Sana
Fatima
a,
Panayiotis C.
Varras
ab,
Atia-tul-Wahab
*c,
M. Iqbal
Choudhary
acd,
Michael G.
Siskos
b and
Ioannis P.
Gerothanassis
*abc
aH.E.J. Research Institute of Chemistry, International Center for Chemical and Biological Sciences, University of Karachi, Karachi 7527, Pakistan
bSection of Organic Chemistry & Biochemistry, Department of Chemistry, University of Ioannina, Ioannina, GR-45110, Greece. E-mail: igeroth@uoi.gr
cDr Panjwani Center for Molecular Medicine and Drug Research, International Center for Chemical and Biological Sciences, University of Karachi, Karachi 7527, Pakistan. E-mail: Atiatulwahab@gmail.com
dDepartment of Biochemistry, Faculty of Science, King Abdul-Aziz University, Jeddah-21589, Saudi Arabia
First published on 12th July 2021
NMR and DFT studies of phenol compounds as molecular sensors were carried out to investigate H2O/DMSO eutectic mixtures at a molecular level. The experimental 1H NMR chemical shifts of the OH groups, δexp(OH), of phenol, paracoumaric acid, and vanillic acid show maximum deshielding and, thus, hydrogen bond interactions in the range of mole fractions 0.20 < χ(DMSO) < 0.33. In the mole fractions χ(DMSO) < 0.2, a progressive decrease in δexp(OH) was observed which demonstrates a decrease in hydrogen bond interactions at infinite dilution in H2O, despite the increase in the number of available hydrogen bond acceptor and donor sites. DFT calculated δcalc(OH) of minimum energy solvation clusters were shown to be in reasonable agreement with the pattern in experimental δexp(OH) data. The chemical shift deshielding and, thus, increased hydrogen bond interactions in the natural product + DMSO + nH2O (n = 2, 3) solvation clusters, relative to complexes in DMSO or H2O solutions, cannot be attributed to a single structural parameter of the cooperative interactions between H2O and DMSO molecules with the phenol OH groups of the natural products. The minimum energy conformers of phenol compounds + 2H2O + DMSO complexes are in excellent agreement with a recent low temperature neutron diffraction experiment of 3D2O + DMSO and demonstrate a general structural motif of solvation complexes. The combined use of 1H NMR and DFT studies with emphasis on δ(OH) of phenol compounds, as molecular sensors, can provide an effective method for the study of solute–solvent interactions at the atomic level.
For a system consisting of water and DMSO, extreme deviation from additivity was observed. At a mole fraction χ(DMSO) ≃ 0.33, a very low freezing point was measured (−140 °C). A stable DMSO/H2O (1:
2) cluster was suggested to be responsible for this unusually low melting point.12 Each such subunit is rather loosely hydrogen bounded to the next subunit, thus, resulting in an overall structure that prevents crystallization. Further research revealed the formation of a rather complex phase diagram, and the dominance of stable DMSO/H2O (1
:
3) clusters.13
Numerous experimental and theoretical studies have been reported on DMSO/water aqueous mixtures including neutron scattering,14,15 investigation of the structure and dynamics of hydrogen bonding,16 solvation dynamics and vibrational relaxation.17 MD simulations at ambient temperature have been carried out to evaluate three models of pure liquid DMSO and in DMSO/H2O mixtures.18 Statistical analysis of the hydrogen bond network revealed the presence of DMSO + 2H2O and not DMSO + 3H2O complexes. MD simulation estimated two regions of co-solvent DMSO–water clusters: χ(DMSO) = 0.12–0.17 and χ(DMSO) = 0.27–0.35. In the second region, there is an anomalous solvation trend observed due to high viscosity while in the first region the solvation increases due to the formation of strong hydrogen bonded DMSO/H2O (1:
2) complexes.19 DFT and polarizable force field have been utilized to investigate energetic and structural properties of DMSO–water clusters.20 Quantum chemical studies were also utilized to investigate inter-cluster interaction modes of the freezing behavior, particularly of DMSO/H2O (1
:
3) clusters.21 Water dynamics in H2O/DMSO binary mixtures were studied with the use of polarization selective pump–probe experiments, 2D IR vibrational echo spectroscopy and IR absorption spectroscopy.22 IR spectroscopy of the S
O stretching mode combined with MD and quantum chemistry models were used to quantify hydrogen-bond populations in DMSO/H2O mixtures.23 MD studies of a DMSO/H2O (1
:
3) cluster using the AMBER force field suggested a value of ∼2.89 ± 0.11 water molecules around each DMSO molecule.24 Nevertheless, the driving forces behind non-covalent interactions in hydration of DMSO which is a dipolar aprotic solvent with both hydrophilic and hydrophobic properties, are not sufficiently understood.25
The phenols are major constituents of many biological and naturally occurring compounds and are ubiquitous in the plant and animal kingdom. Interestingly, around 7% of the current drugs contain phenol groups. Furthermore, it is well known that experimental 1H NMR chemical shifts, δexp(OH), of phenol –OH groups are very sensitive to OH⋯O hydrogen bond interactions.26–32 Linear correlation between X-ray diffraction r(O⋯O) distances and δexp(OH) 1H NMR chemical shifts was suggested for a variety of β-diketone enol compounds with intramolecular O–H⋯O hydrogen bonds.33 A linear relationship has also been presented between δexp(OH) and H⋯O distances, determined from neutron diffraction.34 Very good linear correlations between experimental and computed δ(OH) 1H NMR chemical shifts and between hydrogen bond distances O⋯H(O) and δ(OH) have been suggested by employing a variety of quantum chemical methods.35–42
