Next Article in Journal
Portable Electrochemical System for the Monitoring of Biogenic Amines in Soil as Indicators in Assessment of the Stress in Plant
Previous Article in Journal
1,4-Butane-Sultone Functionalized Graphitic Carbon Nitride as a Highly Efficient Heterogeneous Catalyst for the Synthesis of 2,3-Dihydroquinazolines Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Hydrogen Bond Binding of Water to Two Cholic Acid Residues †

by
María Pilar Vázquez-Tato
1,
Julio A. Seijas
1,*,
Francisco Meijide
2,
Santiago de Frutos
2 and
José Vázquez Tato
2
1
Departamento de Química Orgánica, Facultade de Ciencias, Universidade de Santiago de Compostela, Campus Terra, 27080 Lugo, Spain
2
Departamento de Química Física, Facultade de Ciencias, Universidade de Santiago de Compostela, Campus Terra, 27080 Lugo, Spain
*
Author to whom correspondence should be addressed.
Presented at the 26th International Electronic Conference on Synthetic Organic Chemistry, 15–30 November 2022; Available online: https://sciforum.net/event/ecsoc-26.
Chem. Proc. 2022, 12(1), 95; https://doi.org/10.3390/ecsoc-26-13555
Published: 14 November 2022

Abstract

:
Cholic acid is a trihydroxy bile acid with three hydroxy groups at C-3, C-7 and C-12 carbon atoms; two methyl groups at C-10 and C-13 carbon atoms of the steroid nucleus; and a carboxylic group at C24 of the side alkyl chain. The distance between the oxygen atoms linked to C-7 and C-12 (~4.5 Å) perfectly matches with the edge distance between oxygen atoms in ice. This leads to the design of a cholic acid dimer in which one water molecule is encapsulated between two cholic residues, resembling an ice-like structure. The water molecule participates in four hydrogen bonds, the water simultaneously being acceptor from the O12-H hydroxy groups (two bonds with lengths of 2.177 Å and 2.114 Å) and the donor towards the O-7-H groups (two bonds with lengths of 1.866 Å and 1.920 Å). Regarding this communication, we present the application of the “atoms in molecules” (AIM) theory to the tetrahedral structure. The analysis of the calculated electron density, ρ, is performed using its gradient vector, ∇ρ, and the Laplacian, ∇2ρ. The calculation of the complexation energy used correction of the basis set superposition error (BSSE) and the counterpoise method. As expected, four critical (3,−1) points located in the HO bond paths were identified. All calculated parameters are in concordance with those of similar systems and obey the proposed criteria for hydrogen bonds. The total energy for the interaction is −12.67 kcal/mol and is analysed using proposed energy/electron density equations.

1. Introduction

Bile acids have a bifacial polarity since the hydroxy groups (up to three at C-3, C-7 and C-12 carbon atoms) lie beneath the plane of the steroid nucleus (hydrophilic α-side). In what follows, these oxygen atoms are identified with a superscript (for instance, O7 or O12). Two methyl groups are at the other side of the nucleus at the C-10 and C-13 carbon atoms (hydrophobic β-side). As an example, Figure 1 shows the structure of cholic acid (CA). This explains their amphiphilic behaviour in water-forming aggregates above a critical concentration, commonly referred to as critical micelle concentration (cmc). The biological importance of bile acids has been well described in the literature.
The characterization of the crystal structures of BA and their derivatives using X-ray analysis has been a topic of interest for years [1,2,3,4,5,6,7,8,9]. A common factor of all crystal structures is that the hydroxy groups are always involved in the formation of hydrogen bonds (HB) with another BA monomer, the solvent or both species. Nowadays, interesting and promising applications have emerged, mainly related to the ability of BA to form inclusion compounds in the solid state. This may be used for the resolution of racemates [10].
When analysing BA crystals, for accepting the formation of a hydrogen bond, the geometric criteria (bond lengths and angle) [11] have been used on an exclusively basis. Although enough for most cases, additional criteria may be required in special ones.
With the aim of illustrating the application of the “atoms in molecules” (AIM) theory [12,13] for the analysis of hydrogen bonds in crystal structures of BAs, we have chosen a BA crystal in which a single water molecule is encapsulated between two cholic residues in an ice-like structure [14].

