Leonardo Saravia F.a,
Jorge Gutiérrez-Flores
b,
Eduardo H. Huerta
*a,
Jorge Garza
b,
Rubicelia Vargas
*b and
Marcos Hernández-Rodriguez
*a
aInstituto de Química, Universidad Nacional Autónoma de México, Circuito Exterior, Ciudad Universitaria, Alcaldía Coyoacán C.P. 04510, Ciudad de México, Mexico. E-mail: e.hh@quimica.unam.mx; marcoshr@unam.mx
bDepartamento de Química, División de Ciencias Básicas e Ingeniería, Universidad Autónoma Metropolitana Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, C.P. 09340, Iztapalapa, CDMX, Mexico. E-mail: rvargas@izt.uam.mx
First published on 9th October 2025
Hyperconjugative interactions are widely recognized as stabilizing electronic effects in molecular systems. While their involvement in chemical transformations has been suggested, it remains as an open question whether their influence pertains to thermodynamic stabilization or if they directly affect intrinsic kinetic parameters (also referred to as the intrinsic activation barrier). In this work, we address a fundamental question: can hyperconjugation directly shape the intrinsic barrier in an asymmetric SN2 reaction? To explore this, we present a systematic computational study using the MP2-SMD(THF)/cc-pVTZ level of theory, evaluating how
interactions influence the intrinsic activation barrier in reactions of the type Y− + CH3–Cl. It is worth noting that energy differences were also evaluated using the CCSD(T)-SMD(THF) method. Intrinsic barriers were extracted from the theoretical activation barrier using Marcus' theory to isolate the required energy for the nuclear reorganization free from thermodynamic bias. The donor–acceptor interactions were quantified through natural bond orbital (NBO) analysis; moreover, quantum theory of atoms in molecules (QTAIM) descriptors provided a complementary, orbital-independent perspective. In the studied systems, a strong correlation between the
stabilization energies and the intrinsic activation barriers was observed, one that is not evident when considering apparent barriers. This distinction underscores, within the present framework, the importance of isolating intrinsic contributions when evaluating fundamental reactivity trends. QTAIM descriptors, such as the electron density at the bond critical point (ρBCP) and the |VBCP|/GBCP ratio, captured aspects of the local electronic environment, particularly electronegativity and polarizability, that were consistent with the intrinsic reactivity observed across the specific nucleophile families examined. In systems bearing α-substituents, the presence of secondary stabilizing interactions, likely involving non-covalent contacts between the nucleophile and C–H bonds of the electrophile, may contribute to lowering the intrinsic barrier. Together, these findings demonstrate that both NBO and QTAIM analyses are robust tools to determine the electronic contributions that govern reactivity and selectivity in the analyzed SN2 reactions. Furthermore, they position hyperconjugation not merely as a passive stabilizing effect but also as an active modulator capable of tuning intrinsic reactivity in organic reactions.
