Next Article in Journal
Photoinduced DNA Cleavage and Photocytotoxic of Phenanthroline-Based Ligand Ruthenium Compounds
Next Article in Special Issue
The Genus Haplophyllum Juss.: Phytochemistry and Bioactivities—A Review
Previous Article in Journal
Investigation of Antifungal Mechanisms of Thymol in the Human Fungal Pathogen, Cryptococcus neoformans
Previous Article in Special Issue
NMR and Computational Studies as Analytical and High-Resolution Structural Tool for Complex Hydroperoxides and Diastereomeric Endo-Hydroperoxides of Fatty Acids in Solution-Exemplified by Methyl Linolenate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Calculations of 1H NMR Chemical Shifts of Geometric Isomers of Conjugated Linolenic Acids, Hexadecatrienyl Pheromones, and Model Triene-Containing Compounds: Structures in Solution and Revision of NMR Assignments

by
Themistoklis Venianakis
,
Christina Oikonomaki
,
Michael G. Siskos
*,
Alexandra Primikyri
and
Ioannis P. Gerothanassis
*
Section of Organic Chemistry and Biochemistry, Department of Chemistry, University of Ioannina, GR-45110 Ioannina, Greece
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(11), 3477; https://doi.org/10.3390/molecules26113477
Submission received: 18 May 2021 / Revised: 2 June 2021 / Accepted: 4 June 2021 / Published: 7 June 2021
(This article belongs to the Special Issue Theme Issue in Honor of Professor Atta-Ur-Rahman, FRS)

Abstract

:
A DFT study of the 1H NMR chemical shifts, δ(1H), of geometric isomers of 18:3 conjugated linolenic acids (CLnAs), hexadecatrienyl pheromones, and model triene-containing compounds is presented, using standard functionals (B3LYP and PBE0) as well as corrections for dispersion interactions (B3LYP-D3, APFD, M06–2X and ωB97XD). The results are compared with literature experimental δ(1H) data in solution. The closely spaced “inside” olefinic protons are significantly more deshielded due to short-range through-space HH steric interactions and appear close to or even beyond δ-values of aromatic systems. Several regularities of the computational δ(1H) of the olefinic protons of the conjugated double bonds are reproduced very accurately for the lowest-energy DFT-optimized single conformer for all functionals used and are in very good agreement with experimental δ(1H) in solution. Examples are provided of literature studies in which experimental resonance assignments deviate significantly from DFT predictions and, thus, should be revised. We conclude that DFT calculations of 1H chemical shifts of trienyl compounds are powerful tools (i) for the accurate prediction of δ(1H) even with less demanding functionals and basis sets; (ii) for the unequivocal identification of geometric isomerism of conjugated trienyl systems that occur in nature; (iii) for tackling complex problems of experimental resonance assignments due to extensive signal overlap; and (iv) for structure elucidation in solution.

Graphical Abstract

1. Introduction

Conjugated linolenic acids (CLnAs) are a group of positional and geometric isomers of octadecatrienoic acids (C18:3) that contain three conjugated double bonds primarily in positions Δ9,11,13 and Δ8,10,12 [1,2]. α-Eleostearic acid (α-ESA) has three conjugated double bonds (C18:3Δ9 cis, 11 trans, 13 trans) (Figure 1) and it is produced and stored in the seed oil of plants such as A. fordii, M. charantia, Parinarium spp., and P. mahaleb. α-ESA was reported to have anticancer properties on several tumor cell lines, such as A549 (lung), MCF-7 (breast), DLD1 (colorectal), MKN-7 (stomach), and HepG2 (hepatoma). Lipid peroxidation has been suggested as the underlying mechanism [3]. β-Eleostearic acid (β-ESA) (C18:3Δ9 trans, 11 trans, 13 trans) (Figure 1) is present in pomegranate, bitter gourd, and catalpa [4]. β-ESA was reported to have a stronger antiproliferative effect than the geometric isomer α-ESA. A decrease in Bcl-2 and an increase in Bax mRNA expression along with DNA fragmentation were observed, which indicates different signaling pathways than their cis isomers [5]. Punicic acid (PA) (C18:3Δ9 cis, 11 trans, 13 cis; Figure 1) is a fatty acid derived from the fruit of P. granatum (aspomegranate) and from T. kirilowii. PA has been reported to have several health benefits such as anticancer activity and the prevention of obesity and insulin resistance in mice [6,7,8]. The cytotoxic properties on human monocytic leukemia cells have been attributed to lipid peroxidation [9].
Molecules with three conjugated double bonds have also been identified in hexadecatrienyl systems (Figure 1) from several lepidopterous species [10]. These sex hormones from females attract male moths and, thus, have significant roles in forest ecology, human health, and communication biology [11]. Pheromone-baited traps, therefore, have been widely used for the detection, population monitoring, and control of pest species and for assessing the efficacy of eradication efforts [12]. Isolation and identification of sex pheromones, however, is a difficult task due to the very small amount of pheromone produced by the insects and the small number of individuals available for the analysis of the pheromone gland content [12]. Initially, the isomeric configuration of the natural trienals was unknown and their identity was investigated by synthesis and the use of NMR spectroscopy [13,14,15]. Several geometric isomers, however, are not amenable to complete resonance assignment due to extensive signal overlap.
Hexadecatrienyl pheromones and their synthetic analogues [11,13,14,15], conjugated linolenic acids (CLnAs) [16,17,18,19,20], and model triene-containing compounds [21,22,23,24] have been extensively investigated using 1H and 13C NMR. 2D-NOESY experiments, 13C–1H COSY correlations, and analysis of spin–spin coupling constants with the use of homonuclear decoupling techniques were used to identify the geometric configurations of the trienyl conjugated double bonds. However, in several cases, severe overlap in the NMR spectra leads to equivocal signal assignments even when using 2D spectra and, consequently, to ambiguities in the spectral interpretations.
Several studies have been published that combine experimental NMR chemical shifts with computations for tackling the complex problems of resonance assignment, resonance reassignment, and structural revision [25,26,27,28,29,30,31,32,33,34,35,36,37] and for investigating high-resolution structures in solution [38,39,40,41,42,43,44,45,46,47]. DFT calculations of NMR chemical shifts in conjugated systems, however, are limited to 1H and 13C NMR chemical shifts of retinal isomers [48], and geometric isomers of diene-containing compounds [47]. Since no X-ray structures of conjugated linolenic acids (CLnAs) and hexadecatrienyl pheromones have so far been published, it would be of interest to use quantum chemical calculations of δ(1Η) for structure elucidation in solution. In this paper, we discuss (i) the effect of various functionals and basis sets on the accuracy of the DFT calculation of 1H NMR chemical shifts using the GIAO [49] technique on several geometric conjugated linolenic acids, hexadecatrienyl pheromones, and model triene-containing compounds (Figure 1); (ii) the use of δ(1H) for the unequivocal assignment of geometric isomerism and revision of literature experimental assignments; and (iii) the use of δ(1H) as a tool for structure elucidation in solution.