2. Crystal Structure

The crystal structure of the reference system was published previously [14]. Cif files (CCDC 867499) contain the supplementary crystallographic data for the C-suc-C crystal (the acronym provided in that paper). These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 13 November 2022).
It was shown that water accommodates in a hydrophilic region, establishing hydrogen bonds with the hydroxy groups O7-H and O12-H of two steroid residues belonging to two CA dimers. These four oxygen atoms form a distorted tetrahedral structure with a water molecule at its centre (Figure 2) as evidenced by the fact that their centroid is exactly located on the oxygen of water molecule (Ow). The average tetrahedral angle is 110 ± 12°. The edge length values are shown in the figure, the average value being 4.61 ± 0.39 Å, which favourably compares with the one in ice [15]. Table 1 shows the values for the four hydrogen bonds in which water is involved since there are other hydrogen bonds in the crystal involving different groups, but they are not considered here. All the values fulfil the geometric criteria of the existence of a hydrogen bond [11]. In ice, the hydrogen bond distance is 2.76 Å [16]. From the Cambridge Structural Database, Steiner [17] obtained average values of 1.880(2) Å for the water dimer, HO-HOH2, and 2.825(2) Å for HOH2 and OO distances, respectively.

2.1. Computational Details

Given the high number of atoms involved in the two bile acid dimers, to analyse the interaction with the water molecule, we simplified the system by reducing the number of atoms in the bile acid unit whilst keeping the same geometric parameters of the remaining atoms. Thus, A and D rings were suppressed, and their carbon atoms linked to B and C rings were replaced by hydrogen atoms. To keep the original interatomic distances obtained from the X-ray resolution of the complex, no minimization of the energy of the complex was carried out.
Calculations of the complexation energy used correction of the basis set superposition error (BSSE) with the counterpoise method implemented in Gaussian 19 [18]. Laplacian of electronic density and critical points (AIM) were calculated using Multiwfn_3.8_dev software [19].

