We have quantum chemically analyzed the trends in bond dissociation enthalpy (BDE) of H3C–XHn single bonds (XHn = CH3, NH2, OH, F, Cl, Br, I) along three different dissociation pathways at ZORA-BLYP-D3(BJ)/TZ2P: (i) homolytic dissociation into H3C + XHn, (ii) heterolytic dissociation into H3C+ + XHn, and (iii) heterolytic dissociation into H3C + +XHn. The associated BDEs for the three pathways differ not only quantitatively but, in some cases, also in terms of opposite trends along the C–X series. Based on activation strain analyses and quantitative molecular orbital theory, we explain how these differences are caused by the profoundly different electronic structures of, and thus bonding mechanisms between, the resulting fragments in the three different dissociation pathways. We demonstrate that the nature and strength of a chemical bond are only fully defined when considering both (i) the molecule in which the bond exists and (ii) the fragments from which it forms or into which it dissociates.

The chemical bond is a cornerstone of chemistry.1 Its length, stability, and electronic structure play a decisive role in determining molecular structure, stability, and reactivity. Therefore, a thorough understanding of trends in chemical bond strengths is quintessential for advancing chemical theory and the rational design of novel molecular systems.2,3 One notable trend is that polar covalent bonds are stronger when the polarity of the bond increases. Until recently, this phenomenon was ascribed to the increasing difference in electronegativity between the elements that constitute this polar covalent bond.4 For example, upon going down group 17, from H3C–F to H3C–I, the chemical bond weakens, which is generally explained by the decreasing electronegativity difference between carbon and the consistently heavier halogen.1 

Recently, we have shown, by studying the homolytic bond dissociation of various polar covalent bonds, that the correlation between polar covalent bond strength and differences in electronegativity of the elements involved in bonding is not in all situations causal.5 Indeed, along period 2, from H3C–CH3 to H3C–F, the chemical bond strengthens because the electron-pair bonding orbital interactions become more stabilizing as the electronegativity difference across the bond is larger. However, going up group 17, from H3C–I to H3C–F, the carbon–halogen bond strengthens due to decreasing carbon–halogen Pauli repulsion and not due to the commonly assumed strengthening in the carbon–halogen electron-pair bonding orbital interaction. Interestingly, along this series, the bond, in fact, becomes weaker if one compares the bonds at consistent bond distances. The decreasing carbon–halogen Pauli repulsion going up group 17 arises from the smaller effective size of the lighter halogen atoms. In addition to stabilizing the carbon–halogen bond, the reduction in Pauli repulsion also allows it to assume a shorter equilibrium bond distance. At that shorter equilibrium bond distance, all interactions, including the electron-pair bonding orbital interactions, are stronger. This leads to a correlation between the carbon–halogen electron-pair bond strength and the decreasing electronegativity difference, which is erroneously invoked as a causal correlation in most textbooks.1(b),4

Herein, we extend our previous work on the C–X bond strength in H3C–XHn from homolytic bond dissociation into H3C + XHn to two alternative, heterolytic bond dissociation pathways, namely, (i) into H3C+ + XHn (“hetero 1” in Scheme 1) and (ii) into H3C + +XHn (“hetero 2” in Scheme 1). Our model systems cover the variation of XHn along period 2 (XHn = CH3, NH2, OH, F) and going up group 17 (XHn = I, Br, Cl, F). All dissociation pathways were explored using the same level of relativistic, dispersion-corrected density functional theory at ZORA-BLYP-D3(BJ)/TZ2P, including the homolytic pathway to guarantee a set of consistent data.

SCHEME 1.

Homolytic and heterolytic bond dissociation pathways studied in this work.

SCHEME 1.

Homolytic and heterolytic bond dissociation pathways studied in this work.

Close modal

Our purpose is to understand how H3C–XHn bond strengths, and trends therein, depend on the actual bond dissociation pathway.6 Interestingly, we find that the H3C–XHn bond dissociation enthalpies (BDE) differ pronouncedly for the three pathways, not only quantitatively but, in some cases, also in terms of opposite trends upon varying XHn. Our analyses using the activation strain model (ASM)7 in conjunction with quantitative Kohn–Sham molecular orbital theory (KS-MO)8 and a matching energy decomposition analysis (EDA)9 reveal how these differences are caused by the profoundly different electronic structures of, and thus bonding mechanisms between, the resulting fragments in the three different dissociation pathways.3(c),5 Our findings illustrate that it is essential for fully understanding a chemical bond to consider both (i) the molecule in which the bond exists and (ii) the fragments from which it forms or into which it dissociates.

All calculations were carried out using the Amsterdam Density Functional (ADF) program.10 The geometries and energies were calculated at the BLYP level of the generalized gradient approximation (GGA): the exchange functional developed by Becke (B) and the GGA correlation functional developed by Lee, Yang, and Parr (LYP).11 The DFT-D3(BJ) method developed by Grimme and co-workers,12 which contains the damping function proposed by Becke and Johnson,13 is used to describe non-local dispersion interactions. Scalar relativistic effects are accounted for using the zeroth-order regular approximation (ZORA).14 A large, uncontracted, optimized TZ2P Slater-type orbitals (STOs) basis set containing diffuse functions was used. The TZ2P all-electron basis set,15 with no frozen-core approximation, is of triple-ζ quality for all atoms and has been augmented with two sets of polarization functions. This level is referred to as ZORA-BLYP-D3(BJ)/TZ2P and provides bond dissociation enthalpies that are in close agreement with the experimentally determined values (vide infra). The accuracies of the integration grid (Becke grid) and the fit scheme (Zlm fit) were set to VERYGOOD.16 The potential energy surfaces are obtained by performing a relaxed potential energy surface scan, whereby the H3C–XHn bond distance is elongated from 1.10 to 2.70 Å in 33 equidistant steps. The computed potential energy surfaces were analyzed using the PyFrag 2019 program.17 Optimized structures were illustrated using CYLview.18 