Beyond its structural role, hyperconjugation raises fundamental questions about its influence in chemical reactivity and selectivity, particularly through its potential to modulate the energy barriers that govern molecular transformations, a phenomenon supported by various mechanistic studies across distinct reaction types.14–16 These questions naturally lead to analyzing the relationship between kinetic and thermodynamic parameters in chemical reactions, an essential aspect of physical organic chemistry. To address this issue, several theoretical models have been proposed to decompose the activation energy, also known as observable or apparent activation energy, (ΔE‡), into conceptually distinct components: a thermodynamic term dependent on the reaction energy (ΔE), defined as the energy difference between reactants and products; and an intrinsic barrier (ΔE‡0), which reflects the energetic cost (associated with nuclear reorganization during the chemical transformation) of an idealized isoenergetic transformation (ΔE = 0), and consequently depends on the chemical nature of the interacting species. Within this framework, models such as the Bell–Evans–Polanyi model17–19 and its extension by Laffler20 describe the activation energy as a linear function of the reaction energy, ΔE‡ = ΔE‡0 + αΔE, where ΔE‡0 (intercept) and α (Brønsted slope) are treated as intrinsic properties characteristic of each reaction family. However, when ΔE varies significantly, such linear approximations become inadequate, and more general formulations, such as Marcus' theory, are required. Marcus' theory introduces a quadratic dependence and systematically separates thermodynamic effects from purely structural and, therefore, electronic factors.21 Originally developed to describe classical electron transfer processes, this framework has proven particularly effective in rationalizing a wide range of organic reactions where thermodynamic and kinetic factors are tightly coupled.22–24
In this context, it is worth considering how certain intrinsic electronic factors, particularly those related to the local organization of electron density, may contribute to the intrinsic barrier itself. Among these contributions, hyperconjugation emerges as an outstanding candidate whose influence may extend beyond the thermodynamic stabilization of reactants or products to actively modulate the energy profile of the reaction (Fig. 1c).25–29 If such influence on the intrinsic barrier is confirmed, hyperconjugation would cease to be viewed merely as a passive stabilizing factor and instead be recognized as a determinant of reactivity and selectivity in organic systems.
As a model system to investigate the influence of electronic interactions on molecular reactivity, this work presents a theoretical study of the bimolecular nucleophilic substitution (SN2) reaction. Beyond its classical role as a mechanistic paradigm in physical organic chemistry, the SN2 reaction remains a subject of ongoing research due to its synthetic utility and the emergence of new mechanistic questions and conceptual approaches.30–36 Its well-defined mechanistic framework allows for a meaningful dissection of both thermodynamic and intrinsic contributions to the activation barrier, making it a powerful platform for probing fundamental reactivity principles. This process makes it an ideal model to evaluate how specific electronic interactions, particularly hyperconjugation of the type n → σ*, can influence these distinct energetic components (thermodynamic and intrinsic barriers). We hypothesize that such interactions not only stabilize the involved stationary states, but also lower the intrinsic barrier itself, thereby acting as a key electronic determinant of reactivity. Our theoretical modeling considers the interactions between a series of 11 nucleophiles with chloromethane as a model electrophile to test this hypothesis (Fig. 1d). Analysis of these systems will allow us to establish a relationship between the strength of hyperconjugative interactions and intrinsic barriers, offering quantitative insight into the direct impact of hyperconjugation on reactivity. Despite the numerous elegantly designed quantum chemical studies that have furthered our understanding of the SN2 cornerstone reaction,37–39 to the best of our knowledge, no systematic analysis has been conducted to quantify the influence of specific electronic interactions, such as n → σ* hyperconjugation, on the intrinsic barrier ΔE‡0 in reactions of the type Y− + CH3–Cl. Most studies have focused on identity reactions, X− + CH3–X,40–42 which limits the ability to assess and highlight the role of thermodynamic contributions in modulating reactivity. The knowledge generated in this work will not only offer tools for a deeper understanding of chemical transformations but will also position hyperconjugation as an additional electronic degree of freedom for the rational design of more reactive substrates and the precise modulation of reactivity and selectivity in organic transformations.