2.2. Analysis of Electron Density

Electron density, ρ, is the starting point of the AIM theory. Its topology is easily deduced from the gradient vector, ∇ρ, and the Laplacian, ∇2ρ. The electron density is usually visualized by drawing contour lines that connect electron density points with an identical value. Figure 3 and Figure 4 show two examples for the present system. In Figure 3, the plane is defined by the nuclei of O57 and O25 (which are O7), and O1 (Ow) oxygen atoms, while in Figure 4 the plane is defined by O24 and O56 (which are O12) and Ow oxygen atoms. The thin grey lines are defined by infinitesimal gradient vectors, thus describing gradient paths.
When ∇2ρ < 0, the electronic charge is locally concentrated, this being the case in covalent bonds [20]. When ∇2ρ > 0, the electronic charge is locally depleted [13], the interactions are called closed-shell. This occurs in hydrogen bonds (HB) in which the charge concentrations are separately localized in the basin of the neighbouring atoms [21]. This kind of analysis leads to the proposition of criteria for characterizing hydrogen bonds (see below). Figure 5 shows bond critical points of covalent bonds (located, for instance, between two carbon atoms), while those with numbers 17, 43, 50 and 59 (located between hydrogen and oxygen atoms) correspond to hydrogen bond critical points (HBCP) where the gradient ∇ρ vanishes, being (3,−1) saddle points.
In Figure 3, the contour lines of the electronic density around the water oxygen (Ow) basin resembles a Micky Mouse profile. This is a consequence in which the two hydrogen atoms of water (Hw) form covalent bonds with Ow. In other words, Ow behaves as an HB donor. The basins of O7 of the hydroxy O7-H groups clearly have a circle shape. Similarly, Figure 4 shows the contour lines of the hydrogen bonds between water and the two O12-H hydroxy groups. The plane in the figure is defined by these three oxygen atoms. Now, the contour around the Ow shows a basin while the profiles around the O12-H groups resemble those of peanuts. Now, the two O12 are donors and Ow is the acceptor. Furthermore, the bond paths of the four hydrogen bonds link the expected two atoms, the hydrogen and the acceptor. It is evident that the first condition of the criteria for characterizing a hydrogen bond is fulfilled. These criteria have been published by Koch and Popelier [22] and resumed by Popelier [20] as a table.
The second condition refers to the electron density at HBCP, ρb. According to Popelier [20], the ρb values should be in the range 0.002–0.035 au. Table 1 shows the calculated values for the four HB, all of them being within the expected range. These values are about one order of magnitude smaller than those found for a covalent bond (ρb = 0.391 au, for O–H in H2O) [23]. On the other hand, it may be noticed that the values when water is the donor are almost double than when it is the acceptor.
There is a correlation between the O–O length and ρb: the shorter the former, the higher the latter. Relationships between the hydrogen bond length and electron density have been published [24,25]. The values obtained here differ by less than ±0.004 au from those obtained using the equation ρ b = 2.38 × e x p 2.38 × r O H , [25] ( r O H in Å, from crystal data).
A third criterion refers to the Laplacian of the charge density evaluated at the bond critical point, where charge density is a local minimum along the bond path, i.e., ρb is locally depleted with respect to neighbouring points along the bond path. The range values (Table 2) are also within the range of expected values from 0.024 to 0.139 au. ∇2ρb follows the same dependence as ρb with the O–O length.
The water dimer is a system of two water molecules bound by a single hydrogen bond. Because of this, it is often used as the paradigmatic system, being the object of many experimental and theoretical investigations [26,27]. The equilibrium geometry and dissociation energy of the water dimer are well known. The dimer has a “trans-linear” structure, and the O–O distance was first measured from the microwave spectrum by Dyke et al. [28,29,30], the value being rOO = 2.98 ± 0.04 Å. The O–H distances depend on the role of the water molecules, whether they are the donor (rOH = 95.8 pm Å) or the acceptor (rOH = 0.95 Å) [26].
Previous ρb and ∇2ρb values may be compared with those for the water dimer, H–O–HOH2. Bader et al. [23] ascertained that ρb (au) and ∇2ρb are 0.0199 and 0.0624 (data in au), respectively (Table 3). From the four values of ρb obtained for the pseudo C–H2O–C crystal structure (Table 2), an average value of 0.020 au was obtained, which perfectly matches the one for the water dimer. The ∇2ρb value for H–O–HOH2 is closer to those in which the oxygens (O12) of hydroxy groups are donors and Ow is the acceptor. It should be noted that, in these two cases, the rOO lengths are also closer to the one of the H–O–HOH2 dimer. Other published values for the water dimer are shown in Table 3.
Another criterion consists in the estimation of the mutual penetration of the hydrogen (H) and acceptor atom (B, oxygen) upon hydrogen bond formation. In the literature, this criterion is often considered as a necessary and sufficient condition for classifying the intermolecular interaction as hydrogen bonding [36]. It is estimated as Δ r i = r i r i o (i are the atoms involved in the hydrogen bond, B or H), the superscript referring to nonbonded radii, r i o , of these atoms and its absence the bonded radii, ri [22]. The nonbonded radius is the distance from a nucleus to a given electron density contour (usually 0.001 au) in the absence of interaction. This value is used because this yields atomic diameters in good agreement with gas-phase van der Waals radii [22]. The bonded radius is the distance from a nucleus to the bond critical point in question. Table 2 shows the HBCP...O and HBCP...H lengths calculated for the crystal. It may be noticed that the sum of both lengths coincides with the imposed one from crystal. The HBCP...H (or rH) length for the hydrogen bonds with Ow as acceptor are larger (>0.1 Å) than those for Ow being the donor. All of them are considerably smaller than this distance for the water dimer in the gas phase (=1.34 Å). Accepting that r H o = r v d W H = 1.1 Å [37], Δ r H   < 0 in all cases. Similarly, if r v d W O = r O o = 1.58 Å [37], then Δ r O   < 0. Also r v d W O + r v d W H = r O H . These data evidence a mutual penetration of hydrogen and oxygen atoms, a conclusion which may be raised from checking the contour electron density values in Figure 3 and Figure 4.
It should be noted that Isaev defined [36] Δ r i = r i o r i , i.e., Δ r O = r v d W o r O H B C P and Δ r H = r v d W H r H H B C P . In all cases, Δ r H > Δ r O , meaning that the hydrogen atom is more penetrated than the acceptor one.
Remaining criteria proposed by Popelier [20] are not analysed here. They involve the analysis of properties in which the hydrogen atom is the focus of the analysis: increased net charge, energetic destabilization, decrease in dipolar polarization, and decrease in the atomic volume.

