Next Article in Journal
Overexpression of Receptor Tyrosine Kinase EphB4 Triggers Tumor Growth and Hypoxia in A375 Melanoma Xenografts: Insights from Multitracer Small Animal Imaging Experiments
Previous Article in Journal
Controlled and Efficient Polymerization of Conjugated Polar Alkenes by Lewis Pairs Based on Sterically Hindered Aryloxide-Substituted Alkylaluminum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Behavior of the E–E’ Bonds (E, E’ = S and Se) in Glutathione Disulfide and Derivatives Elucidated by Quantum Chemical Calculations with the Quantum Theory of Atoms-in-Molecules Approach

Faculty of Systems Engineering, Wakayama University, 930 Sakaedani, Wakayama 640-8510, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2018, 23(2), 443; https://doi.org/10.3390/molecules23020443
Submission received: 30 January 2018 / Revised: 10 February 2018 / Accepted: 14 February 2018 / Published: 17 February 2018
(This article belongs to the Section Computational and Theoretical Chemistry)

Abstract

:
The nature of the E–E’ bonds (E, E’ = S and Se) in glutathione disulfide (1) and derivatives 23, respectively, was elucidated by applying quantum theory of atoms-in-molecules (QTAIM) dual functional analysis (QTAIM-DFA), to clarify the basic contribution of E–E’ in the biological redox process, such as the glutathione peroxidase process. Five most stable conformers ae were obtained, after applying the Monte-Carlo method then structural optimizations. In QTAIM-DFA, total electron energy densities Hb(rc) are plotted versus Hb(rc) − Vb(rc)/2 at bond critical points (BCPs), where Vb(rc) are potential energy densities at BCPs. Data from the fully optimized structures correspond to the static nature. Those containing perturbed structures around the fully optimized one in the plot represent the dynamic nature of interactions. The behavior of E–E’ was examined carefully. Whereas E–E’ in 1a3e were all predicted to have the weak covalent nature of the shared shell interactions, two different types of S–S were detected in 1, depending on the conformational properties. Contributions from the intramolecular non-covalent interactions to stabilize the conformers were evaluated. An inverse relationship was observed between the stability of a conformer and the strength of E–E’ in the conformer, of which reason was discussed.

1. Introduction

E–E’ bonds (E, E’ = S and Se) play a crucial role in biological redox processes [1]. High energy levels of HOMO and low energy levels of LUMO of the E–E’ bond must be the driving force for the high reactivity in the redox processes. The HOMO and LUMO of E–E’ would correspond to np(E/E’) and σ* (E–E’), respectively, where np(E/E’) denote the p-type lone pair orbitals of E and/or E’, while σ* (E–E’) corresponds to the σ*-orbital of E–E’. Glutathione disulfide (GSSG: 1) has been widely used as a redox reagent in vitro. To facilitate the protein folding process, a mixture of 1 and glutathione (GSH) is often confirmed as the optimum condition if concentrations similar to those observed in vivo [2] are employed [3,4,5,6]. The reduced form of ribonuclease A will undergo disulfide-coupled folding and gain in structural stability in the presence of 1, for example [7]. The detoxification of hydroperoxides in the glutathione peroxidase (GPx) process must be one of most important biological redox processes [8,9,10,11,12,13]. Scheme 1 summarizes a catalytic mechanism proposed for the antioxidant activity of GPx, which is a typical example of the intervention of E–E’ (E, E’ = S, Se) in biological reactions. According to this mechanism, two equivalents of GSH are oxidized to the corresponding oxidized disulfide in the overall process, while the hydroperoxide is reduced to water [14,15].
The behavior of the S–S, S–Se and Se–Se bonds should be clarified, bearing in mind the role of these bonds in the antioxidant mechanism. It is highly important to elucidate the behavior of the S–S bond in 1 together with S–Se and Se–Se in two derivatives of 1 (compounds 2 and 3, respectively). Scheme 2 illustrates the structures of 13. There are many possibilities for the formation of intramolecular hydrogen bonds (HBs) in 13, although the intermolecular HBs of the solute-solute and solute-solvent interactions must also be important in the real system. HBs in 13 must be considered in assessing the basic properties of 13 based on the calculated results, rather than performing calculations for a single molecule in vacuum. Scheme 2 also shows the structures of R-cystine and its derivatives 46 and MeEE’Me (compounds 79).
The structures of 1 and 4, determined by X-ray crystallographic analysis, have been reported, although 4 is in the di-protonated form. Figure 1 shows these structures. The structure of 1 was observed as a half-extended form close to C2 symmetry with the formation of zwitterions [16]. The structure of 4 was reported as an extended form [17]. The extended form in the observed structure of 4 may be the result of the electrostatic repulsion of the positive charges developed on 42+. Many conformers must exist in such compounds, primarily due to the intramolecular HBs.
Reactions of 1 and/or 4 in vivo proceed under conditions with very large and highly complex species. However, the essence of the elementary processes is expected to be close to that of the typical chemical reactions. Therefore, it would be instructive to start with less complex species to clarify the behavior of the E–E’ bonds (E, E’ = S and Se). We reported the dynamic and static behavior of S–S in R-cystine (4) and S–Se and Se–Se in the derivatives of 4 (5 and 6, respectively), together with MeEE’Me 79 as references [18]. It is challenging to clarify the nature of the E–E’ bonds (E, E’ = S and Se) in glutathione disulfide and its derivatives (13), although the structures of 13 are considerably more complex relative to 46, respectively. Structures 13 will have much more plausible HBs than 46.
The QTAIM approach, introduced by Bader [19,20,21,22,23,24,25,26,27,28], enables us to analyze the nature of chemical bonds and interactions [11,12,13,14,15,16]. A bond critical point (BCP, ∗) is an important concept in QTAIM. BCP is a point where ρ(r) (charge density) reaches a minimum along the interatomic (bond) path, while it is a maximum on the interatomic surface separating the atomic basins. ρ(r) at BCP is denoted by ρb(rc) and other QTAIM functions are denoted in a similar way. Interactions seem to be defined by the corresponding bond paths (BPs), but we must be careful to use the correct terminology with the concept. Interactions would be easily imaged by means of QTAIM if they can be defined as the corresponding BPs, especially for experimental chemists. However, it is demonstrated that the detection of the BPs between two atoms in a molecule emerging from natural alignment of the gradient vector held of the one-electron density of a molecule is neither necessary nor a sufficient condition for the presence of a chemical bond between those atoms [29,30,31,32,33,34]. In this connection, it is pointed out that the terms line paths (LPs) and line critical points (LCPs) should be used in place of BPs and BCPs, respectively [30]. Consequently, the dynamic and static nature in this work should be regarded as the investigation performed at LCPs on LPs corresponding to the E–E’ interactions. Nevertheless, the interactions expected for E–E’ are clearly detected by BPs with BCPs, which is another reason to use BPs and BCPs in this work. The structures of species can be described by molecular graphs, which are the sets of attractors (atoms) and BPs, together with BCPs, ring critical points (RCPs) and cage critical points (CCPs). We recently proposed QTAIM–DFA [35,36,37,38,39], as a tool for experimental chemists to analyze their own results concerning chemical bonds and interactions using their own images. QTAIM-DFA provides an excellent possibility to evaluate, classify, characterize and understand weak to strong interactions in a unified manner [40,41,42,43,44,45]. Hb(rc) are plotted versus Hb(rc) − Vb(rc)/2 in QTAIM-DFA, where Hb(rc) and Vb(rc) are the total electron energy densities and potential energy densities, respectively, at BCPs. The QTAIM-DFA treatment can incorporate the classification of interactions based on the signs of Hb(rc) and ∇2ρb(rc) (Laplacian ρ), as shown in Scheme S1 of the Supplementary Materials, since the signs of Hb(rc) − Vb(rc)/2 must be equal to those of ∇2ρb(rc), where (ћ2/8m) ∇2ρb(rc) = Hb(rc) − Vb(rc)/2 (= Gb(rc) + Vb(rc)/2 while Hb(rc) = Gb(rc) + Vb(rc), Gb(rc): kinetic energy densities at BCPs) (see Equations (S1), (S2) and (S2’) of the Supplementary Materials). In our treatment, data for the perturbed structures around fully optimized structures are employed for the plots in addition to data for the fully optimized structures [35,36,37,38,39]. Data from the fully optimized structures are analyzed by the polar coordinate (R, θ) representation, which corresponds to the static nature of interactions. Data from the perturbed structures and a fully optimized structure are used to construct a curve. Each curve is analyzed in terms of the (θp, κp) parameters: θp corresponds to the tangent line of the plot and κp is the curvature. We proposed the concept of the “dynamic nature of interactions” based on (θp, κp) [35,36,37,38,39]. QTAIM-DFA is applied to typical chemical bonds and interactions. Rough criteria have been established that distinguish the chemical bonds and interactions in question from others. QTAIM-DFA and the criteria are explained in the Supplementary Materials, employing Schemes S1 and S2, Figure S1 and Equations (S1)–(S7). The basic concept of the QTAIM approach is also described in the Supplementary Materials.
The behavior of E–E’ (E, E’ = S and Se) in 13 is expected to be related to that in the glutathione peroxidase (GPx) process. The dynamic and static nature of E–E’ in 13 is elucidated by applying QTAIM-DFA. The same method is applied to E–E’ in 46 and 79 to reexamine the nature of these bonds. We present the results of the theoretical elucidation of the nature of the E–E’ bonds in 16 with QTAIM-DFA to better understand the role of E–E’ in the antioxidant activity of GPx. Quantum chemical (QC) calculations are also applied to examine the structural features of 16. The E–E’ bonds in 16 are classified and characterized by employing the criteria and the behavior of the bonds in 79 as references.