Enthalpies at 298.15 K and 1 atm (∆H298) were calculated from electronic bond energies (∆E) and vibrational frequencies using standard thermochemistry relations for an ideal gas, according to the following equation:19 
(1)
Here, ΔEtrans,298, ΔErot,298, and ΔEvib,0 are the differences between the reactants and products in translational, rotational, and zero-point vibrational energy, respectively. Δ(ΔEvib,0)298 is the change in the vibrational energy difference as one goes from 0 to 298.15 K. The vibrational energy corrections are identical to our frequency calculations. The molar work term Δ(pV) is (Δn)RT; Δn = +1 for one reactant dissociating into the two products. Thermal corrections for the electronic energy are neglected.
The activation strain model (ASM)7 is a fragment-based approach in which the electronic bond energy, ∆E, is decomposed into two components,
(2)
In this equation, the strain energy, ∆Estrain, is the penalty that needs to be paid in order to deform the fragments from their equilibrium structure to the geometry they acquire in the overall system. The interaction energy, ∆Eint, accounts for all mutual chemical interactions that occur between the geometrically deformed fragments in the overall system.
The interaction energy between the deformed reactants is further analyzed by means of the canonical energy decomposition analysis (EDA) within the framework of the Kohn–Sham molecular orbital (KS-MO) theory.8,9 The EDA decomposes the ∆Eint into the following four physically meaningful energy terms:
(3)
The term ∆Velstat corresponds to the quasi-classical electrostatic interaction between the unperturbed charge distributions of the fragments in the geometry they possess in the complex. The Pauli repulsion, ∆EPauli, between these fragments comprises the destabilizing interactions associated with Pauli’s principle for fermions between occupied orbitals and is an indicator of steric repulsion. The orbital interactions, ∆Eoi, between the fragments account for electron-pair bonding [formation of bonding combinations between the singly occupied molecular orbitals (SOMOs) of either fragment], charge transfer, i.e., donor–acceptor interactions (mixing of occupied and unoccupied orbitals between different fragments), and polarization (mixing of occupied and unoccupied orbitals on one fragment due to the presence of another fragment). Non-local dispersion interactions ∆Edisp are described using the aforementioned D3(BJ) correction.12 

The activation strain and energy decomposition analyses, as well as the Kohn–Sham molecular orbital analysis, were performed on all 34 points acquired from the relaxed potential energy surface scan as a function of the HnC–XHn bond distance that defines the bond formation or dissociation process (vide supra).

Table I shows the homolytic and heterolytic bond dissociation enthalpies (BDE or ΔHBDE) of the H3C–XHn bonds for XHn = CH3, NH2, OH, F, Cl, Br, and I, along the three dissociation pathways defined in Scheme 1. Three key observations can be made: First, not surprisingly, the H3C–XHn BDE increases from homolytic dissociation (H3C + XHn) to heterolytic dissociation with the negative charge on the more electronegative fragment (H3C+ + XHn) to heterolytic dissociation with the negative charge on the less electronegative fragment (H3C + +XHn). For example, the H3C–XHn BDE increases from 111.3 to 263.8 to 605.1 kcal mol1 when dissociating into H3C + F, H3C+ + F, and H3C + F+, respectively (see Table I). Note that our computed homolytic BDEs, that is, the lowest energy dissociation pathway, are in close agreement with the experimentally determined BDEs. Second, the BDE values along each of the two series of bonds, i.e., along period 2 and up group 17, span a larger range if we go along these three respective dissociation pathways. For example, while the homolytic BDE along C–C to C–F changes by 26 kcal mol1 (from 85.2 to 111.3 kcal mol1), this range stretches over nearly 300 kcal mol1 (from 315.5 to 605.1 kcal mol1) for heterolytic dissociation into H3C + +F. Third, most interestingly, the BDE increases, with one exception, for all three dissociation pathways along period 2, from C–C to C–F, and up group 17, from C–I to C–F. For example, the homolytic BDE from C–C to C–F increases from 85.2 to 111.3 kcal mol1, and from C–I to C–F it strengthens from 61.0 to 111.3 kcal mol1. The exception occurs for the more favorable of the two heterolytic dissociation pathways, that is, H3C–XHn → H3C+ + XHn, and only for the series along period 2, from C–C to C–F, for which the heterolytic BDE decreases, not increases, from 315.5 (H3C+ + CH3) to 263.8 kcal mol1 (H3C+ + F).

TABLE I.

Bond length (rH3C–XHn; in Å) and bond dissociation enthalpies (ΔHBDE; in kcal mol1) of the homolytic and heterolytic dissociation of H3C–XHn.a

H3C + XHnbH3C+ + XHncH3C + +XHncExperimentd
H3C–XHnrH3C–XHnΔHBDEΔHBDEΔHBDEΔHBDE
H3C–CH3 1.538 85.2 315.5 315.5 90.2 ± 0.2 
H3C–NH2 1.479 80.3 299.1 374.7 85.1 ± 0.5 
H3C–OH 1.443 89.2 282.7 464.2 92.2 ± 0.5 
H3C–F 1.414 111.3 263.8 605.1 110.0 ± 0.5 
H3C–Cl 1.821 80.9 229.7 428.0 83.7 ± 0.4 
H3C–Br 1.988 71.2 221.3 386.0 70.3 ± 0.5 
H3C–I 2.189 61.0 215.2 338.6 57.1 ± 0.6 
H3C + XHnbH3C+ + XHncH3C + +XHncExperimentd
H3C–XHnrH3C–XHnΔHBDEΔHBDEΔHBDEΔHBDE
H3C–CH3 1.538 85.2 315.5 315.5 90.2 ± 0.2 
H3C–NH2 1.479 80.3 299.1 374.7 85.1 ± 0.5 
H3C–OH 1.443 89.2 282.7 464.2 92.2 ± 0.5 
H3C–F 1.414 111.3 263.8 605.1 110.0 ± 0.5 
H3C–Cl 1.821 80.9 229.7 428.0 83.7 ± 0.4 
H3C–Br 1.988 71.2 221.3 386.0 70.3 ± 0.5 
H3C–I 2.189 61.0 215.2 338.6 57.1 ± 0.6 
a

Computed at ZORA-BLYP-D3(BJ)/TZ2P.

b

Values taken from Ref. 5.

c

Bond dissociation enthalpies computed at ZORA-BLYP-D3(BJ)/QZ4P can be found in Table S1.

d

Values taken from Ref. 20.

