Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Can we predict specific numbers of catalytically important molecules of water in H/D exchange in aromatic systems? A combined NMR and DFT study

Photini Chalkidou , Themistoklis Venianakis , George Papamokos , Michael Siskos * and Ioannis P. Gerothanassis *
Section of Organic Chemistry and Biochemistry, Department of Chemistry, University of Ioannina, Ioannina, GR 45110, Greece. E-mail: msiskos@uoi.gr; igeroth@uoi.gr

Received 22nd July 2024 , Accepted 3rd February 2025

First published on 3rd February 2025


Abstract

Base-catalyzed H/D exchange reactions through keto–enol tautomeric equilibrium are a textbook example in mechanistic organic chemistry. The pH effect of H2O catalysis, however, is largely unknown. We report, herein, variable temperature and pD 1H NMR studies of the experimental activation enthalpy image file: d4nj03276d-t1.tif, entropy image file: d4nj03276d-t2.tif, and Gibbs free energy image file: d4nj03276d-t3.tif of H/D exchange reactions of the H-6 and H-8 protons belonging to ring A of the natural product taxifolin. The experimental image file: d4nj03276d-t4.tif values range from ∼25 to 23 kcal mol−1 for pD values of 6.1 to 9.6 and a buffer concentration in the range of 25 to 1000 mM. Differences in image file: d4nj03276d-t5.tif values of neutral and anionic taxifolin and phloroglucinol were found to be very small (≤1.5 kcal mol−1). The experimental data of taxifolin and phloroglucinol were compared with DFT calculations with two up to four H2O molecules explicitly present, which demonstrate a unique catalytic role of H2O of over 35 kcal mol−1. Excellent agreement between image file: d4nj03276d-t6.tif and DFT calculated Gibbs free activation energies, image file: d4nj03276d-t7.tif, was obtained with the use of three molecules of H2O for the neutral state of phloroglucinol (with the “in–in” configuration of the phenol OH groups) and taxifolin. In the ionic form of phloroglucinol, the mechanistic pathway with two molecules of H2O in the transition state (one of which involves the C[double bond, length as m-dash]O moiety) showed very good agreement with the experimental data. For the anionic form of taxifolin, the mechanistic pathway with three molecules of H2O in the transition state showed excellent agreement with the experimental image file: d4nj03276d-t8.tif values. Among the various functionals used, the APFD/6-31+G(d) and B3LYP/6-31+G(d)/GD3BJ resulted in optimum agreement with image file: d4nj03276d-t9.tif. The enthalpic term image file: d4nj03276d-t10.tif is considerably larger than the entropic term image file: d4nj03276d-t11.tif, in agreement with the experimental data. This indicates a dissociative mechanism of the loosely bound activated complex. The present results demonstrate the unique catalytic role of two and/or three molecules of H2O, through keto–enol tautomerization, with minor contribution of base-catalysis, in H/D exchange reactions in aromatic systems.


1. Introduction

The significant role of H2O in enhancing the rate/selectivity in the aqueous phase or adsorbed H2O molecular environment has been extensively emphasized and reviewed in recent literature.1–5 The cooperative hydrogen bond network formed by H2O molecules can significantly affect the structural and electronic properties of solute molecules. Thus, H2O solvation can result in ionic dissociation and classical acid–base catalysis of organic reactions as a proton donor or an acceptor reaction partner. However, addressing how discrete solvation molecules of H2O affect chemical reactivity at the atomic level is a complex issue requiring an appropriate combination of experimental and computational methods.1–6

Acid- and base-catalyzed H/D exchange in aromatic systems is a textbook example in mechanistic organic chemistry. Thus, acid-catalyzed H/D electrophilic aromatic substitution reactions have been extensively utilized since the 1960s with the use of strong deuterated Brønsted or Lewis acids7–11 in the formation of a non-classical σ-carbocation. Base-catalyzed H/D exchange reactions can also provide an alternative method, which has been extensively utilized since the 1960s for the exchange of aromatic protons through keto–enol equilibrium.10,12–14 To the best of our knowledge, the first report demonstrating the catalytic role of H2O appeared in 2015 by Mehr et al.15 in the deuteration of aromatic rings through keto-enamine tautomeric equilibrium. The computational activation energy was found to be significantly reduced with the incorporation of a chain of five H2O molecules. Bonaldo et al.16 reported 1H NMR kinetic measurements of the H/D exchange process on ring A of several classes of flavonoids. It was suggested that the reaction mechanism involves a slow, solvent-mediated, C–D bond formation/O–D bond breaking followed by a fast solvent-mediated C–H bond breaking/O–D bond formation resulting in the aromaticity of ring A. DFT calculations with a single solvation molecule of H2O resulted in a significant reduction in the activation energy which, however, was found to be ∼30 kJ mol−1 higher than the experimental values. Fayaz et al.17 reported the direct activation of aromatic C–H bonds in polyphenolic compounds in a single step, using D2O, at neutral pD and near ambient temperatures. NMR and DFT calculations supported the significant catalytic role of two H2O molecules in a keto–enol tautomerization process.