2. Methodological Details in Calculations

The structures were optimized employing the Gaussian 09 programs [46] unless otherwise noted. For each species 16, five conformers were optimized with the 6-311+G(3d) basis sets for S and Se, and with the 6-311++G(d, p) basis sets for O, N, C and H [47,48,49,50]. The basis set system is called BSS-A in this paper. The DFT level of M06-2X [51] is applied to the calculations. Before the final optimizations, the full conformer search with the Monte-Carlo method in Spartan 02 [52] was applied to each of 16. At least six thousand and five hundred conformers were generated for each of 13 with the Merck Molecular Force Field (MMFF) method [53]. The most stable thirty independent conformers from the Monte-Carlo method were optimized using the 3–21G basis sets at the B3LYP level [54,55] for each of 13. Next, the most stable fifteen conformers were optimized with M06-2X/6-31G(d) for each conformer, as predicted with B3LYP/3-21G of the Gaussian09 program. The most stable five conformers from M06-2X/6-31G(d) were further optimized with BSS-A at the M06-2X level (M06-2X/BSS-A). The final five optimized conformers were confirmed by the frequency analysis for each of 13. These five conformers are called a, b, c, d and e, where conformer a is the most stable among the five, followed by b, then c, then d and then e. In the case of 46, seven hundred and twenty conformers were generated for each species with the PM3 method [56]. Similar to the case for 13, the five most stable conformers (ae) were determined for each of 46. The structures of 7 and 9 are optimized retaining the C2 symmetry, while that of 8 is retaining the C1 symmetry. The population analysis has also been performed by the natural bond orbital method [57] at M06-2X/BSS-A level of theory using natural bond orbital (NBO) program [58].
QTAIM functions were calculated using the Gaussian 09 program package at the same level of DFT theory (M06-2X/BSS-A), and the data were analyzed with the AIM2000 program [19,59]. Normal coordinates of internal vibrations (NIV) obtained by the frequency analysis were employed to generate the perturbed structures [38,39]. This method is called NIV and explained in Equation (1). The k-th perturbed structure in question (Skw) was generated by the addition of the normal coordinates of the k-th internal vibration (Nk) to the standard orientation of a fully optimized structure (So) in the matrix representation. The coefficient fkw in Equation (1) controls the difference in structures between Skw and So: fkw is determined to satisfy Equation (2) for an interaction in question, where r and ro show the interaction distances in question in the perturbed and fully optimized structures, respectively, with ao representing the Bohr radius (0.52918 Å) [12,13]. The perturbed structures with NIV correspond to those where r has been elongated or shortened by 0.05ao or 0.1ao, relative to ro, in the fully optimized structures [60]. The selected motion must be most effectively localized on the interaction in question among the zero-point internal vibrations. Nk of five digits are used to predict Skw:
S k w = S o + f k w N k
r = r o + w a o   ( w = ( 0 ) , ± 0.05   and   ± 0.1 ;   a o = 0.52918   Å )
y = c o + c 1 x + c 2 x 2 + c 3 x 3   ( R c 2 :   square   of   correlation   coefficient )
Hb(rc) are plotted versus Hb(rc) − Vb(rc)/2 for data from the five points of w = 0, ±0.05 and ±0.1 in Equation (2) in QTAIM-DFA. Each plot is analyzed using a regression curve of the cubic function as shown in Equation (3), where (x, y) = (Hb(rc) − Vb(rc)/2, Hb(rc)) (Rc2 (square of correlation coefficient) > 0.99999, usually) [61].

3. Results and Discussion

3.1. Optimized Structures for Conformers of 16 with M06-2X/BSS-A, Together with 79

The five conformers (ae) for each of 16 are optimized with M06-2X/BSS-A, which are called 1a1e, 2a2e, 3a3e, 4a4e, 5a5e and 6a6e, respectively. The whole set of species is also described by 1a6e, if necessary. Each conformer is optimized as a non-extended form. The total energies evaluated for nx (n = 16; x = ae) (E(nx)) are defined to satisfy Equation (4). The relative energies for the conformers of 1a1e [Erel(1x: x = ae)] are evaluated from 1a with M06-2X/BSS-A, so are Erel(nx: n = 26; x = ae). The structures of zwitterions are confirmed between the amino and carboxyl groups at the terminal positions of the main chains in 1a3e, except for 1b and 1e. Only one conformer was optimized for each of 79 with M06-2X/BSS-A, as expected. Table 1 collects the structural parameters of the r(E, E’) distances and the torsional angles of φ(CEE’C) (= φA) for 1a6a and 79, optimized with M06-2X/BSS-A, together with the relative energies Erel (= E(nx) – E(na) (n = 16; x = ae)):
E ( n a ) E ( n b ) E ( n c ) E ( n d ) E ( n e )   ( n = 1 6 )
The r(S, S) values for 1a1e are predicted to be larger than those of 7. The differences in r(S, S) for 1a1er(S, S: 1x) = r(S, S: 1x) − r(S, S: 7), where x = ae) are 0.02 Å < Δr(S, S: 1x) < 0.03 Å for 1a1c, Δr(S, S: 1x) < 0.01 Å for 1d and Δr(S, S: 1x) ≈ 0.20 Å for 1e. Similarly, the Δr(E, E’) values are less than or very close to 0.01 Å for nane (n = 26), except for Δr(E, E’) ≈ 0.015 Å for 2e, 4a, 5d and 5e with Δr(E, E’) ≈ 0.03 Å for 3d. There must be a specific reason for the unexpectedly large value of 0.20 Å for Δr(E, E’: 1e). Three S, S and O atoms align linearly in 1e, which is explained by assuming the formation of hypervalent interactions of S2O σ(3c–4e) of the np(O)→σ*(S–S) type (see Figure 4). In this interaction, σ*(S–S) accepts electrons from np(O). As a result, the S–S bond must be unexpectedly elongated relative to the usual length, and the O---S distance will be substantially shortened relative to the sum of the vdW radii. The O---S distance is predicted to be 2.7714 Å, which is shorter than the sum of the vdW radii by 0.55 Å. The S2O σ(3c–4e) model explains the predicted result for 1e, reasonably well.
In the case of φ(CEE’C) (= φA), the values for 1a6e from the corresponding values of 79 are given by ΔφA(E, E’: nx) = φA(E, E’: nx) − φA(E, E’: MeEE’Me), where n = 16; x = ae; and E, E’ = S and Se. The absolute values of φA will be used to estimate ΔφA. The magnitudes of the values are ΔφA(E, E’: nx) ≈ 58° for 3d, 31° < ΔφA(E, E’: nx) < 35° for 1a1c and 1e, ΔφA(E, E’: nx) ≈ 25° for 2b, and 10° ≤ ΔφA(E, E’: nx) ≤ 20° for 1d, 3e, 4e and 6d. The magnitudes of ΔφA(E, E’: nx) are less than 10° (−10° ≤ ΔφA(E, E’: nx) ≤ 10°) for others, except for ΔφA(E, E’: nx) ≈ –12° for 5d and 6d and ΔφA(E, E’: nx) ≈ −20° for 2e, 4a and 5e. The results must be the reflection from the easy deformation in φA. To evaluate the energy for the deformation in φA, 79 were optimized assuming φA = 0° and 180°, in addition to the fully optimized structures (85° ≤ φA ≤ 86°). They were optimized to be 7 (C2v), 8 (Cs) and 9 (C2v) at φA = 0° and 7 (C2h), 8 (Cs) and 9 (C2h) at φA = 180°. In the case of 9, the structures were further optimized with φA fixed every 15° for 0° ≤ φA ≤ 180°. The results are summarized in Table S1 of the Supplementary Materials. Figure 2 shows the plot of the energies for the optimized structures versus φA. The energy seems less than 15 kJ mol−1 for 45° ≤ φA ≤ 135° in 9. The energy for the deformation of φA in 7 and 8 seems comparable to that in 9. The very easy deformation in φA is well demonstrated, exemplified by 79, which supports the results shown in Table 1. Such easy deformation in φA is also reported for some dichalcogenides [62].