Next, we unveil the physical mechanisms behind the predominant trend that BDEs, in nearly all cases, become stronger from H3C–CH3 to H3C–F and from H3C–I to H3C–F. Furthermore, we uncover the mechanism behind the exception for the heterolytic dissociation of the H3C–XHn bond into H3C + +XHn, where the BDE is weaker from H3C–CH3 to H3C–F.

In this section, we examine the origin of the trends in homolytic BDE (“homo” in Scheme 1), which show a pronounced strengthening along both series, from C–C to C–F and from C–I to C–F. The reasons for these similar trends are, however, quite different for the two series, as has been reported recently.5 On one hand, from C–C to C–F, the bond becomes stronger (with an anomaly from C–C to C–N)21 because of the larger electronegativity difference across the bond, that is, due to stronger electron-pair bond interactions. From a molecular orbital perspective (Scheme 2), this can be ascribed to an increasingly larger SOMO–SOMO orbital energy gap (Δε), which results in a deeper drop in the energy of the radical electron stemming from the higher-energy SOMO and, hence, in stronger electron-pair bonding orbital interaction. On the other hand, from C–I to C–F, a different mechanism is at play because the bond is stronger due to the decreasing steric repulsion with the smaller halogen atom. Upon going up in group 17, the occupied valence orbitals of XHn are spatially less extended and bear fewer electrons in a reduced number of closed-shell AOs. This leads to a decrease in repulsive occupied–occupied orbital overlap (S2), making the C–X bond effectively stronger and shorter. This happens despite a weaker electron-pair bonding orbital interaction caused by a smaller bond overlap along the series.

SCHEME 2.

Schematic representation of the bonding mechanisms and qualitative strengths8(b) along the homolytic and heterolytic dissociation pathways studied in this work. The green arrows indicate the stabilizing electron-pair bonding and donor–acceptor interactions, where the leading term in the energy of the bonding orbital, relative to the FMOs involved, of “homo” and “hetero 2” are Δε + F2/Δε ∼ Δε + C · S2/Δε. The red arrows indicate the destabilizing Pauli’s repulsion. Both the two-center four-electron splitting (double-headed, black/red arrow) and Pauli repulsion (red arrow) are proportional to S2.

SCHEME 2.

Schematic representation of the bonding mechanisms and qualitative strengths8(b) along the homolytic and heterolytic dissociation pathways studied in this work. The green arrows indicate the stabilizing electron-pair bonding and donor–acceptor interactions, where the leading term in the energy of the bonding orbital, relative to the FMOs involved, of “homo” and “hetero 2” are Δε + F2/Δε ∼ Δε + C · S2/Δε. The red arrows indicate the destabilizing Pauli’s repulsion. Both the two-center four-electron splitting (double-headed, black/red arrow) and Pauli repulsion (red arrow) are proportional to S2.

Close modal

In the following sections, we analyze the two heterolytic dissociation pathways and compare the two extremes, along period 2 and going up group 17, that is, H3C–CH3 vs H3C–F and H3C–I vs H3C–F. We can do this due to the fragment-based approach of our analysis methods, i.e., the activation strain model and energy decomposition analysis (see Sec. II C). This approach gives us the freedom to select our fragments, which allows us to accurately analyze how a chemical bond arises along different association pathways (i.e., from different fragments) and, vice versa, how it fades out when it is broken along different dissociation pathways (i.e., into different fragments).22 Table S2 shows that the trends in bond dissociation enthalpies ΔHBDE are set by the electronic bond dissociation energies ΔEBDE, allowing us to further analyze the electronic bond energy ΔEE = −ΔEBDE) of the bond formation process of H3C–XHn between ionic fragments along the bond-forming axis in more detail. The analysis of the complete set of systems along period 2 and down group 17 can be found in Figs. S1–S4 of the supplementary material.

First, we study the heterolytic dissociation pathway of the H3C–CH3 and H3C–F bonds into the methyl cation H3C+ and the respective anionic species CH3 and F (“hetero 1” in Scheme 1), using the activation strain model (ASM) [Fig. 1(a)].7 In contrast to the situation for the homolytic dissociation pathway (“homo” in Scheme 1), we find that, along this heterolytic dissociation pathway (“hetero 1” in Scheme 1), the bond becomes weaker, not stronger, from C–C to C–F because of a less stabilizing (i.e., less negative) interaction energy ΔEint. The strain energy ΔEstrain is not responsible for the overall trend in ∆E. It runs counter to the trend in ΔEint, namely, it is slightly less destabilizing (i.e., less positive) from the stronger C–C to the weaker C–F bond. The reason is that the XHn = CH3 fragment has a small deformation energy upon C–C bond formation (associated with a slight modification in the extent of pyramidalization of the methyl anion fragment), whereas the XHn = F cannot undergo any geometrical deformation. By means of energy decomposition analyses (EDAs),9 we find that the difference in interaction energy predominantly originates from the orbital interactions ∆Eoi, which are significantly less stabilizing if we go from C–C to C–F [Fig. 1(b)]. The electrostatic interaction shows the same trend, although less pronounced so, and is slightly less stabilizing from C–C to C–F. The Pauli repulsion is not responsible for the overall trend in ∆E; it runs counter and is less destabilizing from C–C to C–F.

FIG. 1.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = CH3 and F) into H3C+ and XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C+ and XHn as well as the XHn orbital energies; and (d) contour plots (10 contour lines) of the XHn orbitals, where the black vertical line indicates the maximum spatial extent of the 3A1–CH3 atomic orbital, contour plots from 0.035 to 0.3 au. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

FIG. 1.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = CH3 and F) into H3C+ and XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C+ and XHn as well as the XHn orbital energies; and (d) contour plots (10 contour lines) of the XHn orbitals, where the black vertical line indicates the maximum spatial extent of the 3A1–CH3 atomic orbital, contour plots from 0.035 to 0.3 au. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