From the above, it is evident that: (i) H2O molecules can have a very significant role in the transition states and activation energy barriers and also in the selection of a reaction pathway through H2O-mediated hydrogen/deuterium reaction mechanisms,1–6,16–26 and (ii) the traditional and widespread rational behind the primary role of acid or basic catalysis of H/D exchange in aromatic systems is incomplete. Along these lines, we report herein detailed variable temperature and pD dependent 1H NMR and DFT computational studies of neutral and ionic states of taxifolin and phloroglucinol (Scheme 1) in an effort to investigate whether specific numbers of H2O molecules6,27,28 can play a significant catalytic role in the H/D exchange process in polyphenolic aromatic systems.


image file: d4nj03276d-s1.tif
Scheme 1 Structures of taxifolin and phloroglucinol.

2. Results and discussion

2.1. NMR studies

The H/D exchange reaction of taxifolin was investigated using 1D 1H NMR spectroscopy with variable temperatures, pD and concentrations of phosphate buffer. The assignment of the resonances was confirmed using 2D 1H–13C HSQC and HMBC experiments. Fig. 1 illustrates 1H NMR spectra of the H-6 and H-8 protons of taxifolin recorded at pD = 6.00 and various temperatures and time intervals. The aromatic protons of the ring B do not undergo H/D exchange and, thus, were used as an integration reference along with the reference compound TSP-d4. The H/D exchange at C-8 and C-6 protons of ring A, in contrast, was clearly observed at pD = 6.0, 7.6 and 9.6 and at various temperatures and buffer concentrations. The decay of the NMR signals of the C-6 and C-8 protons as a function of time was recorded at pD values 6.0 (Fig. 2 and 3) and 9.6 (Fig. 4 and 5). At pD = 6.0 taxifolin exists essentially in the neutral form and at pD = 9.6 in the mono-deprotonated form.29 The relative integrals of the C-8 and C-6 protons follow first-order kinetics with reactivity order C-6 > C-8 protons, with an increase in the exchange rate at pD = 9.60 as compared with pD = 6.00.
image file: d4nj03276d-f1.tif
Fig. 1 1D 1H NMR (500 MHz) spectral region illustrating H-6 and H-8 H/D exchange of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), pD = 6.00. The spectra A(a) (T = 313 K), B(a) (T = 323 K), and C(a) (T = 328 K) were recorded at a time interval of 0 h. The spectra A(b) (T = 313 K), B(b) (T = 323 K), and C(b) (T = 328 K) were recorded at a time interval of 14 h.

image file: d4nj03276d-f2.tif
Fig. 2 H-6 H/D exchange kinetic curves of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), and pD = 6.00 at various temperatures.

image file: d4nj03276d-f3.tif
Fig. 3 H-8 H/D exchange kinetic curves of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), and pD = 6.00 at various temperatures.

image file: d4nj03276d-f4.tif
Fig. 4 H-6 H/D exchange kinetic curves of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), and pD = 9.60 at various temperatures.

image file: d4nj03276d-f5.tif
Fig. 5 H-8 H/D exchange kinetic curves of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), and pD = 9.60 at various temperatures.

The Gibbs energy of activation, image file: d4nj03276d-t12.tif (kcal mol−1), is given by:

 
image file: d4nj03276d-t13.tif(1)
where image file: d4nj03276d-t14.tif is the enthalpy of activation (kcal mol−1) and image file: d4nj03276d-t15.tif is the entropy of activation (kcal mol−1 K−1). According to the Eyring equation:
 
image file: d4nj03276d-t16.tif(2)
where kB is the Boltzmann's constant, T is the absolute temperature in Kelvin (K), K is the rate constant in s−1, R is the ideal gas constant, and h is the Planck's constant. The values for image file: d4nj03276d-t17.tif and image file: d4nj03276d-t18.tif, therefore, can be determined from kinetic data, obtained from a image file: d4nj03276d-t19.tif plot, with a negative slope, image file: d4nj03276d-t20.tif, and a y-intercept image file: d4nj03276d-t21.tif. Fig. 6 shows Eyring plots of the H-6 and H-8 protons of taxifolin (2.5 mM) in D2O, phosphate buffer 25 mM and at pD = 6.00 with very good linear regression correlation coefficients (R2 = 0.988 for H-6 and 0.982 for H-8).


image file: d4nj03276d-f6.tif
Fig. 6 Eyring plots of the H-6 (R2 = 0.988) and H-8 (R2 = 0.982) of taxifolin (2.5 mM) in D2O, phosphate buffer (25 mM), and pD = 6.00.