3.2. Structural Feature of 1a6a and 79

Figure 3 illustrates the molecular graphs of 1a3a, drawn on the optimized structures, together with the optimized structures containing the non-covalent interactions. Figure 4 shows the molecular graphs of 1b1e, drawn on the optimized structures. Molecular graphs of 2b2e and 3b3e are drawn in Figures S2 and S3 of the Supplementary Materials, respectively. Figure 5 illustrates the molecular graphs of 4a, 5a, 6a and 4b4e, drawn on the optimized structures and the optimized structure. Molecular graphs of 5b5e, 6b6e and 79 are drawn in Figures S4–S6 of the Supplementary Materials, respectively.
The structural features of 79 are described, first. Only classical chemical bonds are detected in the molecular graphs of 79, as shown in Figure S6 of the Supplementary Materials. Namely, no interactions other than the classical chemical bonds contribute to the interactions in 79. The structural features of 46 are examined next. Various types of intramolecular non-covalent interactions are detected in 4a6e, which are the HB type of O–H---O, O–H---N, N–H---O, N–H---N, O–H---E(E’) and N–H---E(E’), where E, E’ = S and Se. The E---π type of C = O---E(E’) and O = C---E(E’) are also detected. The conformers must be stabilized through the energy lowering effect by the formation of the intramolecular attractive interactions. The HB and E---π type non-covalent interactions contribute to stabilize the conformers. The HB and E---π type interactions in the molecular graphs are drawn on the optimized structures with different colors for the different interaction types to aid in visualization. The interactions are drawn for O–H---O in red, O–H---N and N–H---O in pink, N–H---N in blue, O–H---E(E’), N–H---E(E’) and O = C---E(E’) in olive and C = O---E(E’) in grey (see, Figure 5). The numbers of the intramolecular non-covalent interactions are counted separately by the interaction types, which are differentiated by the colors. The results are collected in Table S2 of the Supplementary Materials. Indeed, the stability of the conformers is expected to relate to the numbers, but they must be stabilized by the total energy of the intramolecular non-covalent interactions. The C–H---X interactions are also detected, however, they are neglected, since they would not make a significant contribution to stabilizing the conformers.
The molecular graphs for 1a3e are very complex, with many intramolecular non-covalent interactions. An effort is made to classify the interactions in 1a3e, as in 4a6e. The non-covalent interactions appearing in the molecular graphs of 1a3e are similarly drawn on the optimized structures, as shown in Figure 3. The numbers of the intramolecular non-covalent interactions in 1a3e are also counted separately based on the type of interaction. The results are collected in Table S2 of the Supplementary Materials. The numbers seem to be correlated to the stability of the conformers. However, it must be difficult to estimate numerically the stability of the conformers based on the numbers. The stability must be controlled by the total energy of the intramolecular interactions.
Nevertheless, it is very important to understand how Erel for the conformers are determined by the intramolecular non-covalent interactions in the conformers of 1a3e, as a whole. How can Erel be evaluated based on the contributions from the non-covalent interactions? We searched for a method to evaluate the stability of the conformers based on the overall intramolecular non-covalent interactions. Then, we devised a method to evaluate the contributions, which is discussed next.

3.3. Factors Determining the Relative Energies of 1a6a

The proposed method is explained in Scheme 3 with Equations (5)–(7). The evaluation process is as follows: (i) GEE’G is fully optimized; (ii) E and E’ in the optimized GEE’G are replaced by H and H; (iii) The structural parameters for the two replaced H atoms are (partially) optimized, with other atoms fixed at the fully optimized geometry; (iv) The structural parameters of CEE’C are fixed at the fully optimized positions, and H atoms are added on each side of CEE’C in place of the organic ligands to give H3CEE’CH3; (v) The structural parameters of the six H atoms are optimized:
E ( GEE G ) opt = 2 E ( G fix C fix H opt ) + E [ ( H opt ) 3 C fix E fix E fix C fix ( H opt ) 3 ] 2 E ( CH 4 ) opt + α   ( α :   almost   constant )
E rel ( GEE G ) opt = 2 E rel ( G fix H opt ) + E rel [ ( H opt ) 3 C fix E fix E fix C fix ( H opt ) 3 ] + α rel
E rel ( GEE G ) opt 2 E rel ( G fix H opt ) + E rel [ ( H opt ) 3 C fix E fix E fix C fix ( H opt ) 3 ]
The method shown in Scheme 3 will evaluate the intramolecular G---G interactions and the deformation energies around CEE’C but not the steric factor around the E–E’ moiety. Equation (5) shows the relationship between E(GEE’G) for the fully optimized structure and the energies evaluated by the proposed method, where α shows the errors in energy between E(GEE’G) and the components, which contains the steric factor around the E–E’ moiety. The relationship for Erel(GEE’G) is shown in Equation (6), where E(CH4) disappears. As shown in Equation (7), Erel(GEE’G)opt can be approximated as 2Erel(Gfix–Hopt) + Erel[(Hopt)3CfixEfix–E’fixCfix(Hopt)3] if α is almost constant. The Erel values are given as the values from the most stable conformers in 1a6a, if applied to 1a6e, respectively. The results of the calculations for 1a6e are collected in Table S3 of the Supplementary Materials, where 2Erel(Gfix–Hopt) and Erel[(Hopt)3CfixEfix–E’fixCfix(Hopt)3] are abbreviated as Erel(2GH)p-opt and Erel(MeSSMe)p-opt, respectively (p-opt: partially optimizations).
Figure 6 shows the plot of Erel(2GH + MeEE’Me)p-opt for 1a3e, together with Erel(GEE’G)opt. The Erel(2GH + MeEE’Me)p-opt values seem to match the Erel(GEE’G)opt values for 1a3e, except for 2b and 2d. Indeed, the relative stabilities of the conformers for the S–S and Se–Se species are explained well by the treatment, but they are not explained well by treatment for the S–Se species, especially for 2b and moderately for 2d. Other factors, such as the steric factor around the S–Se moiety, could be important in this case. The very large magnitude of φA in 2b (110.1°) relative to other species (65.0–85.6°) is responsible for this result. The deviation in 2b, due to Erel(2GH + MeSSeMe)p-opt (−18.4 kJ mol−1) versus Erel(GSSeG)opt (1.0 kJ mol−1), is due to the high stability of Erel(2GH)p-opt (–22.7 kJ mol−1), which is due to the reflection of the C–H optimizations in 2GH from the unstable position of C–H by φA for 2b (110.1°). The smaller magnitudes in Erel for 2a2e may lead to greater deviations (see Table S3 of the Supplementary Materials). A similar plot for 4a6e is shown in Figure S7 of the Supplementary Materials. The relationship between Erel(2CysH + MeEE’Me)p-opt and Erel(CysEE’Cys)opt seems unclear for 4a6e. After clarification of the structural features of 1a6e, the contour plots and negative Laplacians are examined next.

3.4. Contour Plots and Negative Laplacian around the E-∗-E’ Bonds in 1a6e

Figure 7 shows the contour plots of ρ(r), exemplified by 1a1e, which are drawn on an SSC plane of 1ae. The plots show that each BCP on E-∗-E’ exists at the three-dimensional saddle point of ρ(r). Figure 8 illustrates the Negative Laplacian, exemplified by 1a1e, similarly drawn on an SSC plane. All BCP on E-∗-E’ of 1a1e exist in the red area of the plots, which means that the BCPs are all in the range of ∇2ρb(rc) < 0. Therefore, E-∗-E’ of 1a1e are classified by SS (shared shell) interactions. (See also Figure S8 of the Supplementary Materials for the trajectory plots for 1a1e.)