2.3. Energy of Hydrogen Bonds

The energy of hydrogen bond interactions lie between covalent bonds and weak van der Waals interactions, although there are no sharp borders between the three types of interactions [11]. Jeffrey [38] defined three categories of hydrogen bonds according to the energy involved: weak 0.1–1, moderate 5–15 and strong 20–60 (data in kcal/mol), but other classifications have been published. For instance, Emamian et al. [33] proposed four ranges for the interactions: very weak (>−2.5), weak to medium (−2.5 to −14), medium (−11.0 to −15) and strong (<15.0)—all data in kcal mol−1.
Ruscic [39], from a new partition function for water, obtained dissociation enthalpy values for the water dimer, the values being 13.220 ± 0.096 kJ mol−1 and 15.454 ± 0.074 kJ mol−1 (3.693 ± 0.018 kcal mol−1) at 0 K and 298.15 K, respectively. The experimental result determined from the thermal conductivity of the vapor was −3.59 ± 0.5 kcal/mol [40], and Rocher-Casterline et al. [41] used state-to-state vibrational predissociation measurements following excitation of the bound OH stretch fundamental of the donor unit of the dimer to obtain an accurate value of 13.2 ± 0.12 kJ/mol for the dissociation energy. This value was compared with the theoretical one of 13.2 ± 0.05 kJ/mol, obtained by Shank et al. [42].
Some other values are shown in Table 3. As the complexing moieties of the water molecule are hydroxy groups, we also recompiled the values of the methanol–water system that was theoretically studied by Moin et al. [43]. In the gas phase, the obtained values were OmethHOw 1.96–2.04 Å and OwHOmeth 1.94–2.02 Å for the H..O distances, while the hydrogen bond energies were in the ranges −4.87/−6.44 kcal/mol (OmethHOw) and −5.06/−7.00 kcal/mol (OwHOmeth), the values depending on the level of the theory.
For a series of hydrogen-bonded complexes between nitrites and hydrogen chloride, Boyd and Choi [44] noticed a correlation between the electron density at the HBCP ρ(rb) and the energy of the hydrogen bond. The energies ranged from 10 kJ/mol (NCCN…HC1) to 38 kJ/mol (LiCNHCl) while the range of ρ(ρ(rb) was 0.01103–0.02391. Many other equations have been proposed, and Rozenberg [45] reviewed the subject. After analysing 24 equations, he obtained the following relationship:
E (kJ·mol−1) = − (6.6 ± 8.0) + (1215 ± 440) ρ
The application of this equation to the present system provides the results shown in Table 4. Once E H B has been calculated for each HB, the total energy of the four hydrogen bonds was obtained by summation, the value being −16.951 kcal mol−1, although its standard deviation was very high. By using different equations, the range of calculated values was from −11.8 to −28.5 kcal/mol. Ganthy et al. [46] carried out an analysis of the hydrogen bond of the water dimer but with the geometry found in ice, and compared it with the one in the gas phase. Let us recall that the nearest neighbour O–O distance in ice is 2.75 Å compared to 2.98 Å in a gas phase dimer. As we previously noticed, these two values are close to the values observed in our C-suc-C crystal (Table 1), depending on whether Ow is the donor (Ow-H...O7, average 2.724 Å) or the acceptor (O12-H... Ow, average 2.935 Å). Depending on the theory, Ganthy et al. obtained values of −3.6 (rOO = 2.98 Å, gas phase length) and −2.5 kcal mol−1 (rOO = 2.75 Å, ice phase length) (Hartree.Fock), and −3.9 kcal mol−1 (rOO = 2.98 Å) DFT theory). Ganthy et al. noticed that from the rOO distance in ice, the calculations indicate a net antibonding contribution to energy from overlap effects.
The value calculated for the C–H2O–C system is −12.67 kcal/mol, although it cannot be exclusively ascribed to hydrogen bond interactions. However, it is the most important contribution to the interactions existing in the crystal. The value is higher than that predicted using the average in Equation (1), which probably is too low, as the value calculated for H–OHOH2 suggests, when compared with the most accepted experimental and theoretical ones.

3. Conclusions

There are two main oxygen–oxygen (rOO) distances when a hydrogen bond is formed between water molecules, the one observed in the gas phase in the formation of a dimer (rOO = 2.98 Å) and the one in ice (rOO = 2.75 Å). Both lengths are observed in the C-suc-C crystal, in which a water molecule is encapsulated by four hydroxy groups belonging to two cholic acid dimers. The shorter one corresponds to hydrogen bonds in which the water oxygen is donor and the larger one when it is acceptor. The application of the AIM theory to a simplified system (C–H2O–C) confirms the existence of saddle critical points (HBCP) in all these four hydrogen bonds. The estimated interaction energy in the formation of the complex (−12.67 kcal/mol) is in adequate agreement with the summation of the energies of each hydrogen bond (EHB) estimated from the electron density (ρb) of HBCP and published equations for EHB–ρb linear relationships.