2. Results and Discussion

2.1. DFT-Calculated vs. Experimental 1H NMR Chemical Shifts of Model Compounds in Solution: Effects of Various Functionals and Basis Sets

The experimental 1H NMR chemical shifts of the trienyl model compounds (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene and (E,E,E)-2,4,6-octatriene (Figure 1 and Table 1) exhibit resonances of the olefinic protons of the conjugated double bonds in the range of 5.0 to 6.8 ppm. The =CH2 protons are shielded with respect to the rest of the olefinic protons and appear in a narrow range of 5.05 to 5.23 ppm. Alkyl substitution of the H1a proton, as in the case of (E,Z,E)-2,4,6-octatriene, results in a deshielding of ~0.5 ppm, which is in excellent agreement with 1H additive contribution to ethylene of Δδgem = 0.45 ppm [50]. Interestingly, the closely spaced “inside” olefinic 2,5 protons of (Z)-1,3,5-hexatriene and 3,6 protons of (E,Z,E)-2,4,6-octatriene (Figure 1) are the most deshielded and appear close to or even beyond δ-values of aromatic systems. This deshielding effect can be attributed to rigid geometries and significant HH van der Waals repulsive effects of specific protons [22]. It has been demonstrated that hydrogen atoms, which are subject to significant steric compression, exhibit a deshielding that is dependent on the geometrical relationship between the H–C bond and the interacting proximate hydrogen [51]. This occurs because the symmetry of the electron cloud about the proton is disturbed, thus resulting in reduced electronic charge and diamagnetic shielding [52,53]. Trienyl model compounds, therefore, are ideally suited for the study of steric effects on δ(1H) using various functionals and basis sets. Energy minimization of the structures was performed with standard functionals (B3LYP and PBE0) as well as with corrections for dispersion interactions (B3LYP-D3, APFD, M06–2X, and ωB97XD). Inclusion of nonlocal van der Waals density functionals has been shown to improve the accuracy of standard DFT functionals [54,55]. The 6–31+G(d) and 6–311++G(d,p) basis sets were used to compare the results of a low computational cost basis set with a medium-size one. Computations of δcalc(1H) were performed (a) using a single methodology at the GIAO/B3LYP/6–311+G(2d, p)/CPCM level (Figure S1 and Figure S2 and Table S1), as recommended in [28], and (b) using the same level of theory as geometry optimization (Figure 2, Figure S3, and Table S2).
For all the functionals and basis sets used, the literature experimental chemical shifts of the closely spaced “inside” olefinic protons with δ(H (3,6)) = 5.83 ppm and δ(H (4,5)) = 6.50 ppm of (E,Z,E)-2,4,6-octatriene [23] strongly deviate from linearity (Figure S2A and Figure S3A). Revision of the literature experimental chemical shift data so that the inner H(3,6) protons to be deshielded to a larger degree than H(4,5) (Figure S2B and Figure S3B) results in very significant improvements of the regression coefficients and standard deviations for all the functionals and basis sets used (Table S3 and Table S4). The use of several DFT functionals with distinct contributions (long-range corrections or dispersion interactions) would provide distinct results for the geometry. The use, however, of a single methodology at the GIAO/B3LYP/6–311+G(2d,p) level for chemical shift computations results in identical correlations coefficients and mean square errors for all the functionals and basis sets used (Table S3). The use of the same level of theory in δcalc (1H) as geometry optimization results in similar but not identical statistical data (Table S4). The quality of the linear regression procedure is a criterion whether the computational method is able to reproduce the experimental chemical shifts free from random error [28]. Furthermore, the extent to which the slope of the correlation line deviates from unity and the intercept from zero is a measure of the overall systematic error. Among the functionals with corrections for dispersion interactions, the ωB97XD with the 6–31+G(d) basis set performs better (intercept: 0.003, slope: 1.046; Table S4). The M06–2X functional is less suitable for shielding calculations of the selected molecules in agreement with literature data on several small molecules [56].
From the above, it can be concluded that the accuracy of the DFT calculations can provide a practical guide for future experimental assignments and help in the reassignment of literature experimental 1H NMR chemical shifts of triene-containing compounds based on a typical workflow that includes the following steps:
(i)
the 1H NMR spectra are recorded in, e.g., a CDCl3 solution at 298 K, and a preliminary assignment is performed using a variety of 1D and 2D NMR experiments;
(ii)
the 1H NMR chemical shifts are computed with the CPCM model at the same level of theory as geometry optimization or at the GIAO/B3LYP/6–311+G(2d,p) level, even with less demanding functionals and basis sets;
(iii)
a very good linear correlation between experimental NMR chemical shifts, δexp, and calculated shifts, δcalc, provides a strong indication that the assignment procedure is correct.

2.2. Out-of-Plane Deformation of the Trienyl System and Steric Effects in Model Compounds

Variation of the torsion angle φ(C1,C2,C3,C4) of the trienyl system of (Z)-1,3,5-hexatriene, in steps of 10° in the range of 180° to 0°, with energy minimization at the B3LYP/6–31+G(d) level, results in significant changes in the electronic energy (ΔE) (Figure 3A). A second minimum at φ = 40° was observed with ΔG = 2.37 kcal·mol−1 higher than that of the φ = 180° conformer. The significant ΔG energy difference of the two low-energy conformers with φ = 40° and φ = 180.0° shows that the δcalc(1H) data (at the GIAO/B3LYP/6–311+G(2d,p) level), weighted by the respective Boltzmann factor, are nearly identical with those of the φ = 180.0° conformer. The effect of the population of the φ = 40° conformer can, thus, be neglected. The effect of variation of the torsion angle φ on 1H NMR chemical shifts is shown in Figure 3B and Table S5. The δcalc(1H) data indicate similar behavior of the “inside” H2 and H5 protons with a pronounced shielding at 100° < φ < 180° (Figure 3B). In the range of torsion angles 0°< φ < 100°, the H5 proton shows a strong deshielding with a maximum value of δ = 7.32 ppm at φ = 0°. This clearly demonstrates that the strong, through-space, steric interaction with the H1b is the primary factor that results in δ values in the range of aromatic protons. The effect of variation of the torsion angle φ of (Z)-1,3,5-hexatriene on calculated olefinic 1H NMR chemical shifts using the same level of theory as geometry optimization (APFD/6–311++G(d,p)) is shown in Figure S4 and Table S6. The results are very similar to those of Figure 3 and Table S5; however, the ΔΕ and ΔG values are slightly smaller and the deshielding effect is slightly more pronounced due to the dispersion interaction of the APFD functional.
The electronic energy ΔΕ (kcal/mol) and shielding changes due to the variation in the torsion angle φ1 (C2C3C4C5) of the (E,Z,E)-2,4,6-octatriene (Figure 4, Table S7) are very similar to those due to the variation in the torsion angle φ of (Z)-1,3,5-hexatriene. Thus, the δcalc(1H) data of the low-energy conformers, weighted by the respective Boltzmann factor, are identical with those of the φ1 = 180° conformer. Again, the effect of variation of the φ1 torsion angle on δcalc(1H), using the same level of theory as geometry optimization (APFD/6–311++G(d,p)), is to decrease slightly the ΔΕ and ΔG values and increase the deshielding effect (Figure S5 and Table S8).
The effect of the variation in the torsion angle φ2(C1C2C3C4) of (E)-1,3,5-hexatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔΕ(kcal·mol−1), and δcalc(1H), at the GIAO/B3LYP/6–311+G(2d,p) level, are shown in Figure 5. A second broad minimum in the range of φ2 = 30° to 0° was observed with ΔE=3.66 kcal·mol−1G = 3.28 kcal·mol−1) higher than that of the φ2 = 180.0° conformer. The δcalc (1Η) values of the low-energy conformers with φ2 = 30° to 0° and φ2 = 180.0° (Table S9), weighted by the respective Boltzmann factors, are essentially the same as those of the φ2 = 180.0° conformer. This is due to the negligible population of the higher minimum conformers. Of particular interest is the parallel behavior of the H4 and H1b protons as a function of the torsion angle φ2, with strong deshielding in the range of 0° < φ2 < 90°. Again, the effect of the variation in the φ2 torsion angle on δcalc(1H), using the same level of theory as geometry optimization (APFD/6–311++G(d,p)), is to decrease slightly the ΔΕ and ΔG values and increase the deshielding effect (Figure S6 and Table S10).
Figure 6 shows the effect of variation in the torsion angle φ3(C2C3C4C5) of the (E,E,E)-2,4,6-octatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔΕ, and the computational olefinic 1H NMR chemical shifts (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3). As in the case of (E)-1,3,5-hexatriene, a second broad minimum in the range of φ3 = 30° to 0° was observed with ΔE = 3.68 kcal·mol−1G = 3.42 kcal·mol−1). The δexp(1H) data of the olefinic protons (Table S1) are in excellent agreement with the δcalc(1H) values of the minimum energy conformer with φ3 = 180.0o. The δcalc(1H) values of the low-energy conformers (Table S11), weighted by the respective Boltzmann factor, are essentially the same as those of the φ3 = 180.0° conformer. This demonstrates that the weights of the higher minimum conformers are of minor importance. As in the previous cases, the effect of variation in the φ3 torsion angle on δcalc(1H) of the olefinic protons, using the same level of theory as geometry optimization (APFD/6–311++G(d,p)), is to decrease slightly the ΔΕ and ΔG values and increase the deshielding effect (Figure S7 and Table S12).
It can, therefore, be concluded that δexp(1H) are reproduced very accurately for the lowest-energy DFT-optimized single conformer. The other low-energy conformers have negligible effects on the δcalc(1H) data of the conjugated olefinic protons. Furthermore, the minor functional dependence can result in the levels of accuracy that are necessary for structure elucidation in solution (see below).
Abraham et al. [50] showed that there is a deshielding effect above the C=C bond at short distances, due to the van der Waals term, and a deshielding effect in the plane of the C=C bond [57]. Figure S8 illustrates a plot of NBO bond order of the olefinic C-H bonds vs. δcalc(1H). The resulting very poor correlation (R2 = 0.554) shows that the NBO bond order is not a significant factor that determines δcalc(1H). Furthermore, no functional relationship was found for NBO charge densities of the olefinic C-H protons vs. δcalc(1H). Α detailed experimental and computational (with the CHARGE 7 model) study by Abraham et al. [50,57] showed a steric deshielding effect on both alkene and aromatic ring protons, contrary to a shielding effect of HH steric interactions in alkanes. All these steric interactions were described by a simple r−6 dependence,
δsteric = as/r6
where as is a constant of the given hybridization of the atom. Figure S9 and Figure S10 show that, in most cases, an r-n dependence is observed, which for HH distances shorter than those of the van der Waals radius of the hydrogen atom (~1.2 Å), becomes approximately a linear correlation. According to the semi empirical model of Cheney [51], a much better agreement with the experimental deshielding effect was obtained by correlation with f(ri, θi) than with the exponential exp(−ari) term, where ri is the distance between the interacting pair of hydrogens and θi is the angle between HH internuclear line and the H–C bond of interest. The above results clearly demonstrate that the deshielding effect due to non-bonded HH repulsive interaction is a primary factor affecting δ(1H) in the trienyl systems reported therein, as in the case of simple alkenes, conjugated dienyl, and aromatic systems [47,50,51,57].