The direct influence of thermodynamic factors on activation barriers can be evaluated through Marcus' theory,61–65 which decomposes the activation energy (ΔE‡) into two components: an intrinsic activation barrier (Δ0E‡) and a thermodynamic contribution (ΔE). The intrinsic barrier represents the minimum energy required to reach the transition state under thermoneutral conditions, reflecting the total cost of nuclear reorganization. The thermodynamic term reflects the driving force based on the relative stability between reactants and products. Within this framework, the activation energy increases when the reaction energy (ΔE) is positive and decreases when ΔE is negative. Assuming that the potential energy surfaces of reactants and products can be approximated by parabolas, the activation energy is given using the following classical Marcus expression:61,66
![]() | (1) |
![]() | (2) |
Two complementary approaches were employed to reveal key electronic aspects of the reaction to interpret intrinsic barriers: natural bond orbital (NBO) analysis67–75 and topological analysis of the electron density based on the quantum theory of atoms in molecules (QTAIM).76
NBO analysis has emerged as a valuable theoretical framework for rationalizing reactivity trends in SN2 reactions, particularly in cases where activation barriers deviate from classical expected patterns based on nucleophilicity, leaving group ability, or steric effects associated with substituents on the electrophilic carbon.77–80 This analysis provides a quantitative estimate of the electronic stabilization arising from donor–acceptor interactions, particularly those involving the transfer of electron density from a filled (donor) orbital to an empty or antibonding (acceptor) orbital. This stabilization is estimated using second-order perturbation theory:
![]() | (3) |
QTAIM analysis, on the other hand, relies on the topological characterization of the electron density ρ(), identifying critical points (CPs) where its gradient vanishes (∇ρ(
) = 0), and bond paths (BPs), which represent trajectories of maximum electron density connecting pairs of nuclei. CPs are classified by their topological index ranges and curvature (ω, σ); of particular interest in this study are bond critical points (BCPs), characterized as (3, −1), which indicate interactions between two nuclei, such as those between the nucleophile and the electrophile. The sign of the Laplacian of the electron density at the BCP, ∇2ρ(
), determines whether the interaction is shared-shell (covalent) if negative, or closed-shell (non-covalent) if positive. In addition, Espinosa et al.81 proposed that |VBCP|/GBCP, the ratio between the magnitude of the potential energy density and the kinetic energy density at the BCP, acts as a local energetic descriptor of the stabilization/delocalization balance. According to this criterion, an interaction is closed-shell if |VBCP|/GBCP < 1, partially covalent if 1 < |VBCP|/GBCP < 2, and shared-shell if |VBCP|/GBCP > 2. This parameter thus identifies the threshold beyond which an interaction becomes energetically stabilizing, associated with local density accumulation and the possible bond formation. Furthermore, this method offers an orbital-independent description based on the real-space charge distribution, enabling the inference of the physical nature of the interactions involved in SN2 reactions.33
Together, these descriptors offer a robust basis for correlating electronic and orbital contributions with intrinsic activation barriers and understanding the nature of the nucleophilic interaction from both energetic and structural perspectives. NBO analysis was performed using the Gaussian16 implementation (NBO v3), while QTAIM calculations were carried out with the GPUAM software.82–84
Before analyzing activation barriers and reaction energies, the relative stability of the intermediate complexes, specifically the reactant and product complexes, was examined based on their relative electronic energies (Table S1). In the halide nucleophile series, the RC stability followed the order F− (−4.7 kcal mol−1) > Br− (−2.7 kcal mol−1) > Cl− (7.2 kcal mol−1), a trend that remained consistent in the product complexes: F− (−28.2 kcal mol−1) > Br− (−4.4 kcal mol−1) > Cl− (7.2 kcal mol−1). For Cl−, both RC and PC stabilities are the same because this process corresponds to an identity reaction. For simple nucleophiles (H2N−, HO−, and HS−) and their substituted counterparts (MeHN−, MeO−, and MeS−), the oxygen-containing RCs were more stable than their nitrogen- and sulfur-based analogues. In contrast, the trend reversed for the PCs: nitrogenated derivatives were the most stable, followed by the oxygenated and sulfur-based ones. Regarding nucleophiles with α-substituents, HOO− was more stable (−7.6 kcal mol−1) than HSHN− (−4.4 kcal mol−1) in the RC, although both displayed similar stability in the PC (−58.5 kcal mol−1 and −58.9 kcal mol−1, respectively). These findings reveal distinct stabilization patterns in both RC and PC, strongly dependent on the electronic nature of the nucleophile. Notably, the product complexes are more stable in all cases than the reactant complexes, resulting in negative reaction energies (Table 1), both in terms of electronic energy and Gibbs free energy, thereby characterizing the overall process as exothermic and exergonic. This observation aligns with the increased strength of the bond formed after substitution, except for Br− as a nucleophile, suggesting that in this set of SN2 reactions, thermodynamic factors may contribute significantly to lowering the activation barrier. Such a driving force could potentially contribute to lowering the activation barrier.