Author Contributions

Conceptualization, J.A.S. and J.V.T.; methodology, J.A.S. and J.V.T.; formal analysis, M.P.V.-T., F.M., J.A.S. and J.V.T.; investigation, M.P.V.-T., F.M. and S.d.F.; writing—original draft preparation, writing—review and editing, M.P.V.-T., F.M., J.A.S. and J.V.T.; funding acquisition, M.P.V.-T., F.M., J.A.S. and J.V.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Ministerio de Ciencia y Tecnología, Spain (Project MAT201786109P).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meijide, F.; de Frutos, S.; Soto, V.H.; Jover, A.; Seijas, J.A.; Vázquez-Tato, M.P.; Fraga, F.; Vázquez Tato, J. A standard structure for bile acids and derivatives. Crystals 2018, 8, 86. [Google Scholar] [CrossRef]
  2. Giglio, E. Structural aspect of inclusion compounds formed by organic host lattices. In Inclusion Compounds of Deoxycholic Acid; Atwood, J.L., Davies, J.E.D., MacNicol, D.D., Eds.; Academic Press: London, UK, 1984; pp. 207–229. [Google Scholar]
  3. Miyata, M.; Sada, K. Deoxycholic acid and related hosts. In Comprehensive Supramolecular Chemistry; MacNicol, D.D., Bishop, R.T.F., Eds.; Elsevier: Oxford, UK, 1996; Volume 6, pp. 147–176. [Google Scholar]
  4. Campanelli, A.R.; Candeloro de Sanctis, S.; Giglio, E.; Viorel Pavel, N.; Quagliata, C. From crystal to micelle: A new approach to the micellar structure. J. Incl. Phenom. Macrocyclic. Chem. 1989, 7, 391–400. [Google Scholar] [CrossRef]
  5. Sada, K.; Sugahara, M.; Kato, K.; Miyata, M. Controlled Expansion of a Molecular Cavity in a Steroid Host Compound. J. Am. Chem. Soc. 2001, 123, 4386–4392. [Google Scholar] [CrossRef] [PubMed]
  6. Sugahara, M.; Sada, K.; Miyata, M. A robust structural motif in inclusion crystals of nor-bile acids. Chem. Commun. 1999, 3, 293–294. [Google Scholar] [CrossRef]
  7. Sugahara, M.; Sada, K.; Hirose, J.; Miyata, M. Inclusion abilities of bile acids with different side chain length. Mol. Cryst. Liquid Cryst. 2001, 356, 155–162. [Google Scholar] [CrossRef]
  8. Hishikawa, Y.; Aoki, Y.; Sada, K.; Miyata, M. Selective inclusion phenomena in lithocholamide crystal lattices; design of bilayered assemblies through ladder-type hydrogen bonding network. Chem. Lett. 1998, 27, 1289–1290. [Google Scholar] [CrossRef]
  9. Miyata, M.; Tohnai, N.; Hisaki, I. Supramolecular chirality in crystalline assemblies of bile acids and their derivatives; three-axial, tilt, helical, and bundle chirality. Molecules 2007, 12, 1973–2000. [Google Scholar] [CrossRef]
  10. Miragaya, J.; Jover, A.; Fraga, F.; Meijide, F.; Vázquez Tato, J. Enantioresolution and Chameleonic Mimicry of 2-Butanol with an Adamantylacetyl Derivative of Cholic Acid. Crystal Growth Des. 2010, 10, 1124–1129. [Google Scholar] [CrossRef]
  11. Grabowski, S.J. Hydrogen Bond—Definitions, Criteria of Existence and Various Types. In Understanding Hydrogen Bonds: Theoretical and Experimental Views; Theoretical and Computational Chemistry Series; Royal Society of Chemistry: London, UK, 2020; Chapter 1; pp. 1–40. ISBN 978-1-83916-040-0. [Google Scholar] [CrossRef]
  12. Popelier, P.L.A. On the full topology of the Laplacian of the electron densityOn the full topology of the Laplacian of the electron density. Coord. Chem. Rev. 2000, 197, 169–189. [Google Scholar] [CrossRef]
  13. Bader, R.F.W. Atoms in Molecules. Acc. Chem. Res. 1985, 18, 9–15. [Google Scholar] [CrossRef]
  14. Soto, V.H.; Alvarez, M.; Meijide, F.; Trillo, J.V.; Antelo, A.; Jover, A.; Galantini, L.; Vázquez Tato, J. Ice-like encapsulated water by two cholic acid moieties. Steroids 2012, 77, 1228–1232. [Google Scholar] [CrossRef] [PubMed]