3.5. Application of QTAIM-DFA to the E–E’ Bonds in 1a6e

QTAIM functions are calculated for 1a6e and 79. Table 2 collects the results for 1a3e and 79. Figure 9 shows the plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 1a1d and 7, where the data for 1e do not appear in the plotted area. Figure 10 displays the plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 2a3e, 8 and 9. Figure 9 and Figure 10 also contain magnified representations of the data around the fully optimized structures.
All data in Table 2 and the perturbed structures of 1a3e and 79 are plotted in Figure S9 of the Supplementary Materials. All data for the fully optimized structures of 1a6e and 79 appear in the range of Hb(rc) < 0 and Hb(rc) − Vb(rc)/2 < 0. Therefore, the E–E’ interactions of 1a6e and 79 are all classified by SS (shared shell) interactions, irrespective of the substantial elongation of the S-∗-S bond length by the perturbation that occurred in the conformers, such as S–S in 1e. The plots are analyzed according to Equations (S3)–(S6) of the Supplementary Materials, by applying QTAIM-DFA.
The QTAIM-DFA parameters of (R, θ) and (θp, κp) are also collected in Table 2, together with the frequencies (ν) and force constants (kf) corresponding to the E-∗-E′ bonds in question. While the data for 4a4e are plotted in Figure S10, those for 5a6e are in Figure S11 of the Supplementary Materials, together with those for 79. Similarly, the plots are analyzed to give the QTAIM-DFA parameters of (R, θ) and (θp, κp). The parameters are collected in Table S4 of the Supplementary Materials, together with the frequencies (νn) and force constants (kf) corresponding to the E-∗-E′ bonds in question.

3.6. Nature of the E–E’ Bonds in 1a6e

The E-∗-E′ bonds in 1a6e are classified and characterized based on R, θ and θp values, employing those of the standard interactions given in Scheme S2 of the Supplementary Materials as a reference. Before discussing the nature of the E-∗-E’ bonds in 1a6e and 79, it would be instructive to survey the related criteria. Interactions will be classified as SS or CS interactions for θ > 180° or θ < 180°, respectively, which correspond to Hb(rc) − Vb(rc)/2 < 0 and Hb(rc) − Vb(rc)/2 > 0, respectively. The θp values play an important role in characterizing the interactions. For SS interactions with θ > 180°, θp > 190° is tentatively assigned, where θp = 190° corresponds to θ = 180° for typical interactions. The covalent interactions will be sub-divided depending on the values of R. The (classical) covalent interactions will be called strong (Covstrong) if R > 0.15 au; therefore, they should be called weak (Covweak) when R < 0.15 au.
The R value of 0.076 au (<0.15 au) is predicted for MeS-∗-SMe (7), and those for S-∗-S, S-∗-Se and Se-∗-Se in 1a6e, 8 and 9, examined in this work, are all less than 0.076 au. Therefore, the Covstrong interactions are not detected in this work. As shown in Table 2, the (θ, θp) values for S-∗-S in 1a1e are (188.8–189.6°, 197.2–197.6°) for 1a1d with (181.9°, 193.8°) for 1e. The value for 1e is apparently smaller than those for 1a1d due to the elongation of S-∗-S by the formation of S2O σ(3c–4e) in 1e. This means that S-∗-S in 1e should be (much) weaker than those in 1a1d. Nevertheless, the S-∗-S interaction in 1e is classified by the SS interaction and characterized as Covweak in nature (SS/Covweak). All S-∗-S interactions in 1a1d are, of course, predicted to have the nature of (SS/Covweak). The (θ, θp) values for S-∗-Se in 2a2e are (184.7–184.9°, 188.1–188.7°). Therefore, the S-∗-Se interactions are also classified by the SS interactions and characterized as Covweak in nature (SS/Covweak), although the θp values are slightly less than 190°. In the case of Se-∗-Se in 3a3e, the (θ, θp) values are (185.9–186.4°, 189.0–189.8°). The Se-∗-Se interactions are predicted to have the nature of (SS/Covweak), similar to the cases of 1a2e. Note that S-∗-S in 1e is predicted to be weaker than S-∗-Se in 2a2e and Se-∗-Se in 3a3e by R and θ, although the trend is reversed for θp. Indeed, Se-∗-Se in 3a3e is predicted to be stronger than S-∗-Se in 2a2e by θ and θp, and the inverse trend is true for R. The E-∗-E′ bonds in 4a6e are all predicted to have the nature of (SS/Covweak), as are the bonds in 79, similar to the case for 1a3e.

3.7. Factors that Stabilize the E–E’ Bonds and the Conformers

The S–S bonds of 1a1e are predicted to be less stable than that of 7. The S-∗-S bond in 1a1e is predicted to be weaker in the order shown in Equation (8), where 1e is significantly destabilized due to the elongation by the formation of S2O σ(3c–4e). The order for the strength of S-∗-S seems to exhibit almost the reverse trend of stability compared with the conformers. A similar order is predicted for S-∗-Se of 2a2e with 8. Equation (9) shows the order for S-∗-Se, where 2a seems substantially destabilized. On the other hand, the trend is not so clear for Se-∗-Se in 3a3e. The predicted order for Se-∗-Se is given in Equation (10). The strength of E-∗-E’ seems to show a trend with almost inverse stability compared with the conformers containing E-∗-E’, as mentioned above. While the trend seems rather clear for 1a1e with 7 and 2a2e with 8, the trend seems unclear for 3a3e with 9. This trend would be clear if the data were plotted in a narrow range, whereas it would not be clear if they were plotted in a wider range, although the mechanism is not clear:
S -∗- S   in   7 > 1 d > 1 b > 1 a 1 c > > 1 e
S -∗- Se   in   8 > 2 b > 2 c 2 d > 2 a > 2 e
Se -∗- Se   in   3 c > 9 > 3 b > 3 a > 3 e > 3 d
S -∗- S   in   7 > 4 b > 4 c > 4 d > 4 e   > >   4 a
S -∗- Se   in   8 5 b > 5 a 5 c 5 d > 5 e
Se -∗- Se   in   9 > 6 e > 6 c > 6 b > 6 d > 6 a
In the case of S–S in 4a4e with 7, the S-∗-S bond becomes less stable in the order shown in Equation (11). The order for the strength of S-∗-S is almost the reverse of the stability of the conformers, with the species are divided into four groups of 7, 4b4d, 4e and 4a. The order for the strength of Se-∗-Se is also almost inverse, with the species divided into four groups of 9, 6b6d, 6e and 6a, as shown in Equation (13). However, the trend in S-∗-Se is not as clear for 5a5e with 8, as predicted in Equation (12). As shown in Figures S10 and S11 of the Supplementary Materials, the data for 4a4e with 7 are plotted in a narrow range, as are those for 6a6e with 9, which seems to exhibit a clear trend. However, the data for 5a5e with 8 are plotted over a wider range, and the trend seems unclear, similar to the cases of 1a3e, although the mechanism remains unclear.
The trends shown in Equations (8)–(13) are also confirmed through the analysis of ρb(rc) and bond orders evaluated based on the natural atomic orbitals, for E–E′ in 1a6e and 79. The plot of ρb(rc) versus the bond orders is shown in Figure S12 of the Supplementary Materials. See also Table S5 of the Supplementary Materials for bond orders of 1a6e and 79.
While the results could be explained in a variety of ways, our explanation is as follows: the intramolecular attractive interactions in 1a6c stabilize the species, but E–E’ would be destabilized through the distortion. This is because the E–E’ bonds operate to relax the excess deformation brought about by intramolecular attractive interactions such as HBs. This destabilization would increase the stability of the species. As a result, the E–E’ bonds will be predicted to be less stable if they exist in more stable species. The E–E’ bonds could be predicted to be rather stable, if the intramolecular attractive interactions do not affect the structures around the E–E’ bonds.
The nature of the E–E’ bonds in 19 is well-described by applying QTAIM-DFA based on the dynamic nature with (θp, κp) and the static nature with (R, θ).

