Next Article in Journal
Computational Study of Methane C–H Activation by Main Group and Mixed Main Group–Transition Metal Complexes
Next Article in Special Issue
Structural and Conformational Aspects in the Chemistry of Heterocycles
Previous Article in Journal
Inhibition of Photoconversion Activity in Self-Assembled ZnO-Graphene Quantum Dots Aggregated by 4-Aminophenol Used as a Linker
Previous Article in Special Issue
Enthalpies of Combustion and Formation of Severely Crowded Methyl-Substituted 1,3-dioxanes. The Magnitudes of 2,4- and 4,6-diaxial Me,Me-Interactions and the Chair–2,5-twist Energy Difference
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Reactions of 1-Substituted Silatranes and Germatranes, 2,2-Disubstituted Silocanes and Germocanes, 1,1,1-Trisubstituted Hyposilatranes and Hypogermatranes with Alcohols (Methanol, Ethanol): DFT Study

by
Denis Chachkov
1,
Rezeda Ismagilova
2 and
Yana Vereshchagina
2,*
1
Kazan Department of Joint Supercomputer Center of Russian Academy of Sciences–Branch of Federal State Institution “Scientific Research Institute for System Analysis of the RAS”, Lobachevskogo 2/31, 420111 Kazan, Russia
2
Department of Physical Chemistry, A.M. Butlerov Institute of Chemistry, Kazan Federal University, Kremlevskaya 18, 420008 Kazan, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(12), 2803; https://doi.org/10.3390/molecules25122803
Submission received: 20 May 2020 / Revised: 11 June 2020 / Accepted: 15 June 2020 / Published: 17 June 2020
(This article belongs to the Special Issue Structure and Conformational Analysis of Heterocyclic Compounds)

Abstract

:
The mechanism of reactions of silatranes and germatranes, and their bicyclic and monocyclic analogues with one molecule of methanol or ethanol, was studied at the Density Functional Theory (DFT) B3PW91/6-311++G(df,p) level of theory. Reactions of 1-substituted sil(germ)atranes, 2,2-disubstituted sil(germ)ocanes, and 1,1,1-trisubstituted hyposil(germ)atranes with alcohol (methanol, ethanol) proceed in one step through four-center transition states followed by the opening of a silicon or germanium skeleton and the formation of products. According to quantum chemical calculations, the activation energies and Gibbs energies of activation of reactions with methanol and ethanol are close, their values decrease in the series of atranes–ocanes–hypoatranes for interactions with both methanol and ethanol. The reactions of germanium-containing derivatives are characterized by lower activation energies in comparison with the reactions of corresponding silicon-containing compounds. The annular configurations of the product molecules with electronegative substituents are stabilized by the transannular N→X (X = Si, Ge) bond and different intramolecular hydrogen contacts with the participation of heteroatoms of substituents at the silicon or germanium.

Graphical Abstract

1. Introduction

Heterocyclic compounds of hypervalent atoms, primarily intra-complex compounds of triethanolamine–metallatranes, possess unique physical and chemical properties that are due to their unusual trigonal-bipyramidal structure with a transannular nitrogen–element bond and the presence of an axial fragment. Triethanolamine derivatives of silicon, germanium, tin, carbon, boron, phosphorus, and arsenic, as well as bismuth, titanium, vanadium, aluminum and iron are currently known [1,2]. The most studied compounds of the elements of group 14 are silatranes and germatranes, as well as their bicyclic analogues–ocanes and monocyclic–hypoatranes. The diverse biological activity of atrane systems depends on the nature of the substituents, and such compounds can be both effective immunostimulants and adaptogens, and immunosuppressants, thus providing a stimulating or inhibiting effect on the vital activity of micro- and macroorganisms [3,4,5,6,7,8,9,10]. In addition, atranes are used as important functional reagents [11,12,13,14,15], catalysts for the formation of polyurethanes [16,17].
Recent achievements, challenges and prospects for using atranes are given in the reviews [18,19,20]. Atranes, ocanes and hypoatranes contain an intramolecular coordination bond nitrogen→element. Theoretical studies on the structure and hypervalent intramolecular coordination N→X (X = C, Si, Ge, B, P) in some atrane systems were carried out in [21,22,23,24,25,26,27,28,29,30,31]; information on the structure of some ocanes is given in [32,33,34,35,36,37,38,39], and hypoatranes in [36,40,41,42,43].
The length and strength of the intramolecular transannular bond N→X is determined by the number and nature of electronegative substituents at the central atom X: An increase in the number of highly electronegative substituents at the silicon or germanium atom significantly reduces the length of the N→X bond and, therefore, increases its strength [33].
Despite the long history of the chemistry of atrane systems, issues affecting their spatial and electronic structure, as well as the reaction mechanisms of these compounds, do not lose their relevance, and the development of quantum chemistry methods contributes a lot to this. These new results allow the available experimental data to be substantiated. Various aspects of the reactivity of atranes are considered in reviews [19,44], the study of the electron-donating ability of heteroatoms of the silatranyl group by calculation methods is described in [45,46,47]. There are a few publications concerned with reaction mechanisms of metallatranes and their bicyclic and tricyclic analogs with nucleophiles. The results of theoretical studies of the hydrolysis reactions of some atrane, ocane and hypoatrane systems are described in [22,43,48,49,50,51,52].
In the present work, we performed a theoretical study of the reactions of 1-substituted atranes, 2,2-disubstituted ocanes, and 1,1,1-trisubstituted hypoatranes with nucleophilic reagents—alcohols (methanol and ethanol)—and compared the results with the available data for hydrolysis reactions of these compounds.