  15. Fletcher, N.H. The Chemical Physics of Ice. In The Chemical Physics of Ice; Cambridge University Press: Cambridge, UK, 1970. [Google Scholar]
  16. Bergmann, U.; Di Cicco, A.; Wernet, P.; Principi, E.; Glatzel, P.; Nilsson, A. Nearest-neighbor oxygen distances in liquid water and ice observed by x-ray Raman based extended x-ray absorption fine structure. J. Chem. Phys. 2007, 127, 174504. [Google Scholar] [CrossRef] [PubMed]
  17. Steiner, T. The Hydrogen Bond in the Solid State. Angew. Chem. Int. Ed. 2002, 41, 48–76. [Google Scholar] [CrossRef]
  18. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 RC; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  19. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  20. Popelier, P.L.A. Characterization of a Dihydrogen Bond on the Basis of the Electron Density. J. Phys. Chem. A 1998, 102, 1873–1878. [Google Scholar] [CrossRef]
  21. Carroll, M.T.; Chang, C.; Bader, R.F.W. Prediction of the structures of hydrogen-bonded complexes using the laplacian of the charge density. Mol. Phys. 1988, 63, 387–405. [Google Scholar] [CrossRef]
  22. Koch, U.; Popelier, P.L.A. Characterization of C-H-O Hydrogen Bonds on the Basis of the Charge Density. J. Phys. Chem. A 1995, 99, 9747–9754. [Google Scholar] [CrossRef]
  23. Bader, R.F.W.; Carroll, M.T.; Cheeseman, J.R.; Chang, C. Properties of Atoms in Molecules: Atomic Volumes. J. Am. Chem. Soc. 1987, 109, 7968–7979. [Google Scholar] [CrossRef]
  24. Alkorta, I.; Rozas, I.; Elguero, J. Bond Length–Electron Density Relationships: From Covalent Bonds to Hydrogen Bond Interactions. Struct. Chem. 1998, 9, 243–247. [Google Scholar] [CrossRef]
  25. Tang, T.-H.; Deretey, E.; Jensen, S.J.K.; Csizmadia, I.G. Hydrogen bonds: Relation between lengths and electron densities at bond critical points. Eur. Phys. J. D 2006, 37, 217–222. [Google Scholar] [CrossRef]
  26. Mukhopadhyay, A.; Cole, W.T.; Saykally, R.J. The water dimer I: Experimental characterization. Chem. Phys. Lett. 2015, 633, 13. [Google Scholar] [CrossRef]
  27. Mukhopadhyay, A.; Xantheas, S.S.; Saykally, R.J. The water dimer II: Theoretical investigations. Chem. Phys. Lett. 2018, 700, 163–175. [Google Scholar] [CrossRef]
  28. Dyke, T.R.; Muenter, J.S. Microwave spectrum and structure of hydrogen bonded water dimer. J. Chem. Phys. 1974, 60, 2929. [Google Scholar] [CrossRef]
  29. Dyke, T.R.; Mack Kenneth, M.; Muenter, J.S. The structure of water dimer from molecular beam electric resonance spectroscopy. J. Chem. Phys. 1977, 66, 498. [Google Scholar] [CrossRef]
  30. Odutola, J.A.; Dyke, T.R. Partially deuterated water dimers: Microwave spectra and structure. J. Chem. Phys. 1980, 72, 5062. [Google Scholar] [CrossRef]
  31. Grabowski, S.J. High-Level Ab Initio Calculations of Dihydrogen-Bonded Complexes. J. Phys. Chem. A 2000, 104, 5551–5557. [Google Scholar] [CrossRef]
  32. Grabowski, S.J. Ab Initio Calculations on Conventional and Unconventional Hydrogen Bonds Study of the Hydrogen Bond Strength. J. Phys. Chem. A 2001, 105, 10739–10746. [Google Scholar] [CrossRef]
  33. Emamian, S.; Lu, T.; Kruse, H.; Emamian, H. Exploring Nature and Predicting Strength of Hydrogen Bonds: A Correlation Analysis Between Atoms-in-Molecules Descriptors, Binding Energies, and Energy Components of Symmetry-Adapted Perturbation Theory. J. Comput. Chem. 2019, 40, 2868–2881. [Google Scholar] [CrossRef]
  34. Gu, Y.; Kar, T.; Scheiner, S. Fundamental Properties of the CH···O Interaction:  Is It a True Hydrogen Bond? J. Am. Chem. Soc. 1999, 121, 9411–9422. [Google Scholar] [CrossRef]