The objective of the present study was to address three important issues. First to investigate the importance of hydrogen bonds in DMSO/H2O eutectic mixtures by using, as molecular sensors, phenol (Ph, 1), paracoumaric acid (PCA, 2), and vanilic acid (VA, 3) (Fig. 1). As discussed above, δ(OH) of phenol compounds are very sensitive to hydrogen bonding interactions and, furthermore, they are significantly deshielded relative to the chemical shifts of H2O. Overlapping, therefore, with the strong H2O peak can be avoided which results in accurate δexp(OH) values, contrary to the case of e.g., alcohol type OH groups. Second, to utilize the agreement between experimental and DFT computed 1H NMR chemical shifts of phenol OH groups as a tool for investigating solute–solvent interactions and accurate structural models at the atomic level. Third to investigate whether there is a general structural pattern of phenol compound + (2, 3) H2O + DMSO solvation complexes. The molecules selected are secondary metabolites derived from plants and contain a single phenol OH group. Compounds with multiple phenol OH groups were not used in order to avoid strong overlapping of the resonances and, thus, uncertainties in the determination of the chemical shifts. Phenol (1), occurs in the essential oil from tobacco leaves (Nicotiannatabacum, Solanaceace) and pine needles (Pinus Sylvestris, Pinaceae)43 and has the ability to encounter the gut microbes.44 Paracoumaric acid (2) was found in many plants, such as plumbaginaceae's root, leaf and flower,45 and has been used to treat diabetes, hypertension and cardiovascular diseases.46 Vanillic acid has been isolated from origanum vulgare and acts as anti-melanogenesis agent.47
![]() | ||
Fig. 1 Chemical structures of the compounds used in the present work: phenol, Ph(1), paracoumaric acid, PCA(2), and vanillic acid, VA(3). |
The use of 1H NMR of hydroxyl protons in H2O or H2O/organic solvent mixtures, however, presents experimental challenges due to chemical exchange among hydroxyl protons, and protic solvents. Fig. 2(Aa) illustrates the OH 1H NMR spectral region of (1) for various mole fractions, χ(DMSO), of DMSO/H2O mixtures at 300 K. For mole fractions χ(DMSO) = 1–0.35 relatively sharp resonances were observed which allowed the accurate determination of chemical shifts. A linear correlation of δexp(OH) vs. χ(DMSO) was observed for χ(DMSO) ≥ 0.35 which shows an increase in deshielding and, thus, in hydrogen bond interactions upon increasing the concentration of H2O. This is to be expected since an increasing number of water molecules will increase the number of available sites of hydrogen bond acceptors, and donors. In the range of mole fractions 0.2 ≤ χ(DMSO) ≤ 0.33, a broad maximum of δexp(OH) and, thus, in hydrogen bond strength was observed. This range of mole fractions corresponds to mole ratios of H2O/DMSO of 2/1–4/1 (Table S1, ESI†). For mole fraction χ(DMSO) < 0.2, a progressive decrease in δ(OH) was observed which demonstrates a decrease in hydrogen bond interactions at infinite dilution in H2O. The OH resonance at mole fractions χ(DMSO) < 0.1 cannot be distinguished from the baseline.
Similar experiments were performed at 280 K and 270 K (Fig. 2(Ba) and (Ca)) in order to reduce the proton exchange rate and, thus, linewidths which would allow a better definition of the OH chemical shifts for mole fractions χ(DMSO) < 0.3. For mole fractions χ > 0.88 and 0.83 the solutions were frozen at 280 K and 270 K, respectively, therefore the OH chemical shifts were obtained through extrapolation of the linear dependence of δexp(OH) vs. T31 in the temperature range of 292–318 K. By decreasing the temperature, the proton exchange is reduced, however, even at 270 K the resonances are still very broad which prohibit accurate estimation of the chemical shifts for χ(DMSO) ≤ 0.1.
From the above it is evident that, in the three temperatures investigated, δexp(OH) values strongly deviate from linearity for χ(DMSO) < 0.35 and the extrapolated value of δexp(OH) at infinite dilution in H2O is smaller than that in the range of mole fractions 0.2 ≤ χ(DMSO) ≤ 0.33. This demonstrates weaker hydrogen bond interactions at infinite dilution in H2O, despite the increase in the available hydrogen bond acceptor and donor sites, provided that δexp(OH) is not affected by proton exchange. Assuming that the spin ½ nucleus, without J coupling, jumps instantaneously from one site to the other, then, the rate constant τOH−1 is the probability of the phenol OH nucleus jumping per unit time and τH2O−1 is that of the H2O site. If the exchange is slow, that is when τOH−1, τH2O−1 ≪ |νOH − νH2O|, then, the line shapes reduce to two Lorentzian lines with maxima at νOH, the resonance frequency of the phenol OH site, and νH2O, the resonance frequency of the H2O, with intensities POH and PH2O and widths τOH−1 and τH2O−1. In this limit the only effect of the exchange is to produce a life-time broadening.50,51 As the exchange gets more rapid and τOH−1, τH2O−1 become comparable with νOH − νH2O, then, the peaks at νOH and νH2O move towards each other. Since the natural line width, especially at low temperatures, is dominated by field inhomogeneity broadening, it appears plausible to extract it from the internal reference signal of DSS and subtract it from the exchange-broadened lines, at a given temperature, using the equation τOH−1 = Δν1/2(OH) − Δν1/2(DSS) where Δν1/2(OH) and Δν1/2(DSS) are the linewidths of the OH group and DSS, respectively.50 Tables S2 and S3 (ESI†) show that in all cases Δν/τOH−1 > 10 which demonstrates that the slow exchange regime τOH−1 ≪ |νOH − νH2O| is fulfilled and no effect of chemical exchange on δexp(OH) is expected. The minor effect of increasing the water content on τOH−1, for 0.2 < χ(DMSO) < 0.4 (Table S3 in ESI†) can be attributed to increased hydrogen bonding interactions of Ph(1) + nH2O + DMSO (n = 2–4) subunits, which are loosely bound to each other, thus, resulting in reduced proton exchange (see also discussion on DFT calculations).