4. Conclusions

The dynamic and static nature of E–E’ (E, E’ = S and Se) in glutathione disulfide (1) and derivatives 23, respectively, is elucidated by applying QTAIM-DFA together with R-cystine and its derivatives (46) and MeEE’Me (79). Five conformers ae for each of 16 are optimized with M06-2X/BSS-A. The conformers are called 1a1e, which are defined to satisfy E(1a) < E(1b) < E(1c) < E(1d) < E(1e), for example. No intramolecular non-covalent interactions are detected in 79, while many such interactions operate to stabilize 1a3e. Among such interactions, the formation of S2O σ(3c–4e) of the np(O)→σ*(S–S) type detected in 1e elongates r(S, S) by 0.20 Å. The contribution from the intramolecular non-covalent interactions to the stability of the conformers of GEE’G is estimated by separately calculating the G---G part as 2G–H and the E–E’ part (as MeE–E’Me), under suitable conditions. The Erel(2GH + MeEE’Me) values explain the Erel values well for 1a1e and 3a3e but not so for 2a2e.
QTAIM-DFA is applied to E-∗-E’ in 1a6e and 79 by plotting Hb(rc) versus Hb(rc) − Vb(rc)/2 for the data from the fully optimized structures and the perturbed structures at BCPs. The QTAIM-DFA parameters of (R, θ) and (θp, κp) are obtained by analyzing the plots. The (θ, θp, R) values for S-∗-S in 1a1e are (188.8–189.6°, 197.2–197.6°, 0.0673–0.0744 au) for 1a1d with (181.9°, 193.8°, 0.0343 au) for 1e. The values for 1e are apparently smaller than those for 1a1d, due to the elongation of S-∗-S by the formation of S2O σ(3c–4e) in 1e. Nevertheless, the S-∗-S interactions in 1a1e are all predicted to have the (SS/Covweak) nature. Similarly, the E–E’ interactions in 2a6e and 79 are all predicted to have the (SS/Covweak) nature. The S–S bonds of 1a1e are predicted to be less stable than those of 7, and the S-∗-S bond in 1a becomes weaker than those in 1b, 1d and 7, although 1a is the most stable among 1a1e. This inverse trend between the stability of the conformers and the strength of E-∗-E’ is widely observed. The intramolecular non-covalent attractive interactions in 1a6e stabilize the species but destabilize the E–E’ bonds through distortion, where the E–E’ bonds act to relax the excess deformation caused by the attractive interactions. These predictions of bond behavior help to understand the reactivity of E–E’ in chemical and biological processes.

Supplementary Materials

The following are available online at https://www.mdpi.com/1420-3049/23/2/443/s1, Scheme S1: QTAIM-DFA: Plot of Hb(rc) versus Hb(rc) − Vb(rc)/2 for weak to strong interactions, Scheme S2: Rough classification of interactions by θ and θp, together with kb(rc) (= Vb(rc)/Gb(rc)), Equations (S1)–(S7), Table S1: The energies for 79 of the optimized structures and partially optimized structures with φA, fixed suitably, with M06-2X/BSS-A, Table S2: Number of the O–H---O (Int A), O–H---N, N–H---O (Int B), N–H---N (Int C), HBs with the E(E’)---H–O(N) and E(E’)---C = O (Int D) and E(E’)---O = C and E(E’)---NH–C = O (Int E) interactions in 16, evaluated with M06-2X/BSS-A, Table S3: The relative energies (Erel) of REE’R, (R–H + R–H) and MeE–E’Me for 4a6e, evaluated with M06-2X/BSS-A, Table S4: QTAIM-DFA parameters and QTAIM functions at BCPs for the E–E’ bonds in 4a6e and 79, a together with the frequencies (ν) and force constants (kf), corresponding to the E-∗-E′ bonds in question, Table S5: NAO bond orders for 1a6e and 78, Figure S1: Polar (R, θ) coordinate representation of Hb(rc) versus Hb(rc) − Vb(rc)/2, with (θp, κp) parameters, Figure S2: Molecular graphs of 2b2e, drawn on the optimized structures, Figure S3: Molecular graphs of 3b3e, drawn on the optimized structures, Figure S4: Molecular graphs of 5b5e, drawn on the optimized structures, Figure S5: Molecular graphs of 6b6e, drawn on the optimized structures, Figure S6: Molecular graphs of 79, drawn on the optimized structures, Figure S7: Plots of Erel of REE’R and (2R–H + MeEE’Me) for 4ae (CysSSCys), 5ae (CysSSeCys) and 6ae (CysSeSeCys), evaluated with M06-2X/BSS-A, Figure S8: Trajectory plots of ρb(rc) drawn on the S–S–C planes of 1a1e, similarly to the case of Figure 6 in the text. Color and marks are same as those in Figure 6, Figure S9: Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 1a3e and 79, Figure S10: Plots of Hb(rc) versus Hb(rc) – Vb(rc)/2 for 4a4e and 7, Figure S11: Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 5a6e and 78, Figure S12: Plots of ρb(rc) versus NAO bond orders for 1a6e and 78 and Cartesian coordinates and energies of all the species involved in the present work.

Acknowledgments

This work was partially supported by a Grant-in-Aid for Scientific Research (No. 26410050 and 17K05785) from the Ministry of Education, Culture, Sports, Science and Technology of Japan.