  35. Kumar, P.S.V.; Raghavendra, V.; Subramanian, V. Bader’s Theory of Atoms in Molecules (AIM) and its Applications to Chemical Bonding. J. Chem. Sci. 2016, 128, 1527–1536. [Google Scholar] [CrossRef]
  36. Isaev, A.N. Ammonia and phosphine complexes with proton donors. Hydrogen bonding from the backside of the N(P) lone pair. Comp. Theor. Chem. 2018, 1142, 28–38. [Google Scholar] [CrossRef]
  37. Rowland, R.S.; Taylor, R. Intermolecular nonbonded contact distances in organic crystal structures: Comparison with distances expected from van der Waals radii. J. Phys. Chem. 1996, 100, 7384–7391. [Google Scholar] [CrossRef]
  38. Jeffrey, G.J. An Introduction to Hydrogen Bonding; Oxford University Press: New York, NY, USA, 1997; p. 12. [Google Scholar]
  39. Ruscic, B. Active Thermochemical Tables: Water and Water Dimer. J. Phys. Chem. A 2013, 117, 11940–11953. [Google Scholar] [CrossRef] [PubMed]
  40. Feyereisen, M.W.; Feller, D.; Dixon, D.A. Hydrogen Bond Energy of the Water Dimer. J. Phys. Chem. 1996, 100, 2993–2997. [Google Scholar] [CrossRef]
  41. Rocher-Casterline, B.E.; Ch’ng, L.C.; Mollner, A.K.; Reisler, H. Communication: Determination of the Bond Dissociation Energy (Do) of the Water Dimer, (H2O)2, by Velocity Map Imaging. J. Chem. Phys. 2011, 134, 211101. [Google Scholar] [CrossRef] [PubMed]
  42. Shank, A.; Wang, Y.; Kaledin, A.; Braams, B.; Bowman, J. Accurate ab initio and "hybrid" potential energy surfaces, intramolecular vibrational energies, and classical ir spectrum of the water dimer. J. Chem. Phys. 2009, 130, 144314. [Google Scholar] [CrossRef]
  43. Moin, S.T.; Hofer, T.S.; Randolf, B.R.; Rode, B.M. Structure and dynamics of methanol in water: A quantum mechanical charge field molecular dynamics study. J. Comput. Chem. 2011, 32, 886–892. [Google Scholar] [CrossRef]
  44. Boyd, R.J.; Choi, S.C. Hydrogen bonding between nitriles and hydrogen halides and the topological properties of molecular charge distributions. Chem. Phys. Lett. 1986, 129, 62–65. [Google Scholar] [CrossRef]
  45. Rozenberg, M. The hydrogen bond-practice and QTAIM theory. RSC Adv. 2014, 4, 26928–26931. [Google Scholar] [CrossRef]
  46. Ghanty, T.K.; Staroverov, V.N.; Koren, P.R.; Davidson, E.R. Is the Hydrogen Bond in Water Dimer and Ice Covalent? J. Am. Chem. Soc. 2000, 122, 1210–1214. [Google Scholar] [CrossRef]
Figure 1. Structure of bile acids.
Figure 1. Structure of bile acids.
Chemproc 12 00095 g001
Figure 2. Oxygen–oxygen distances (green lines; data in Å) of the tetrahedron formed by the O7-H and O12-H hydroxy atoms of the two steroid residues encapsulating a water molecule located at their centroid [14]. The four hydrogen bonds are indicated with blue lines.
Figure 2. Oxygen–oxygen distances (green lines; data in Å) of the tetrahedron formed by the O7-H and O12-H hydroxy atoms of the two steroid residues encapsulating a water molecule located at their centroid [14]. The four hydrogen bonds are indicated with blue lines.
Chemproc 12 00095 g002
Figure 3. Electron density contour of the pseudo CA–H2O–CA complex (thin black lines). O25 and O57 are O7 oxygen atoms and O1 is the oxygen atom (Ow) of the water molecule. Thin grey lines correspond to the electron density gradient.
Figure 3. Electron density contour of the pseudo CA–H2O–CA complex (thin black lines). O25 and O57 are O7 oxygen atoms and O1 is the oxygen atom (Ow) of the water molecule. Thin grey lines correspond to the electron density gradient.
Chemproc 12 00095 g003
Figure 4. Electron density contour of the pseudo CA–H2O–CA complex (thin black lines). O24 and O56 are O12 oxygen atoms and O1 is the oxygen atom of the water molecule (Ow). Thin grey lines correspond to the gradient of the electron density.