Fig. 3 illustrates the OH 1H NMR spectral region and correlation of δexp(OH) vs. χ(DMSO) of PCA(2) at 300 K, 280 K and 270 K. The resulting linewidths were found to be broader, than those obtained with Ph(1). This can be attributed to the presence of the carboxyl group which increases proton exchange rate constant and, thus, linewidths. For mole fractions χ(DMSO) = 1–0.35, a linear correlation of δ(OH) vs. χ(DMSO) is evident. In the range of mole fractions 0.20 ≤ χ(DMSO) ≤ 0.33, a broad maximum of δ(OH) was observed. Since the slow exchange regime τOH−1 ≪ |νOH − νH2O| is fulfilled (Table S4, ESI†), it can be concluded that the broad maximum of δ(OH) corresponds to a maximum in hydrogen bonding interactions, as in the case of Ph(1). Again, the extrapolated value of δexp(OH) at infinite dilution in H2O is smaller than that in the region of the mole fractions 0.20 ≤ χ(DMSO) ≤ 0.33. Similar behavior has been observed for VA(3) (Fig. S1 and Table S5, ESI†). The resulting linewidths were found to be broader than in the case of Ph(1) due to the presence of the carboxyl group which enhances proton exchange rate constants. Nevertheless, again a broad maximum at χ(DMSO) = 0.35–0.2 was observed. The OH resonance at mole ratios χ(DMSO) < 0.12 cannot be distinguished from the baseline, therefore, the extrapolated value of δexp(OH) at infinite dilution in H2O could not be obtained accurately.
The present computational study was performed by using the Gaussian 09 with the DFT method54 with the structures optimized in the gas phase by using the standard functional B3LYP as well as the APFD with corrections for dispersion interactions, with the 6-31+G(d) basis set. The 1H NMR chemical shifts were calculated with the GIAO (Gauge-Independent Atomic Orbital) and CSGT (Continuous Set of Gauge Transformations)55 methods by using the B3LYP/6-311+G(2d,p) level in the gas phase.
The optimized structures of the phenol molecules with various discrete water and DMSO molecules, were obtained successively starting with the phenol type molecules by adding various numbers of water molecules followed by DMSO molecules in various possible positions, in such a way that they act either as donors or acceptors. The resulting molecular geometries were verified as energy minima on the potential energy surface (PES) by performing frequency calculations at the same level (zero imaginary frequencies). Four solvation clusters were examined for each of the three natural products of Fig. 1: (i) with a single solvation molecule of DMSO, (ii) with DMSO + 2H2O, (iii) with DMSO + 3H2O, and (iv) with 3H2O. In the case of VA(3), due to the proximity of hydroxy and methoxy groups at the ortho position, the importance of intramolecular hydrogen bonding was investigated also in VA + nH2O (n:1, 2) solvation clusters.
Due to the existence of different configurations at each mole ratio, it is necessary to identify the key hydrogen bond interactions, particularly in the minimum energy configurations. Selected structural and conformational properties of Ph(1) + solvent complexes, energy minimized using the APFD/6-31+G(d) and B3LYP/6-31+G(d) methods in the gas phase, are shown in Fig. 4 and Fig. S2, (ESI†) respectively. The minimum energy configuration of Ph(1) with a single molecule of DMSO (Fig. 4) is characterized by a rather strong intermolecular hydrogen bond of O1–H1⋯OD = 1.706 Å and an angle O1–H1⋯OD = 160.4°. These values are comparable to those found for the phenol OH⋯DMSO intermolecular hydrogen bond interaction in the natural product emodin.38 The torsion angles O1–H1⋯OD–S, and C1–O1–H1⋯OD were found to be 10.26° and 68.37°, respectively, which demonstrate a non-planar arrangement. It can be noted the long-range interactions of the CH3 groups of DMSO with O1 (O1⋯HDa = 2.735 Å and O1⋯HDb = 2.519 Å).