Author Contributions

S.H. and W.N. formulated the project. Y.T. optimized all compounds 1a6e after conformer search, applying the Monte-Carlo method. S.H., Y.T. and W.N. calculated the QTAIM functions and evaluated the QTAIM-DFA parameters and analyzed the data. S.H. and W.N. wrote the paper, while Y.T. organized the data to assist the writing.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Nakanishi, W.; Hayashi, S.; Morinaka, S.; Sasamori, T.; Tokitoh, N. Extended hypervalent E’···E–E···E’ 4c–6e (E, E’ = Se, S) interactions: Structure, stability and reactivity of 1-(8-PhE’C10H6)EE(C10H6E’Ph-8’)-1’. New J. Chem. 2008, 32, 1881–1889. [Google Scholar] [CrossRef]
  2. Hwang, C.; Sinskey, A.J.; Lodish, H.F. Oxidized redox state of glutathione in the endoplasmic reticulum. Science 1992, 257, 1496–1502. [Google Scholar] [CrossRef] [PubMed]
  3. Wetlaufer, D.B.; Branca, P.A.; Chen, G.-X. The oxidative folding of proteins by disulfide plus thiol does not correlate with redox potential. Protein Eng. 1987, 1, 141–146. [Google Scholar] [CrossRef] [PubMed]
  4. Gilbert, H.F. Thiol/disulfide exchange equilibria and disulfidebond stability. Methods Enzymol. 1995, 251, 8–28. [Google Scholar] [PubMed]
  5. Lyles, M.M.; Gilbert, H.F. Catalysis of the oxidative folding of ribonuclease A by protein disulfide isomerase: Dependence of the rate on the composition of the redox buffer. Biochemistry 1991, 30, 613–619. [Google Scholar] [CrossRef] [PubMed]
  6. Beld, J.; Kenneth, W.J.; Hilvert, D. Selenoglutathione: Efficient Oxidative Protein Folding by a Diselenide. Biochemistry 2007, 46, 5382–5390. [Google Scholar] [CrossRef] [PubMed]
  7. Konishi, Y.; Ooi, T.; Scheraga, H.A. Regeneration of ribonuclease A from the reduced protein. 1. Rate-limiting steps. Biochemistry 1982, 21, 4734–4740. [Google Scholar] [CrossRef] [PubMed]
  8. Hendrickson, W.A. Determination of macromolecular structures from anomalous diffraction of synchrotron radiation. Science 1991, 254, 51–58. [Google Scholar] [CrossRef] [PubMed]
  9. See references cited in [18].
  10. Kumakura, F.; Mishra, B.; Priyadarsini, K.I.; Iwaoka, M. A Water-Soluble Cyclic Selenide with Enhanced Glutathione Peroxidase-Like Catalytic Activities. Eur. J. Org. Chem. 2010, 440–445. [Google Scholar] [CrossRef]
  11. Arai, K.; Dedachi, K.; Iwaoka, M. Rapid and Quantitative Disulfide Bond Formation for a Polypeptide Chain Using a Cyclic Selenoxide Reagent in an Aqueous Medium. Chem. Eur. J. 2011, 17, 481–485. [Google Scholar] [CrossRef] [PubMed]
  12. Yoshida, S.; Kumakura, F.; Komatsu, I.; Arai, K.; Onuma, Y.; Hojo, H.; Singh, B.G.; Priyadarsini, K.I.; Iwaoka, M. Antioxidative Glutathione Peroxidase Activity of Selenoglutathione. Angew. Chem. Int. Ed. 2011, 50, 2125–2128. [Google Scholar] [CrossRef] [PubMed]
  13. Manna, D.; Mugesh, G. Regioselective Deiodination of Thyroxine by Iodothyronine Deiodinase Mimics: An Unusual Mechanistic Pathway Involving Cooperative Chalcogen and Halogen Bonding. J. Am. Chem. Soc. 2012, 134, 4269–4279. [Google Scholar] [CrossRef] [PubMed]
  14. Flohe, L.; Günzler, W.A.; Schock, H.H. Glutathione peroxidase: A selenoenzyme. FEBS Lett. 1973, 32, 132–134. [Google Scholar] [CrossRef]
  15. Flohe, L. Glutathione Peroxidase Brought into Focus. Free Radic. Biol. 1982, 5, 223–254. [Google Scholar]
  16. Jelsch, C.; Didierjean, C. The oxidized form of glutathione. Acta Cryst. C 1999, 55, 1538–1540, CSD code: BERLOZ. [Google Scholar] [CrossRef]
  17. Leela, S.; Ramamurthi, K. Private Communication, 2007. CSD code: CYSTCL03.
  18. Tsubomoto, Y.; Hayashi, S.; Nakanishi, W. Dynamic and static behavior of the E–E’ bonds (E, E’ = S and Se) in cystine and derivatives, elucidated by AIM dual functional analysis. RSC Adv. 2015, 5, 11534–11540. [Google Scholar] [CrossRef]
  19. Bader, R.F.W. Atoms in Molecules. A Quantum Theory; Oxford University Press: Oxford, UK, 1990. [Google Scholar]
  20. Matta, C.F.; Boyd, R.J. An Introduction to the Quantum Theory of Atoms in Molecules in the Quantum Theory of Atoms in Molecules: From Solid State to DNA and Drug Design; WILEY-VCH: Weinheim, Germany, 2007. [Google Scholar]
  21. Biegler-König, F.; Schönbohm, J. Update of the AIM2000-Program for atoms in molecules. J. Comput. Chem. 2002, 23, 1489–1494. [Google Scholar] [CrossRef] [PubMed]
  22. Biegler-König, F.; Schönbohm, J.; Bayles, D. AIM2000. J. Comput. Chem. 2001, 22, 545–559. [Google Scholar]
  23. Bader, R.F.W. A Bond Path: A Universal Indicator of Bonded Interactions. J. Phys. Chem. A 1998, 102, 7314–7323. [Google Scholar] [CrossRef]
  24. Bader, R.F.W. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–926. [Google Scholar] [CrossRef]
  25. Bader, R.F.W. Atoms in molecules. Acc. Chem. Res. 1985, 18, 9–15. [Google Scholar] [CrossRef]
  26. Tang, T.H.; Bader, R.F.W.; MacDougall, P. Structure and bonding in sulfur-nitrogen compounds. Inorg. Chem. 1985, 24, 2047–2053. [Google Scholar] [CrossRef]
  27. Bader, R.F.W.; Slee, T.S.; Cremer, D.; Kraka, E. Description of conjugation and hyperconjugation in terms of electron distributions. J. Am. Chem. Soc. 1983, 105, 5061–5068. [Google Scholar] [CrossRef]
  28. Biegler-König, F.; Bader, R.F.W.; Tang, T.H. Calculation of the average properties of atoms in molecules. J. Comput. Chem. 1982, 3, 317–328. [Google Scholar] [CrossRef]
  29. Bader, R.F.W. Bond Paths Are Not Chemical Bonds. J. Phys. Chem. A 2009, 113, 10391–10396. [Google Scholar] [CrossRef] [PubMed]
  30. Foroutan-Nejad, C.; Shahbazian, S.; Marek, R. Toward a Consistent Interpretation of the QTAIM: Tortuous Link between Chemical Bonds, Interactions, and Bond/Line Paths. Chem. Eur. J. 2014, 20, 10140–10152. [Google Scholar] [CrossRef] [PubMed]
  31. Garcıa-Revilla, M.; Francisco, E.; Popelier, P.L.A.; Pendás, A.M. Domain-Averaged Exchange-Correlation Energies as a Physical Underpinning for Chemical Graphs. ChemPhysChem 2013, 14, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  32. Keyvani, Z.A.; Shahbazian, S.; Zahedi, M. To What Extent are “Atoms in Molecules” Structures of Hydrocarbons Reproducible from the Promolecule Electron Densities? Chem. Eur. J. 2016, 22, 5003–5009. [Google Scholar] [CrossRef] [PubMed]
  33. Poater, J.; Sol, M.; Bickelhaupt, F.M. Hydrogen–Hydrogen Bonding in Planar Biphenyl, Predicted by Atoms-In-Molecules Theory, Does Not Exist. Chem. Eur. J. 2006, 12, 2889–2895. [Google Scholar] [CrossRef] [PubMed]
  34. Shahbazian, S. Why Bond Critical Points Are Not “Bond” Critical Points. Chem. Eur. J. 2018. [Google Scholar] [CrossRef] [PubMed]
  35. Nakanishi, W.; Hayashi, S.; Narahara, K. Polar Coordinate Representation of Hb(rc) versus (ћ2/8m)∇2ρb(rc) at BCP in AIM Analysis: Classification and Evaluation of Weak to Strong Interactions. J. Phys. Chem. A 2009, 113, 10050–10057. [Google Scholar] [CrossRef] [PubMed]
  36. Nakanishi, W.; Hayashi, S. Atoms-in-Molecules Dual Parameter Analysis of Weak to Strong Interactions: Behaviors of Electronic Energy Densities versus Laplacian of Electron Densities at Bond Critical Points. J. Phys. Chem. A 2008, 112, 13593–13599. [Google Scholar] [CrossRef] [PubMed]
  37. Nakanishi, W.; Hayashi, S. Atoms-in-Molecules Dual Functional Analysis of Weak to Strong Interactions. Curr. Org. Chem. 2010, 14, 181–197. [Google Scholar] [CrossRef]
  38. Nakanishi, W.; Hayashi, S. Dynamic Behaviors of Interactions: Application of Normal Coordinates of Internal Vibrations to AIM Dual Functional Analysis. J. Phys. Chem. A 2010, 114, 7423–7430. [Google Scholar] [CrossRef] [PubMed]
  39. Nakanishi, W.; Hayashi, S.; Matsuiwa, K.; Kitamoto, M. Applications of Normal Coordinates of Internal Vibrations to Generate Perturbed Structures: Dynamic Behavior of Weak to Strong Interactions Elucidated by Atoms-in-Molecules Dual Functional Analysis. Bull. Chem. Soc. Jpn. 2012, 85, 1293–1305. [Google Scholar] [CrossRef]