2.3. DFT-Calculated vs. Experimental 1H NMR Chemical Shifts: Structure Elucidation in Solution of Geometric Isomers of 18:3 Conjugated Linolenic Acids and Hexadecatrienyl Pheromones

Table 2 shows conformational properties, ΔG values (kcal·mol−1) and % populations of various low-energy conformers of the 18:3 ω-5 CLA with energy minimization at the B3LYP/6–31+G(d) and APFD/6–31+G(d) levels. The structures of five low-energy conformers (A), (B), (C), (D), and (E) of β-eleostearic acid and the definition of various torsion angles are shown in Figure 7 and Figure 8, respectively. The general tendency of the torsion angles φ(C7C8C9C10) and φ(C13C14C15C16), which involve the allylic C(8) and C(14) carbons, respectively, is to adopt a low-energy gauche conformation, also known as skew (S = 120°) or skew′ (S′ = −120°). On the contrary, the eclipsed syn conformation (φ = 0°) results in high energy and, thus, low population (Table 2). The results are very similar for both B3LYP/6–31+G(d) and APFD/6–31+G(d) levels. An anti-zigzag conformational behavior was observed for the two polymethylene (CH2)n groups, which are attached to the conjugated trienyl system, for all conformers of the geometric isomers of 18:3 ω-5 CLA for both B3LYP/6–31+G(d) and APFD/6–31+G(d) levels (Table 2 and Figure S11).
Crystallographic data for cis-monounsaturated fatty acids [58,59,60] revealed that the zigzag (anti) conformation of the polymethylene chains (CH2)n is the most rigid and stable, whereas the gauche conformer is less stable, in agreement with our computational data. For the allylic carbons, the most stable conformations are skew (S) and skew′ (S′) in the γ-crystallization form, again in agreement with our computational data of Table 2. However, the anti-cis-anti and skew-cis-anti conformations have been observed in the α- and β-crystallization forms, respectively [58]. The formation of these high-energy conformers can be attributed to specific crystal packing interactions, which are absent in solution.
Figure S12 shows a graphical presentation of the calculated 1H NMR chemical shifts (at the GIAO/B3LYP/6–311+G(2d,p) (CPCM, CHCl3) level), weighted by the respective Boltzmann factors of the various conformers of Table 2 vs. experimental chemical shifts (Table 1 and Table S13) of the three geometric isomers of the 9,11,13-CLA. The experimental chemical shifts of the H11 and H12 of β-eleostearic acid [18] deviate from linearity. Revision of the literature assignment so that H12 is deshielded to a larger degree than H11 results in very good agreement with the computational data and significant improvement of the statistical data (Table S13 and Table S14). Similar results were obtained with calculations of 1H NMR chemical shifts using the same level of theory as geometry optimization (APFD/6–31+G(d) level) (Table S14).
Table 2 shows conformational properties, ΔG values (kcal·mol−1), and % populations of various low-energy conformers of the 10,12,14-conjugated hexadecatrienyl acetate geometric isomers with geometry optimization at the B3LYP/6–31+G(d) and APFD/6–31+G(d) levels. As in the case of CLnAs, the torsion angle φ(C7C8C9C10) adopts a low-energy skew (120°) or skew′ (120°) conformation. On the contrary, the eclipsed syn conformation with φ angles around 0o results in high energy and, thus, low population. Figure S13 and Table S15 and Table S16 show an improvement in the correlation of δcalc vs. δexp when the literature experimental chemical shifts of H13 and H15 of 9Z,11Z,13E and H13 and H14 of 9Z,11E,13E-conjugated hexadecatrienyl acetates (Table 1, [13]) have been revised (Figure S14). Similar results were obtained with the calculation of 1H NMR chemical shifts using the same level of theory as geometry optimization (APFD/6–31+G(d) level) (Table S15 and Table S16).
Figure 9 and Table 3 show that a significant improvement can be achieved in the linear regression correlation coefficient and mean square error of the correlation δcalc vs. δexp of the olefinic protons of all the compounds of Figure 1 when the literature experimental chemical shifts of the compounds shown in Figure S14 have been revised. This clearly demonstrates that the accuracy of the DFT 1H NMR shift prediction can be used to resolve ambiguities in resonance assignment.

3. Computational Methods

The computational study was performed using the Gaussian 09 Rev. D.01 software [61]. The initial structures of the model triethyl compounds (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene, and (E,E,E)-2,4,6-octatriene were drawn using the GaussView program and were energy minimized at the B3LYP/6–31+G(d), B3LYP/6–311++G(d,p), B3LYP-D3/6–31+G(d), B3LYP-D3/6–311++G(d,p) (the D3 version of Grimme dispersion with Becke–Johnson damping), APFD/6–31+G(d), APFD/6–311++G(d,p), PBE0/6–31+G(d), PBE0/6–311++G(d,p), M06–2X/6–31+G(d), M06–2X/6–311++G(d,p), ωB97XD/6–31+G(d), and ωB97XD/6–311++G(d,p) levels (gas phase). In the case of α-oleostearic acid (9Z, 11E, 13E), β-oleostearic acid (9E, 11E, 13E), punicic acid (9Z, 11E, 13Z), (10E, 12E, 14Z)-, (10E,12Z,14Z)-, (10Z, 12Z, 14E)-, and (10Z, 12E, 14E)-hexatrienyl acetates, a conformational analysis was, firstly, performed by rotation of the aliphatic carbons of the (CH2)n chains, which constitute with the rigid conjugated double bonds a flat system. The resulting conformers were energy minimized with DFT at the B3LYP/6–31+G(d) and APFD/6–31+G(d) levels, while the subsequent energy minimization of the selected stable structures was carried out at the same level of calculation. The optimized geometries were verified by performing frequency calculation at the same level (zero imaginary frequencies). The scanning of torsional angles was performed using the redundant coordinates in Gaussian 09 [47]. ΔG values were calculated from the thermochemistry results, either between the stable conformers or between the stable conformers and the corresponding transition states. The computed proton chemical shifts were performed using (i) a single methodology at the GIAO/B3LYP/6–311+G(2d,p) level and (ii) GIAO at the same level of theory as geometry optimization. In both cases, the conductor-like polarizable continuum model (CPCM) was used in CHCl3 or CCl4, for experimental values obtained in CDCl3 and CCl4, respectively (Table S1 and Table S2). δcalc(1Η) were referenced with respect to the standard TMS, which was optimized at the same level (Table S17).