In the minimum energy solvation complex Ph + 2H2O + DMSO, a water molecule (W1) is hydrogen bonded with the OH group of Ph (H1⋯Ow1 = 1.697 Å and O1–H1⋯Ow1 = 170.7°) (Fig. 4b). W1 is hydrogen bonded with the second molecule of water (W2) (Ow2⋯Hw1 = 1.687 Å and Ow2⋯Hw1–Ow1 = 170.9°). W2 is hydrogen bonded with the DMSO molecule (Hw2⋯OD = 1.655 Å and Ow2–Hw2 ⋯OD = 175.1°). This preferential coordination of the DMSO molecule is stabilized also because of secondary interactions of the methyl groups with the OH group of Ph (HDb⋯O1 = 2.328 Å) and the water molecule W2 (HDa⋯Ow2 = 2.852 Å).
In the minimum energy configuration of Ph + 3H2O + DMSO solvation cluster (Fig. 4c), two molecules of water, W1 and W3, are hydrogen bonded with the OH group of Ph (W1: Ow1⋯H1 = 1.670 Å and Ow1⋯H1– O1 = 165.4°; W3: Hw3⋯O1 = 1.819 Å and Ow3–Hw3⋯O1 = 171.4°). W1 is also hydrogen bonded with W2 (Hw1⋯Ow2 = 1.624 Å and Ow1– Hw1⋯Ow2 = 171.0°). W2 is also hydrogen bonded with DMSO (Hw2⋯OD = 1.662 Å and Ow2– Hw2 ⋯OD = 171.6°). It can be noted that the observed coordination of DMSO is stabilized also because of secondary interactions with the water molecule W3 (HD1b⋯Ow3 = 2.434 Å and HD1a⋯Ow1 = 2.378 Å) and the OH group of Ph (HD1a′⋯O1 = 2.823 Å).
The minimum energy configuration of the Ph + 3H2O solvation cluster is shown in Fig. 4d. The three water molecules and the OH group of Ph are hydrogen bonded in a cyclic order with hydrogen bond lengths in the range of 1.677–1.764 Å and hydrogen bond angles 164.6°–169.1°.
Selected structural and conformational properties of Ph(1) + solvent complexes with energy minimization using the B3LYP functional with the 6-31+G(d) basis set are shown in Fig. S2 (the ESI†). The hydrogen bond properties of the Ph + DMSO complex are very similar to those obtained by using the APFD functional. There are minor differences in bond angles and torsion angles; the hydrogen bond lengths are invariably shorter using the APFD functional due to the empirical dispersion correction term for noncovalent intermolecular interactions (Fig. 4 and Fig. S2, ESI†).
Table 1 shows calculated, δcalc(OH) 1H NMR chemical shifts of Ph + DMSO, Ph + 2H2O + DMSO, Ph + 3H2O + DMSO, and Ph + 3H2O solvation complexes, with energy minimization at the B3LYP/6-31+G(d) and APFD/6-31+G(d) levels. In practice, the various configurations of each solvation complex that exist within a 0.0 to 3.0 kcal mol−1 Gibbs energy window, can contribute to the overall computational chemical shift. For a given solvation species of Table 1 a Boltzmann analysis was performed in order to determine the relative contribution of each configuration to the overall computational chemical shift, δcalc,Wg, using the equation56,57
![]() | (1) |
Composition of solvation clusters | B3LYP/6-31+G(d) | APFD/6-31+G(d) | |||||||||
---|---|---|---|---|---|---|---|---|---|---|---|
ΔG | δ calc,g (ppm) | δ calc,Wg (ppm) | δ calc,g (ppm) | δ calc,Wg (ppm) | ΔG | δ calc,g (ppm) | δ calc,Wg (ppm) | δ calc,g (ppm) | δ calc,Wg (ppm) | δ exp | |
(A) GIAO | (B) CSGT | (A) GIAO | (B) CSGT | ||||||||
a Denotes that H2O molecules are hydrogen bonded with the OH group of Ph(1). b Denotes that the DMSO molecule is hydrogen bonded with the OH group of Ph(1). | |||||||||||
Ph + DMSO | 0.33 | 8.61 | 8.81 | 7.65 | 7.90 | 0.00 | 9.69 | 9.35 | 8.93 | 8.66 | 9.53 |
Ph + DMSO | 0.00 | 9.02 | 8.15 | 1.11 | 9.07 | 8.43 | |||||
Ph + 2H2O + DMSOa | 0.00 | 10.50 | 10.41 | 9.93 | 9.76 | 0.00 | 10.84 | 10.84 | 10.29 | 10.27 | 9.99 |
Ph + 2H2O + DMSOa | 1.46 | 9.15 | 8.47 | 1.76 | 9.52 | 8.93 | |||||
Ph + 2H2O + DMSOa | 1.32 | 10.71 | 10.05 | 2.11 | 11.45 | 10.85 | |||||
Ph + DMSO + 2H2Ob | 2.66 | 9.11 | 8.59 | 2.95 | 10.09 | 9.51 | |||||
Ph + DMSO + 2H2Ob | 0.57 | 10.19 | 9.57 | 1.28 | 11.31 | 10.64 | |||||
Ph + DMSO + 3H2Ob | 5.50 | 10.23 | 10.19 | 9.84 | 9.84 | 2.65 | 11.67 | 11.19 | 11.21 | 10.78 | 10.00 |
Ph + 3H2O + DMSOa | 0.00 | 11.19 | 9.84 | 0.00 | 11.18 | 10.78 | |||||
Ph + 3H2O | 0.00 | 10.43 | 9.95 | 9.79 | 9.87 | 0.12 | 11.14 | 11.11 | 10.46 | 10.44 | 9.91 |
Ph + 3H2O | 0.73 | 8.32 | 10.14 | 0.00 | 11.08 | 10.42 |
The δcalc(OH) 1H NMR chemical shifts by the GIAO method, weighted by the respective Boltzmann factors (Table 1), δcalc,Wg, were aligned in accord to the tendencies of the experimental chemical shifts. Thus, a maximum in the chemical shift was obtained with the Ph + 3H2O + DMSO solvation cluster, and a small reduction in the chemical shift of the Ph + 3H2O solvation cluster. By using the CSGT method, the Ph + 3H2O + DMSO solvation cluster has a slightly smaller chemical shift (9.84 ppm) than the Ph + 3H2O cluster (9.87 ppm), contrary to the experimental data. With minimization of the structures at the APFD/6-31+G(d) level, the calculated OH 1H NMR chemical shifts, for both the GIAO and CSGT methods, weighted by the respective Boltzmann factors (Table 1), δcalc,Wg, were aligned in accord to the tendencies of the experimental chemical shifts, δexp. Comparison of the Ph + 3H2O and Ph + 3H2O + DMSO solvation complexes using both B3LYP and APFD functionals shows that: (i) the incorporation of the DMSO molecule increases the size of the cycling structure and (ii) the increase in the chemical shift in the Ph + 3H2O + DMSO complex cannot be attributed to a single structural parameter of the cooperative interactions between H2O and DMSO molecules with the phenol OH group.