Nu | MP2a | MP2a | CCSD(T)b | HFc | ||||
---|---|---|---|---|---|---|---|---|
Δε‡(1) | Δε(1) | ΔG‡(1) | ΔG(1) | Δε‡(2) | Δε(2) | Δε‡(3) | Δε(3) | |
a MP2-SMD(THF)/cc-pVTZ.b CCSD(T)-SMD(THF)/cc-pVTZ//MP2-SMD(THF)/cc-pVTZ.c HF-SMD(THF)/cc-pVTZ//MP2-SMD(THF)/cc-pVTZ. | ||||||||
Halide nucleophiles | ||||||||
F− | 10.9 | −23.5 | 13.2 | −21.7 | 9.4 | −25.0 | 12.8 | −30.2 |
Cl− | 20.6 | 0.0 | 21.8 | 0.3 | 19.4 | 0.0 | 23.0 | 0.0 |
Br− | 17.9 | −1.8 | 19.2 | −1.6 | 16.9 | −1.5 | 21.2 | 0.0 |
Simple nucleophiles | ||||||||
H2N− | 5.9 | −68.4 | 7.2 | −64.9 | 5.1 | −68.2 | 9.8 | −77.2 |
HO− | 6.4 | −54.3 | 7.2 | −52.2 | 5.3 | −55.3 | 5.6 | −67.8 |
HS− | 14.4 | −23.9 | 16.3 | −22.4 | 13.7 | −23.3 | 18.5 | −26.2 |
Substituted nucleophiles | ||||||||
CH3HN− | 3.9 | −72.9 | 5.8 | −68.9 | 3.4 | −72.4 | 9.0 | −82.3 |
CH3O− | 7.1 | −50.0 | 6.3 | −46.2 | 6.3 | −50.5 | 7.8 | −63.2 |
CH3S− | 10.8 | −34.3 | 11.9 | −32.8 | 10.4 | −33.4 | 15.8 | −37.8 |
Nucleophiles with α-substituents | ||||||||
HOO− | 4.1 | −50.9 | 4.6 | −49.7 | 3.2 | −50.3 | 6.9 | −56.7 |
HSHN− | 7.6 | −54.5 | 10.6 | −50.1 | 7.2 | −54.2 | 11.5 | −63.1 |
The quantitative analysis of activation barriers shows significant trends that deserve closer examination (see Table 1). In the halogenated nucleophile series, an unusual ascending order was observed: F− (10.9 kcal mol−1) < Br− (17.9 kcal mol−1) < Cl− (20.6 kcal mol−1). This pattern, observed in this specific halide series, does not align with expectations based on periodic trends like nucleophile electronegativity or polarizability. Among oxygen-, nitrogen-, and sulfur-based nucleophiles, both simple and substituted, the following patterns were identified: H2N− (5.9 kcal mol−1) < HO− (6.4 kcal mol−1) < HS− (14.4 kcal mol−1), and MeNH− (3.9 kcal mol−1) < MeO− (7.1 kcal mol−1) < MeS− (10.8 kcal mol−1). Again, within the nucleophile subsets analyzed here, these trends do not adhere to predictable patterns based solely on periodic properties. In contrast, nucleophiles with α-substituents exhibited divergent behavior relative to their parent nucleophiles: HOO− showed a lower barrier (4.1 kcal mol−1) than both the simple and substituted oxygen nucleophiles, whereas HSHN− exhibited a higher barrier (7.6 kcal mol−1) than its nitrogen-based analogues. This variation suggests that, for the studied systems, the presence of α-substituents can significantly alter the relative reactivity of the nucleophilic center,85 in line with previous reports on α-substituent effects. Despite the barriers calculated under the continuum approximation coinciding with the experimental trends observed for certain nucleophiles,86,87 the lack of a systematic correlation in the activation barriers, which remains consistent even after single-point refinement at the CCSD(T) level (Table 1), indicates that, within the present set of nucleophiles, the atomic identity of the nucleophile does not solely determine reactivity but is likely strongly modulated by global thermodynamic factors, as reflected in the reaction energies, calculated both as electronic and Gibbs free energies, which vary significantly across the nucleophile series (see Table 1).