Close modal

By performing a Kohn–Sham molecular orbital (KS-MO) analysis as a function of the C–X bond distance,8 we find that it is both the less favorable HOMO–LUMO orbital energy gap (which becomes larger from C–C to C–F) and the less favorable HOMO–LUMO overlap (which becomes smaller from C–C to C–F) that makes the orbital interactions less stabilizing from C–C to C–F. Figure 1(c) shows the progression of the donor orbital energy and the stabilizing donor–acceptor orbital overlap along the bond-forming process between ⟨3A1H3C+|3A1–CH3⟩ and ⟨3A1H3C+|2p–F⟩, where 3A1H3C+ is the accepting 2pσ-derived LUMO of H3C+, 3A1–CH3 is the donating 2pσ-derived HOMO of CH3, and 2p–F is the donating valence atomic 2pσ-HOMO of F. The donating 2p–F orbital is lower in energy than the 3A1–CH3 orbital, leading to a larger and, hence, less stabilizing HOMO–LUMO orbital energy gap for the former. In addition, the orbital overlap is also smaller and, therefore, less favorable for C–F compared to C–C. The difference in orbital overlap can be traced back to the spatial extent of the occupied, donating orbital of XHn. In Fig. 1(d), we have visualized the occupied orbitals of CH3 and F, i.e., 3A1–CH3 and 2p–F, that engage in the donor–acceptor interaction with H3C+. It becomes clear that, going from 3A1–CH3 to 2p–F, the HOMO is more contracted to the nucleus, and hence, extends less toward the incoming H3C+.23 

Next, we examine why the chemical bond is stronger from H3C–I to H3C–F when heterolytically dissociating into H3C+ and anionic species I and F, respectively. Figure 2(a) shows that the bond becomes stronger from C–I to C–F because of a more stabilizing interaction energy ΔEint and a less destabilizing strain energy. The difference in strain energy originates from the deformation of the intrinsically planar H3C+ fragment, which needs to pyrimidalize to accommodate the chemical bond formation with the XHn fragment. The extent of pyramidalization, and hence the magnitude of the strain energy, stems from the effective size of the XHn fragment. A large XHn, such as I, engages in a more steric repulsion with the hydrogen atoms of H3C+ and, therefore, forces this fragment to pyrimidalize to a larger extent than the small F. The more stabilizing interaction energy for C–F, on the other hand, originates from a less steep rise of destabilizing Pauli repulsion compared to C–I [Fig. 2(b)]. The orbital interactions show an opposite trend, namely, they are less stabilizing from the weaker C–I to the stronger C–F bond.

FIG. 2.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = F and I) into H3C+ and XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity. (c) Repulsive occupied–occupied orbital overlaps (S) between H3C+ and XHn; (d) contour plots (10 contour lines) of np–XHn atomic orbitals, where the black vertical line indicates the maximum spatial extent of the 2p–F atomic orbital; and (e) and graphical illustrations of ⟨2A1H3C+|2p–F⟩ and ⟨2A1H3C+|5p–I⟩ orbital overlaps (2 contour lines) at a H3C⋯XHn distance of 1.50 Å, contour plots from 0.035 to 0.3 au. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

FIG. 2.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = F and I) into H3C+ and XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity. (c) Repulsive occupied–occupied orbital overlaps (S) between H3C+ and XHn; (d) contour plots (10 contour lines) of np–XHn atomic orbitals, where the black vertical line indicates the maximum spatial extent of the 2p–F atomic orbital; and (e) and graphical illustrations of ⟨2A1H3C+|2p–F⟩ and ⟨2A1H3C+|5p–I⟩ orbital overlaps (2 contour lines) at a H3C⋯XHn distance of 1.50 Å, contour plots from 0.035 to 0.3 au. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

Close modal

By analyzing the KS-MOs, we find that upon bond formation, C–F builds up significantly less repulsive occupied–occupied orbital overlap between the two approaching fragments than H3C–I. Figure 2(c) shows the evolution of the repulsive occupied–occupied orbital overlap between ⟨2A1H3C+|2p–F⟩ for C–F and ⟨2A1H3C+|5p–I⟩ for C–I, where 2A1H3C+ is the occupied all in-phase σ-orbital of H3C+ and 2p–F and 5p–I are the valence atomic npσ-HOMO of F and I, respectively. A small and, hence, less destabilizing orbital overlap is found for C–F, whereas a large and, therefore, more destabilizing orbital overlap is found for C–I. The reduction of repulsive orbital overlap can be explained by the difference in spatial extent of the occupied valence atomic orbitals of XHn, that is, 2p–F and 5p–I. The occupied 2p–F of C–F points, due to the intrinsic nature of this valence atomic orbital, less toward the approaching H3C+ fragment than the large 5p–I. This difference in spatial extent can be graphically illustrated by visualizing these occupied orbitals [Fig. 2(d)]. Where 2p–F spans until the black vertical line, 5p–I expands far beyond this line, making the latter have a better spatial match with the 2A1H3C+, which amplifies the repulsive orbital overlap even further [Fig. 2(e)]. In addition, the heavier halogen also has more subvalence shells, which further raise the number of repulsive occupied–occupied orbital interactions [see also Ref. 5(a)].

Now, we move to the second heterolytic dissociation pathway, where, along period 2, the H3C–CH3 and H3C–F bonds are dissociated into the methyl anion H3C and the respective cationic species +CH3 and +F (“hetero 2” in Scheme 1). In this case, the bond becomes stronger, i.e., it has a more stabilizing ΔE, from C–C to C–F, thus paralleling the same trend along the homolytic dissociation pathway for C–X bonds. This trend for pathway “hetero 2” originates from a more stabilizing interaction energy ΔEint [Fig. 3(a)]. The strain energy ΔEstrain plays only a minor role, and the difference stems from the deformation of the +CH3 fragment, which needs to pyrimidalize upon forming the C–X bond. The trend in interaction energy, and hence bond energy, is dictated by the orbital interactions, which are significantly more stabilizing from C–C to C–F. In contrast, the electrostatic interaction and Pauli repulsion are less important for the trend; they differ hardly from C–C to C–F, with nearly superimposed curves.