2. Results and Discussion

2.1. Methodology

A theoretical study of the reactions of 1-substituted silatranes 1ag and germatranes 2ag, 2,2-disubstituted silocanes 3ag and germocanes 4ag, 1,1,1-trisubstituted hyposilatranes 5ag and hypogermatranes 6ag (Scheme 1) with nucleophilic reagents methanol and ethanol was carried out by quantum chemical calculations.
The Density Functional Theory (DFT) with B3PW91 hybrid functional [53,54] and 6-311++G(df,p) [55] extended basis set (calculations of the molecules in vacuum) using GAUSSIAN 09 software package [56] were earlier successfully used to study hydrolysis reactions of atrane, ocane and hypoatrane systems [43,52]. The choice of the B3PW91 method was also based on the data [57]. In all cases, the geometric parameters of the molecules were fully optimized. The critical points on the potential energy surfaces were identified as energy minima by the absence of imaginary frequencies in the corresponding Hessian matrices and as transition states by the presence of one imaginary frequency therein. Transition states were assigned to reaction paths by descent toward starting molecule and reaction product. Hessian was generated analytically in all cases, when searching for minimum energy it was calculated once to confirm that a minimum energy has been entered. In the case of a search for transition state, the Hessian was analytically calculated at each iteration. The parameter GRID = UltraFine was used. The solvent has not been taken into account.
Reactions of 1-substituted silabicyclo[3.3.3]undecanes 1ag and 1-germabicyclo[3.3.3]undecanes 2ag, 2,2-disubstituted 1,3,6,2-dioxazasilocanes 3ag and 1,3,6,2-dioxazagermocanes 4ag, hyposilatranes 5ag and hypogermatranes 6ag with one methanol molecule or one ethanol molecule were simulated.
First, we carried out theoretical conformational analysis of all starting compounds 16. Namely, the conformational analysis of all reaction participants is a necessary step in studying their reactivity and establishing the reaction mechanisms. We found energetically preferred conformers for each of compounds 16. These structures were used as the starting reagents in the simulation of the mechanism of their reactions with alcohol (methanol or ethanol). The energy difference between the conformers was 5–15 kJ/mol. The barrier of the transition between the conformers was noticeably higher: 18–46 kJ/mol. The reactions studied can proceed not only for the most favorable conformer of the starting reagent, but also for those which are not advantageous. The reaction barrier for unfavorable conformers was 5–25 kJ/mol higher than for the most favorable conformation of the reagent.
According to our previous quantum chemical calculations [52], 1-substituted compounds 1ag, 2ag are classical atrane systems where the silicon or germanium atom has a flattened tetrahedral configuration, and the nitrogen moiety is also flattened; the substituent R at the X atom occupies the axial position. Eight-membered silicon-nitrogen-oxygen-containing cycles in them have crown conformations (chair–chair or chair–bath), the molecules have a transannular N→X bond (X = Si, Ge).
The most preferred conformation of bicyclic analogues of atranes–2,2-disubstituted ocanes, 3ag, 4ag, is symmetrical crown (chair–chair) where the silicon (germanium) and nitrogen atoms are close, the transannular interaction N→X (X = Si, Ge) is observed [52]. The silicon and germanium atoms are tetrahedra with one axial and one equatorial substituent.
In the preferred conformers of monocyclic hypoatranes 5ag, 6ag [43], the silicon (germanium) and nitrogen atoms are flattened tetrahedra where oxygen atom and two substituents at the Si (Ge) are in equatorial positions, and third substituent is axial. In hypoatranes with perchlorate (5e, 6e), nitrate (5f, 6f), and thiocyanate (5g, 6g) substituents, intramolecular contacts between one of the hydrogen atoms of the amino group and the heteroatom (O or N) of one of the functional groups at the Si and/or Ge, in halogen derivatives (5bd, 6bd), contacts between the hydrogen atoms of the amino group and the halogen atoms are possible. The Endo-configuration of the molecules 5ag, 6ag is determined by the transannular interaction N→Si or N→Ge as well as for atranes 1ag, 2g and ocanes 3ag, 4ag.

2.2. Reactions of 1-Substituted Atranes with Methanol

The mechanism of reactions of atranes 1ag, 2ag with one methanol molecule according to quantum chemical calculations is shown in Scheme 2. The obtained energy characteristics of the reactions and some geometric parameters of the reaction participants are given in Table 1 and Table S1. As an example, the reaction schemes for some atranes according to theoretical studies are presented in Figure 1.
The interaction of atranes 1a–g, 2a–g with methanol proceeds in one step. Initially, prereaction complexes are formed between the molecules of atrane and methanol, in which the N→Si (N→Ge) bond length decreases, a contact arises between the proton of the hydroxyl group of methanol and the O1 atom of the atrane skeleton (Table S1). The prereaction complexes are transformed into transition states in which the N→Si (N→Ge) distance is further reduced, the Si(Ge)‒O1 and O4‒H1 bonds are stretched in methanol, and the H1O1 contact is reduced (Scheme 2, Table S1). The third stage of interaction is the opening of one of the half rings of the atrane skeleton, as a result, the Si(Ge)‒O1 bond is cleaved, and the Si(Ge)‒O4 and O1‒H1 bonds are formed. The products of methanol addition to atranes are substituted ocanes—2-R-6-(2-hydroxyethyl)-1,3,2,6-dioxazasil(germ)ocane-2-ols 7a–g, 8a–g.
According to the quantum chemical calculations, the possible intramolecular hydrogen bonds O4H1O1 as well as the transannular interactions N→Si or N→Ge (Scheme 2, Table S1, and Figure 1) stabilize the structure of molecules 7a–g, 8a–g. The ocane fragments of product molecules have crown conformation—a chair–boat for silocane 7a and germocanes 8ag, and symmetrical boat–boat for silocanes 7bg. The silicon (germanium) atom is flattened tetrahedron, and the nitrogen-containing fragment is also flattened. The N→Si (N→Ge) bond in the molecules of products 7ag, 8ag is longer (Table S1) than in the molecules of the starting atranes, which is generally consistent with the change in the N→Si (N→Ge) bond length in similar atranes and their hydrolysis products [52]. In the series of halogen-substituted products 7bd and 8bd, the distance N→Si (N→Ge) increases from F to Br. The smallest distance is observed for perchlorate derivatives 7e and 8e. The intramolecular hydrogen bonds O4H1O1 in molecules 7ag, 8ag are quite strong, as evidenced by their lengths (Table S1); the strongest ones are in hydroxyl 8a and fluorine 8b derivatives, moreover the angle O4H1O1 is 161°–176°.
The analysis of Table 1 shows that the reactions of compounds 1ag, 2ag with methanol are controlled by thermodynamic factors; the contribution of the enthalpy component is most significant, since the activation enthalpy and the Gibbs activation energy change symbiotically. This conclusion is valid not only for activation parameters, but also for the enthalpy of the reaction (thermal effect) and for the Gibbs energy of the reaction.
In the series of halogenated silatranes and germatranes (F, Cl, Br), the activation energy increases (Table 1) naturally with a decrease in the electronegativity of the substituents [33,58]. The reactions of germatranes 2ag proceed with slightly lower activation energies compared with reactions of silatranes 1ag. In transition states, the smallest distance N‒X (X = Si, Ge) and the lowest activation energy (Table 1, Table S1) are observed for perchlorate-substituted atranes 1e and 2e.