Figure 4. Electron density contour of the pseudo CA–H2O–CA complex (thin black lines). O24 and O56 are O12 oxygen atoms and O1 is the oxygen atom of the water molecule (Ow). Thin grey lines correspond to the gradient of the electron density.
Chemproc 12 00095 g004
Figure 5. Bond critical points (BCP) and hydrogen bond critical points (HBCP) obtained for the model complex formed using two steroid residues and water. HBCPs are identified by numbers 17, 43, 50 and 59.
Figure 5. Bond critical points (BCP) and hydrogen bond critical points (HBCP) obtained for the model complex formed using two steroid residues and water. HBCPs are identified by numbers 17, 43, 50 and 59.
Chemproc 12 00095 g005
Table 1. Hydrogen bonds distances measured for the C-suc-C crystal. Data for Å.
Table 1. Hydrogen bonds distances measured for the C-suc-C crystal. Data for Å.
Oxygen of Water Is DonorOxygen of Water Is Acceptor
Ow-H...O7O12-H...Ow
O–O distance/Å2.7102.7382.9352.936
Table 2. Lengths involved in the formation of hydrogen bonds from the crystal structure and calculated electron density and Laplacian values.
Table 2. Lengths involved in the formation of hydrogen bonds from the crystal structure and calculated electron density and Laplacian values.
Oxygen of Water Is DonorOxygen of Water Is Acceptor
Property at HBCPOw-HO7Ow-HO7O12-HOwO12-H… Ow
CP43CP59CP17CP50
a O–O length/Å crystal2.7102.7382.9352.936
a O…H length/Å crystal1.8661.9202.1772.114
Electron density ρb (au)0.02700.02390.01380.0154
ρb (au) calculated from eq of Ref. [25] (see text)0.03070.02700.01450.0170
Laplacian of the electron density at HBCP, ∇2ρb (au)0.1180.1060.06160.0667
HBCP…O length/Å, r11.2171.2421.3831.355
HBCP…H length/Å, r20.6500.6790.7950.759
r1 + r2 = O…H length/Å1.8671.9212.1782.114
Δ r O = r v d W o r O H B C P 0.3630.3380.1970.225
Δ r H = r v d W H r H H B C P / Å 0.4500.4210.3050.341
a These data are in excellent agreement with data reported by by Steiner [17].
Table 3. Electron density, Laplacian of the electron density, hydrogen acceptor length, and hydrogen bond energy reported by several authors for the water dimer.
Table 3. Electron density, Laplacian of the electron density, hydrogen acceptor length, and hydrogen bond energy reported by several authors for the water dimer.
ρb (au)2ρbrH…YEHB, kcal/molRef.
0.01990.0624 −5.5[23]
1.949−4.45[31]
1.825−10.97[31]
0.0230.0911.950−4.45[32]
0.0170.0752.056−4.25[32]
0.0259−4.93[33]
−4.96[33]
−4.32 ± 0.31[34]
0.02190.03961.949−4.33[35]
Table 4. Hydrogen bond energies calculated using different equations.
Table 4. Hydrogen bond energies calculated using different equations.
HBCP17435059Total 4 HBH–OHOH2
E H B /kcal mol−1 Equation (1) [45]−2.430−6.263−2.895−5.363−16.951−4.201
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vázquez-Tato, M.P.; Seijas, J.A.; Meijide, F.; de Frutos, S.; Tato, J.V. Hydrogen Bond Binding of Water to Two Cholic Acid Residues. Chem. Proc. 2022, 12, 95. https://doi.org/10.3390/ecsoc-26-13555

AMA Style

Vázquez-Tato MP, Seijas JA, Meijide F, de Frutos S, Tato JV. Hydrogen Bond Binding of Water to Two Cholic Acid Residues. Chemistry Proceedings. 2022; 12(1):95. https://doi.org/10.3390/ecsoc-26-13555

Chicago/Turabian Style

Vázquez-Tato, María Pilar, Julio A. Seijas, Francisco Meijide, Santiago de Frutos, and José Vázquez Tato. 2022. "Hydrogen Bond Binding of Water to Two Cholic Acid Residues" Chemistry Proceedings 12, no. 1: 95. https://doi.org/10.3390/ecsoc-26-13555

Article Metrics

Back to TopTop