Fig. 3 shows the relationship between the activation barriers and the reaction energies for the series of nucleophiles studied, clearly revealing an approximately linear trend. Within the studied series of nucleophiles, exothermic and exergonic reactions tend to exhibit lower activation barriers, whereas less thermodynamically favorable processes present higher ones; moreover, the degree of exothermicity varies systematically with the electronic nature of the nucleophile. A linear fit based on the model proposed by Laffler yielded the characteristic parameters for this (Δε‡0 and ΔG‡0) and the Brønsted slope (α). These results indicate that, for the analyzed chloromethane-based SN2 reactions, thermodynamic factors significantly influence the observed reactivity, as evidenced by the strong linear correlation between barriers and reaction energies (R2 ≈ 0.90) observed within this specific dataset.
A logical strategy to isolate the effect of the hyperconjugative interaction would involve repeating this analysis while explicitly turning off such donor–acceptor interactions, thereby evaluating their impact on Δε‡0, ΔG‡0, and α intrinsic parameters. However, the methodological tools required to selectively deactivate electronic interactions are only functionally available within the HF framework, which lacks the electronic correlation effects essential for accurately describing anionic systems such as those investigated here. To further support this limitation, we estimated activation barriers and reaction energies at the HF level (Table 1), which yielded significantly overestimated values compared to correlated methods. For this reason, we adopted the quadratic Marcus' model, as it offers a more general and physically consistent framework for estimating intrinsic barriers, allowing the measure of reactivity free from thermodynamic bias. This approach ensures that the derived trends reflect intrinsic electronic and structural effects rather than artifacts introduced by insufficient correlation treatment.
Nu | Δ0ε‡ | Δcorra | Δ0G‡ | Δcorrb |
---|---|---|---|---|
a Δcorr = Δε‡0 − Δε‡.b Δcorr = ΔG‡0 − ΔG‡. | ||||
Halide nucleophiles | ||||
F− | 21.0 | 10.1 | 22.8 | 9.6 |
Cl− | 20.6 | 0.0 | 21.7 | −0.1 |
Br− | 18.8 | 0.9 | 20.0 | 0.8 |
Simple nucleophiles | ||||
H2N− | 30.5 | 24.6 | 31.2 | 24.0 |
HO− | 26.6 | 20.2 | 27.0 | 19.8 |
HS− | 24.9 | 10.5 | 26.3 | 10.0 |
Substituted nucleophiles | ||||
CH3HN− | 28.9 | 25.0 | 30.6 | 24.7 |
CH3O− | 26.1 | 19.0 | 23.8 | 17.5 |
CH3S− | 25.0 | 14.2 | 25.7 | 13.8 |
Nucleophiles with α-substituents | ||||
HOO− | 22.3 | 18.2 | 22.7 | 18.0 |
HSHN− | 28.2 | 20.7 | 30.5 | 19.9 |
The intrinsic barriers obtained revealed coherent trends within the analyzed nucleophile families, emphasizing how their electronic character influences the inherent reactivity in the chloromethane-based SN2 context. In the halide series, a consistent trend was noted for both electronic energies and Gibbs free energies: Br− (18.8 kcal mol−1 for electronic energy and 20.0 kcal mol−1 for Gibbs free energy) < Cl− (20.6 kcal mol−1 and 21.7 kcal mol−1) < F− (21.0 kcal mol−1 and 22.8 kcal mol−1). This behavior contrasts with that observed for the apparent barriers but, across this subset, appears to align with periodic trends such as electronegativity and polarizability, suggesting a consistent electronic contribution. For simple nucleophiles derived from nitrogen, oxygen, and sulfur, a descending trend was identified down the group: H2N− (30.5 kcal mol−1) > HO− (26.6 kcal mol−1) > HS− (24.9 kcal mol−1) when using electronic energies, and H2N− (31.2 kcal mol−1) > HO− (27.0 kcal mol−1) > HS− (26.3 kcal mol−1) when using Gibbs free energies. This suggests that periodic properties may influence the magnitude of the intrinsic barrier. A similar trend was observed in the substituted analogues, where the values were 28.9, 26.1, and 25.0 kcal mol−1 for MeNH−, MeO−, and MeS−, respectively, when considering electronic energy. However, when Gibbs free energies were used, the relative order changed to MeNH− (30.6 kcal mol−1) > MeS− (25.7 kcal mol−1) > MeO− (23.8 kcal mol−1).