2.3. Reactions of 2,2-Disubstituted Ocanes with Methanol

According to DFT calculations, the mechanism of reactions of ocanes 3ag, 4ag with one methanol molecule is presented in Scheme 3. The obtained energy characteristics of the reactions and some geometric parameters of the reaction participants are given in Table 2 and Table S2. As an example, the reaction schemes for some ocanes according to theoretical studies are presented in Figure 2.
The interaction of ocanes 3ag, 4ag with methanol also proceeds in one step. Initially, prereaction complexes are formed between the molecules of ocanes and methanol, in which the N‒Si (N‒Ge) bond length decreases, a contact arises between the proton of the hydroxyl group of methanol and the O1 atom. Prereaction complexes are transformed into transition states, in which N‒Si (N‒Ge) distance is further reduced, while the Si‒O1 (Ge‒O1) bond is extended (Table S2), the silicon(germanium)- and nitrogen-containing fragments are flattened, which leads to distortion of crown conformation. The O3‒H2 bond is stretched, and the O1‒H2 contact arises.
The nucleophilic attack of a methanol molecule leads to the cleavage of the Si‒O1 (Ge‒O1) bond in the molecules of ocanes 3ag, 4ag and the appearance of a new Si‒O3 (Ge‒O3) bond. The annular configuration of product molecules 9ag, 10ag is stabilized due to the dative interaction N→Si (N→Ge) and the intramolecular hydrogen bond O2H3 (Table S2). The possibility of a transannular interaction N→Si (N→Ge) is evidenced by both the spatial factor (the distance N‒Si (N‒Ge) is less than the sum of the van der Waals radii of these elements) and the presence of electronegative substituents at the acceptor silicon or germanium atom. In addition, short contacts between the hydrogen (N‒H) and the heteroatom of the axial substituent at the Si (Ge) atom contribute to the stabilization of the crown-like 10-membered configuration of molecules: oxygen in hydroxyl substituted 9a and 10a, perchlorate 9e and 10e, nitro 9f and 10f, halogen in 9bd and 10bd, nitrogen of the thiocyanate group in 9g and 10g.
The silicon, germanium and nitrogen atoms in the molecules 9ag, 10ag are flattened tetrahedra. The N–Si (N‒Ge) distance in the molecules of the products 9ag, 10ag is less than in the reagents, as well as in the hydrolysis reactions of these compounds [52], with the exception of dichloro- and dibromo-substituted germocanes 10c and 10d (Table S2). Silocanes are characterized by a regular decrease in the N–X distance as the electronegativity of the exocyclic halogen substituent increases in a series—F, Cl, Br; however, the opposite regularity is observed for germocanes (Table S2).
The reactions of ocanes 3ag, 4ag are controlled by thermodynamic factors. The activation energy in a series of halogen-substituted ocanes 9bd and 10bd (F, Cl, Br) increases naturally with a decrease in the electronegativity of the substituents (Table 2). In the case of reactions of germocanes 4ag, the activation energies and Gibbs activation energies are lower than those of silocanes 3ag. The smallest N–Si (N‒Ge) distances and the smallest values of the activation energy (Gibbs activation energy) are observed for the diperchlorate derivatives 9e and 10e (Table 2, Table S2). The same regularity was revealed for hydrolysis reactions of similar ocanes [52].

2.4. Reactions of 1,1,1-Trisubstituted Hypoatranes with Methanol

The mechanism of reactions of 1,1,1-substituted hyposilatranes 5ag and hypogermatranes 6ag with one methanol molecule according to quantum chemical calculations is shown in Scheme 4. The obtained energy characteristics of the reactions and some geometric parameters of the reaction participants are given in Table 3 and Table S3. As an example, the reaction schemes for some ocanes according to theoretical studies are presented in Figure 3.
The interaction of hypoatranes 5ag, 6ag with methanol proceeds in several stages, as in the case of atranes 1ag, 2ag and ocanes 3ag, 4ag. At the first stage, prereaction complexes are formed in which the reagent molecules come together, the distance N‒X (X = Si, Ge) is reduced (Scheme 4, Table S3). Then, the prereaction complexes are transformed into transition states, the distances N‒X (X = Si, Ge) and O1H3 are shortened, and the X‒O1 bond is stretched, a new contact arises between the silicon (germanium) atom and the oxygen atom of methanol (Scheme 4, Table S3). At the final stage, the Si‒O1 (Ge‒O1) bond breaks and a new Si‒O2 (Ge‒O2) bond arises, resulting in reaction products–complexes of 2-aminoethanol with trisubstituted methoxysilanes 11ag and methoxygermanes 12ag, in which the N–Si or N‒Ge distance is shorter than in the reagent molecules (Table S3).
The configuration of molecules 11ag, 12ag is due to the transannular bond N→Si or N→Ge, the dative interaction O→Si or O→Ge (11bg, 12bg), and the bonding between the oxygen of methoxyl substituent at the Si (Ge) and a hydrogen of the hydroxyl group of 2-aminoethanol (with the exception of trinitroderivative 12f) (Scheme 4, Table S3). In addition, as in the starting molecules 5bg and 6bg, in complexes 11bg and 12bg hydrogen contacts between the hydrogen atoms of the amino group and heteroatoms (halogen, oxygen of perchlorate or nitro groups, and nitrogen of thiocyanate groups) of substituents at the Si or Ge are possible. In complexes 11ag and 12ag, the silicon or germanium polyhedron expands: The tetravalent Si or Ge atom is six-coordinated and has the structure of a distorted tetragonal bipyramid. Two substituents at the silicon (germanium) and oxygen atoms lie in the equatorial plane of this bipyramid, and the nitrogen and third substituent at the Si (Ge) occupy axial positions (Scheme 4, Figure 3). Apparently, the preferred configurations of the complexes 11ag, 12ag are stabilized precisely due to the dative interactions and intramolecular hydrogen bonds arising because of the favorable orientation of the substituents at the silicon or germanium.
The reactions of hypoatranes 5ag, 6ag with methanol, as in the case of ocanes 3ag, 4ag and atranes 1ag, 2ag, are controlled by thermodynamic factors. In series of halogen-substituted (F, Cl, Br) hyposilatranes and hypogermatranes, the activation energy increases (Table 3), which is naturally associated with a decrease in the electronegativity of the substituents and is consistent with the data for the corresponding atranes and ocanes. For the reactions of hypogermatranes 6ag, the activation energies are lower in comparison with the reactions of hyposilatranes 5ag (Table 3). The smallest N‒X (X = Si, Ge) distance in the transition states and the lowest activation energy are observed for the reactions of triperchlorate derivatives 5e and 6e (Table 3, Table S3).
The energy diagrams of the reactions of hydroxy-, fluoro-, and perchlorate-substituted atranes, ocanes, and hypoatranes are shown in Figure 4. To estimate the thermal effect of the reactions, for zero we chose the energies of the starting reagent and methanol molecules infinitely distant from each other.
A comparison of the results obtained for 1-substituted atranes, 2,2-disubstituted ocanes and 1,1,1-trisubstituted hypoatranes (Table 1, Table 2 and Table 3) shows that the values of activation energy and Gibbs activation energy decrease in the series of atranes—ocanes—hypoatranes. A decrease in the number of cycles in the molecule in this series and an increase in the number of highly electronegative substituents (having negative mesomeric and inductive effects) at the silicon or germanium atom led to a decrease in the N‒Si or N‒Ge bond length (Tables S1–S3) and, consequently, an increase in its strength. Probably, this fact, along with the presence of dative and intramolecular hydrogen interactions, explains the high stability of trisubstituted hypoatranes obtained in an aqueous–alcoholic medium at elevated temperatures [36,40,41,42].