The resulting image file: d4nj03276d-t22.tif, image file: d4nj03276d-t23.tif and image file: d4nj03276d-t24.tif, for H/D exchange of taxifolin for various pD and phosphate buffer concentrations, are shown in Table 1. In all cases, the enthalpy term image file: d4nj03276d-t25.tif makes a major contribution to image file: d4nj03276d-t26.tif values with a minor entropic image file: d4nj03276d-t27.tif term contribution. The image file: d4nj03276d-t28.tif values for H-8 and H-6 protons were found to be in the range of 24.67 to 25.01 and 22.81 to 24.53 kcal mol−1, respectively. The effect of phosphate buffer concentration (25 mM to 1 M) on image file: d4nj03276d-t29.tif values was ≤ 0.7 kcal mol−1, thus, its catalytic role is of minor importance.

Table 1 Activation enthalpy image file: d4nj03276d-t30.tif, activation entropy image file: d4nj03276d-t31.tif and Gibbs activation energy image file: d4nj03276d-t32.tif for H/D exchange reactions of the H-8 and H-6 protons of taxifolin (2.5 mM) at various pD values and phosphate buffer concentrations
Compound Buffer concentration pD H-8 H-6

image file: d4nj03276d-t33.tif

(kcal mol−1)

image file: d4nj03276d-t34.tif

(kcal mol−1)

image file: d4nj03276d-t35.tif

(kcal mol−1)

image file: d4nj03276d-t36.tif

(kcal mol−1)

image file: d4nj03276d-t37.tif

(kcal mol−1)

image file: d4nj03276d-t38.tif

(kcal mol−1)
a Ref. 17.
Taxifolin 25 mM 6.0 20.31 ± 1.57 4.70 ± 0.79 25.01 18.91 ± 1.18 5.62 ± 0.59 24.53
25 mM 7.6 18.79 ± 1.59 5.80 ± 0.80 24.59 14.01 ± 1.47 9.17 ± 0.82 23.18
50 mM 7.6 19.34 ± 1.18 5.05 ± 0.65 24.39 20.63 ± 1.26 2.24 ± 0.49 22.87
25 mM 9.6 18.43 ± 1.79 6.24 ± 0.72 24.67 16.67 ± 0.55 6.15 ± 0.55 22.81
50 mM 9.6 19.62 ± 0.77 4.07 ± 0.65 23.69 16.96 ± 0.54 5.51 ± 0.58 22.46
1 M 9.6 15.96 ± 1.22 7.98 ± 0.78 23.94 11.76 ± 0.75 10.38 ± 0.98 22.14
Phloroglucinola H-2,4,6
25 mM 6.9 17.46 ± 0.30 3.50 ± 0.09 20.96
25 mM 7.9 16.05 ± 0.78 3.69 ± 0.26 19.74
25 mM 8.9 16.55 ± 1.15 2.86 ± 0.29 19.41


