Kevin
Lynch
a,
Adam
Maloney
a,
Austin
Sowell
a,
Changwei
Wang
b,
Yirong
Mo
*b and
Joel M.
Karty
*a
aDepartment of Chemistry, Elon University, Elon, North Carolina, USA. E-mail: jkarty@elon.edu
bDepartment of Chemistry, Western Michigan University, Kalamazoo, Michigan, USA. E-mail: yirong.mo@wmich.edu
First published on 7th October 2014
Two different and complementary computational methods were used to determine the contributions by inductive/field effects and by electron-delocalization effects toward the enhancement of the gas-phase deprotonation enthalpy of sulfuric acid over ethanol. Our alkylogue extrapolation method employed density functional theory calculations to determine the deprotonation enthalpy of the alkylogues of sulfuric acid, HOSO2–(CH2CH2)n–OH, and of ethanol, CH3CH2–(CH2CH2)n–OH. The inductive/field effect imparted by the HOSO2 group for a given alkylogue of sulfuric acid was taken to be the difference in deprotonation enthalpy between corresponding (i.e., same n) alkylogues of sulfuric acid and ethanol. Extrapolating the inductive/field effect values for the n = 1–6 alkylogues, we obtained a value of 51.0 ± 6.4 kcal mol−1 for the inductive/field effect for n = 0, sulfuric acid, leaving 15.4 kcal mol−1 as the contribution by electron-delocalization effects. Our block-localized wavefunction method was employed to calculate the deprotonation enthalpies of sulfuric acid and ethanol using the electron-localized acid and anion species, which were compared to the values calculated using the electron-delocalized species. The contribution from electron delocalization was thus determined to be 18.2 kcal mol−1, which is similar to the value obtained from the alkylogue extrapolation method. The two methods, therefore, unambiguously agree that both inductive/field effects and electron-delocalization effects have significant contributions to the enhancement of the deprotonation enthalpy of sulfuric acid compared with ethanol, and that the inductive/field effects are the dominant contributor.
Despite its vast chemical importance, the origin of sulfuric acid's enhanced acid strength over alcohols is not well understood, in much the same way that the origin of the enhanced acid strength of carboxylic acids was not well understood until the turn of the twenty-first century. The couple decades prior to that saw a robust controversy9–30 surrounding the contributions by resonance and inductive/field effects to a carboxylic acid's enhanced acidity. The inductive/field effects arise from the electron-withdrawing C
O group attached directly to the reaction center, which serve to stabilize the negative charge that develops on the reaction-center oxygen in the carboxylate anion (Fig. 1a). Resonance, on the other hand, stabilizes that negative charge by allowing the charge to be shared over both oxygen atoms of the carboxylate anion (Fig. 1b, bottom), made possible by the conjugation of the orbitals of π symmetry from the O
C–O− system (Fig. 1b, top).
Much like in carboxylic acids, the enhanced acid strength of sulfuric acid should have a major contribution from inductive/field effects. In the case of sulfuric acid, those inductive/field effects are provided by the HOSO2 group attached to the OH reaction center (Fig. 2a). The HOSO2 group is expected to be quite electron withdrawing inductively, similar to the methylsulfonyl group (CH3SO2),31 owing to the presence of the highly electronegative oxygen atoms and the moderately electronegative sulfur atom. Thus, the HOSO2 group should inductively stabilize the negative charge generated on the reaction-center oxygen in HSO4−.
It might also seem that resonance should have a significant contribution to the acid strength of sulfuric acid, given that resonance structures can be drawn to show the delocalization of the negative charge that develops in HSO4− (Fig. 2b). Drawing these resonance structures assumes that sulfur is hypervalent, where the π bond from each S
O bond is constructed using a d orbital from sulfur. However, several recent studies have shown that the d orbitals from sulfur do not participate substantially in such π bonding,32–35 suggesting that this kind of resonance stabilization in HSO4− should be insignificant.
Nevertheless, there are other orbital interactions in HSO4− that can serve to delocalize the negative charge, thus leading to significant stabilization. These are the negative hyperconjugation interactions involving a lone pair of electrons from O− and the σ* orbitals of geminal S–O bonds (i.e., n → σ* interactions), as shown in Fig. 3.33
![]() | ||
| Fig. 3 Negative hyperconjugation in HSO4−. A lone pair of electrons on O− is delocalized into the σ* orbital of an S–O bond. | ||
The question thus becomes, how much of sulfuric acid's enhancement in acid strength is attributed to these orbital interactions, and how much is attributed to inductive/field effects? Here in this study, we investigate this question using two different and complementary computational methods. One is the alkylogue extrapolation method, and the second is the block-localized wavefunction method. Both of these methods are described later.