4. Conclusions

From the DFT data of the trienyl conjugated compounds of Figure 1 reported therein, it can be concluded that:
(a)
Very good linear correlations can be obtained between DFT-calculated and experimental 1H NMR chemical shifts of the olefinic protons of the lowest-energy DFT-optimized single conformer using standard functionals (B3LYP and PBE0) as well as corrections for dispersion interactions (B3LYP-D3, APFD, M06–2X and ωB97XD). The ωB97XD performs slightly better, but again the accuracy of the functionals used was rather similar.
(b)
Through-space HH steric interaction is the primary factor that results in strong deshielding of closely spaced trienyl olefinic protons, in excellent agreement with literature data on alkenes and aromatic systems [50,51,57].
(c)
The accuracy of computational 1H NMR chemical shifts can facilitate (i) the unequivocal assignment of the geometric isomerism in conjugated trienyl systems of biological systems, such as CLnAs and hexadecatrienyl pheromones, and especially in the case of problematic resonance assignment due to extensive signal overlap, and (ii) structure elucidation in solution [28,36].

Supplementary Materials

Figure S1: Calculated 1H NMR chemical shifts vs. experimental chemical shifts, Figure S2: Calculated 1H NMR chemical shifts vs. experimental chemical shifts, Figure S3: Calculated 1H NMR chemical shifts vs. experimental chemical shifts, Figure S4: Effect of variation of the torsion angle φ (C1C2C3C4), Figure S5: Effect of variation of the torsion angle φ1(C2C3C4C5), Figure S6: Effect of variation of the torsion angle φ2(C1C2C3C4), Figure S7: Effect of variation of the torsion angle φ3 (C2C3C4C5), Figure S8: ΝΒO bond order of the olefinic C-H bonds vs. calculated 1H NMR chemical shifts, Figure S9: The dependence of δ(1H) vs. distance, Figure S10: The dependence of δ(1H) vs. distance, Figure S11: Structures of various conformers of the punicic acid, Figure S12: Graphical presentation of calculated 1H NMR chemical shifts vs. experimental values, Figure S13 Graphical presentation of calculated 1H NMR chemical shifts vs. experimental values, Figure S14: Experimental 1H NMR and calculated chemical shifts, Table S1: Experimental and computational 1H NMR chemical shifts, Table S2: Experimental and computational 1H NMR chemical shifts, Table S3: Statistical data of calculated vs. experimental 1H chemical shifts, Table S4: Statistical data of calculated vs. experimental 1H chemical shifts, Table S5: Effect of variation of the torsion angle φ(1C2C3C4) on calculated 1H NMR chemical shifts, Table S6: Effect of variation of the torsion angle φ(C1C2C3C4) on calculated 1H NMR chemical shifts, Table S7: Effect of variation of the C2C3C4C5 torsion angle on calculated 1H NMR chemical shifts, Table S8: Effect of variation of the C2C3C4C5 torsion angle on calculated 1H NMR chemical shifts, Table S9: Effect of variation of the C1C2C3C4 torsion angle on calculated 1H NMR chemical shifts, Table S10: Effect of variation of the C1C2C3C4 torsion angle on calculated 1H NMR chemical shifts, Table S11: Effect of variation of the C2C3C4C5 torsion angle on calculated 1H NMR chemical shifts, Table S12: Effect of variation of the C2C3C4C5 torsion angle on calculated 1H NMR chemical shifts, Table S13: Calculated and experimental 1H NMR chemical shifts, Table S14: Statistical analysis of the chemical shift data, Table S15: Calculated and experimental 1H NMR chemical shifts, Table S16: Statistical analysis of the chemical shift data, Table S17: Calculated shielding values of the reference TMS molecule.

Author Contributions

T.V., C.O., and M.G.S. performed DFT calculations; A.P. performed literature search; M.G.S. and I.P.G. conceived and designed the study and contributed to the writing of the manuscript. All authors have read and approved the final manuscript.

Funding

I.P.G. received funding from the Hellenic Foundation for Research and Innovation (H.F.R.I.) under the “First Call for H.F.R.I. Research Projects to support Faculty members and Researchers and the procurement of high-cost research equipment grant” (Project Number: 2050).