2.5. Reactions of Atranes, Ocanes, and Hypoatranes with Ethanol

We studied the reactions of substituted atranes and their analogues with ethanol using the reactions of hydroxy derivatives 1a, 2a, 3a, 4a, 5a, 6a and halogen derivatives 1bd, 2bd, 3bd, 4bd, 5bd, 6bd to compare the interactions with methanol. The mechanism of reactions of 1-substituted silatranes and germatranes, 2,2-disubstituted silocanes and germocanes, 1,1,1-trisubstituted hyposilatranes and hypogermatranes with one ethanol molecule according to quantum chemical calculations is shown in Scheme 5. The calculated data are given in Table 4, Tables S4–S6.
As in the case of interaction with methanol, the reactions with ethanol proceed in one step: Initially, prereaction complexes are formed that transform into transition states, as a result of the cleavage of one Si‒O (Ge‒O) bond and the arising of a new Si‒O(CH2CH3) or Ge‒O(CH2CH3) bond, the reaction products are formed.
The transition states are characterized by a decrease in the N‒Si (N‒Ge) distance, the Si(Ge)‒O1 and O‒H bonds are stretched, and the H1O1 contact is reduced (Scheme 5, Tables S4–S6). Molecules of the products of ethanol addition reactions have a configuration similar to the configuration of the corresponding product molecules formed upon interaction with methanol.
The products of reactions of atranes are substituted silocanes 13ad and germocanes 14ad. According to the quantum chemical calculations, the ocane fragments of product molecules have crown conformation—symmetrical boat–boat for silocanes and chair–boat for germocanes. The silicon (germanium) and nitrogen-containing fragments are flattened. The transannular interaction N→Si or N→Ge and the possible intramolecular hydrogen bonds O4H1O1 stabilize these structures (Table S4, Scheme 5). The N→Si (N→Ge) bond in the molecules of products 13ad and 14ad is longer than in the molecules of the starting atranes (Table S4).
The annular configuration of product molecules 15ad and 16ad is also stabilized due to the dative interaction N→Si (N→Ge) and the intramolecular hydrogen bond O2H3 (Table S5, Scheme 5). In addition, short contacts between the hydrogen atom at the nitrogen and the heteroatom of the axial substituent at the silicon (germanium) atom contribute to the stabilization of the crown-like 10-membered configuration of molecules 15ad and 16ad: oxygen in the hydroxy-substituted 15a and 16a or halogen in 15bd and 16bd. The N‒Si (N‒Ge) distance in the molecules of products 15ad and 16ad is shorter than in the molecules of the starting ocanes, as in the case of reactions with methanol, and this distance increases in 15b and 16d (Table S5).
The configuration of the molecules 17ad and 18ad is due to the transannular interaction N→Si or N→Ge and hydrogen bonding O2H3 between oxygen atom of the ethoxyl substituent at the silicon (germanium) and hydrogen atom of the hydroxy group of 2-aminoethanol (Scheme 5, Table S6). In addition, as in the starting molecules 5ad and 6ad, hydrogen contacts between the hydrogen atoms of the amino group and the heteroatoms (oxygen or halogen) of the substituents at the silicon or germanium are possible in products 17ad and 18ad. In complexes 17ad and 18ad, the tetravalent six-coordinated silicon or germanium atom has the structure of a distorted tetragonal bipyramid, in the equatorial plane of which there are two substituents at the silicon or germanium and oxygen atoms, and the nitrogen and the third substituent at the Si (Ge) occupy the axial positions.
The transannular interaction N→Si (N→Ge) in the product molecules is due to both the spatial factor and the presence of electronegative substituents at the acceptor silicon or germanium atom. In all product molecules, sufficiently strong hydrogen bonds are formed between the oxygen atom of the ethoxyl substituent at the silicon (germanium) and the hydrogen atom of the hydroxy group (Scheme 5, Tables S4–S6).
The analysis of obtained data shows that the reactions with ethanol are controlled by thermodynamic factors (Table 4, Figure 5). The value of activation energy naturally increases in the series of halogen-substituted (F, Cl, Br) compounds 1bd, 2bd, 3bd, 4bd, 5bd, and 6bd by a decrease in the electronegativity of the substituents [33,58]. Reactions of germanium-containing derivatives proceed with slightly lower activation energies as compared with reactions of silicon-containing derivatives. In addition, the activation energy decreases in the series of atranes–ocanes–hypoatranes (Table 4, Figure 5). The same regularities were revealed for hydrolysis reactions of similar atranes, ocanes [52], and hypoatranes [43].
A comparison of the results for reactions of atranes, ocanes and hypoatranes with ethanol and the corresponding data for reactions with methanol allowed us to conclude that the activation energies are close; moreover, in the case of ocanes they are slightly higher in reactions with ethanol.

