.00.05.674.7541.3
1bE, syn1.50.45.445.1535.3
1cZ, syn2.98.21.721.771.5
1dE, anti3.71.65.535.8421.5
1eZ, anti5.611.53.743.400.4
2aZ, syn0.02.15.484.7423
2bE, syn1.60.05.225.3355
2cZ, syn3.09.82.352.051
2dZ, syn3.45.75.714.776
2eE, syn5.16.05.545.515
2fE, anti5.44.35.305.8010
3aZ, syn0.01.85.444.7521
3bE, syn1.50.05.185.4343
3cZ, syn2.99.32.342.281
3dZ, syn3.45.05.694.326
3eE, syn5.03.85.555.449
3fE, anti5.32.05.275.8020
Table 3. Selected dihedral angles (degree) for the preferred conformers of 1–3 according to DFT B3PW91/6-311++G(df,p): α O1 = P1–C1–C2, β O1 = P1–C3–C4, γ O1 = P1–C6–C7, δ P1–C6–C7 = O3, ε O3 = C7–N–C15, ζ C7–N–C15–C16, η C15–N–C8–C9, θ N–C8–C9–P2, ι C8–C9–P2 = O2, κ O2 = P2–C10–C11, λ O2 = P2–C13–C14.
Table 3. Selected dihedral angles (degree) for the preferred conformers of 1–3 according to DFT B3PW91/6-311++G(df,p): α O1 = P1–C1–C2, β O1 = P1–C3–C4, γ O1 = P1–C6–C7, δ P1–C6–C7 = O3, ε O3 = C7–N–C15, ζ C7–N–C15–C16, η C15–N–C8–C9, θ N–C8–C9–P2, ι C8–C9–P2 = O2, κ O2 = P2–C10–C11, λ O2 = P2–C13–C14.
αβγδεζηΘικλ
1a−12282930-−84168−65−147
1b6247599−178-100−17960831
1c59124789−2-−73−16641725
1d72577104173-−94178−61−30−9
1e6115711003-80134−583617
2a−1228292087−84167−65−157
2b6257497−179−97100−17860831
2c59125976−3−84−92−15349725
2d−2227995−1−77−79168−67−157
2e7267498−179−97100−17960931
2f62477100−179−102−109179−60−30−8
3a−1228192−187−84168−65−147
3b6257497−179−96100−17860831
3c58136076−3−84−91−15349705
3d−1237995−1−77−79167−68−156
3e82675103179121108−17960831
3f62577100−179−101−108179−60−29−7
Table 4. Selected geometry characteristics of hydrogen bonds (distance, Å; angle, degrees) according to DFT B3PW91/6-311++G(df,p).
Table 4. Selected geometry characteristics of hydrogen bonds (distance, Å; angle, degrees) according to DFT B3PW91/6-311++G(df,p).
H5⋯O1C9–H5⋯O1H6⋯O1C12–H6⋯O1
1a2.1941532.339164
1b2.247 *151 *
1c2.1601442.305157
1d2.236152
1e2.175 **152 **2.422154
2a2.1891542.341164
2b2.271 *153 *
2c2.1521622.255172
2d2.2481472.323163
2e2.259 *153 *
2f2.1891542.350163
3a2.272 *153 *
3b2.1621612.248172
3c2.2501472.323163
3d2.260 *153 *
3f2.1941532.339164
* H7⋯O1, C15–H7⋯O1; ** H4⋯O1, C8–H4⋯O1.
Table 5. Selected vibration frequencies of the phosphoryl and carbonyl groups of 13; theoretical values are listed for conformers a–f.
Table 5. Selected vibration frequencies of the phosphoryl and carbonyl groups of 13; theoretical values are listed for conformers a–f.
P=OC=O
123123
νexp solid1179, 118811871181164816421632
νexp melt119411941195164316401641
νexp solution118411841182163916391634
νtheor a1178, 11871177, 1182, 11941179, 1184, 1187165016441644
νtheor b1187, 11921190, 11931188, 1192, 1193164216391639
νtheor c11931162, 11751162, 1175164116521652
νtheor d1171, 11811179, 1181, 11911179, 1184164716431643
νtheor e1180, 11831190, 11941191, 1194165916321632
νtheor f-1190, 11931192, 1193-16351635
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kuznetsova, A.; Chachkov, D.; Artyushin, O.; Bondarenko, N.; Vereshchagina, Y. Conformational Analysis of N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of Diphenylphosphorylacetic Acid: Dipole Moments, IR Spectroscopy, DFT Study. Molecules 2021, 26, 4832. https://doi.org/10.3390/molecules26164832

AMA Style

Kuznetsova A, Chachkov D, Artyushin O, Bondarenko N, Vereshchagina Y. Conformational Analysis of N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of Diphenylphosphorylacetic Acid: Dipole Moments, IR Spectroscopy, DFT Study. Molecules. 2021; 26(16):4832. https://doi.org/10.3390/molecules26164832

Chicago/Turabian Style

Kuznetsova, Anastasiia, Denis Chachkov, Oleg Artyushin, Natalia Bondarenko, and Yana Vereshchagina. 2021. "Conformational Analysis of N-Alkyl-N-[2-(diphenylphosphoryl)ethyl]amides of Diphenylphosphorylacetic Acid: Dipole Moments, IR Spectroscopy, DFT Study" Molecules 26, no. 16: 4832. https://doi.org/10.3390/molecules26164832

Article Metrics

Back to TopTop