  40. QTAIM-DFA is successfully applied to analyze weak to strong interactions in gas phase. It could also be applied to the interactions in crystals and to those in larger systems, containing bioactive materials, the methodological improvement is inevitable to generate the perturbed structures suitable for the systems.
  41. Hayashi, S.; Matsuiwa, K.; Kitamoto, M.; Nakanishi, W. Dynamic Behavior of Hydrogen Bonds from Pure Closed Shell to Shared Shell Interaction Regions Elucidated by AIM Dual Functional Analysis. J. Phys. Chem. A 2013, 117, 1804–1816. [Google Scholar] [CrossRef] [PubMed]
  42. Sugibayashi, Y.; Hayashi, S.; Nakanishi, W. Dynamic and static behavior of hydrogen bonds of the X–H···π type (X = F, Cl, Br, I, RO and RR’N; R, R’ = H or Me) in the benzene π-system, elucidated by QTAIM dual functional analysis. Phys. Chem. Chem. Phys. 2015, 17, 28879–28891. [Google Scholar] [CrossRef] [PubMed]
  43. Hayashi, S.; Sugibayashi, Y.; Nakanishi, W. Dynamic and static behavior of the H···π and E···π interactions in EH2 adducts of benzene π-system (E = O, S, Se and Te), elucidated by QTAIM dual functional analysis. Phys. Chem. Chem. Phys. 2016, 18, 9948–9960. [Google Scholar] [CrossRef] [PubMed]
  44. Hayashi, S.; Matsuiwa, K.; Nishizawa, N.; Nakanishi, W. Transannular E···E’ Interactions in Neutral, Radical Cationic, and Dicationic Forms of cyclo-[E(CH2CH2CH2)2E’] (E, E’ = S, Se, Te, and O) with Structural Feature: Dynamic and Static Behavior of E···E’ Elucidated by QTAIM Dual Functional Analysis. J. Org. Chem. 2015, 80, 11963–11976. [Google Scholar] [CrossRef] [PubMed]
  45. Sugibayashi, Y.; Hayashi, S.; Nakanishi, W. Behavior of Halogen Bonds of the Y–X···π Type (X, Y = F, Cl, Br, I) in the Benzene π System, Elucidated by Using a Quantum Theory of Atoms in Molecules Dual-Functional Analysis. Chem. Phys. Chem. 2016, 17, 2579–2589. [Google Scholar] [CrossRef] [PubMed]
  46. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09 (Revision D.01); Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  47. Binning, R.C.; Curtiss, L.A. Compact contracted basis sets for third-row atoms: Ga–Kr. J. Comput. Chem. 1990, 11, 1206–1216. [Google Scholar] [CrossRef]
  48. Curtiss, L.A.; McGrath, M.P.; Blaudeau, J.-P.; Davis, N.E.; Binning, R.C., Jr.; Radom, L. Extension of Gaussian-2 theory to molecules containing third-row atoms Ga–Kr. J. Chem. Phys. 1995, 103, 6104–6113. [Google Scholar] [CrossRef]
  49. McGrath, M.P.; Radom, L. Extension of Gaussian-1 (G1) theory to bromine-containing molecules. J. Chem. Phys. 1991, 94, 511–516. [Google Scholar] [CrossRef]
  50. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Schleyer, P.V.R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  51. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  52. Spartan ’02 Windows, Tutorial and User’s Guide; Wavefunction, Inc.: Irvine, CA, USA, 2001.
  53. Halgren, T.A. Merck molecular force field. I. Basis, form, scope, parameterization, and performance of MMFF94. J. Comput. Chem. 1996, 17, 490–519. [Google Scholar] [CrossRef]
  54. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  55. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  56. Stewart, J.J.P. Optimization of parameters for semiempirical methods I. Method. J. Comput. Chem. 1989, 10, 209–220. [Google Scholar] [CrossRef]
  57. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  58. Glendening, E.D.; Landis, C.R.; Weinhold, F. NBO 6.0: Natural bond orbital analysis program. J. Compt. Chem. 2013, 34, 1429–1437. [Google Scholar] [CrossRef] [PubMed]
  59. The AIM2000 program (Version 2.0) is employed to analyze and visualize atoms-in-molecules: Biegler-König, F. Calculation of atomic integration data. J. Comput. Chem. 2000, 21, 1040–1048. [Google Scholar]see also refs 19 and 20
  60. The values of w = (0), ±0.1, and ±0.2 in r = ro + wao were employed for the perturbed structures in [33,34], which covers roughly the observed range of the interaction distances in molecular complexes. The bond orders become 2/3 and 3/2 times larger at w = +0.2 and –0.2, respectively, relative to the original values at w = 0. However, it seems better to employ the perturbed structures closer to the fully optimized one, which will reduce the calculation errors from the perturbed structures. Therefore, w = (0), ±0.05, and ±0.1 have been used for r = ro + wao in the usual cases of the CS interactions. The shorter range of w must be more suitable for the SS interactions, although w = (0), ±0.05, and ±0.1 are employed in this paper.
  61. Nakanishi, W.; Hayashi, S. Role of dG/dw and dV/dw in AIM Analysis: An Approach to the Nature of Weak to Strong Interactions. J. Phys. Chem. A 2013, 117, 1795–1803. [Google Scholar] [CrossRef] [PubMed]
  62. Nakanishi, W.; Hayashi, S.; Hashimoto, M.; Arca, M.; Aragoni, M.C.; Lippolis, V. Recent advances of structural chemistry of organoselenium and organotellurium compounds. In The Chemistry of Organic Selenium and Tellurium Compounds; Rappoport, Z., Ed.; Wiley: New York, NY, USA, 2013; Volume 4, Chapter 11; pp. 885–972. [Google Scholar]
Scheme 1. Catalytic mechanism, proposed for the antioxidant activity of GPx.
Scheme 1. Catalytic mechanism, proposed for the antioxidant activity of GPx.
Molecules 23 00443 sch001
Scheme 2. Structures of glutathione disulfide (1) and its derivatives (2 and 3) and R-cystine (4) and its derivatives (5 and 6), together with MeEE’Me (79).
Scheme 2. Structures of glutathione disulfide (1) and its derivatives (2 and 3) and R-cystine (4) and its derivatives (5 and 6), together with MeEE’Me (79).
Molecules 23 00443 sch002
Figure 1. Structures of 1 (E = E’ = S) (a) and di-protonated form of 4 (E = E’ = S) (b), determined by X-ray analysis.
Figure 1. Structures of 1 (E = E’ = S) (a) and di-protonated form of 4 (E = E’ = S) (b), determined by X-ray analysis.
Molecules 23 00443 g001
Figure 2. Plots of deformation energy (ΔE) versus φA. For 79 (a) and for 9 (b).
Figure 2. Plots of deformation energy (ΔE) versus φA. For 79 (a) and for 9 (b).
Molecules 23 00443 g002
Figure 3. Molecular graphs of 1a, 2a and 3a, drawn on the optimized structures (top) and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures. The red and blue circles show the zwitter ionic –NH3+ and –COO moieties, respectively (bottom). The energies of 1a, 2a and 3a are employed as the standards for 1a1e, 2a2e and 3a3e, respectively.
Figure 3. Molecular graphs of 1a, 2a and 3a, drawn on the optimized structures (top) and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures. The red and blue circles show the zwitter ionic –NH3+ and –COO moieties, respectively (bottom). The energies of 1a, 2a and 3a are employed as the standards for 1a1e, 2a2e and 3a3e, respectively.
Molecules 23 00443 g003
Figure 4. Molecular graphs of 1b1e, drawn on the optimized structures and the molecular orbital of HOMO-3 for 1e with the orbital interaction map explaining the HOMO-3.
Figure 4. Molecular graphs of 1b1e, drawn on the optimized structures and the molecular orbital of HOMO-3 for 1e with the orbital interaction map explaining the HOMO-3.
Molecules 23 00443 g004
Figure 5. Molecular graphs of 4a6a, drawn on the optimized structures and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures (top two). Molecular graphs of 4b4e, drawn on the optimized structures and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures (bottom two).
Figure 5. Molecular graphs of 4a6a, drawn on the optimized structures and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures (top two). Molecular graphs of 4b4e, drawn on the optimized structures and the intramolecular non-covalent interactions, corresponding to BPs in the molecular graphs, drawn on the optimized structures (bottom two).