Conflicts of Interest

The authors declare that they have no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Yuan, G.F.; Chena, X.E.; Li, D. Conjugated linolenic acids and their bioactivities: A review. Food Funct. 2014, 5, 1360–1368. [Google Scholar] [CrossRef]
  2. Dubey, K.K.D.; Sharma, G.; Kumar, A. Conjugated linolenic acids: Implication in cancer. J. Agric. Food Chem. 2019, 67, 6091–6101. [Google Scholar] [CrossRef] [PubMed]
  3. Ιgarashi, M.; Miyazawa, T. Newly recognized cytotoxic effect of conjugated trienoic fatty acids on cultured human tumor cells. Cancer Lett. 2000, 148, 173–179. [Google Scholar] [CrossRef]
  4. Özgül-Yücel, S. Determination of conjugated linolenic acid content of selected oil seeds grown in Turkey. J. Am. Oil Chem. Soc. 2005, 82, 893–897. [Google Scholar] [CrossRef]
  5. Yasui, Y.; Hosokawa, M.; Kohno, H. Growth inhibition and apoptosis induction by all-trans-conjugated linolenic acids on human colon cancer cells. Anticancer Res. 2006, 1855–1860. [Google Scholar]
  6. Grossmann, M.E.; Mizuno, N.K.; Schuster, T.; Cleary, M.P. Punicic acid is an ω-5 fatty acid capable of inhibiting breast cancer proliferation. Int. J. Oncol. 2009, 36, 421–426. [Google Scholar]
  7. Lansky, E.P.; Newman, R.A. Punica granatum (Pomegranate) and its potential for prevention and treatment of inflammation and cancer. J. Ethnopharmacol. 2007, 109, 177–206. [Google Scholar] [CrossRef]
  8. Aruna, P.; Venkataramanamma, D.; Singh, A.K.; Singh, R.P. Health benefits of punicic acid: A review. Compr. Rev. Food Sci. Food Saf. 2016, 15, 16–27. [Google Scholar] [CrossRef]
  9. Suzuki, R.; Noguchi, R.; Ota, T.; Abe, M.; Miyashita, K.; Kawada, T. Cytotoxic effect of conjugated trienoic fatty acids on mouse tumor and human monocytic leukemia cells. Lipids 2001, 36, 477–482. [Google Scholar] [CrossRef]
  10. Seol, K.Y.; Honda, H.; Usui, K.; Ando, T.; Matsumoto, Y. 10,12,14-Hexadecatrienyl acetate: Sex pheromone of the mulberry pyralid, Glyphodes pyloalis walker (lepidoptera: Pyralidae). Agric. Biol. Chem. 1987, 51, 2285–2287. [Google Scholar] [CrossRef]
  11. Gries, R.; Reckziegel, A.; Bogenschütz, H.; Kontzog, H.-G.; Schlegel, C.; Francke, W.; Millar, J.G.; Gries, G. (Z,Z)-11,13-Hexadecadienyl acetate and (Z,E)-11,13,15-hexadecatrienyl acetate: Synergistic sex pheromone components of oak processionary moth, Thaumetopoea processionea (Lepidoptera: Thaumetopoeidae). Chemoecology 2004, 14, 95–100. [Google Scholar] [CrossRef]
  12. Millar, J.G.; McElfresh, J.S.; Romero, C.; Vila, M.; Mari-Mena, N.; Lopez-Vaamonde, C. Identification of the sex pheromone of a protected species, the Spanish moon moth Graellsia isabellae. J. Chem. Ecol. 2010, 36, 923–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ando, T.; Ogura, Y.; Koyama, M.; Kurane, M.; Uchiyama, M.; Yaul Seol, K.Y. Syntheses and NMR analyses of eight geometrical isomers of 10,12,14-hexadecatrienyl acetate, sex pheromone candidates of the mulberry pyralid. Agric. Biol. Chem. 1988, 52, 2459–2468. [Google Scholar]
  14. Doolittle, R.E.; Brabham, A.; Tumlinson, J.H. Sex phermonex of Manduca Sexta (L) stereoselective synthesis of (10E, 12E, 14Z)-10,12,14-hexadecatrienal and isomers. J. Chem. Ecol. 1990, 16, 1131–1153. [Google Scholar] [CrossRef] [PubMed]
  15. Ando, T.; Ohsawa, H. Sex pheromone candidates with a conjugated triene system: Synthesis and chemical characterization. J. Chem. Ecol. 1993, 19, 119–132. [Google Scholar] [CrossRef]
  16. Tulloch, A.P.; Bergter, L. Analysis of the conjugated trienoic acid containing oil from fevillea trilobata by 13C nuclear magnetic resonance spectroscopy. Lipids 1979, 14, 996–1002. [Google Scholar] [CrossRef]
  17. Cao, Y.; Gao, H.-L.; Chen, J.-N.; Chen, Z.-Y.; Yang, L. Identification and characterization of conjugated linolenic acid isomers by Ag+-HPLC and NMR. J. Agric. Food. Chem. 2006, 54, 9004–9009. [Google Scholar] [CrossRef]
  18. Cao, Y.; Yang, L.; Gao, H.-L.; Chen, J.-N.; Chen, Z.-Y.; Ren, Q.-S. Re-characterization of three conjugated linolenic acid isomers by GC–MS and NMR. Chem. Phys. Lipids 2007, 145, 128–133. [Google Scholar] [CrossRef]
  19. Sassano, G.; Sanderson, P.; Franx, J.; Groot, P. van Straalen, J.; Bassaganya-Riera, J. Analysis of pomegranate seed oil for the presence of jacaric acid. J. Sci. Food Agric. 2009, 89, 1046–1052. [Google Scholar] [CrossRef]
  20. Alexandri, E.; Ahmed, R.; Siddiqui, H.; Choudhary, M.; Tsiafoulis, C.; Gerothanassis, I.P. High resolution NMR spectroscopy as a structural and analytical tool for unsaturated lipids in solution. Molecules 2017, 22, 1663. [Google Scholar] [CrossRef]
  21. Albriktsen, P.; Harris, R.K. NMR Studies of conjugated linear trienes. Acta Chem. Scand. 1973, 27, 1875–1882. [Google Scholar] [CrossRef] [Green Version]
  22. Brouwer, A.M.; Bezemer, L.; Jacobs, H.J.C. Steric perturbation of the conjugated triene chromophore. Conformational analysis of (E)- and (2)-3-methyI-1,3,5-hexatriene and (2)-3-tert-butyl- 1,3,5-hexatriene. Red. Trav. Chim. Pays-Bas 1992, 111, 138–143. [Google Scholar] [CrossRef] [Green Version]
  23. Brouwer, A.M.; Bezemer, L.; Jacobs, H.J.C. Kinetics of thermal interconversion between cis, cis -1,3,5-octatriene, cis, cis, cis -2,4,6-octatriene, and cis, cis, trans-2,4,6-octatriene. Red. Trav. Chim. Pays-Bas 1998, 53, 1132–1137. [Google Scholar]
  24. Townsend, E.M.; Schrock, R.R.; Hoveyda, A.H. Z-Selective metathesis homocoupling of 1,3-dienes by molybdenum and tungsten monoaryloxide pyrrolide (MAP) complexes. J. Am. Chem. Soc. 2012, 134, 11334–11337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bifulco, G.; Dambruoso, P.; Gomez-Paloma, L.; Riccio, R. Determination of relative configuration in organic compounds by NMR spectroscopy and computational methods. Chem. Rev. 2007, 107, 3744–3779. [Google Scholar] [CrossRef] [PubMed]
  26. Smith, S.G.; Goodman, J.M. Assigning the stereochemistry of pairs of diastereoisomers using GIAO NMR shift calculations. J. Org. Chem. 2009, 74, 4597–4607. [Google Scholar] [CrossRef] [PubMed]
  27. Saielli, G.; Nicolaou, K.C.; Ortiz, A.; Zhang, H.; Bagno, A. Addressing the stereochemistry of complex organic molecules by density functional theory-NMR: Vannusal B. in retrospective. J. Am. Chem. Soc. 2011, 133, 6072–6077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Lodewyk, M.W.; Siebert, M.R.; Tantillo, D.J. Computational prediction of 1H and 13C chemical shifts: A useful tool for natural product, mechanistic, and synthetic organic chemistry. Chem. Rev. 2012, 112, 1839–1862. [Google Scholar] [CrossRef]
  29. Tantillo, D.J. Walking in the woods with quantum chemistry –applications of quantum chemical calculations in natural products research. Nat. Prod. Rep. 2013, 30, 1079–1086. [Google Scholar] [CrossRef] [PubMed]
  30. Siskos, M.G.; Kontogianni, V.G.; Tsiafoulis, C.G.; Tzakos, A.G.; Gerothanassis, I.P. Investigation of solute–solvent interactions in phenol compounds: Accurate ab initio calculations of solvent effects on 1H NMR chemical shifts. Org. Biomol. Chem. 2013, 11, 7400–7411. [Google Scholar] [CrossRef]