Finally, nucleophiles bearing α-substituents preserved the reactivity order previously observed: HOO− showed intrinsic barriers of 22.3 kcal mol−1 (electronic) and 22.7 kcal mol−1 (Gibbs), while HSHN− exhibited higher values of 28.2 and 30.5 kcal mol−1. This difference supports the conclusion that, in the context of the chloromethane-based SN2 reactions studied, HOO− is more reactive than HSHN−, both in terms of apparent and intrinsic barriers. Moreover, both nucleophiles displayed lower intrinsic barriers than their simple and substituted analogues, suggesting that, for the considered nucleophiles, the electronic effects associated with α-substituents contribute favorably to the activation process.
Altogether, these findings indicate that, in the present set of SN2 reactions, thermodynamic corrections not only change the absolute values of activation barriers but can also affect the relative reactivity order among different nucleophiles. The shift in trends observed before and after deconvolution highlights the importance of distinguishing between intrinsic effects and thermodynamic contributions when interpreting activation profiles. Access to intrinsic barriers offers a direct measure of the energy cost associated with reaching the transition state and consistently reflects the electronic nature of the nucleophilic center, representing a valuable step toward understanding the electronic factors that govern reactivity in SN2 reactions involving nucleophiles with diverse electronic profiles.
Reactant complex (RC) | Transition state (TS) | |||||||
---|---|---|---|---|---|---|---|---|
dNu–C | dC–Cl | θ | ϕ | dNu–C | dC–Cl | θ | ϕ | |
The geometric parameters of isolated chloromethane calculated at the same level of theory are dC–Cl = 1.786 Å and ϕ = 33.5°. | ||||||||
Halide nucleophiles | ||||||||
F− | 3.337 | 1.791 | 179.9 | 33.9 | 1.956 | 2.138 | 179.8 | 7.6 |
Cl− | 3.569 | 1.794 | 179.6 | 33.5 | 2.279 | 2.279 | 179.9 | 0.0 |
Br− | 3.474 | 1.800 | 179.8 | 33.6 | 2.424 | 2.272 | 179.9 | 1.3 |
Simple nucleophiles | ||||||||
H2N− | 3.376 | 1.798 | 177.8 | 33.5 | 2.376 | 2.027 | 179.9 | 19.1 |
HO− | 2.853 | 1.806 | 115.5 | 32.4 | 2.229 | 2.010 | 179.8 | 20.2 |
HS− | 3.733 | 1.794 | 177.9 | 33.6 | 2.474 | 2.182 | 179.9 | 8.1 |
Substituted nucleophiles | ||||||||
CH3HN− | 3.230 | 1.800 | 171.2 | 33.2 | 2.369 | 2.001 | 175.8 | 20.0 |
CH3O− | 2.923 | 1.800 | 112.4 | 32.4 | 2.150 | 2.028 | 179.2 | 16.8 |
CH3S− | 3.512 | 1.799 | 168.3 | 33.4 | 2.502 | 2.134 | 176.3 | 11.3 |
Nucleophiles with α-substituents | ||||||||
HOO− | 2.836 | 1.812 | 158.6 | 33.6 | 2.157 | 2.027 | 176.9 | 17.7 |
HSHN− | 3.248 | 1.798 | 177.4 | 33.4 | 2.248 | 2.063 | 177.6 | 15.0 |
In the analyzed systems, the structural analysis of reactant complexes typically reveals a spatial arrangement consistent with an SN2 mechanism. In most cases, the Nu–C–Cl angle approaches 180°, which corresponds to the expected colinear orientation for bimolecular nucleophilic substitution. However, in systems with oxygen-based nucleophiles such as HO−, MeO−, and HOO−, a significant deviation from this ideal geometry is observed (the optimized geometries are shown in Fig. S1). For HO− and MeO−, the unexpected