Conformational analysis of PCA(2) has been performed with a variety of DFT methodologies by several groups.58,59 Four minimum energy conformers A, B, C and D were obtained with PES calculations of the torsion angles C5–C4–C3–C2 and C3–C2–C1–O1a (Fig. 5). The global minimum energy conformer is A; conformer B, with energy very similar to that of A, has been observed in the X-ray crystal structure of the PCA.60 The stronger intramolecular interaction between the electron lone pair of the carboxylic oxygen, O1a, and the olefinic hydrogen,
C3–H, is responsible for the stabilization of conformers A and B vs. C and D. A similar stabilization is also observed for most of the configurations of the solvation complexes PCA + DMSO, PCA + 2H2O + DMSO, PCA + 3H2O + DMSO and PCA + 3H2O.
![]() | ||
Fig. 5 The four minimum energy conformers of PCA(2) in the gas phase with minimization at the APFD/6-31+G(d) level. |
Selected structural and conformational properties of PCA + solvent complexes, using the APFD/6-31+G(d) and B3LYP/6-31+G(d) energy minimization methods, in the gas phase, are shown in Fig. 6 and Fig. S3 (the ESI†), respectively. The PCA + DMSO hydrogen bond interaction (SO(D)⋯H7 = 1.673 Å, O(D)⋯H7–O7 = 162.5°) is very similar to that of the Ph + DMSO complex. Secondary interactions of the two methyl groups of DMSO with the phenol OH oxygen of PCA (HD1a⋯O7 = 2.735 Å and HD1b⋯O7 = 2.607 Å) are within the limits of the van der Waals radii definition.
In the PCA + 2H2O + DMSO solvation complex, a water molecule (W1) forms a hydrogen bond interaction with the OH group of PCA ((O7)H7⋯Ow1 = 1.666 Å and O7–H7⋯Ow1 = 170.9°). W1 is, also, hydrogen bonded with the second molecule of water (W2) (Ow2⋯Hw1 = 1.668 Å and Ow1–Hw1⋯Ow2 = 176.3°). W2 is hydrogen bonded to the (S)O of the DMSO molecule (Hw2⋯OD = 1.688 Å, Ow2–Hw2⋯OD = 175.7°). It can be noted that this preferential coordination of the DMSO molecule is stabilized, also, because of secondary interactions HD1a⋯O7 = 2.411 Å, and HD1b⋯O7 = 3.512 Å (Fig. 6b).
In the PCA + 3H2O + DMSO solvation cluster, two molecules of water, W1 and W3, are hydrogen bonded with the OH group of PCA (W1: H7⋯Ow1 = 1.652 Å and O7–H7⋯Ow1 = 177.4°; W3: O7⋯Hw3 = 1.836 Å and O7⋯Hw3–Ow3 = 166.5°). W1 is also hydrogen bonded to W2 (Hw1⋯Ow2 = 1.690 Å and Ow1–Hw1⋯Ow2 = 170.7°). W2 is also hydrogen bonded to DMSO (Hw2⋯OD = 1.645 Å and Ow2–Hw2⋯OD = 170.5 Å). It can be noted that this coordination of the DMSO is, furthermore, stabilized because of secondary interactions with the water molecule W3 (HD1b⋯Ow3 = 2.506 Å, HD1a⋯Ow3 = 2.529 Å).
The minimum energy configuration of the PCA + 3H2O solvation cluster is shown in Fig. 6d. The three water molecules and the OH group of the PCA molecule are hydrogen bonded in a cyclic order with hydrogen bond lengths in the range of 1.655 Å to 1.806 Å and hydrogen bond angles 164.5° to 170.3°.
Selected structural and conformational properties of PCA + solvent complexes, using the B3LYP/6-31+G(d) energy minimization method, in the gas phase, are shown in Fig. S3 (the ESI†). The hydrogen bond lengths are longer with the B3LYP functional, as in the case of phenol (1) + solvent complexes.