3. Conclusions

Thus, we have studied the mechanism of reactions of 1-sustituted sil(germ)atranes, 2,2-disustituted sil(germ)ocanes, and 1,1,1-trisubstituted hyposil(germ)atranes with one molecule of methanol or ethanol at the DFT B3PW91/6-311++G(df,p) level of theory. These reactions with alcohol (methanol, ethanol), as well as with water [43,52], proceed in one step through four-center prereaction complexes and transition states followed by the opening of heteroatom (silicon or germanium) skeleton and the formation of products.
According to DFT calculations, the activation energies and Gibbs energies of activation of reactions with methanol and ethanol are close, their values decrease in the series of atranes-ocanes-hypoatranes for reactions with both methanol and ethanol. Reactions of germanium-containing derivatives are characterized by lower activation energies in comparison with silicon-containing compounds.
The results of theoretical conformational analysis show that the annular configurations of the product molecules with electronegative substituents are stabilized by the transannular N→X (X = Si, Ge) interaction and different intramolecular hydrogen contacts with the participation of heteroatoms of substituents at the silicon or germanium.

Supplementary Materials

The following are available online, Table S1: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of atranes 1ag, 2ag (a OH, b F, c Cl, d Br, e OClO3, f ONO2, g SCN) with methanol. For X‒N distances the Mayer bond order in the NAO basis are given in parentheses, Table S2: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of ocanes 3ag, 4ag (a OH, b F, c Cl, d Br, e OClO3, f ONO2, g SCN) with methanol. For X‒N distances the Mayer bond order in the NAO basis are given in parentheses, Table S3: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of hypoatranes 5ag, 6ag (a OH, b F, c Cl, d Br, e OClO3, f ONO2, g SCN) with methanol. For X‒N distances the Mayer bond order in the NAO basis are given in parentheses, Table S4: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of atranes 1ad, 2ad (a OH, b F, c Cl, d Br) with ethanol, Table S5: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of ocanes 3ad, 4ad (a OH, b F, c Cl, d Br) with ethanol, Table S6: Selected interatomic distances (Å) in the reagents, prereaction complexes, transition states, and products of the reactions of hypoatranes 5ad, 6ad (a OH, b F, c Cl, d Br) with ethanol.