configuration can be attributed to the formation of a C–H⋯O hydrogen bond (1.733 Å and 1.818 Å for HO− and MeO−, respectively). This interaction is hypothesized based on the optimized geometry adopted by the nucleophile relative to the electrophile in these systems. In the case of HOO−, although no specific interaction of this kind can be assumed due to its lower basicity and H acceptor capacity, the angular deviation may reflect the presence of an alternative stabilizing effect. In all analyzed cases, the reactant complexes exhibit incipient interaction between the nucleophile and the electrophilic carbon, as evidenced by moderately shortened Nu–C distances (ranging from 2.836 to 3.733 Å) and elongation of the C–Cl bond relative to free chloromethane (see note in Table 3), indicating early polarization of the reactive center. This variability in dNu–C depends not only on the size of the nucleophile but also on its electronic nature and ability to establish attractive interactions. Introducing a methyl group into O-, S-, and N-based nucleophiles leads to a systematic shortening of the Nu–C distance compared to their unsubstituted analogues, which may suggest enhanced nucleophilic efficiency in the context of the studied systems. Among the systems analyzed, HOO− exhibits the shortest Nu–C distance and the greatest C–Cl elongation, which may reflect a more favorable structural preactivation in terms of geometry. Finally, analysis of the improper dihedral angle produced values comparable to free chloromethane. This indicates that, in the reactant complexes examined, a significant geometric reorganization toward a pseudoplanar configuration has not yet occurred at this stage.
Upon reaching the transition state, the optimized structures exhibit substantial geometric changes in the above mentioned parameters. The Nu–C distance shortens significantly, with values ranging from 1.956 to 2.424 Å, indicating considerable progress in forming the new bond. Complementarily, the C–Cl distance increases relative to the reactant complex, exceeding 2.001 Å, consistent with the progressive rupture of the leaving bond. These observations are consistent with a concerted process as expected for the SN2 mechanism in the studied systems. Comparing the bond lengths of Nu–C and C–Cl shows significant asynchronicity in most systems, except for the Cl− nucleophile. In all the analyzed cases, the Nu–C–Cl angle converges toward values near 180°, reflecting a colinear arrangement in the transition region. Similarly, the improper dihedral angle decreases to values below 20.2°, indicating a transition from tetrahedral geometry to a pseudoplanar configuration. This reduction in tetrahedral character aligns with the expected geometry of the transition state, and the resulting degree of planarity varies depending on the type of nucleophile involved which is consistent with the reaction's exothermicity. As postulated by Hammond's principle, more exothermic reactions tend to exhibit transition states that resemble the reactants more closely.88 This trend is observed in the analyzed systems, where increased exothermicity correlates with a lower degree of planarity in the SN2 transition state.