The calculated 1H-NMR chemical shifts, δcalc,g, of the four conformational states A, B, C and D (Fig. 5) and various solvation configurations of PCA, in the gas phase, are shown in Table 2. The 1H NMR chemical shifts with the GIAO and CSGT methods, of the PCA + DMSO + 3H2O and PCA + 3H2O complexes, weighted by the respective Boltzmann factors, were aligned in accord to the tendencies of δexp, with energy minimization at the APFD/6-31+G(d) level. Similar tendencies were obtained at the B3LYP36-31+G(d)/CSGT level (Table 2). Comparison of the PCA + 3H2O and PCA + 3H2O + DMSO solvation complexes using both B3LYP and APFD functionals shows that: (i) the incorporation of the DMSO molecule increases the size of the cycling structure and (ii) the increase in the chemical shift in the PCA + 3H2O + DMSO complex cannot be attributed to a single structural parameter of the cooperative interactions between H2O and DMSO molecules with the PCA OH group.
B3LYP/6-31+G(d) | APFD/6-31+G(d) | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|
Composition of solvation conformersa | ΔG | δ calc,g (ppm) | δ calc,Wg (ppm) | δ calc,g (ppm) | δ calc,Wg (ppm) | ΔG | δ calc,g (ppm) | δ calc,Wg (ppm) | δ calc,g (ppm) | δ calc,Wg (ppm) | δ exp |
(A) GIAO | (B) CSGT | (A) GIAO | (B) CSGT | ||||||||
a For the definition of conformers A, B, C, and D see Fig. 5. b Denotes that H2O molecules are hydrogen bonded with the OH group of PCA. c Denotes that the DMSO molecule is hydrogen bonded with the OH group of PCA. | |||||||||||
PCA+ DMSO conformer A | 0.03 | 9.97 | 9.97 | 9.04 | 9.05 | 0.00 | 10.58 | 10.57 | 9.88 | 9.86 | 10.19 |
PCA+ DMSO conformer B | 0.00 | 9.96 | 9.05 | 0.12 | 10.56 | 9.85 | |||||
PCA + DMSO conformer C | 0.78 | 9.98 | 9.06 | 0.84 | 10.58 | 9.87 | |||||
PCA + DMSO conformer D | 0.77 | 9.96 | 9.04 | 0.77 | 10.57 | 9.86 | |||||
PCA + 2H2O + DMSOb conformer A | 0.00 | 11.37 | 11.35 | 10.76 | 10.73 | 0.12 | 11.68 | 11.65 | 11.13 | 11.09 | 10.64 |
PCA + 2H2O + DMSOb conformer B | 0.60 | 11.39 | 10.76 | 0.00 | 11.68 | 11.12 | |||||
PCA + 2H2O + DMSOb conformer C | 0.59 | 11.37 | 10.75 | 0.73 | 11.69 | 11.13 | |||||
PCA + 2H2O + DMSOb conformer D | 1.13 | 11.40 | 10.77 | 0.89 | 11.69 | 11.14 | |||||
PCA + 2H2O + DMSOb conformer A | 2.34 | 9.97 | 9.28 | 2.42 | 10.29 | 9.65 | |||||
PCA + 2H2O + DMSOb conformer B | 2.51 | 9.92 | 9.25 | 2.15 | 10.17 | 9.53 | |||||
PCA + 2H2O + DMSOb conformer C | 3.02 | 9.99 | 9.25 | 2.42 | 10.19 | 9.51 | |||||
PCA + 2H2O + DMSOb conformer D | 2.77 | 9.95 | 9.27 | 2.99 | 10.24 | 9.56 | |||||
PCA + H2O + DMSO + H2Ob conformer A | 2.04 | 11.62 | 11.00 | 2.69 | 12.24 | 11.74 | |||||
PCA + H2O + DMSO + H2Ob conformer B | 2.09 | 11.61 | 11.00 | 2.12 | 12.24 | 11.73 | |||||
PCA + H2O + DMSO + H2Ob conformer C | 2.54 | 11 | 11.00 | 3.33 | 12.21 | 11.72 | |||||
PCA + H2O + DMSO + H2Ob conformer D | 2.64 | 11.62 | 11.01 | 3.37 | 12.26 | 11.75 | |||||
PCA + DMSO + 2H2Oc conformer A | 4.93 | 10.58 | 9.81 | 3.83 | 11.61 | 11.00 | |||||
PCA + DMSO + 2H2Oc conformer B | 5.84 | 10.60 | 9.81 | 3.71 | 11.61 | 10.97 | |||||
PCA + DMSO + 2H2Oc conformer C | 6.93 | 10.65 | 10.06 | 4.60 | 11.59 | 10.98 | |||||
PCA + DMSO + 2H2Oc conformer D | 5.60 | 10.56 | 9.82 | 4.67 | 11.64 | 10.92 | |||||
PCA +3H2O + DMSOb conformer A | 0.00 | 11.07 | 11.09 | 10.55 | 10.55 | 0.00 | 12.02 | 12.00 | 11.45 | 11.40 | 10.64 |
PCA +3H2O + DMSOb conformer B | 0.20 | 11.13 | 10.56 | 0.18 | 11.99 | 11.42 | |||||
PCA +3H2O + DMSOb conformer C | 0.61 | 11.09 | 10.55 | 0.98 | 12.02 | 11.44 | |||||