To our knowledge, we are the first to provide an accurate evaluation of these kinds of contributions to the acid strength of sulfuric acid at the ab initio level. Denehy et al.,33 however, applied second-order perturbation analysis (as part of their natural bond order and natural resonance theory analyses) on HOSO3CH3 and HOSO3− to estimate the stabilization afforded each species by delocalizing interactions. They estimated 26.9 kcal mol−1 of stabilization in HOSO3CH3 and 93.7 kcal mol−1 of stabilization in HOSO3−. Assuming H2SO4 and HSO4− have similar stabilizations as HOSO3CH3 and HOSO3−, respectively, this would predict the orbital interactions to contribute about 67 kcal mol−1 to sulfuric acid's strength, which is essentially the entirety of the enhancement in deprotonation enthalpy of sulfuric acid over ethanol (68 kcal mol−1). As we show later, we believe this value to be greatly overestimated, by roughly a factor of 4.
The difference in deprotonation enthalpy between a particular alkylogue of sulfuric acid and the corresponding (i.e., same value of n) alkylogue of ethanol, therefore, is attributed essentially entirely to inductive/field effects imposed by the HOSO2 group. After determining the inductive/field contributions for several different values of n, extrapolation of these values back to n = 0 yields the inductive/field effect contributed by the HOSO2 group in sulfuric acid. Finally, the component of sulfuric acid's enhancement in deprotonation enthalpy attributed to the direct orbital interactions between the reaction center and the HOSO2 group is determined by taking the difference between the extrapolated n = 0 inductive/field effect and the total enhancement in deprotonation enthalpy of sulfuric acid over ethanol.
The deprotonation enthalpy of each alkylogue of sulfuric acid should reflect some residual negative hyperconjugation between the reaction center and the bonds of the attached alkyl carbon. However, this effect should be essentially the same in the deprotonation enthalpy of the corresponding alkylogue of ethanol, so when the inductive/field effect is computed as described above, the impact that this residual negative hyperconjugation has should be effectively cancelled. Indeed, when this type of method was applied to para-substituted phenols,36 the results were in quite good agreement with the well-known Swain–Lupton resonance and inductive/field parameters.31
In VB theory, a Lewis structure is defined by a Heitler–London–Slater–Pauling (HLSP) function,44 which can be expanded into 2N Slater determinants (N is the number of bonds). The basic idea of the BLW method is to reduce the number of Slater determinants for a VB function, thus dramatically reducing the computational costs. In the BLW method, we partition the molecular system into several blocks (or fragments or groups), and limit the block-localized MOs (BL-MOs) to expand within only one block and to be doubly occupied in closed-shell cases. In such a way, a HLSP function can be reduced to only one Slater determinant ΨL. BL-MOs in the same block are constrained to be orthogonal like in MO methods, but among different blocks they are non-orthogonal like in VB theory. If we allow all orbitals to expand in the whole space of primitive orbitals, the BLW will be upgraded to the Hartree–Fock (HF, or Kohn–Sham within the density functional theory, DFT) wave function ΨD, corresponding to a delocalized state, which is implicitly a superposition of all electron-localized states. Thus, the energy difference between ΨL and ΨD is defined as the electron delocalization energy (DE), according to eqn (1).
| DE = E(ΨD) − E(ΨL) | (1) |
For each alkylogue, the alkyl chain was in the all-anti conformation. This ensures that the torsional strain in each alkyl chain is minimized and that artifacts imposed by the addition of an alkyl chain are consistent from one alkylogue to another.