Author Contributions

Conceptualization, D.C. and Y.V.; investigation, D.C., R.I., and Y.V.; methodology, Y.V.; software, D.V.; project administration, Y.V.; supervision, Y.V.; visualization, R.I. and Y.V.; writing—original draft, D.C., R.I., and Y.V.; writing—review & editing, D.C. and Y.V. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly performed under financial support by the state subsidy allocated to Kazan (Volga Region) Federal University to increase its competitiveness among world leading scientific educational centers.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Voronkov, M.G.; Zelčans, G.; Mažeika, I.; Baltkajs, J.; Jegorov, J.; Pestunovich, V. Silatranes: Intracomplex heterocyclic compounds of pentacoordinated silicon. Pure Appl. Chem. 1966, 13, 35–59. [Google Scholar] [CrossRef]
  2. Verkade, J.G. Main group atranes: Chemical and structural features. Coord. Chem. Rev. 1994, 137, 233–295. [Google Scholar] [CrossRef]
  3. Voronkov, M.G.; Baryshok, V.P. Antitumor activity of silatranes (a review). Pharm. Chem. J. 2004, 38, 3–9. [Google Scholar] [CrossRef]
  4. Voronkov, M.G.; Baryshok, V.P. Atranes as a new generation of biologically active substances. Herald Russ. Acad. Sci. 2010, 80, 514–521. [Google Scholar] [CrossRef]
  5. Menchikov, L.G.; Ignatenko, M.A. Biological activity of organogermanium compounds (a review). Pharm. Chem. J. 2013, 46, 635–638. [Google Scholar] [CrossRef]
  6. Voronkov, M.G.; Baryshok, V.P. Silatranes in medicine and agriculture; Publishing House SB RAS: Novosibirsk, Russia, 2005; p. 255. [Google Scholar]
  7. Garabadzhiu, A.V.; Voronkov, M.G.; Nyanikova, G.G.; Samokhin, G.S.; Vrazhnov, D.V.; Kochina, T.A. The influence of silatranes, germatranes, protatranes, and triethanolamine on vital functions of microorganisms. Dokl. Biol. Sci. 2011, 439, 264–266. [Google Scholar] [CrossRef]
  8. Singh, G.; Saroa, A.; Girdhar, S.; Rani, S.; Sahoo, S.; Choquesillo-Lazarte, D. Synthesis, characterization, electronic absorption and antimicrobial studies of N-(silatranylpropyl)phthalimide derived from phthalic anhydride. Inorg. Chim. Acta 2015, 427, 232–239. [Google Scholar] [CrossRef]
  9. Ye, F.; Song, X.; Liu, J.; Xu, X.; Wang, Y.; Hu, L.; Wang, Y.; Liang, G.; Guo, P.; Xie, Z. Design, synthesis, and biological evaluation of γ-aminopropyl silatrane–acyclovir hybrids with immunomodulatory effects. Chem. Biol. Drug Des. 2015, 86, 905–910. [Google Scholar] [CrossRef]
  10. Singh, G.; Arora, A.; Mangat, S.S.; Rani, S.; Kaur, H.; Goyal, K.; Sehgal, R.; Maurya, I.K.; Tewari, R.; Choquesillo-Lazarte, D.; et al. Design, synthesis and biological evaluation of chalconyl blended triazole allied organosilatranes as giardicidal and trichomonacidal agents. Eur. J. Med. Chem. 2016, 108, 287–300. [Google Scholar] [CrossRef]
  11. Singh, R.; Mutneja, R.; Kaur, V.; Wagler, J.; Kroke, E. Derivatization of 3-aminopropylsilatrane to introduce azomethine linkage in the axial chain: Synthesis, characterization and structural studies. J. Organomet. Chem. 2013, 724, 186–191. [Google Scholar] [CrossRef]
  12. Brennan, B.J.; Gust, D.; Brudvig, G.W. Organosilatrane building blocks. Tetrahedron Lett. 2014, 55, 1062–1064. [Google Scholar] [CrossRef]
  13. Mutneja, R.; Singh, R.; Kaur, V.; Wagler, J.; Felsc, S.; Kroke, E. Schiff base tailed silatranes for the fabrication of functionalized silica based magnetic nano-cores possessing active sites for the adsorption of copper ions. New J. Chem. 2016, 40, 1640–1648. [Google Scholar] [CrossRef] [Green Version]
  14. Vlasova, N.N.; Sorokin, M.S.; Oborina, E.N. Carbofunctional sulfur-containing organosilicon compounds. Appl. Organomet. Chem. 2017, 31, e3668. [Google Scholar] [CrossRef]
  15. Singh, G.; Arora, A.; Rani, S.; Kalra, P.; Aulakh, D.; Wriedt, M. A family of silatrane-armed triazole-encapped salicylaldehyde-derived Schiff bases: Synthesis, spectral analysis, and antimicrobial and quantum chemical evaluation. Appl. Organomet. Chem. 2017, 31, e3728. [Google Scholar] [CrossRef]
  16. Krause, J.; Richter, S. New catalysts and use thereof in the production of polyurethanes. Patent DE 10 2008 021 980 A1, 18 April 2009. [Google Scholar]
  17. Zöller, T.; Iovkova-Berends, L.; Dietz, C.; Berends, T.; Jurkschat, K. On the reaction of elemental tin with alcohols: A straightforward approach to tin(II) and tin(IV) alkoxides and related tinoxo clusters. Chem. Eur. J. 2011, 17, 2361–2364. [Google Scholar] [CrossRef]
  18. Karlov, S.S.; Zaitseva, G.S. Germatranes and their analogs. Synthesis, structure, and reactivity. Chem. Heterocycl. Comp. 2001, 37, 1325–1357. [Google Scholar] [CrossRef]
  19. Puri, J.K.; Singh, R.; Chahal, V.K. Silatranes: A review on their synthesis, structure, reactivity and applications. Chem. Soc. Rev. 2011, 40, 1791–1840. [Google Scholar] [CrossRef]
  20. Adamovich, S.N. New atranes and similar ionic complexes. Synthesis, structure, properties. Appl Organomet. Chem. 2019, 33, e4940. [Google Scholar] [CrossRef]
  21. Milov, A.A.; Minyaev, R.M.; Minkin, V.I. Hypervalent intramolecular X←N (X = C, Si, Ge) coordination in atranes: Quantum-chemical study. Russ. J. Org. Chem. 2003, 39, 340–347. [Google Scholar] [CrossRef]
  22. Chernyshev, E.A.; Knyazev, S.P.; Kirin, V.N.; Vasilev, I.M.; Alekseev, N.V. Structural features of silatranes and germatranes. Russ. J. Gen. Chem. 2004, 74, 58–65. [Google Scholar] [CrossRef]
  23. Zabalov, M.V.; Karlov, S.S.; Zaitseva, G.S.; Lemenovskii, D.A. The molecular and electronic structure features of silatranes, germatranes, and their carbon analogs. Russ. Chem. Bull. 2006, 55, 464–476. [Google Scholar] [CrossRef]
  24. Feshin, V.P.; Feshina, E.V. Nature of coordination bond in silatranes and its formation dynamics according to the ab initio calculations. Russ. J. Gen. Chem. 2014, 84, 70–74. [Google Scholar] [CrossRef]
  25. Korlyukov, A.A. Coordination compounds of tetravalent silicon, germanium and tin: The structure, chemical bonding and intermolecular interactions in them. Russ. Chem. Rev. 2015, 84, 422–440. [Google Scholar] [CrossRef]
  26. Marín-Luna, M.; Alkorta, I.; Elguero, J. Theoretical study of the geometrical, energetic and NMR properties of atranes. J. Organomet. Chem. 2015, 794, 206–215. [Google Scholar] [CrossRef]
  27. Karlov, S.S.; Tyurin, D.A.; Zabalov, M.V.; Churakov, A.V.; Zaitseva, G.S. Quantum chemical study of group 14 elements pentacoordinated derivatives–metallatranes. J. Mol. Struct. THEOCHEM. 2005, 724, 31–37. [Google Scholar] [CrossRef]
  28. Belogolova, E.F.; Sidorkin, V.F. Correlation among the gas-phase, solution, and solid-phase geometrical and NMR parameters of dative bonds in the pentacoordinate silicon compounds. 1-Substituted silatranes. J. Phys. Chem. A. 2013, 117, 5365–5376. [Google Scholar] [CrossRef]
  29. Ignatyev, I.S.; Samokhin, G.S.; Kochina, T.A.; Belyaeva, V.V.; Khaikin, S.Y.; Montejo, M.; López González, J.J.; Voronkov, M.G. Vibrational spectra and electronic structure of germatranols (HO)4-n Ge(OCH2CH2)nNR3-n (R = H; n = 1–3) with transannular GeN bonding. J. Organomet. Chem. 2013, 747, 62–68. [Google Scholar] [CrossRef]
  30. Voronkov, M.G.; Zel’bst, E.A.; Belyaeva, V.V. Intramolecular inductive through-space interaction between nitrogen and oxygen atoms in silatranes, quasisilatranes, protatranes, triethanolamine, and diethanolamine. Dokl. Chem. 2014, 433, 167–170. [Google Scholar] [CrossRef]
  31. Glowacki, B.; Lutter, M.; Schollmeyer, D.; Hiller, W.; Jurkschat, K. Novel stannatrane N(CH2CMe2O)2(CMe2CH2O)SnO-t-Bu and related oligonuclear tin(IV) oxoclusters. Two isomers in one crystal. Inorg. Chem. 2016, 55, 10218–10228. [Google Scholar] [CrossRef]
  32. Lermontova, E.K.; Selina, A.A.; Karlov, S.S.; Churakov, A.V.; Howard, J.A.K.; Oprunenko, Y.F.; Antipin, M.Y.; Sundermeyer, J.; Zaitseva, G.S. The transannular interaction germanium–nitrogen in germocanes: The influence of substituents. J. Organomet. Chem. 2006, 691, 5710–5724. [Google Scholar] [CrossRef]
  33. Voronkov, M.G.; Korlyukov, A.A.; Zelbst, E.A.; Grebneva, E.A.; Trofimova, O.M.; Antipin, M.Y. Molecular structure of 1,1-difluoroquasisilatrane (N→Si) F2Si(OCH2CH2)2NH. Dokl. Chem. 2008, 418, 27–29. [Google Scholar] [CrossRef]
  34. Ignat’ev, I.S.; Voronkov, M.G.; Kochina, T.A.; Vrazhnov, D.V. Quantum-chemical study of the stereoelectronic structure of 1-fluorosilatrane, 1,1-difluoroquasisilatrane, 1,1,1-trifluorohyposilatrane, and cations formed thereof. Russ. J. Gen. Chem. 2010, 80, 2274–2282. [Google Scholar] [CrossRef]
  35. Alekseev, N.V.; Chernyshev, E.A. A quantum chemical study of the electronic structure of substituted germocanes. J. Struct. Chem. 2010, 51, 419–427. [Google Scholar] [CrossRef]