2.2. DFT mechanistic studies

The catalytic role of discrete molecules of H2O in the neutral and ionic forms of taxifolin and phloroglucinol was investigated using various DFT levels of theory (B3LYP, PBE0, APFD, M06-2X, ωB97X-D, B3LYP/GD3BJ, PBE0/GD3BJ, and CAM-B3LYP/GD3BJ) and the 6-31+G(d) basis set.30–32 Basis sets of this size give better results than large basis sets due to the cancellation of errors.33 At the highest level of theory and basis set, it is evident that all protons move simultaneously.
2.2.1. The neutral form of phloroglucinol and taxifolin-effects of two-to-four molecules H2O. Various mechanistic pathways of the neutral form of phloroglucinol were examined with two to four molecules of H2O in various “in–in”, and “in–out” orientations of the phenol OH groups (Table 2 and Table S1, ESI). The computational data with two molecules of H2O in the transition state resulted in image file: d4nj03276d-t39.tif values in the range of 23.15 to 27.11 kcal mol−1 which deviate from the experimental value of 20.96 kcal mol−1 (Table 2 and Table S1, ESI). The only exceptions are the computational data of the APFD/6-31+G(d) method with image file: d4nj03276d-t40.tif, which is in excellent agreement with the experimental value (Fig. 7). Of particular interest is the lengthening of the C(1)O–H bond from 0.983 Å in the ground state to 1.540 Å in the transition state and the shortening of the HO–H⋯C(2)H distance from 2.425 Å in the ground state to 1.523 Å in the transition state. The computational data demonstrate that the enthalpic activation energy image file: d4nj03276d-t41.tif is significantly larger than the entropic activation energy image file: d4nj03276d-t42.tif which is in excellent agreement with the experimental data (image file: d4nj03276d-t43.tif and image file: d4nj03276d-t44.tif). The minor role of the image file: d4nj03276d-t45.tif term could be attributed to the relatively small increase of entropy in the transition state, which indicates a dissociative mechanism of the loosely bound activated complex.
Table 2 Selected computed Gibbs activation energy image file: d4nj03276d-t53.tif (kcal mol−1) for neutral and ionic forms of phloroglucinol and taxifolin for various molecular H2O solvation species and computational methods (for complete set of data see Tables S1–S3, ESI). In parenthesis are the Boltzmann populations
Neutral phloroglucinol +2H2O +3H2O Exp.
“in–out” “in–in”
“in–in A”
a A transition state could not be determined.
APFD/6-31+G(d) 20.59 26.28 (10.94%) 20.13 (58.52%) 20.96
19.75 (30.51%)
B3LYP/6-31+G(d)/GD3BJ 23.15 27.52 (14.19%) 21.66 (46.80%)
21.54 (38.19%)
PBE0/6-31+G(d)/GD3BJ 23.16 27.76 (9.65%) 21.65 (51.47%)
21.48 (38.75%)
ωB97XD/6-31+G(d) 26.19 30.86 (17.12%) 25.08 (52.04%)
24.76 (30.36%)
CAM-B3LYP/6-31+G(d)/GD3BJ 24.68 29.71 (16.10%) 23.99 (46.39%)
23.62 (37.37%)

Neutral taxifolin +2H2O +3H2O Exp.
C-6 C-8 C-6 C-8 C-6 C-8
APFD/6-31+G(d) 20.79 20.77 22.98 23.55 24.53 25.01
B3LYP/6-31+G(d)/GD3BJ 20.81 21.93 25.63 24.72
PBE0/6-31+G(d)/GD3BJ 20.80 21.28 25.18 24.67
ωB97XD/6-31+G(d) 25.08 24.19 27.38 27.37
CAM-B3LYP/6-31+G(d) 24.78 25.67 29.12 28.81
CAM-B3LYP/6-31+G(d)/GD3BJ 22.69 23.59 27.26/27.05 27.66

Ionic phloroglucinol +2H2O +3H2O Exp.
on OH C[double bond, length as m-dash]O (A) C[double bond, length as m-dash]O (B)
APFD/6-31+G(d) 11.83 19.15 11.26 19.74
B3LYP/6-31+G(d)/GD3BJ 13.01 (0.26%) 19.91 (33.85%) 19.86 (65.89%) 11.63
PBE0/6-31+G(d)/GD3BJ 13.05 19.71 11.48
ωB97XD/6-31+G(d) 15.21 (0.37%) 22.16 (41.71%) 21.37 (57.92%) 14.07
CAM-B3LYP/6-31+G(d) 14.39 (0.32%) 21.69 (60.95%) 20.58 (38.73%) 13.36
CAM-B3LYP/6-31+G(d)/GD3BJ 14.08 (0.30%) 21.32 (39.49%) 20.47 (60.20%) 12.95

Ionic taxifolin +2H2O +3H2O Exp.
C-6 C-8 C-6 C-8
B3LYP/6-31+G(d)/GD3BJ 23.83 24.27 23.18 24.59
PBE0/6-31+G(d)/GD3BJ 24.76 26.00
ωB97XD/6-31+G(d) 24.51 26.18
CAM-B3LYP-D/6-31+G(d) 23.44 22.80



image file: d4nj03276d-f7.tif
Fig. 7 The mechanistic pathway of the aromatic hydrogen exchange process of the complex of the neutral phloroglucinol (Flu) with two molecules of H2O with the APFD/6-31+G(d) method.