  31. Willoughby, P.H.; Jansma, M.J.; Hoye, T.R. A guide to small-molecule structure assignment through computation of (1H and 13C) NMR chemical shifts. Nat. Protoc. 2014, 9, 643–660. [Google Scholar] [CrossRef]
  32. Jaremko, L.; Jaremko, M.; Buczek, A.; Broda, A.M.; Kupka, T.; Jackowski, K. 1H and 13C shielding measurements in comparison with DFT calculations performed for 2-(acetylamino)-N,N-dimethyl-3-phenylacrylamide isomers. Chem. Phys. Lett. 2015, 627, 1–6. [Google Scholar] [CrossRef]
  33. Navarro-Vázquez, A. State of the art and perspectives in the application of quantum chemical prediction of 1H and 13C chemical shifts and scalar couplings for structural elucidation of organic compounds. Magn. Reson. Chem. 2017, 55, 2–32. [Google Scholar] [CrossRef] [PubMed]
  34. Tarazona, G.; Benedit, G.; Fernández, R.; Pérez, M.; Rodriguez, J.; Jiménez, C.; Cuevas, C. Can stereoclusters separated by two methylene groups be related by DFT studies? The case of the cytotoxic meroditerpeneshalioxepines. J. Nat. Prod. 2018, 81, 343–348. [Google Scholar] [CrossRef]
  35. Krivdin, L.B. Computational protocols for calculating 13C NMR chemical shifts. Progr. NMR Spectrosc. 2019, 112–113, 103–156. [Google Scholar] [CrossRef] [PubMed]
  36. Krivdin, L.B. Computational 1H NMR: Part 1. Theoretical background. Magn. Reson. Chem. 2019, 57, 897–914. [Google Scholar] [CrossRef]
  37. Semenov, V.A.; Krivdin, L.B. Computational 1H and 13C NMR of strychnobaillonine: On the way to larger molecules calculated at lower computational costs. Magn. Reson. Chem. 2021, 59, 108–116. [Google Scholar] [CrossRef] [PubMed]
  38. Siskos, M.G.; Tzakos, A.G.; Gerothanassis, I.P. Accurate ab initio calculations of O–HO and O–HO proton chemical shifts: Towards elucidation of the nature of the hydrogen bond and prediction of hydrogen bond distances. Org. Biomol. Chem. 2015, 13, 8852–8868. [Google Scholar] [CrossRef]
  39. Siskos, M.G.; Choudhary, M.I.; Tzakos, A.G.; Gerothanassis, I.P. 1H NMR chemical shift assignment, structure and conformational elucidation of hypericin with the use of DFT calculations–The challenge of accurate positions of labile hydrogens. Tetrahedron 2016, 72, 8287–8293. [Google Scholar] [CrossRef]
  40. Siskos, M.G.; Choudhary, M.I.; Gerothanassis, I.P. Hydrogen atomic positions of O–HO hydrogen bonds in solution and in the solid state: The synergy of quantum chemical calculations with 1H-NMR chemical shifts and X-ray diffraction methods. Molecules 2017, 22, 415. [Google Scholar] [CrossRef]
  41. Siskos, M.G.; Choudhary, M.I.; Gerothanassis, I.P. Refinement of labile hydrogen positions based on DFT calculations of 1H NMR chemical shifts: Comparison with X-ray and neutron diffraction methods. Org. Biomol. Chem. 2017, 15, 4655–4666. [Google Scholar] [CrossRef]
  42. Siskos, M.G.; Choudhary, M.I.; Gerothanassis, I.P. DFT-calculated structures based on 1H NMR chemical shifts in solution vs. structures solved by single-crystal X-ray and crystalline-sponge methods: Assessing specific sources of discrepancies. Tetrahedron 2018, 74, 4728–4737. [Google Scholar] [CrossRef]
  43. Torralba, M.P.; Sanz, D.; Claramunt, R.M.; Alkorta, I.; Dardonville, C.; Elguero, J. The structure of fosfomycin salts in solution and in the solid state by nuclear magnetic resonance spectroscopy and DFT calculations. Tetrahedron 2018, 74, 3937–3942. [Google Scholar] [CrossRef]
  44. Mari, S.H.; Varras, P.C.; Wahab, A.-t.; Choudhary, I.M.; Siskos, M.G.; Gerothanassis, I.P. Solvent-dependent structures of natural products based on the combined use of DFT calculations and 1H-NMR chemical shifts. Molecules 2019, 24, 2290. [Google Scholar] [CrossRef] [Green Version]
  45. Siskos, M.G.; Varras, P.C.; Gerothanassis, I.P. DFT calculations of O-H....O 1H NMR chemical shifts in investigating enol–enol tautomeric equilibria: Probing the impacts of intramolecular hydrogen bonding vs. stereoelectronic interactions. Tetrahedron 2020, 76, 130979. [Google Scholar] [CrossRef]
  46. Ahmed, R.; Varras, P.C.; Siskos, M.G.; Siddiqui, H.; Choudhary, M.I.; Gerothanassis, I.P. NMR and computational studies as analytical and high-resolution structural tool for complex hydroperoxides and endo-hydroperoxides of fatty acids in solution. Molecules 2020, 25, 4902. [Google Scholar] [CrossRef] [PubMed]
  47. Venianakis, T.; Oikonomaki, C.; Siskos, M.G.; Varras, P.C.; Primikyri, A.; Alexandri, E.; Gerothanassis, I.P. DFT Calculations of 1H- and 13C-NMR chemical shifts of geometric isomers of conjugated linoleic acid (18:2 ω-7) and model compounds in solution. Molecules 2020, 25, 3660. [Google Scholar] [CrossRef] [PubMed]
  48. Touw, S.I.E.; de Groot, H.J.M.; Buda, F. DFT calculations of the 1H NMR chemical shifts and 13C chemical shifts tensors of retinal isomers. J. Mol. Struct. (Theochem) 2004, 711, 141–147. [Google Scholar] [CrossRef]
  49. Ditchfield, R. Self-consistent perturbation theory of diamagnetism I. A gauge-invariant LCAO method for N.M.R. chemical shifts. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  50. Abraham, R.J.; Canton, M.; Grifiths, L. Proton chemical shifts in alkenes and anisotropic and steric effects in double bond. Magn. Reson. Chem. 2001, 39, 421–431. [Google Scholar] [CrossRef]
  51. Cheney, B.V. Magnetic deshielding of protons due to intramolecular steric interactions with proximate hydrogens. J. Am. Chem. Soc. 1968, 90, 5386–5390. [Google Scholar] [CrossRef]
  52. Marshall, T.W.; Pople, J.A. Nuclear magnetic shielding of a hydrogen atom in an electric field. Mol. Phys. 1958, 1, 199–202. [Google Scholar] [CrossRef]
  53. Cobb, T.B.; Memory, J.B. High-resolution NMR spectra of polycyclic hydrocarbons. II. Pentacyclic compounds. J. Phys. Chem. 1967, 47, 2020–2025. [Google Scholar] [CrossRef]
  54. Grimme, S.; Diedrich, C.; Korth, M. The importance of inter- and intramolecular van der Waals interactions in organic reactions: The dimerization of anthracene revisited. Angew. Chem. Int. Ed. 2006, 118, 641–645. [Google Scholar] [CrossRef]
  55. Hujo, W.; Grimme, S. Performance of the van der Waals density functional VV10 and (hybrid) GGA variants for thermochemistry and noncovalent interactions. J. Chem. Theory Comput. 2011, 7, 3866–3871. [Google Scholar] [CrossRef] [PubMed]
  56. Kupka, T.; Stachow, M.; Nieradka, M.; Kaminsky, J.; Pluta, T. Convergence of nuclear magnetic shieldings in the Kohn-Sham limit of several small molecules. J. Chem. Theory Comput. 2010, 6, 1580–1589. [Google Scholar] [CrossRef]
  57. Abraham, R.J.; Mobli, M. Modelling 1H NMR Spectra of Organic Compounds Theory, Applications and NMR Prediction Software; John Wiley & Sons Ltd.: Chichester, UK, 2008. [Google Scholar]
  58. Kaneko, F.; Yano, J.; Sato, K. Diversity in the fatty-acid conformation and chain packing of cis-unsaturated lipids. Curr. Opin. Struct. Biol. 1998, 8, 417–425. [Google Scholar] [CrossRef]