Comparing the structural parameters of the reactant complex and the transition state, in these systems, illustrates the molecular rearrangement associated with the SN2 mechanism. The lack of significant distortion at the electrophilic center in the reactant complex, as indicated by the improper dihedral angle values, suggests that structural reorganization primarily begins when approaching the transition state. Nevertheless, initial changes in parameters such as the Nu–C distance and the attack angle point to the presence of early interactions, which may help facilitate reaching the transition state. In this context, characterizing the electronic nature of these interactions may provide valuable insight into the origin of the trends previously observed in the intrinsic barriers within this set of SN2 reactions.
![]() | ||
Fig. 4 Representative IRC profiles for (a) identity and (b) asymmetric SN2 reactions: electronic energy (ε) and Laplacian curvature (|∇ε|RMS) highlighting the pre-TS point. |
a In some cases, NBO analysis identified two ![]() |
---|
![]() |
a Not determined (n.d.) due to the absence of BCP(S–C) in the QTAIM analysis. |
---|
![]() |
a Not determined (n.d.) due to the absence of BCP(O–C) in the QTAIM analysis. |
---|
![]() |
In summary, the combined NBO-QTAIM analysis for the studied systems reveals the presence of early donor–acceptor interactions in the reactant complexes, which become significantly stronger near the transition state geometry. Fig. 5 illustrates the intrinsic activation barriers alongside the hyperconjugative stabilization energies computed at the pre-transition state geometries. Within the studied reaction set, an inverse trend is observed: systems exhibiting higher activation barriers (red and blue bars) tend to show lower values (green bars), whereas lower barriers are consistently associated with stronger hyperconjugative interactions. This relationship highlights, in the context of the modeled SN2 reactions, the important role of the nucleophile's capacity to modulate the interaction strength with the electrophilic center, thereby compensating for the energetic demands of the structural reorganization occurring at the transition state. In this study, the topological parameters ρBCP and |VBCP|/GBCP further support these findings by providing a complementary perspective on the nature of bonding: ρBCP reflects local electron density concentration, while the |VBCP|/GBCP ratio indicates the degree of electron sharing, highlighting how both electronegativity and polarizability shape the interaction profile.
In the present set of modeled SN2 reactions, thermodynamic factors appear to play a dominant role in determining the apparent activation barriers. This is supported by the strong correlation observed between barriers and reaction energies, a relationship evident in both electronic and Gibbs free energies. Within this dataset, the strength of the hyperconjugative interaction, represented by
, does not correlate with the stability of the reactant complex, suggesting a limited contribution to its thermodynamic stabilization. This may be attributed to the fact that the overall stabilization energy, reflected in electronic and Gibbs free energies, is greater in magnitude than the hyperconjugative contribution alone, implying that other factors govern the complex's stability. Instead, the aforementioned interaction consistently correlates with the intrinsic barriers derived from Marcus' theory. This finding suggests that, in the studied systems, hyperconjugation fundamentally affects intrinsic reactivity by damping the energetic requirements needed to reach the transition state. Furthermore, the observed geometric and energetic trends across different nucleophile families appear to align with periodic properties, such as electronegativity and polarizability, as effectively illustrated by NBO and QTAIM analyses. The examination of reactive complexes and pre-transition states reveals that early donor–acceptor interactions reach their maximum as the system approaches the transition state. The above observation reinforces the idea that, in these reactions, intrinsic electronic organization, rather than mere thermodynamic stabilization, contributes to the barrier height and, consequently, the resulting reactivity trends.
To the best of our knowledge, no systematic analysis has previously quantified the impact of specific electronic interactions, such as hyperconjugation, on the intrinsic activation barrier in non-identity SN2 reactions. The results presented here suggest that, rather than merely providing passive stabilization, such interactions can serve as crucial electronic elements that fine-tune reactivity in these systems. Overall, our findings enhance the fundamental understanding of the factors influencing nucleophilic efficiency and highlighting the potential of hyperconjugation as an additional structural degree of freedom that could be utilized to modulate substrate reactivity and in predicting trends in regioselectivity. This approach opens new ways for developing more efficient and controllable synthetic strategies in organic chemistry.
This journal is © the Owner Societies 2025 |