In the case of the complex of phloroglucinol with three molecules of H2O, four reactions mechanisms were examined with the OH groups in the “in–out” and “in–in” configuration (Fig. 8). In the “in–out” and “in–out A” configurations (with slightly different hydrogen bond network of the molecules of H2O, Fig. 8A) two image file: d4nj03276d-t46.tif values of 26.28 and 23.03 kcal mol−1 were obtained with relative Boltzmann populations of 10.9% and 0.04%, respectively, which deviate significantly from the experimental value image file: d4nj03276d-t47.tif. In contrast, in the “in–in” and “in–in A” configurations, the resulting image file: d4nj03276d-t48.tif values of 20.13 and 19.75 kcal mol−1 with relative Boltzmann populations of 58.5% and 30.5% are in excellent agreement with the experimental value. In the transition state of the “in–in” and “in–in A” configurations, the oxygen atom of the central molecule of H2O forms a closer contact with the C(2)–H bond of 1.451 Å, contrary to 1.600 Å in the “in–out” and in the “in–out A” configurations (Fig. 8). This demonstrates that the relative configuration of the phenol OH groups could have a significant effect on the ΔG values and, thus, on the reaction mechanisms. Again, the image file: d4nj03276d-t49.tif term is significantly larger than the image file: d4nj03276d-t50.tif term, which is in excellent agreement with the experimental data.


image file: d4nj03276d-f8.tif
Fig. 8 The mechanistic pathways of the aromatic hydrogen exchange process of the complex of neutral phloroglucinol with three molecules of H2O and the two phenol –OH groups in the “in–out” (A) and “in–in” (B) configurations at the APFD/6-31+G(d) level. In parenthesis are the Boltzmann populations (Table 2 and Table S1, ESI).

In the case of the complex of phloroglucinol with four molecules of H2O, three reaction mechanisms were examined with the OH groups in the “in–out”, “in–in”. and “in–in A” configuration (Fig. S1, ESI). In the “in–in A” and “in–in” configurations, the image file: d4nj03276d-t51.tif (at the B3LYP/6-31+G(d)/GD3BJ level) with relative Boltzmann populations of 77.9% and 22.6%, respectively, are in excellent agreement with the experimental value. The image file: d4nj03276d-t52.tif value of 21.97 kcal mol−1 with a Boltzmann population of 0.02% deviates from the experimental value. It can be concluded that: (i) the increase in the number of catalytically important molecules of H2O from three to four does not improve significantly the agreement of computational results with the experimental data. (ii) In all complexes with two to four molecules of H2O, the relatively small increase of entropy in the transition state indicates a dissociative mechanism of the loosely bound activated complex.

From the mechanistic point of view, the case of taxifolin is simpler than that of phloroglucinol since the C-6 and C-8 positions were examined with a single orientation of the OH(5) group due to the formation of a strong intramolecular hydrogen bond with the C(4)O group,34–37 which persists in organic and aqueous solutions.34 This significantly reduces the number of the available mechanistic pathways in the transition state. The computational Gibbs activation energies, with two molecules of H2O in the transition state, at the ωB97XD/6-31+G(d) level image file: d4nj03276d-t54.tif and image file: d4nj03276d-t55.tif and at the CAM-B3LYP/6-31+G(d) level (image file: d4nj03276d-t56.tif and image file: d4nj03276d-t57.tif) are in excellent agreement with image file: d4nj03276d-t58.tif and (image file: d4nj03276d-t59.tif (Table 2 and Table S2, ESI)). It is of interest that at the CAM-B3LYP/6-31+G(d) level, the small increase in image file: d4nj03276d-t60.tif (C-8) relative to image file: d4nj03276d-t61.tif (C-6) is reproduced correctly. The other functionals result in smaller image file: d4nj03276d-t62.tif values, such as with the APDF/6-31+G(d) method (Fig. 9A). The use of three molecules of H2O in the transition state results in larger image file: d4nj03276d-t63.tif values than those obtained with two molecules of H2O (Table 2, Table S2 (ESI) and Fig. 9B). At the B3LYP/6-31+G(d)/GD3BJ and PBE0/6-31+G(d)/DG3BJ levels, the image file: d4nj03276d-t64.tif values are in very good agreement with the experimental data, contrary to the case of CAM-B3LYP/6-31+G(d) and CAM-B3LYP/6-31+G(d)/GD3BJ. In all cases, the image file: d4nj03276d-t65.tif values are significantly larger than the image file: d4nj03276d-t66.tif values, in very good agreement with the experimental data. Inclusion of four molecules of water in the transition state does not improve the agreement between image file: d4nj03276d-t67.tif and image file: d4nj03276d-t68.tif for both the C-6 and C-8 hydrogens (Table S2, ESI). Interestingly, the entropic term image file: d4nj03276d-t69.tif reduces significantly and for some functionals, it becomes negative which is contrary to the experimental data (Table 1).


image file: d4nj03276d-f9.tif
Fig. 9 The mechanistic pathway of the aromatic C(8)–H and C(6)–H hydrogen exchange process of the complex of the neutral taxifolin with two (A) and three (B) molecules of H2O at the APFD/6-31+G(d) level.
2.2.2. The ionic form of phloroglucinol and taxifolin-effects of two-to-four molecules of water. While for the neutral molecules, the location of the transition state was obtained by guessing the most probable structure, and the anionic form was far more demanding. Thus, we thoroughly scanned the potential energy surface (PES), defining the protons participating in the transfer as reaction coordinates. Upon inspection, saddle points were located and subsequently tested successfully as possible transition states.