  59. Suzuki, M.; Ogaki, T.; Sato, K. Crystallization and transformation mechanisms of α, β and γ polymorphs of ultra-pure oleic acid. J. Am. Oil. Chem. Soc. 1985, 62, 1600–1604. [Google Scholar] [CrossRef]
  60. Kaneko, F.; Yamazaki, K.; Kitagawa, K.; Kikyo, T.; Kobayashi, M.; Sato, K.; Suzuki, M. Structure and crystallization behavior of the β phase of oleic acid. J. Phys. Chem. B 1997, 101, 1803–1809. [Google Scholar] [CrossRef]
  61. 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 0.9, Revision. D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
Figure 1. Chemical structures of three geometric isomers of the 18:3 ω-5 conjugated linolenic acid, four geometric isomers of the 16:3 ω-2 conjugated hexadecatrienyl acetate, and triene-containing model compounds investigated in the present work.
Figure 1. Chemical structures of three geometric isomers of the 18:3 ω-5 conjugated linolenic acid, four geometric isomers of the 16:3 ω-2 conjugated hexadecatrienyl acetate, and triene-containing model compounds investigated in the present work.
Molecules 26 03477 g001
Figure 2. (A) Calculated, δcalc, 1H NMR chemical shifts with CPCM of the olefinic protons vs. experimental, δexp, chemical shifts using the same level of theory as geometry optimization for the model trienyl compounds (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene, and (E,E,E)-2,4,6-octatriene. (B) The same as in (A), but the literature experimental values of δ(H (3,6)) = 5.83 ppm and δ(H (4,5)) = 6.50 ppm of (E,Z,E)-2,4,6-octatriene [23] have been reversed.
Figure 2. (A) Calculated, δcalc, 1H NMR chemical shifts with CPCM of the olefinic protons vs. experimental, δexp, chemical shifts using the same level of theory as geometry optimization for the model trienyl compounds (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene, and (E,E,E)-2,4,6-octatriene. (B) The same as in (A), but the literature experimental values of δ(H (3,6)) = 5.83 ppm and δ(H (4,5)) = 6.50 ppm of (E,Z,E)-2,4,6-octatriene [23] have been reversed.
Molecules 26 03477 g002aMolecules 26 03477 g002b
Figure 3. Effect of the variation in the torsion angle φ (C1C2C3C4) of (Z)-1,3,5-hexatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A), and δcalc(1H) data of the olefinic protons (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Figure 3. Effect of the variation in the torsion angle φ (C1C2C3C4) of (Z)-1,3,5-hexatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A), and δcalc(1H) data of the olefinic protons (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Molecules 26 03477 g003
Figure 4. Effect of variation in the torsion angle φ1(C2C3C4C5) of (E,Z,E)-2,4,6-octatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on the olefinic δcalc(1H) data (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Figure 4. Effect of variation in the torsion angle φ1(C2C3C4C5) of (E,Z,E)-2,4,6-octatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on the olefinic δcalc(1H) data (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Molecules 26 03477 g004
Figure 5. Effect of variation in the torsion angle φ2(C1C2C3C4) of (E)-1,3,5-hexatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on the olefinic δcalc(1H) data (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Figure 5. Effect of variation in the torsion angle φ2(C1C2C3C4) of (E)-1,3,5-hexatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on the olefinic δcalc(1H) data (at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM in CHCl3) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Molecules 26 03477 g005
Figure 6. Effect of variation in the torsion angle φ3 (C2C3C4C5) of (E,E,E)-2,4,6-octatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on δcalc(1H) data of the olefinic protons (at the GIAO/B3LYP/6–311+G(2d,p) level of theory with CPCM in CCl4) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Figure 6. Effect of variation in the torsion angle φ3 (C2C3C4C5) of (E,E,E)-2,4,6-octatriene, with energy minimization at the B3LYP/6–31+G(d) level, on the electronic energy ΔE (kcal/mol) (characteristic ΔG values are also shown) (A) and on δcalc(1H) data of the olefinic protons (at the GIAO/B3LYP/6–311+G(2d,p) level of theory with CPCM in CCl4) (B). The experimental chemical shift values are denoted with the horizontal dotted lines.
Molecules 26 03477 g006
Figure 7. Structures of various conformers (AE) of the β-eleostearic acid (9E,11E,13E-isomer) with energy minimization in the gas phase at the B3LYP/6–31+G(d) level of theory. ΔG values (kcal·mol−1) and % populations of conformers (AE) are shown in Table 2.
Figure 7. Structures of various conformers (AE) of the β-eleostearic acid (9E,11E,13E-isomer) with energy minimization in the gas phase at the B3LYP/6–31+G(d) level of theory. ΔG values (kcal·mol−1) and % populations of conformers (AE) are shown in Table 2.
Molecules 26 03477 g007
Figure 8. Definition of various conformations of the allylic carbon C8 of β-eleostearic acid (9E,11E,13E isomer).
Figure 8. Definition of various conformations of the allylic carbon C8 of β-eleostearic acid (9E,11E,13E isomer).
Molecules 26 03477 g008
Figure 9. (A) Calculated, δcalc, 1H NMR chemical shifts of the olefinic protons at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM) vs. experimental, δexp, chemical shifts with energy minimization using B3LYP/6–31+G(d) and APFD/6–31+G(d) for (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene, (E,E,E)-2,4,6-octatriene, α-oleostearic acid (9Z, 11E, 13E), β-oleostearic acid (9E, 11E, 13E), punicic acid (9Z, 11E, 13Z), and (10E, 12E, 14Z)-, (10E,12Z,14Z)-, (10Z, 12Z, 14E)-, and (10Z, 12E, 14E)-hexatrienyl acetates. (B) The same as in (A); however, the literature experimental chemical shift data (δexp) of the compounds shown in Figure S14 have been revised.
Figure 9. (A) Calculated, δcalc, 1H NMR chemical shifts of the olefinic protons at the GIAO/B3LYP/6–311+G(2d,p) level with CPCM) vs. experimental, δexp, chemical shifts with energy minimization using B3LYP/6–31+G(d) and APFD/6–31+G(d) for (Z)-1,3,5-hexatriene, (E)-1,3,5-hexatriene, (E,Z,E)-2,4,6-octatriene, (E,E,E)-2,4,6-octatriene, α-oleostearic acid (9Z, 11E, 13E), β-oleostearic acid (9E, 11E, 13E), punicic acid (9Z, 11E, 13Z), and (10E, 12E, 14Z)-, (10E,12Z,14Z)-, (10Z, 12Z, 14E)-, and (10Z, 12E, 14E)-hexatrienyl acetates. (B) The same as in (A); however, the literature experimental chemical shift data (δexp) of the compounds shown in Figure S14 have been revised.
Molecules 26 03477 g009
Table 1. Literature experimental 1H NMR chemical shifts, δexp, and the sample and spectral specifications of the compounds of Figure 1.
Table 1. Literature experimental 1H NMR chemical shifts, δexp, and the sample and spectral specifications of the compounds of Figure 1.
ABCDEFGHIJKLM
Groupδexp
(ppm)
δexp
(ppm)
δexp
(ppm)
δexp
(ppm)
Groupδexp
(ppm)