FIG. 3.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = CH3 and F) into H3C and +XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C and +XHn as well as the +XHn orbital energies; and (d) graphical illustration of ⟨3A1H3C|3A1+CH3⟩ and ⟨3A1H3C|2p+F⟩ orbital overlaps (2 contour lines) at a H3C⋯XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

FIG. 3.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = CH3 and F) into H3C and +XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C and +XHn as well as the +XHn orbital energies; and (d) graphical illustration of ⟨3A1H3C|3A1+CH3⟩ and ⟨3A1H3C|2p+F⟩ orbital overlaps (2 contour lines) at a H3C⋯XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

Close modal

To understand the origin of the difference in orbital interactions, we perform a KS-MO analysis as a function of the bond distance. This analysis reveals that the extremely low energy of the 2p+F LUMO causes the more favorable orbital interaction ∆Eoi and, therefore, a stronger C–F bond. Figure 3(c) shows the relevant HOMO–LUMO overlap integrals and LUMO orbital energies as a function of the bond distance, that is, the overlap ⟨3A1H3C|3A1+CH3⟩ for C–C and ⟨3A1H3C|2p+F ⟩ for C–F, where 3A1H3C is the carbon 2pσ-derived HOMO of H3C, 3A1+CH3 is the carbon 2pσ-derived LUMO of H3C+, and 2p+F is the 2pσ-LUMO of +F. The very compact 2p+F LUMO of the +F fragment is significantly lower in energy than the less compact 3A1+CH3 LUMO of the +CH3 fragment. Already the neutral parent radical F has a lower lying SOMO than CH3, due to the larger effective nuclear charge that binds the electrons more closely to the nucleus, i.e., a larger ionization energy. The drop in orbital energy going from a neutral radical to a cationic species is, as a result, larger for F than for CH3, due to a larger reduction of Coulomb repulsion in the very compact fluorine system.

Interestingly, the 2p+F LUMO is effectively even lower in energy than the 3A1H3C HOMO. This is also reflected by the fact that the 3A1+CH3 + 2p+F bonding MO in the C–F bond has a larger contribution from the +F LUMO than from the H3C HOMO, resulting in a gross Mulliken population of 1.5 electrons in the 2p+F. Thus, the electrons of the 3A1H3C HOMO of the methyl anion drop into the (bonding combination with the) 2p+F LUMO of +F in a way that is reminiscent of the corresponding C–F electron-pair bond in which the electron of the higher-energy methyl 3A1CH3· SOMO drops into the (bonding combination with the) fluorine 2pσ SOMO. This type of HOMO–LUMO interaction gains stabilization not only from the energy-lowering of the bonding MO relative to the lower-energy fragment molecular orbital (FMO) (∝S2/∆ε) but also from the energy gap between the two FMOs involved (∝∆ε). This is qualitatively illustrated in Scheme 2, which shows how the HOMO–LUMO interaction of path “hetero 2” (H3C + +XHn) leads to more stabilization than that of path “hetero 1” (H3C+ + XHn). For a quantitative comparison of the more stabilizing ∆Eoi values of path “hetero 2” than those of path “hetero 1,” see Table S3.

The HOMO–LUMO orbital overlap shows the opposite trend, namely, it becomes significantly less favorable, i.e., smaller, from C–C to C–F [see Fig. 3(c)]. The reason for this is the spatial mismatch between the relatively diffuse H3C HOMO and the very compact +F LUMO [see Fig. 3(d)]. Therefore, the orbital overlap is not responsible for the weakening of the HOMO–LUMO orbital interactions from C–C to C–F.

Finally, we study the trend in bonding upon heterolytically dissociating H3C–I and H3C–F into the methyl anion H3C and the respective cationic species +I and +F (“hetero 2” in Scheme 1). Figure 4(a) shows that the bond strengthens from C–I to C–F because of a more favorable interaction energy ΔEint. The strain energy ΔEstrain, on the other hand, plays essentially no role: It is small, nearly negligible, and the strain curves for C–I and C–F are nearly superimposed. The reason is that, for both systems, the interacting fragments need to deform only very slightly (the methyl anion is already pyramidal, and the halogen fragment is monoatomic). The interaction ∆Eint is significantly more stabilizing from C–I to C–F because the trend in orbital interactions ∆Eoi and ∆EPauli work in the same direction: the former is more stabilizing and the latter less repulsive [Fig. 4(b)]. The electrostatic interaction shows the opposite trend, that is, it is less stabilizing from C–I to C–F.

FIG. 4.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = F and I) into H3C and +XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C and +XHn as well as the +XHn orbital energies; (d) graphical illustration of ⟨3A1H3C|2p+Fσ⟩ and ⟨3A1H3C|5p+⟩ orbital overlaps (2 contour lines) at a H3C∙∙∙XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u.; (e) repulsive occupied–occupied orbital overlaps (S) between H3C+ and XHn; and (f) graphical illustration of ⟨1E1H3C|2p+Fπ⟩ and ⟨1E1H3C|5p+⟩ orbital overlaps (2 contour lines) at a H3C∙∙∙XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

FIG. 4.

(a) Activation strain and (b) energy decomposition analysis of the heterolytic bond dissociation of H3C–XHn (XHn = F and I) into H3C and +XHn, where the equilibrium H3C–XHn bond distances are indicated with a dot and the dispersion interactions are omitted for clarity; (c) donor–acceptor orbital overlaps (S) between H3C and +XHn as well as the +XHn orbital energies; (d) graphical illustration of ⟨3A1H3C|2p+Fσ⟩ and ⟨3A1H3C|5p+⟩ orbital overlaps (2 contour lines) at a H3C∙∙∙XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u.; (e) repulsive occupied–occupied orbital overlaps (S) between H3C+ and XHn; and (f) graphical illustration of ⟨1E1H3C|2p+Fπ⟩ and ⟨1E1H3C|5p+⟩ orbital overlaps (2 contour lines) at a H3C∙∙∙XHn distance of 1.55 Å, contour plots from 0.035 to 0.3 a.u. Computed at ZORA-BLYP-D3(BJ)/TZ2P.