In the phloroglucinol anion, as in the case of the neutral form, several possible mechanistic pathways were investigated with two to four molecules of H2O in the transition state (Table 2 and Table S3, ESI). For the complex of the phloroglucinol anion with two molecules of H2O, the reaction mechanism with one bound molecule of H2O on the OH group resulted in very low Boltzmann population and image file: d4nj03276d-t70.tif values (11.83 to 17.83 kcal mol−1) which strongly deviate from the experimental data image file: d4nj03276d-t71.tif. In contrast, the reaction mechanisms (A and B, Fig. 10) with one bound molecule of H2O on the C[double bond, length as m-dash]O moiety resulted in high Boltzmann populations and very good agreement with the experimental data, especially at the B3LYP/6-31+G(d)/GD3BG level image file: d4nj03276d-t72.tif (Fig. 10b, c, Table 2 and Table S2, ESI).


image file: d4nj03276d-f10.tif
Fig. 10 Three possible mechanistic pathways of the aromatic hydrogen exchange process of the phloroglucinol anion with two molecules of H2O: one molecule of H2O on the phenol –OH group (a), and one molecule of H2O on the C[double bond, length as m-dash]O bond (b) and (c), at the B3LYP/6-31+G(d)/D3BJ level.

The computational data with three molecules of H2O in the transition state image file: d4nj03276d-t73.tif strongly deviate from the experimental data (Table S3, ESI). Similarly, the computational data with four molecules of H2O strongly deviate from the experimental data image file: d4nj03276d-t74.tif. In both cases, the enthalpic activation energy image file: d4nj03276d-t75.tif is strongly underestimated (Table S3, ESI).

For the taxifolin anion, the ring C (Fig. 1) was substituted by a methyl group to facilitate computations. In the case of the complex with two molecules of H2O, a transition state could not be determined using a variety of functionals and basis sets (Table 2 and Table S4, ESI). In contrast, with three molecules of H2O in the transition state, the majority of the functionals and basis set used resulted in image file: d4nj03276d-t76.tif and image file: d4nj03276d-t77.tif, in excellent agreement with the experimental values image file: d4nj03276d-t78.tif and image file: d4nj03276d-t79.tif (Table 2 and Table S4, ESI). Fig. 11 presents the potential energy scan (PES) of the ionic form of taxifolin with three catalytically important molecules of H2O. The PES indicated various values of saddle points of the electronic energy function of the two scan coordinates (SC1 and SC2 shown in Fig. 11B) that represent the systematic proton translocation [Eel = f(SC1, SC2)]. From the systematic optimization of these saddle points, we located the transition state described in Fig. 11B. The adopted approach underlines the importance of the systematic search of transition states which are difficult to locate.


image file: d4nj03276d-f11.tif
Fig. 11 Potential energy scan of the ionic form of taxifolin with three catalytically important molecules of H2O.

3. Computational details

Calculations of the neutral and ionic forms of phloroglucinol and taxifolin were performed with a variety of DFT methods (B3LYP, PBE0 (PBE1PBE), APFD, M06-2X and ωB97XD) using the 6-31+G(d) basis set as well as with the B3LYP/6-31+G(d)/GD3BJ and PBE1PBE/6-31+G(d)/GD3BJ level hybrid functionals as implemented in the Gaussian 09W Package.38 Additional computations were performed at the CAM-B3LYP-D/6-31+G(d) level. In all cases, the polarized continuum model IEF-PCM (integral equation formalism-polarizable continuum model) was used in H2O. In order to investigate the transition state, as well as the structure of the products, calculations were carried out at the same level of theory using the gas phase and IEF-PCM model. The computed geometries were verified as minima by frequency calculations at the same level of theory (no imaginary frequencies).