PCA +3H2O + DMSOb conformer D | 0.68 | 11.09 | 10.54 | 0.96 | 11.99 | 11.43 | |||||
PCA + DMSO + 3H2Oc conformer A | 3.45 | 10.84 | 10.26 | 4.59 | 10.61 | 9.98 | |||||
PCA + DMSO + 3H2Oc conformer B | 3.30 | 10.85 | 10.26 | 4.09 | 10.60 | 9.92 | |||||
PCA + DMSO + 3H2Oc conformer C | 3.83 | 10.84 | 10.27 | 5.34 | 10.59 | 9.94 | |||||
PCA + DMSO + 3H2Oc conformer D | 3.87 | 10.84 | 10.27 | 5.25 | 10.65 | 9.97 | |||||
PCA + 3H2O conformer A | 0.00 | 11.08 | 11.10 | 10.51 | 10.51 | 0.00 | 11.65 | 11.67 | 11.14 | 11.13 | 10.50 |
PCA + 3H2O conformer B | 0.18 | 11.14 | 10.50 | 0.16 | 11.71 | 11.13 | |||||
PCA + 3H2O conformer C | 0.56 | 11.07 | 10.49 | 0.88 | 11.68 | 11.10 | |||||
PCA + 3H2O conformer D | 0.69 | 11.10 | 10.52 | 0.93 | 11.64 | 11.07 |
The proximity of hydroxyl and methoxy groups at the ortho position in VA(3) may result in the formation of intramolecular hydrogen bonds.61,62 Thus, the X-ray structure determination of vanillic acid and theobromide co-crystal hydrate (VA·THBR·2 H2O)63,64 shows that the hydroxyl and methoxy groups of VA adopt an in-out conformation which results in the formation of an intramolecular hydrogen bond. A water molecule acts as a bifurcated acceptor to the OH group of VA (O–H⋯O(W1) = 2.719(2) Å, H⋯O(W1) = 1.859(3) Å and O–H⋯O(W1) = 165(2)°) and to a second water molecule ((W2) O–H⋯O(W1) = 2.868(2) Å, H⋯O(W1) = 2.064 Å and (W2) O–H⋯O(W1) = 157(3)°). The oxygen atom of the second water molecule forms also a hydrogen bond with the second hydrogen atom of the first H2O in a ring motif. The X-ray structure determination of VA(3) and isonicotinamide hydrate (2VA·2INM·2H2O)63,64 shows that the hydroxyl and methoxy groups of VA adopt the out–out conformation with a water molecule playing a bridging role ((W) O–H⋯O(H)VA = 2.945(2) Å, H⋯O(H)VA = 2.296(2) Å and (W) O–H⋯O(H)VA = 133(2)°; (W) O–H⋯O(CH3)VA = 3.051(2) Å, H⋯O(CH3)VA = 2.243(3) Å and (W) O–H⋯O(CH3)VA = 158(2)°). This water molecule also links two VA molecules.
Several solvation species of VA, including intra- and intermolecular hydrogen bonds were investigated (Fig. 7). Selected structural and conformational properties of VA(3) + solvent complexes with energy minimization using the APFD and B3LYP functionals with the 6-31+G(d) basis set are shown in Fig. S4 and S5 (the ESI†), respectively. In the minimum energy complex of vanillic acid + DMSO, the phenol OH hydrogen forms a bifurcated hydrogen bond with DMSO (H4⋯OD = 1.716 Å and O4–H4⋯OD = 153.9°) and O3 of the methoxy group (H4⋯O3 = 2.241 Å). Secondary interactions of the methyl groups of DMSO with the phenol OH group (HD1a⋯O4 = 2.695 Å and HD1b⋯O4 = 2.502 Å) contribute to the stability of this particular configuration. It can be noted that the torsion angle φ(C3C4O4H4) = −39.3°, strongly deviates from planarity.
![]() | ||
Fig. 7 The four minimum energy conformers of VA in the gas phase with energy minimization at the B3LYP/6-31+G(d) level. |
In the VA + 2H2O + DMSO solvation complex (Fig. S4b, ESI†), the hydroxyl and methoxy groups adopt an anti-conformation. A water molecule (W1) forms a hydrogen bond interaction with the OH group of VA (H4⋯Ow1 = 1.689 Å and O4–H4⋯Ow1 = 165.0°). W1 is hydrogen bonded with the second molecule of water (W2) (Ow2⋯Hw1 = 1.685 Å and Ow1–Hw1⋯Ow2 = 176.1°). W2 is hydrogen bonded with the S(O) of the DMSO molecule (Hw2⋯OD = 1.659 Å and Ow2–Hw2⋯OD = 173.4°). It can be noted that this preferential coordination of the DMSO molecule is stabilized also because of secondary interactions HDa⋯O4 = 2.628 Å and HD2⋯O3 = 2.413 Å (Fig. S4b, ESI†).