Molecules 23 00443 g005
Scheme 3. Proposed method to evaluate the contributions from the G---G intramolecular non-covalent interactions in GEE’G. The processes illustrated in the scheme are explained in the text.
Scheme 3. Proposed method to evaluate the contributions from the G---G intramolecular non-covalent interactions in GEE’G. The processes illustrated in the scheme are explained in the text.
Molecules 23 00443 sch003
Figure 6. Plots of Erel of GEE’G and (2G − H + MeEE’Me) for 1ae (GSSG), 2ae (GSSeG) and 3ae (GSeSeG), evaluated with M06-2X/BSS-A.
Figure 6. Plots of Erel of GEE’G and (2G − H + MeEE’Me) for 1ae (GSSG), 2ae (GSSeG) and 3ae (GSeSeG), evaluated with M06-2X/BSS-A.
Molecules 23 00443 g006
Figure 7. Contour plots of ρb(rc) drawn on the SSC planes of 1a1e, together with BCPs (red solid dots on the plane and pink solid dots out of the plane), RCPs (ring critical points: deep green solid squares on the plane and green solid squares out of the plane), CCPs (cage critical points: blue solid dots on the plane and cyan solid dots out of the plane) and bond paths (black solid lines). The contours (eao−3) are at 2l (l = ±8, ±7, …., 0) with the heavy line of 0.0047 for the molecular surface.
Figure 7. Contour plots of ρb(rc) drawn on the SSC planes of 1a1e, together with BCPs (red solid dots on the plane and pink solid dots out of the plane), RCPs (ring critical points: deep green solid squares on the plane and green solid squares out of the plane), CCPs (cage critical points: blue solid dots on the plane and cyan solid dots out of the plane) and bond paths (black solid lines). The contours (eao−3) are at 2l (l = ±8, ±7, …., 0) with the heavy line of 0.0047 for the molecular surface.
Molecules 23 00443 g007
Figure 8. Negative Laplacians drawn on the SSC planes of 1a1e. The negative areas are shown in red and the positive areas are in blue.
Figure 8. Negative Laplacians drawn on the SSC planes of 1a1e. The negative areas are shown in red and the positive areas are in blue.
Molecules 23 00443 g008
Figure 9. Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 1a1e and 7. (a) Whole plot and (b) magnified view of the data around the fully optimized structures.
Figure 9. Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 1a1e and 7. (a) Whole plot and (b) magnified view of the data around the fully optimized structures.
Molecules 23 00443 g009
Figure 10. Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 2a3e, 8 and 9. (a) Whole plot; (b) magnified view of the data around the fully optimized structures of 3a3e and 9; and (c) magnified view of the data around the fully optimized structures of 2a2e and 8.
Figure 10. Plots of Hb(rc) versus Hb(rc) − Vb(rc)/2 for 2a3e, 8 and 9. (a) Whole plot; (b) magnified view of the data around the fully optimized structures of 3a3e and 9; and (c) magnified view of the data around the fully optimized structures of 2a2e and 8.
Molecules 23 00443 g010
Table 1. Optimized r(E, E’) distances, torsional angles φ(CEE’C) (= φA) and Erel values for 1a6e and 79, evaluated with M06-2X/BSS-A 1.
Table 1. Optimized r(E, E’) distances, torsional angles φ(CEE’C) (= φA) and Erel values for 1a6e and 79, evaluated with M06-2X/BSS-A 1.
Speciesr(E, E)φAErelSpeciesr(E, E)φAErel
(Å)(°)(kJ mol−1)(Å)(°)(kJ mol−1)
1a2.0736−117.40.02a2.2002−85.60.0
1b2.0694−116.48.62b2.1963−110.11.0
1c2.0778−119.314.12c2.1982−84.518.0
1d2.0561100.329.32d2.1959−78.423.1
1e2.2454117.997.42e2.2079−65.023.7
4a2.062567.70.05a2.1984−83.90.0
4b2.0471−82.20.35b2.189084.315.7
4c2.052988.50.75c2.201194.017.5
4d2.0541−75.73.25d2.207072.919.6
4e2.051595.78.85e2.2067−66.927.4
72.049185.0282.192385.62
3a2.3252−85.20.0
3b2.3215−82.513.6
3c2.3138−92.534.9
3d2.3546−144.547.9
3e2.3320105.258.8
6a2.327588.50.0
6b2.330393.41.4
6c2.330990.23.3
6d2.335174.93.6
6e2.328693.13.7
92.323686.12
1 BSS-A: The 6-311+G(3d) basis sets for S and Se with the 6-311++G(d,p) basis sets for O, N, C and H. 2 Not applicable.
Table 2. QTAIM-DFA Parameters and QTAIM Functions at BCPs for the E-∗-E′ bonds in 1a3e and 79, 1 together with the frequencies (ν) and force constants (kf) corresponding to the E-∗-E′ species in question.
Table 2. QTAIM-DFA Parameters and QTAIM Functions at BCPs for the E-∗-E′ bonds in 1a3e and 79, 1 together with the frequencies (ν) and force constants (kf) corresponding to the E-∗-E′ species in question.
Compoundρ(rc)c2ρb(rc) 2Hb(rc)R 3θ 4kb(rc) 5νn 6kf 7θp:NIV 8κp:NIV 9Classification/
(Sym: E-∗-E’)(au)(au)(au)(au)(°)(cm−1)(mdyn Å−1)(°)(au−1)Characterization
1a (C1: S-∗-S)0.1378−0.0106−0.0676 0.0684 188.9 −2.460493.30.928197.30.82SS/Covweak
1b (C1: S-∗-S)0.1391−0.0113−0.06920.0701189.2−2.483513.81.718197.30.78SS/Covweak
1c (C1: S-∗-S)0.1368−0.0103−0.06650.0673188.8−2.451489.70.910197.20.84SS/Covweak
1d (C1: S-∗-S)0.1428−0.0124−0.07330.0744189.6−2.512506.00.917197.60.70SS/Covweak
1e (C1: S-∗-S)0.1025−0.0011−0.03450.0345181.9−2.070353.70.447193.83.62SS/Covweak
2a (C1: S-∗-Se)0.1169−0.0043−0.05280.0530184.7−2.195418.70.545188.10.23SS/Covweak
2b (C1: S-∗-Se)0.1178−0.0046−0.05350.0537184.9−2.206434.00.955188.40.15SS/Covweak
2c (C1: S-∗-Se)0.1174−0.0045−0.05320.0534184.8−2.203421.50.634188.30.19SS/Covweak
2d (C1: S-∗-Se)0.1176−0.0045−0.05320.0534184.8−2.203404.60.684188.70.21SS/Covweak
2e (C1: S-∗-Se)0.1158−0.0043−0.05200.0522184.7−2.198415.90.835188.20.20SS/Covweak
3a (C1: Se-∗-Se)0.1027−0.0046−0.04370.0440186.0−2.265302.00.346189.00.83SS/Covweak
3b (C1: Se-∗-Se)0.1035−0.0048−0.04440.0446186.2−2.275310.20.485189.20.78SS/Covweak
3c (C1: Se-∗-Se)0.1048−0.0050−0.04580.0461186.3−2.282316.00.423189.20.84SS/Covweak
3d (C1: Se-∗-Se)0.0988−0.0046−0.04060.0409186.4−2.291304.20.684189.81.31SS/Covweak
3e (C1: Se-∗-Se)0.1022−0.0045−0.04350.0437185.9−2.259300.60.835189.40.99SS/Covweak
7 (C2: S-∗-S)0.1446−0.0131−0.07510.0763189.9−2.535513.72.645197.60.66SS/Covweak
8 (C1: S-∗-Se)0.1189−0.0048−0.05440.0547 185.0−2.213419.72.072188.60.38SS/Covweak
9 (C2: Se-∗-Se)0.1036−0.0050−0.04450.0448186.4−2.291307.72.730189.10.77SS/Covweak
1 6-311+G(3d) basis sets employed for S and Se with the 6-311++G(d,p) basis sets for O, N, C and H at the DFT level of M06-2X. Frequencies and force constants related to the NIV to generate the perturbed structures are also listed. 2 c2ρb(rc) = Hb(rc) − Vb(rc)/2 where c = ħ2/8m. 3 R = [(Hb(rc) − Vb(rc)/2)2 + Hb(rc)2]1/2. 4 θ = 90° − tan−1[Hb(rc)/(Hb(rc) − Vb(rc)/2)]. 5 kb(rc) = Vb(rc)/Gb(rc). 6 Frequency corresponding to the stretching mode of the E-∗-E’ bond, where ∗ means the bond critical point in question. 7 Force constants correspond to νn. 8 θp = 90° − tan−1(dy/dx), where (x, y) = (Hb(rc) − Vb(rc)/2, Hb(rc)). 9 κp = |d2y/dx2|/[1 + (dy/dx)2]3/2.

Share and Cite

MDPI and ACS Style

Hayashi, S.; Tsubomoto, Y.; Nakanishi, W. Behavior of the E–E’ Bonds (E, E’ = S and Se) in Glutathione Disulfide and Derivatives Elucidated by Quantum Chemical Calculations with the Quantum Theory of Atoms-in-Molecules Approach. Molecules 2018, 23, 443. https://doi.org/10.3390/molecules23020443

AMA Style

Hayashi S, Tsubomoto Y, Nakanishi W. Behavior of the E–E’ Bonds (E, E’ = S and Se) in Glutathione Disulfide and Derivatives Elucidated by Quantum Chemical Calculations with the Quantum Theory of Atoms-in-Molecules Approach. Molecules. 2018; 23(2):443. https://doi.org/10.3390/molecules23020443

Chicago/Turabian Style

Hayashi, Satoko, Yutaka Tsubomoto, and Waro Nakanishi. 2018. "Behavior of the E–E’ Bonds (E, E’ = S and Se) in Glutathione Disulfide and Derivatives Elucidated by Quantum Chemical Calculations with the Quantum Theory of Atoms-in-Molecules Approach" Molecules 23, no. 2: 443. https://doi.org/10.3390/molecules23020443

Article Metrics

Back to TopTop