4. Conclusions

In summary, variable temperature and pD 1H NMR studies of the activation enthalpy image file: d4nj03276d-t80.tif, entropy image file: d4nj03276d-t81.tif and Gibbs free energy image file: d4nj03276d-t82.tif of H/D exchange reactions of the H-6 and H-8 protons belonging to ring A of the natural product taxifolin are reported. The image file: d4nj03276d-t83.tif values are in the range of ∼25 to 23 kcal mol−1 for pD values of 6.1 to 9.6 and practically independent of the buffer concentration. Differences in image file: d4nj03276d-t84.tif values of neutral and anionic taxifolin and phloroglucinol were found to be very small (≤1.5 kcal mol−1). Excellent agreement between image file: d4nj03276d-t85.tif and DFT calculated Gibbs free activation energies, image file: d4nj03276d-t86.tif, was obtained, especially at the APFD/6-31+G(d) and B3LYP/6-31+G(d)/GD3BJ levels, with the use of three molecules of H2O in the transition state for the neutral state of phloroglucinol (with the “in–in” configuration of the phenol OH groups) and taxifolin. In the ionic form of phloroglucinol, the mechanistic pathway with two molecules of H2O in the transition state (one of which involves the C[double bond, length as m-dash]O moiety), showed very good agreement with the experimental data. For the anionic form of taxifolin, the mechanistic pathway with three molecules of H2O in the transition state showed excellent agreement with the experimental image file: d4nj03276d-t87.tif values. The enthalpic term image file: d4nj03276d-t88.tif is considerably larger than the entropic image file: d4nj03276d-t89.tif term, which is in excellent agreement with the experimental data. This indicates a dissociative mechanism of the loosely bound activated complex. The present results are important for the following reasons:

(i) The catalytic role of two and/or three molecules of H2O through keto–enol tautomerization, which is a many body cooperative effect of hydrogen-bonded H2O molecules in proton translocation, brought into question the generally accepted acid/base catalyzed H/D exchange mechanism emphasized in undergraduate chemistry textbooks.

(ii) Water catalysis plays a significant role at neutral pH values and near ambient temperatures thus, by definition, is a green approach by avoiding strong acid and basic conditions.

Author contributions

Photini Chalkidou: performed the NMR experiments. Themistoklis Venianakis: performed the NMR experiments. George Papamokos: performed DFT computations, analyzed computational data and prepared the manuscript. Michael Siskos: performed DFT computations, conceived and designed the project, analyzed the experimental and computational data and prepared the manuscript. Ioannis P. Gerothanassis: conceived and designed the project, analyzed the experimental and computational data and prepared the manuscript.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The use of the NMR Core Facility of the University of Ioannina is gratefully acknowledged. This work was supported by Greek Community Support Framework III, Regional Operational, MIS 91629.