δexp
(ppm)
Groupδexp
(ppm)
δexp
(ppm)
δexp
(ppm)
δexp
(ppm)
Groupδexp
(ppm)
δexp (ppm)δexp (ppm)
H1a5.155.15.095.05H11.81.84H92.092.122.182.17H116.106.196.48
H1b5.245.235.175.18H25.75.69H105.705.745.485.40H126.106.406.48
H26.86.366.746.3H35.836.12H116.106.506.435.98H106.046.016.08
H36.06.225.936.16H46.56.10H126.185.986.136.37H136.046.126.08
H46.06.225.936.16H56.56.10H136.416.165.966.16H95.665.45.46
H56.86.366.746.3H65.836.12H146.026.466.526.07H145.665.745.46
H6a5.155.15.095.05H75.75.69H155.475.565.755.70H22.372.402.37
H6b5.245.235.175.18H81.81.84H161.761.771.801.77H82.102.202.22
H152.102.202.22
H31.651.641.65
H41.391.411.39
H51.391.411.39
H61.391.411.39
H71.391.411.39
H161.331.331.32
H171.331.331.32
H180.910.970.94
A (Z)-1,3,5-hexatriene (CDCl3), 1H and 2D 1H–1H COSY NMR (270 MHz) [15]; B (E)-1,3,5-hexatriene (CDCl3), 1H and 2D 1H–1H COSY NMR (270 MHz) [15]; C (Z)-1,3,5-hexatriene (CCl4), 1H NMR (100 MHz) [21,22]; D (E)-1,3,5-hexatriene (CCl4), 1H NMR (100 MHz) [21,22]; E (E,Z,E)-2,4,6-octatriene (CDCl3), 1H NMR (300 MHz) [23]; F (E,E,E)-2,4,6-octatriene (CCl4), 1H NMR (100 MHz) [21]; G 10E,12E,14Z-hexadecatrienyl acetate (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (270 MHz) [13]; H 10E,12Z,14Z-hexadecatrienyl acetate (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (270 MHz) [13]; I 10Z,12Z,14E-hexadecatrienyl acetate (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (270 MHz) [13]; J 10Z,12E,14E-hexadecatrienyl acetate (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (270 MHz) [13]; K β-eleostearic acid (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (400 MHz) [18]; L α-eleostearic acid (CDCl3), 1H, 13C, and 2D 1H-1H COSY NMR (400 MHz) [18]; M punicic acid (CDCl3), 1H, 13C, and 2D 1H–1H COSY NMR (400 MHz) [18].
Table 2. Conformational properties, ΔG values (kcal·mol−1), and % populations of various low-energy conformers of the 9,11,13-conjugated fatty acid and 10,12,14-conjugated hexadecatrienyl acetate geometric isomers with geometry optimization at the B3LYP/6–31+G(d) and APFD/6–31+G(d) levels.
Table 2. Conformational properties, ΔG values (kcal·mol−1), and % populations of various low-energy conformers of the 9,11,13-conjugated fatty acid and 10,12,14-conjugated hexadecatrienyl acetate geometric isomers with geometry optimization at the B3LYP/6–31+G(d) and APFD/6–31+G(d) levels.
CompoundConformerB3LYP/6–31+G(d)APFD/6–31+G(d)
φc7c8c9c10 (°)φc13c14c15c16 (°)ΔG
(kcal/mol)
(% Population)
φc7c8c9c10 (°)φc13c14c15c16 (°)ΔG
(kcal/mol)
(% Population)
9E,11E,13E Isomer
(β-Eleostearic acid)
A120.26
(S)
118.59
(S)
+0.36
(33.41)
119.21
(S)
117.99
(S)
+0.16
(38.68)
B120.11
(S)
−119.97
(S’)
0.00
(61.35)
118.73
(S)
−118.59
(S’)
0.00
(50.67)
C−0.78
(Syn)
−0.03
(Syn)
+3.00
(0.39)
−0.59
(Syn)
−0.11
(Syn)
+2.13
(1.39)
D−117.53
(S’)
−0.83
(Syn)
+1.89
(2.53)
−116.94
(S’)
−0.62
(Syn)
+1.38
(4.94)
E−0.75
(Syn)
−118.55
(S’)
+1.94
(2.32)
−0.58
(Syn)
−117.27
(S’)
+1.46
(4.31)
9Z,11E,13E Isomer
(β-Eleostearic acid)
A116.67
(S)
119.18
(S)
+0.03
(25.07)
111.25
(S)
117.87
(S)
0.00
(36.53)
B−118.92
(S’)
−119.37
(S’)
+0.15
(20.48)
−114.31
(S’)
−117.53
(S’)
+0.48
(16.25)
C116.88
(S)
−119.16
(S’)
+0.02
(25.50)
112.96
(S)
−117.54
(S’)
+0.18
(26.96)
D−118.88
(S’)
119.28
(S)
0.00
(26.38)
−113.83
(S’)
118.32
(S)
+0.47
(16.52)
E−119.43
(S’)
−0.76
(Syn)
+1.38
(2.57)
−113.80
(S’)
−0.22
(Syn)
+1.35
(3.74)
9Z,11E,13Z Isomer
(Punicic acid)
A117.46
(S)
119.33
(S)
0.00
(38.54)
113.45
(S)
114.70
(S)
0.00
(44.84)
B117.76
(S)
−118.63
(S’)
+0.08
(33.67)
112.18
(S)
−112.33
(S’)
+0.31
(26.57)
C−119.22
(S’)
119.55
(S)
+0.27
(24.43)
−113.43
(S’)
113.54
(S)
+0.27
(28.43)
D1.52−121.09
(S’)
+3.83
(1.51)
5.34−110.17
(S’)
+3.77
(0.08)
E117.81
(S)
2.02
(Syn)
+3.63
(1.85)
111.01
(S)
4.55
(Syn)
+3.78
(0.08)
10E,12E,14Z-Hexadecatrienyl acetateA120.04
(S)
+0.28
(25.59)
118.73
(S)
+0.08
(28.88)
B−119.39
(S’)
0.00
(41.05)
−118.73
(S’)
+0.03
(31.42)
C120.03
(S)
+0.20
(29.29)
118.73
(S)
0.00
(33.05)
D1.56
(Syn)
+1.37
(4.07)
1.16
(Syn)
+0.95
(6.65)
10E,12Z,14Z-Hexadecatrienyl acetateA119.67
(S)
+0.44
(22.44)
118.45
(S)
+0.20
(26.06)
B−119.64
(S’)
0.00
(47.15)
−118.86
(S’)
0.00
(36.52)
C119.56
(S)
+0.35
(26.12)
118.28
(S)
+0.12
(29.82)
D1.22
(Syn)
+1.42
(4.29)
0.87
(Syn)
+0.93
(7.60)
10Z,12Z,14E-Hexadecatrienyl acetateA119.01
(S)
0.00
(30.52)
112.30
(S)
+0.04
(20.48)
B−119.51
(S’)
+0.51
(12.91)
−113.47
(S’)
+0.10
(18.51)
C118.80
(S)
0.00
(30.52)
112.24
(S)
0.00
(21.92)
D−119.50
(S’)
+0.48
(13.57)
−113.16
(S’)
+0.02
(21.19)
10Z,12E,14E-Hexadecatrienyl acetateA119.05
(S)
0.00
(31.98)
113.03
(S)
0.00
(33.79)
B−121.15
(S’)
+0.54
(12.85)
−116.62
(S’)
+0.63
(11.67)
C118.73
(S)
+0.04
(29.89)
113.00
(S)
+0.04
(31.59)
D−121.06
(S’)
+0.56
(12.43)
−116.57
(S’)
+0.65
(11.28)
Table 3. Linear regression correlation coefficient, mean square error, intercept, and slope of the calculated vs. experimental olefinic 1H NMR chemical shifts of Figure 9.
Table 3. Linear regression correlation coefficient, mean square error, intercept, and slope of the calculated vs. experimental olefinic 1H NMR chemical shifts of Figure 9.
Minimization MethodCorrelation Coefficient
(R2)
Mean Square ErrorInterceptSlope
B3LYP/6–31+G(d) a0.8730.037−0.0021.062
APFD/6–31+G(d) a0.868 (0.825)0.041(0.044)−0.102 (0.518)1.085 (0.956)
B3LYP/6–31+G(d) b0.9840.005−0.3871.127
APFD/6–31+G(d) b0.982 (0.942)0.005 (0.015)−0.506 (0.138)1.153 (1.020)
a Data of Figure 9A; calculation of 1H NMR chemical shifts at the GIAO/B3LYP/6–311+G(2d,p) level and b data of Figure 9B; calculation of 1H NMR chemical shifts at the GIAO/B3LYP/6–311+G(2d,p) level. The data in parenthesis were obtained at the GIAO with the same level of theory as geometry optimization.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Venianakis, T.; Oikonomaki, C.; Siskos, M.G.; Primikyri, A.; Gerothanassis, I.P. DFT Calculations of 1H NMR Chemical Shifts of Geometric Isomers of Conjugated Linolenic Acids, Hexadecatrienyl Pheromones, and Model Triene-Containing Compounds: Structures in Solution and Revision of NMR Assignments. Molecules 2021, 26, 3477. https://doi.org/10.3390/molecules26113477

AMA Style

Venianakis T, Oikonomaki C, Siskos MG, Primikyri A, Gerothanassis IP. DFT Calculations of 1H NMR Chemical Shifts of Geometric Isomers of Conjugated Linolenic Acids, Hexadecatrienyl Pheromones, and Model Triene-Containing Compounds: Structures in Solution and Revision of NMR Assignments. Molecules. 2021; 26(11):3477. https://doi.org/10.3390/molecules26113477

Chicago/Turabian Style

Venianakis, Themistoklis, Christina Oikonomaki, Michael G. Siskos, Alexandra Primikyri, and Ioannis P. Gerothanassis. 2021. "DFT Calculations of 1H NMR Chemical Shifts of Geometric Isomers of Conjugated Linolenic Acids, Hexadecatrienyl Pheromones, and Model Triene-Containing Compounds: Structures in Solution and Revision of NMR Assignments" Molecules 26, no. 11: 3477. https://doi.org/10.3390/molecules26113477

Article Metrics

Back to TopTop