  36. Voronkov, M.G.; Korlyukov, A.A.; Samokhin, G.S.; Vrazhnov, D.V.; Kochina, T.A. Germatranes and their quasi and hypo analogs with highly electronegative substituent at the Ge atom. Russ. Chem. Bull. 2012, 61, 992–998. [Google Scholar] [CrossRef]
  37. Vereshchagina, Y.A.; Alimova, A.Z.; Chachkov, D.V.; Ishmaeva, E.A.; Kochina, T.A.; Voronkov, M.G. Polarity and structure of 1,3,5,11-tetraoxa-8-aza-4-germaspiro [3,7]undecan-2-one. Russ. J. Org. Chem. 2014, 50, 1225–1226. [Google Scholar] [CrossRef]
  38. Vereshchagina, Y.A.; Alimova, A.Z.; Chachkov, D.V.; Ishmaeva, E.A.; Kochina, T.A. Polarity and structure of 1,1-dihalo-2,8-dioxa-5-azagermocanes. Russ. J. Org. Chem. 2015, 51, 750–752. [Google Scholar] [CrossRef]
  39. Ignatyev, I.S.; Kochina, T.A.; Avrorin, V.V.; Gurzhiy, V.V.; Fundamensky, V.S. Molecular and crystal structures of 2-phenyl-2-hydro-6-methyl-1,3-dioxa-6-aza-2-silacyclooctane. J. Mol. Struct. 2015, 1094, 169–173. [Google Scholar] [CrossRef] [Green Version]
  40. Samokhin, G.S.; Vrazhnov, D.V.; Kochina, T.A.; Voronkov, M.G. Quasigermatranes and their monocyclic analogs with highly electronegative substituents at the germanium atom: A new synthesis strategy. Glass Phys. Chem. 2010, 36, 623–627. [Google Scholar] [CrossRef]
  41. Samokhin, G.S. A new path to Ge-substituted germatranes, quasigermatranes, and hypogermatranes. Ph.D. Thesis, I.V. Grebenshchikov Institute of Silicate Chemistry of RAS, Saint-Petersburg, Russia, 2011. [Google Scholar]
  42. Voronkov, M.G.; Samokhin, G.S.; Vrazhnov, D.V.; Kochina, T.A. New intracomplex compounds of pentacoordinate germanium. Hypogermatranes, Ge-substituted (N→Ge) 2-oxa-5-aza-1-germacyclopentanes. Russ. J. Gen. Chem. 2012, 82, 170–171. [Google Scholar] [CrossRef]
  43. Vereshchagina, Y.A.; Ismagilova, R.R.; Chachkov, D.V. Mechanism of hydrolysis of 1,1,1-trisubstituted hyposilatranes and hypogermatranes. Russ. J. Org. Chem. 2019, 55, 227–233. [Google Scholar] [CrossRef]
  44. Voronkov, M.G.; Belyaeva, V.V.; Abzaeva, K.A. Basicity of silatranes (Review). Chem. Heterocycl. Comp. 2012, 47, 1330–1338. [Google Scholar] [CrossRef]
  45. Sidorkin, V.F.; Shagun, V.A.; Pestunovich, V.A. Stereoelectronic effects and the problem of the choice of model compounds for organic derivatives of a pentacoordinated silicon atom (taking silatranes as an example). Russ. Chem Bul. 1999, 48, 1049–1053. [Google Scholar] [CrossRef]
  46. Yashikawa, A.; Gordon, M.S.; Sidorkin, V.F.; Pestunovich, V.A. Proton affinities of the silatranes and their analogues. Organometallics 2001, 20, 927–931. [Google Scholar] [CrossRef] [Green Version]
  47. Belyaeva, V.V.; Frolov, Y.L.; Abumov, P.F.; Voronkov, M.G. Quantum-chemical study of the electron-donating ability of endocyclic nitrogen and oxygen atoms in silicon-substituted silatranes. Dokl. Phys. Chem. 2004, 397, 150–153. [Google Scholar] [CrossRef]
  48. Sok, S.; Gordon, M.S. A dash of protons: A theoretical study on the hydrolysis mechanism of 1-substituted silatranes and their protonated analogs. Comput. Theor. Chem. 2012, 987, 2–15. [Google Scholar] [CrossRef]
  49. Zoöller, T.; Jurkschat, K. Novel trialkanolamine derivatives of tin of the type [N(CH2CMe2O)2(CH2)nOSnOR]m (m = 1, 2; n = 2, 3; R = t-Bu, 2,6-Me2C6H3) and related tri- and pentanuclear tin(IV) oxoclusters. Syntheses and molecular structures. Inorg. Chem. 2013, 52, 1872–1882. [Google Scholar] [CrossRef]
  50. Vereshchagina, Y.A.; Chachkov, D.V.; Alimova, A.Z.; Ishmaeva, E.A. Mechanism of the hydrolysis reactions of 1-hydroxysilatrane and 1-hydroxygermatrane, 2,2-dihydroxysilocane and 2,2-dihydroxygermocane. Phosphorus, Sulfur Silicon Relat. Elem. 2016, 191, 496–501. [Google Scholar] [CrossRef]
  51. Ignatyev, I.S.; Montejo, M.; Rodriguez Ortega, P.G.; Kochina, T.A.; González, J.J.L. DFT study of the hydrolysis reaction in atranes and ocanes: The influence of transannular bonding. J. Mol. Model. 2016, 22, 3. [Google Scholar] [CrossRef]
  52. Vereshchagina, Y.A.; Chachkov, D.V.; Ismagilova, R.R.; Vedeneeva, E.A. Mechanism of hydrolysis of 2,2-disubstituted silocanes and germocanes and 1-substituted silatranes and germatranes. Russ. J. Org. Chem. 2018, 54, 490–499. [Google Scholar] [CrossRef]
  53. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  54. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  56. 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. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  57. Medvedev, M.G.; Bushmarinov, I.S.; Sun, J.; Perdew, J.P.; Lyssenko, K.A. Density functional theory is straying from the path toward the exact functional. Science 2017, 355, 49–52. [Google Scholar] [CrossRef] [PubMed]
  58. Anglada, J.M.; Bo, C.; Bofill, J.M.; Crehuet, R.; Poblet, J.M. Inductive effects in neutral pentacoordinated silicon compounds containing a Si←N dative bond. A theoretical study. Organometallics 1999, 18, 5584–5593. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Scheme 1. Structures of atranes, ocanes, and hypoatranes.
Scheme 1. Structures of atranes, ocanes, and hypoatranes.
Molecules 25 02803 sch001
Scheme 2. Mechanism of reactions of atranes with methanol.
Scheme 2. Mechanism of reactions of atranes with methanol.
Molecules 25 02803 sch002
Figure 1. Mechanism of reactions of germatranes 2a, 2b, 2e with methanol according to DFT calculations.
Figure 1. Mechanism of reactions of germatranes 2a, 2b, 2e with methanol according to DFT calculations.
Molecules 25 02803 g001
Scheme 3. Mechanism of reactions of ocanes with methanol.
Scheme 3. Mechanism of reactions of ocanes with methanol.
Molecules 25 02803 sch003
Figure 2. Mechanism of reactions of silocanes 3a, 3c, 3f with methanol according to DFT calculations.
Figure 2. Mechanism of reactions of silocanes 3a, 3c, 3f with methanol according to DFT calculations.
Molecules 25 02803 g002
Scheme 4. Mechanism of reactions of hypoatranes with methanol.
Scheme 4. Mechanism of reactions of hypoatranes with methanol.
Molecules 25 02803 sch004
Figure 3. Mechanism of reactions of hypogermatranes 6a, 6d, 6g with methanol according to DFT calculations.
Figure 3. Mechanism of reactions of hypogermatranes 6a, 6d, 6g with methanol according to DFT calculations.
Molecules 25 02803 g003
Figure 4. Energy diagrams for reactions of 1a,b,e; 2a,b,e; 3a,b,e; 4a,b,e; 5a,b,e; 6a,b,e with methanol (ΔE, kJ/mol).
Figure 4. Energy diagrams for reactions of 1a,b,e; 2a,b,e; 3a,b,e; 4a,b,e; 5a,b,e; 6a,b,e with methanol (ΔE, kJ/mol).
Molecules 25 02803 g004aMolecules 25 02803 g004b
Scheme 5. Mechanism of reactions of atranes, ocanes, and hypoatranes with ethanol.
Scheme 5. Mechanism of reactions of atranes, ocanes, and hypoatranes with ethanol.
Molecules 25 02803 sch005
Figure 5. Energy diagrams of formation of products 13a,b; 14a,b; 15a,b; 16a,b; 17a,b; 18a,b.
Figure 5. Energy diagrams of formation of products 13a,b; 14a,b; 15a,b; 16a,b; 17a,b; 18a,b.
Molecules 25 02803 g005
Table 1. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of atranes 1ag, 2ag with methanol.
Table 1. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of atranes 1ag, 2ag with methanol.
ΔE #ΔG #ν1, cm−1
1a91.3149.4877i
1b69.3128.1910i
1c72.4131.9951i
1d73.2132.8966i
1e62.4120.4992i
1f71.9129.1969i
1g78.8135.8972i
2a56.4113.2821i
2b61.5115.6803i
2c67.3121.4815i
2d68.7122.7819i
2e56.1110.9891i
2f66.3119.4851i
2g72.4124.8822i
Table 2. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of ocanes 3ag, 4ag with methanol.
Table 2. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of ocanes 3ag, 4ag with methanol.
ΔE #ΔG #ν1, cm−1
3a67.7125.6835i
3b48.1106.0942i
3c55.1114.0990i
3d55.8113.41020i
3e33.793.31041i
3f48.6104.61000i
3g49.9107.61035i
4a42.197.3802i
4b24.479.8922i
4c38.093.9917i
4d41.597.3928i
4e14.270.81010i
4f37.692.6950i
4g38.595.3952i
Table 3. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of hypoatranes 5ag, 6ag with methanol.
Table 3. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of hypoatranes 5ag, 6ag with methanol.
ΔE #ΔG #ν1, cm−1
5a79.9136.0963i
5b30.8−95.11020i
5c35.592.91049i
5d39.396.31064i
5e−4.756.01145i
5f13.972.61074i
5g28.483.91103i
6a56.9110.7925i
6b13.468.31000i
6c27.784.2995i
6d33.489.9994i
6e−15.144.81120i
6f3.661.21037i
6g21.779.21034i
Table 4. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of 1ad6ad with ethanol.
Table 4. Activation energies, Gibbs activation energies (kJ/mol), and imaginary vibration frequencies (cm−1) of transition states # of reactions of 1ad6ad with ethanol.
ΔE #ΔG #ν1, cm−1
1a76.4137.4893i
1b71.4130.5917i
1c74.8134.5956i
1d75.7134.7973i
2a58.3115.5833i
2b63.3117.9814i
2c69.5124.1820i
2d71.1125.6828i
3a74.1130.9857i
3b50.0108.1951i
3c57.1116.1998i
3d57.9116.51028i
4a53.4107.9905i
4b33.488.3976i
4c43.499.1968i
4d47.6103.1965i
5a78.4131.9914i
5b32.188.81030i
5c36.893.61056i
5d40.697.41072i
6a54.8106.6883i
6b14.669.71012i
6c29.386.11003i
6d35.291.91001i