Close modal

The more stabilizing orbital interactions, from C–I to C–F, are the direct result of the significant lowering of the halogen-cation's LUMO, from +I to +F, as emerges from our analysis of the MO interaction mechanism [see Fig. 4(c)]. In this HOMO–LUMO interaction, the electrons of the carbon 2pσ-derived 3A1 HOMO in the methyl anion CH3 swap into the valence npσ of the halogen cation, as illustrated by the gross Mulliken population of 1.5 electrons in 2p+F and 1.1 electrons in 5p+Iσ. Thus, from C–I to C–F, these electrons are significantly more stabilized as the LUMO energy drops, from the 5p+Iσ LUMO of +I to the 2p+Fσ LUMO of +F.24 Note that the trend in LUMO orbital energy overrules the trend in HOMO–LUMO overlap, which runs counter. Thus, the overlap integral decreases from ⟨3A1H3C|5p+Iσ⟩ to ⟨3A1H3C|2p+Fσ⟩, i.e., from C–I to C–F, as shown in Fig. 4(c). The reason is the better spatial match of the rather diffuse 3A1H3C HOMO of the methyl anion with the likewise relatively diffuse 5p+ LUMO of the large iodine cation than with the more compact 2p+Fσ LUMO of the significantly smaller fluorine cation [see Figs. 4(d) and 4(e)].

The trend in Pauli repulsion stems from the reduction in occupied–occupied orbital overlap upon going from the formation of the C–I bond to the C–F bond, also found for the “homo” and “hetero 1” dissociation pathways. The occupied–occupied overlap integrals as a function of the bond distance, ⟨1E1H3C|5p+⟩ for H3C–I and ⟨1E1H3C|2p+Fπ⟩ for H3C–F, are shown in Fig. 4(e); herein, 1E1H3C represents C–H bonding orbitals of H3C and 5p+ and 2p+Fπ are the π-symmetric valence np-atomic orbitals of +I and +F. The repulsive orbital overlap starts at a longer bond distance and is larger for C–I than for C–F. As discussed earlier, all 5p-atomic orbitals of I+ are larger than the analogous 2p-atomic orbitals of F+ [see Fig. 4(f)]. Due to this difference in spatial extent, the former will start to overlap with the occupied 1E1H3C earlier, and this overlap will rise more steeply than the latter. Furthermore, the heavier halogen also has more subvalence shells, which increases the number of repulsive occupied–occupied orbital interactions, leading to effectively more destabilizing Pauli repulsion [see also Ref. 5(a)]. This, together with the strengthening of the orbital interactions, leads to a contraction and a strengthening of the bond going from C–I to C–F.

The carbon–element bond strength in covalent H3C–XHn single bonds, and trends therein, along various XHn (XHn = CH3, NH2, OH, F, Cl, Br, I) depend profoundly on which homolytic or heterolytic dissociation pathway is considered. As expected, the heterolytic bond dissociation goes, in all cases, with a substantially higher bond dissociation enthalpy (BDE) than homolytic dissociation. Interestingly, heterolytic dissociation into H3C+ + XHn shows along one of the series the opposite trend (weakening from C–C to C–F but still strengthening from C–I to C–F) compared to both homolytic dissociation and heterolytic dissociation into H3C + +XHn (strengthening from C–C to C–F and from C–I to C–F). This follows from our quantum chemical activation strain model (ASM) and quantitative molecular orbital (MO) analyses based on relativistic, dispersion-corrected density functional theory at ZORA-BLYP-D3(BJ)/TZ2P.

Our bonding analyses reveal that, upon homolytic dissociation (“homo” in Scheme 1), the H3C–XHn bond is stronger along period 2, from C–C to C–F, due to an increasingly more favorable electron-pair bond associated with the increased electronegativity difference across the bond. Up group 17, from C–I to C–F, the bond is also stronger but for a different reason, namely, due to decreasing steric repulsion and not due to the generally assumed strengthening of the electron-pair bonding orbital interaction. For the heterolytic dissociation into H3C + +XHn (“hetero 2” in Scheme 1), we observe the same trends for related reasons. Thus, along period 2, from C–C to C–F, the C–X bond becomes stronger because of a strengthening in the HOMO–LUMO orbital interaction, i.e., because the +XHn LUMO drops, and up group 17, from C–I to C–F, the C–X bond strengthens due, again, to the reduction of steric repulsion as X becomes smaller. However, a different trend emerges in the case of heterolytic dissociation into H3C+ + XHn (“hetero 1” in Scheme 1). The C–X bond still strengthens up group 17, from C–I to C–F, due to the decreasing steric repulsion when X is smaller. However, now it weakens instead of strengthening along period 2, from C–C to C–F, because of a weakening in the HOMO–LUMO orbital interactions as the XHn HOMO drops in energy, which increases the HOMO–LUMO orbital energy gap.

The present work highlights what is often overlooked: The nature and strength of a chemical bond are completely defined only if both are considered: (i) the molecule in which the bond features and (ii) the fragments into which it dissociates or from which it forms. A model of the chemical bond that does not account for this is, therefore, incomplete.

The supplementary material contains additional theoretical data and Cartesian coordinates of all structures.

We thank the Netherlands Organization for Scientific Research (NWO) for the financial support. This work was carried out on the Dutch national e-infrastructure with the support of SURF Cooperative and VU BAZIS.

The authors have no conflicts to disclose.

Pascal Vermeeren: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). F. Matthias Bickelhaupt: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article and its supplementary material. Cartesian coordinates can be downloaded from Yoda VU; see Ref. 25.