In the minimum energy configuration of the VA + 3H2O + DMSO solvation cluster (Fig. S4c, ESI†), the hydroxyl and methoxy groups adopt an anti-conformation. Two molecules of water, W1 and W3, are hydrogen bonded with the OH group of VA (W1: Ow1⋯H4 = 1.680 Å and Ow1–H4⋯O4 = 168.8°; W3: Hw3⋯O4 = 1.936 Å and Ow3–Hw3⋯O4 = 137.0°). W1 is also hydrogen bonded with water molecule W2 (Hw1⋯Ow2 = 1.659 Å and Ow1–Hw1⋯Ow2 = 176.1°). W2 is also hydrogen bonded with DMSO (Hw2⋯OD = 1.644 Å and Ow2–Hw2⋯OD = 174.7°). The methyl groups of DMSO form secondary interactions with the water molecule W3 (Ow3⋯HD3b = 2.202 Å) and the methoxy group (O3⋯HD3a = 2.615 Å). The torsion angle φ(C5C4O4H4) = −8.43° is significantly reduced relative to that in the VA + DMSO complex.
In the minimum energy configuration of the VA + 3H2O solvation cluster, the hydroxyl and methoxy groups adopt an anti-conformation (Fig. S4d, ESI†). The three water molecules and the OH group of VA are hydrogen bonded in a cyclic order with hydrogen bond lengths in the range of 1.589–1.763 Å and hydrogen bond angles 152.9°–164.2°. The water molecule W3 forms a very short hydrogen bond with O4 of VA (1.589 Å), however, the hydrogen bond angle Ow3–Hw3⋯O4 strongly deviates from linearity (152.9°). The torsion angle φ(C5C6C4O4) = −14.5° is larger than that of the VA + 3H2O + DMSO solvation cluster.
Selected structural and conformational properties of VA + solvent complexes with energy minimization using the B3LYP functional with the 6-31+G(d) basis set in the gas phase, are shown in Fig. S5 (the ESI†). The structures and hydrogen bond properties are very similar to those obtained by the use of APFD functional. The hydrogen bond lengths are longer and the φ torsion angles smaller by using the B3LYP functional.
Fig. S6 (ESI†) illustrates minimum energy structures of VA + 2H2O and VA + H2O complexes. In the VA + 2H2O solvation cluster the hydroxyl and methoxy groups adopt anti-conformation as in the case of VA + 3H2O. The two molecules of water and the OH group of VA are hydrogen bonded in a cyclic order with hydrogen bond lengths of 1.745–1.957 Å, which are significantly longer than those in the VA + 3H2O complex. This results also in hydrogen bond angles which strongly deviate from linearity (138.2°–152.5°). In the VA + H2O complex the hydroxyl and methoxy groups adopt an in-out conformation which results in a rather weak intramolecular hydrogen bond (O3⋯H4 = 2.376 Å). The water molecule W1 plays a bridging role, forming a hydrogen bond interaction with the OH group (Ow1⋯H4 = 1.759 Å and Ow1⋯H4–O4 = 167.7°) and the methoxy group of VA (Hw1⋯O3 = 1.909 Å and Ow1–Hw1⋯O3 = 136.6°).
From the above it can be concluded that in the case of VA + DMSO complexes the most stable configurations are those with an intramolecular hydrogen bond interaction (Table S6, ESI†). In the case of VA + 2H2O + DMSO solvation complexes most of the open configurations are significantly more stable than those with intramolecular hydrogen bond interactions. Similarly, in VA + 3H2O + DMSO solvation complexes the open configurations are the most stable with negligible population of the molecular clusters with intramolecular hydrogen bond interactions (Table S6, ESI†). The proximity of hydroxyl and methoxy groups in vanillic acid may result also in reduced solvent accessibility. Contrary to the case of VA + 3H2O and VA + 2H2O, low energy configurations of the VA + H2O complex with intramolecular hydrogen bond interactions were obtained.
The computational 1H NMR chemical shifts of the solvation complex VA + 3H2O + DMSO, with energy minimization using the B3LYP/6-31+G(d) and the APFD/6-31+G(d) methods, weighted by the respective Boltzmann factors, δcalc,wg, values were found to be smaller than those of the VA + 3H2O complex for both the GIAO and CSGT methods in the gas phase. This is not in agreement with the tendencies of the experimental chemical shifts (Table S6, ESI†). Since the computational 1H NMR chemical shifts of the VA + 2H2O and VA + H2O complexes are significantly smaller than those of the VA + 3H2O + DMSO complex, very probably, in solution state, there is an equilibrium of various VA + nH2O (n = 3–1) solvation species.
From the comparison of experimental and computational NMR chemical shifts of the phenol OH groups of the natural products of Fig. 1, it appears to be conclusive that hydrogen bond interactions in the solvation clusters of natural product + 3H2O + DMSO are stronger than in DMSO or H2O. The enhanced hydrogen bond interactions can be attributed to the significant cooperative nature of the interaction between the H2O and DMSO molecules and the OH groups of the natural products.
![]() | ||
Fig. 8 Structural similarities of the DMSO + 3D2O complex motif of the neutron powder diffraction data (a),65 with the minimum energy structures of Ph(1) + 2H2O + DMSO (b), PCA(2) + 2H2O + DMSO (conformer D) (c) and VA(3) + 2H2O + DMSO (d) with energy minimization using the B3LYP/6-31+G(d) method in the gas phase. |
δtrue − δobs = ⅓(κsample − κreference) |
Footnote |
† Electronic supplementary information (ESI) available: 1H NMR spectral region, minimum energy structures of solvation species of phenol, paracoumaric acid and vanillic acid, tables of proton exchange rates. See DOI: 10.1039/d0cp05861k |
This journal is © the Owner Societies 2021 |