Notes and references

  1. M. F. Ruiz-Lopez, J. S. Francisco, M. T. C. Martins-Costa and J. M. Anglada, Nat. Rev. Chem., 2020, 4, 459–475 CrossRef CAS.
  2. T. Kitanosono and S. Kobayashi, Chem. Eur. J., 2020, 26, 9408–9429 CrossRef CAS.
  3. L. Lin, Y. Ge, H. Zhang, M. Wang, D. Xiao and D. Ma, J. Am. Chem. Soc. Au, 2021, 1, 1834–1848 CAS.
  4. M. Cortes-Clerget, J. Yu, J. R. A. Kincaid, P. Walde, F. Gallou and B. H. Lipshutz, Chem. Sci., 2021, 12, 4237–4266 RSC.
  5. H. Gröger, F. Gallou and B. H. Lipshutz, Chem. Rev., 2023, 123, 5262–5296 CrossRef PubMed.
  6. G. Li, Y. Dai, Y. Yang and D. E. Resasco, Chem. Catal., 2021, 1, 959–965 CrossRef CAS.
  7. P. M. Yavorsky and E. Gorin, J. Am. Chem. Soc., 1962, 84, 1071–1072 CrossRef CAS.
  8. P. McGeady and R. A. Croteau, J. Chem. Soc., Chem. Commun., 1993, 9, 774–776 RSC.
  9. P. S. Kiuru and K. Wähälä, Steroid, 2006, 71, 54–60 CrossRef CAS.
  10. J. Atzrodt, V. Derdau, T. Fey and J. Zimmermann, Angew. Chem., Int. Ed., 2007, 46, 7744–7765 CrossRef CAS PubMed.
  11. M. Jordheim, T. Fossen, J. Songstad and O. M. Andersen, J. Agric. Food Chem., 2007, 55, 8261–8268 CrossRef CAS PubMed.
  12. A. J. Kresge and Y. Chiang, J. Am. Chem. Soc., 1961, 83, 2877–2885 CAS.
  13. E. S. Hand and R. M. Horowitz, J. Am. Chem. Soc., 1964, 86, 2084–2085 CAS.
  14. P. Ryberg and O. Matsson, J. Org. Chem., 2002, 67, 811–814 CrossRef CAS PubMed.
  15. S. H. M. Mehr, K. Fukuyama, S. Bishop, F. Lelj and M. J. MacLachlan, J. Org. Chem., 2015, 80, 5144–5150 CrossRef CAS PubMed.
  16. F. Bonaldo, F. Mattivi, D. Catorci, P. Arapitsas and G. Guella, Molecules, 2021, 26, 3544 CrossRef CAS.
  17. A. Fayaz, M. G. Siskos, P. C. Varras, M. I. Choudhary, A. T. Wahab and I. P. Gerothanassis, Phys. Chem. Chem. Phys., 2020, 22, 17401–17411 RSC.
  18. R. Breslow, Acc. Chem. Res., 1991, 24, 159–164 CrossRef CAS.
  19. S. Narayan, J. Muldoon, M. G. Finn, V. V. Fokin, H. C. Kolb and K. B. Sharpless, Angew. Chem., Int. Ed., 2005, 44, 3275–3279 CrossRef CAS.
  20. U. M. Lindström, Chem. Rev., 2002, 102, 2751–2772 CrossRef.
  21. A. Chanda and V. V. Fokin, Chem. Rev., 2009, 109, 725–748 CrossRef CAS PubMed.
  22. Y. Jung and R. A. Marcus, J. Am. Chem. Soc., 2007, 129, 5492–5502 CrossRef CAS PubMed.
  23. T. G. Kim, Y. Kim and D.-J. Jang, J. Phys. Chem. A, 2001, 105, 4328–4332 CrossRef CAS.
  24. S. Otto and J. B. Engberts, Org. Biomol. Chem., 2003, 1, 2809–2820 RSC.
  25. N. A. Isley, R. T. Linstadt, S. M. Kelly, F. Gallou and B. H. Lipshutz, Org. Lett., 2015, 17, 4734–4737 CrossRef CAS.
  26. T. Kitanosono, K. Masuda, P. Xu and S. Kobayashi, Chem. Rev., 2018, 118, 679–746 CrossRef CAS PubMed.
  27. G. Alagona, C. Ghio and P. I. Nagy, Phys. Chem. Chem. Phys., 2010, 12, 10173–10188 RSC.
  28. J. M. Saa and A. Frontera, Chem. Phys. Chem., 2020, 21, 313–320 CrossRef CAS PubMed.
  29. J. M. Herrero-Martínez, M. Sanmartin, M. Rosés, E. Bosch and C. Ràfols, Electrophoresis, 2005, 26, 1886–1895 CrossRef PubMed.
  30. S. Grimme, S. Ehrlish and L. Goerigk, J. Comput. Chem., 2011, 32, 1456–1465 CAS.
  31. Y. Zhao and D. G. Truhlar, J. Phys. Chem., 2006, 110, 5121–5129 CAS.
  32. J.-D. Chai and M. Head-Gordon, Phys. Chem. Chem. Phys., 2008, 10, 6615–6620 RSC.
  33. X. Xuefei, I. M. Alecu and G. Donald, J. Chem. Theory Comput., 2011, 7(6), 1667–1676 Search PubMed.
  34. V. Exarchou, A. N. Troganis, I. P. Gerothanassis, M. Tsimidou and D. Boskou, Tetrahedron, 2002, 58, 7423–7429 CrossRef CAS.
  35. P. Charisiadis, V. Exarchou, A. N. Troganis and I. P. Gerothanassis, Chem. Commun., 2010, 46, 3589–3591 RSC.
  36. S. H. Mari, P. C. Varras, A.-T. Wahab, I. M. Choudhary, M. G. Siskos and I. P. Gerothanassis, Molecules, 2019, 24, 2290 CrossRef.
  37. P. Charisiadis, T. Venianakis, C. D. Papaemmanouil, A. Primikyri, A. G. Tzakos, M. G. Siskos and I. P. Gerothanassis, Magn. Reson. Chem., 2025, 63, 170–179 CrossRef CAS.
  38. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. MontgomeryJr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, B. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, D. J. Gaussian 09, revision B.01, Gaussian, Inc., Wallingford, CT, 2010 Search PubMed.

Footnote

Electronic supplementary information (ESI) available: Tables of computational data. See DOI: https://doi.org/10.1039/d4nj03276d

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025
Click here to see how this site uses Cookies. View our privacy policy here.