1.
(a)
M. B.
Smith
,
March’s Advanced Organic Chemistry: Reactions, Mechanisms, and Structure
(
Wiley
,
New York
,
2013
);
(b)
E. V.
Anslyn
and
D. A.
Dougherty
,
Modern Physical Organic Chemistry
(
University Science Books
,
Sausalito
,
2006
);
(c)
J.
Clayden
,
N.
Greeves
, and
S.
Warren
,
Organic Chemistry
, 2nd ed. (
Oxford University Press
,
Oxford
,
2014
).
2.
(a)
G. N.
Lewis
, “
The chemical bond
,”
J. Chem. Phys.
1
,
17
28
(
1933
);
(b)
W.
Kutzelnigg
, “
The physical mechanism of the chemical bond
,”
Angew Chem. Int. Ed. Engl.
12
,
546
562
(
1973
);
(c)
R. S.
Mulliken
, “
Chemical bonding
,”
Annu. Rev. Phys. Chem.
29
,
1
31
(
1978
);
[PubMed]
(d)
G.
Frenking
and
S.
Shaik
,
The Chemical Bond: Chemical Bonding Across the Periodic Table
(
John Wiley & Sons
, 2014), Vol. 2;
(e)
L.
Zhao
,
S.
Pan
,
N.
Holzmann
,
P.
Schwerdtfeger
, and
G.
Frenking
, “
Chemical bonding and bonding models of main-group compounds
,”
Chem. Rev.
119
,
8781
8845
(
2019
).
[PubMed]
3.
(a)
E.
Kraka
,
D.
Setiawan
, and
D.
Cremer
, “
Re-evaluation of the bond length-bond strength rule: The stronger bond is not always the shorter bond
,”
J. Comput. Chem.
37
,
130
142
(
2016
);
(b)
M.
Kaupp
,
B.
Metz
, and
H.
Stoll
, “
Breakdown of bond length-bond strength correlation: A case study
,”
Angew. Chem., Int. Ed.
39
,
4607
4609
(
2000
);
(c)
C.
Esterhuysen
and
G.
Frenking
, “
The nature of the chemical bond revisited. An energy partitioning analysis of diatomic molecules E2 (E = N–Bi, F–I), CO and BF
,”
Theor. Chem. Acc.
111
,
381
389
(
2004
);
(d)
O.
,
M.
Yáñez
,
M.
Eckert-Maksić
,
Z. B.
Maksić
,
I.
Alkorta
, and
J.
Elguero
, “
Periodic trends in bond dissociation energies. A theoretical study
,”
J. Phys. Chem. A
109
,
4359
4365
(
2005
);
(e)
K. T.
Giju
,
F.
De Proft
, and
P.
Geerlings
, “
Comprehensive study of density functional theory based properties for group 14 atoms and functional groups, –XY3 (X = C, Si, Ge, Sn, Pb, Element 114; Y = CH3, H, F, Cl, Br, I, At)
,”
J. Phys. Chem. A
109
,
2925
2936
(
2005
);
[PubMed]
(f)
L.
Zhao
,
S.
Pan
, and
G.
Frenking
, “
The nature of the polar covalent bond
,”
J. Chem. Phys.
157
,
034105
(
2022
).
4.
(a)
L.
Pauling
, “
The nature of the chemical bond IV. The energy of single bonds and the relative electronegativity of atoms
,”
J. Am. Chem. Soc.
54
,
3570
3582
(
1932
);
(b)
H. O.
Pritchard
and
H. A.
Skinner
, “
The concept of electronegativity
,”
Chem. Rev.
55
,
745
786
(
1955
);
(c)
R. T.
Sanderson
, “
Electronegativity and bond energy
,”
J. Am. Chem. Soc.
105
,
2259
2261
(
1983
).
5.
(a)
E.
Blokker
,
X.
Sun
,
J.
Poater
,
M.
van der Schuur
,
T. A.
Hamlin
, and
F. M.
Bickelhaupt
, “
The chemical bond: When atom size instead of electronegativity difference determines trend in bond strength
,”
Chem. - Eur. J.
27
,
15616
15622
(
2021
);
(b)
W.-J.
van Zeist
and
F. M.
Bickelhaupt
, “
Trends and anomalies in H–AHn and CH3–AHn bond strengths (AHn = CH3, NH2, OH, F)
,”
Phys. Chem. Chem. Phys.
11
,
10317
10322
(
2009
);
[PubMed]
6.
(a)
P.
Vermeeren
,
W.-J.
van Zeist
,
T. A.
Hamlin
,
C.
Fonseca Guerra
, and
F. M.
Bickelhaupt
, “
Not carbon s–p hybridization, but coordination number determines C−H and C−C bond length
,”
Chem. - Eur. J.
27
,
7074
7079
(
2021
);
(b)
P.
Vermeeren
and
F. M.
Bickelhaupt
, “
The abnormally long and weak methylidyne C–H bond
,”
Nat. Sci.
3
,
e20220039
(
2023
);
(c)
D.
Rodrigues Silva
,
E.
Blokker
,
J. M.
van der Schuur
,
T. A.
Hamlin
, and
F. M.
Bickelhaupt
, “
Nature and strength of group-14 A–A′ bonds
,”
Chem. Sci.
15
,
1648
1656
(
2024
);
[PubMed]
(d)
C.
Nieuwland
and
C.
Fonseca Guerra
, “
How the chalcogen atom size dictates the hydrogen-bond donor capability of carboxamides, thioamides, and selenoamides
,”
Chem. - Eur. J.
28
,
e202200755
(
2022
);
(e)
L.-J.
Cui
,
Y.-Q.
Liu
,
M.-H.
Wang
,
B.
Yan
,
S.
Pan
,
Z.-H.
Cui
, and
G.
Frenking
, “
Multiple bonding in AeN (Ae = Ca, Sr, Ba)
,”
Chem. - Eur. J.
30
,
e202400714
(
2024
);
(f)
A.
Krapp
,
F. M.
Bickelhaupt
, and
G.
Frenking
, “
Orbital overlap and chemical bonding
,”
Chem. - Eur. J.
12
,
9196
9216
(
2006
).
7.
(a)
P.
Vermeeren
,
S. C. C.
van der Lubbe
,
C.
Fonseca Guerra
,
F. M.
Bickelhaupt
, and
T. A.
Hamlin
, “
Understanding chemical reactivity using the activation strain model
,”
Nat. Protoc.
15
,
649
667
(
2020
);
[PubMed]
(b)
P.
Vermeeren
,
T. A.
Hamlin
, and
F. M.
Bickelhaupt
, “
Chemical reactivity from an activation strain perspective
,”
Chem. Commun.
57
,
5880
5896
(
2021
).
8.
(a)
R.
van Meer
,
O. V.
Gritsenko
, and
E. J.
Baerends
, “
Physical meaning of virtual Kohn–Sham orbitals and orbital energies: An ideal basis for the description of molecular excitations
,”
J. Chem. Theory Comput.
10
,
4432
4441
(
2014
);
(b)
T. A.
Albright
,
J. K.
Burdett
, and
W. H.
Wangbo
,
Orbital Interactions in Chemistry
(
Wiley
,
New York
,
2013
).
9.
(a)
T. A.
Hamlin
,
P.
Vermeeren
,
C.
Fonseca Guerra
, and
F. M.
Bickelhaupt
, in
Complementary Bonding Analysis
, edited by
S.
Grabowsky
(
De Gruyter
,
Berlin
,
2021
), pp.
199
212
;
(b)
F. M.
Bickelhaupt
and
E. J.
Baerends
, in
Reviews in Computational Chemistry
, edited by
K. B.
Lipkowitz
and
D. B.
Boyd
(
Wiley
,
Hoboken
,
2000
), pp.
1
86
.
10.
(a)
G.
te Velde
,
F. M.
Bickelhaupt
,
E. J.
Baerends
,
C.
Fonseca Guerra
,
S. J. A.
van Gisbergen
,
J. G.
Snijders
, and
T.
Ziegler
, “
Chemistry with ADF
,”
J. Comput. Chem.
22
,
931
967
(
2001
);
(b)
C.
Fonseca Guerra
,
J. G.
Snijders
,
G.
te Velde
, and
E. J.
Baerends
, “
Towards an order-N DFT method
,”
Theor. Chem. Acc.
99
,
391
403
(
1998
);
(c)ADF2017.111, SCM Theoretical Chemistry,
Vrije Universiteit
,
Amsterdam, The Netherlands
, www.scm.com.
11.
(a)
A. D.
Becke
, “
Density-functional exchange-energy approximation with correct asymptotic behavior
,”
Phys. Rev. A
38
,
3098
(
1988
);
(b)
C.
Lee
,
W.
Yang
, and
R. G.
Parr
, “
Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density
,”
Phys. Rev. B
37
,
785
(
1988
).
12.
(a)
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
, “
A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu
,”
J. Chem. Phys.
132
,
154104
(
2010
);
(b)
S.
Grimme
,
S.
Ehrlich
, and
L.
Goerigk
, “
Effect of the damping function in dispersion corrected density functional theory
,”
J. Comput. Chem.
32
,
1456
1465
(
2011
).
13.
E. R.
Johnson
and
A. D.
Becke
, “
A post-Hartree–Fock model of intermolecular interactions
,”
J. Chem. Phys.
123
,
024101
(
2005
).
14.
(a)
E.
van Lenthe
,
E. J.
Baerends
, and
J. G.
Snijders
, “
Relativistic regular two-component Hamiltonians
,”
J. Chem. Phys.
99
,
4597
4610
(
1993
);
(b)
E.
van Lenthe
,
E. J.
Baerends
, and
J. G.
Snijders
, “
Relativistic total energy using regular approximations
,”
J. Chem. Phys.
101
,
9783
9792 (
1994
);
(c)
E.
van Lenthe
,
A.
Ehlers
, and
E. J.
Baerends
, “
Geometry optimizations in the zero order regular approximation for relativistic effects
,”
J. Chem. Phys.
110
,
8943
8953 (
1999
).
15.
E.
van Lenthe
and
E. J.
Baerends
, “
Optimized Slater-type basis sets for the elements 1–118
,”
J. Comput. Chem.
24
,
1142
1156
(
2003
).
16.
(a)
M.
Franchini
,
P. H. T.
Philipsen
, and
L.
Visscher
, “
The Becke fuzzy cells integration scheme in the Amsterdam density functional program suite
,”
J. Comput. Chem.
34
,
1819
1827
(
2013
);
(b)
M.
Franchini
,
P. H. T.
Philipsen
,
E.
van Lenthe
, and
L.
Visscher
, “
Accurate Coulomb potentials for periodic and molecular systems through density fitting
,”
J. Chem. Theory Comput.
10
,
1994
2004
(
2014
).
17.
X.
Sun
,
T. M.
Soini
,
J.
Poater
,
T. A.
Hamlin
, and
F. M.
Bickelhaupt
, “
PyFrag 2019—Automating the exploration and analysis of reaction mechanisms
,”
J. Comput. Chem.
40
,
2227
2233
(
2019
).
18.
C. Y.
Legault
, CYLview20,
Université de Sherbrooke
,
2020
; http://www.cylview.org.
19.
(a)
D. A.
McQuarrie
and
J. D.
Simon
,
Physical Chemistry: A Molecular Approach
(
University Science Books
,
Sausalito, CA
,
1997
);
(b)
C. J.
Cramer
,
Essentials of Computational Chemistry: Theories and Models
(
Wiley
,
Chichester, West Sussex, England; Hoboken, NJ
,
2004
);
(c)
F.
Jensen
,
Introduction to Computational Chemistry
(
Wiley
,
Chichester, UK; Hoboken, NJ
,
2017
);
(d)
P.
W Atkins
,
J.
De Paula
, and
J.
Keeler
,
Atkins’ Physical Chemistry
(
Oxford University Press
,
Oxford, UK; New York, NY
,
2018
).
20.
Y.-R.
Luo
,
Comprehensive Handbook of Chemical Bond Energies
(
Taylor & Francis
,
Boca Raton, FL
,
2007
).
21.