Share and Cite

MDPI and ACS Style

Chachkov, D.; Ismagilova, R.; Vereshchagina, Y. Mechanism of Reactions of 1-Substituted Silatranes and Germatranes, 2,2-Disubstituted Silocanes and Germocanes, 1,1,1-Trisubstituted Hyposilatranes and Hypogermatranes with Alcohols (Methanol, Ethanol): DFT Study. Molecules 2020, 25, 2803. https://doi.org/10.3390/molecules25122803

AMA Style

Chachkov D, Ismagilova R, Vereshchagina Y. Mechanism of Reactions of 1-Substituted Silatranes and Germatranes, 2,2-Disubstituted Silocanes and Germocanes, 1,1,1-Trisubstituted Hyposilatranes and Hypogermatranes with Alcohols (Methanol, Ethanol): DFT Study. Molecules. 2020; 25(12):2803. https://doi.org/10.3390/molecules25122803

Chicago/Turabian Style

Chachkov, Denis, Rezeda Ismagilova, and Yana Vereshchagina. 2020. "Mechanism of Reactions of 1-Substituted Silatranes and Germatranes, 2,2-Disubstituted Silocanes and Germocanes, 1,1,1-Trisubstituted Hyposilatranes and Hypogermatranes with Alcohols (Methanol, Ethanol): DFT Study" Molecules 25, no. 12: 2803. https://doi.org/10.3390/molecules25122803

Article Metrics

Back to TopTop