From C–C to C–N, the homolytic BDE becomes weaker, not stronger, because the methyl groups in H3C–CH3 can all move their C–H bonds away from the other fragment, which lowers the steric Pauli repulsion, whereas the amino group in H3C–NH2cannot move its lone pair away, which results in more steric Pauli repulsion [see Ref. 5(b)].

22.
F. M.
Bickelhaupt
,
C.
Fonseca Guerra
,
M.
Motoraj
,
F.
Sagan
,
A.
Michalak
,
S.
Pan
, and
G.
Frenking
, “
Clarifying notes on the bonding analysis adopted by the energy decomposition analysis
,”
Phys. Chem. Chem. Phys.
24
,
15726
15735
(
2022
).
23.
R. J.
Mulder
,
C.
Fonseca Guerra
, and
F. M.
Bickelhaupt
, “
Methyl cation affinities of neutral and anionic maingroup-element hydrides: Trends across the periodic table and correlation with proton affinities
,”
J. Phys. Chem. A
114
,
7604
7608
(
2010
).
24.
W.-J.
van Zeist
,
Y.
Ren
, and
F. M.
Bickelhaupt
, “
Halogen versus halide electronic structure
,”
Sci. China Chem.
53
,
210
215
(
2010
).
25.
P.
Vermeeren
and
F. M.
Bickelhaupt
(2024). “
Cartesian coordinates for ‘Chemical bonding trends depend on dissociation pathways
,’”
Vrije Universiteit Amsterdam
. https://doi.org/10.48338/vu01-qjbb8g