) of sulfuric acid, ethanol, and their n = 1–6 alkylogues are presented in Table 1. The calculated value of the inductive/field effect for each alkylogue of sulfuric acid, denoted as Ind(n), is computed by subtracting the deprotonation enthalpy of the nth alkylogue of sulfuric acid from the deprotonation enthalpy of the corresponding (i.e., same n) alkylogue of ethanol. Also included in the table is the calculated distance between S and O−, r(S–O−), in each geometry-optimized alkylogue of HSO4−, because those values were used for extrapolation of the Ind(n) values. The value for Ind(0) appears in parentheses because it was obtained by extrapolating the n = 1–6 values.
| n | H2SO4 | CH3CH2OH | ||||
|---|---|---|---|---|---|---|
| r (S–O−) (Å) | Ind(n)d | |||||
| a Thermally corrected values calculated at the B3LYP/6-311++G(d,p) level of theory. b Experimental values not found in the NIST database (ref. 48) are indicated by a dash. c Calculated distance between S and O− from the geometry-optimized alkylogues of HSO4−. d Each value of the inductive/field effect is computed by subtracting the deprotonation enthalpy of the nth alkylogue of sulfuric acid from the deprotonation enthalpy of the corresponding (i.e., same n) alkylogue of ethanol. e See ref. 7; gas-phase equilibrium measurement. f See ref. 8; gas-phase equilibrium measurement. g Numbers in parentheses are derived from the extrapolation of the n = 1–6 values. h See ref. 49; gas-phase equilibrium measurement. i See ref. 50; kinetic method. | ||||||
| 0 | 1.47 | 308.1 | 309.6 ± 2.6e | 374.5 | 377.4 ± 2.1f | (51.0)g |
| 1 | 4.04 | 352.9 | — | 373.9 | 375.4 ± 2.1f | 21.0 |
| 2 | 6.57 | 363.8 | — | 373.5 | 374.0 ± 2.1h | 9.7 |
| 3 | 9.12 | 368.5 | — | 373.4 | 374.3 ± 2.1h | 4.9 |
| 4 | 11.68 | 370.5 | — | 373.4 | 372.9 ± 2.0i | 2.8 |
| 5 | 14.24 | 371.5 | — | 373.3 | — | 1.8 |
| 6 | 16.80 | 371.9 | — | 373.3 | — | 1.4 |
Fig. 5 is the plot of Ind(n) against the S–O− distance in each geometry-optimized alkylogue of HSO4−. The dashed curve represents the function derived from the nonlinear least-squares fit of the data, as explained later in the Discussion section.
| E(ΨD) | E(ΨL) | DE | |
|---|---|---|---|
| H2SO4 | −698.05189 | −698.01769 | −21.5 |
| HSO4− | −697.54499 | −697.44582 | −62.2 |
| C2H5OH | −154.09492 | −154.07400 | −13.1 |
| C2H5O− | −153.46222 | −153.40541 | −35.6 |
Experimental deprotonation enthalpies for the n = 1–4 alkylogues of ethanol are also available for comparison (Table 1). The experimental values for n = 1–3 are all derived from gas-phase equilibrium measurements. They all lie within a 1.1 kcal mol−1 range, with the value for n = 1 being the largest. Similarly, the calculated values for n = 1–3 all lie within a range of 0.5 kcal mol−1, with the value for n = 1 being the largest. The experimental value for n = 4 appears to be anomalously low, probably because it was derived from a kinetic method rather than equilibrium measurements.
The calculated deprotonation enthalpies of the ethanol alkylogues decrease monotonically from n = 0–6, owing to the increasingly polarizable alkyl chain. This is explained by the fact that with greater polarizability, the negative charge that is produced in the alkoxide anion enjoys greater internal solvation.52 A saturation effect is exhibited by the leveling off of these values around 373.3 kcal mol−1.
The experimental deprotonation enthalpies for the n = 0–3 alkylogues, on the other hand, exhibit a minimum at n = 2. As Higgins and Bartmess49 point out, this is due to the ability of the alkyl chain to bend, allowing for more effective internal solvation when the chain becomes long enough. Such an effect is not reflected by our calculated numbers because all of the calculated geometries have an extended, all-anti alkyl chain. In fact, for the purpose of our alkylogue method, the geometries with the extended alkyl chains are preferable because, when computing the difference in deprotonation enthalpies between corresponding alkylogues of sulfuric acid and ethanol, artifacts due to differences in this internal solvation are avoided.
Unlike what is observed for the alkylogues of ethanol, the calculated deprotonation enthalpies of the n = 1–6 alkylogues of sulfuric acid increase monotonically. This is because the SO3H group is becoming increasingly distant from the reaction center, so the charge stabilization in the conjugate base is decreasing. More specifically, as shown in Table 1, this is due to the decreasing stabilization by the inductive/field effects, Ind(n).
The contribution by inductive/field effects to the deprotonation enthalpy of sulfuric acid—that is, Ind(0)—has a lower bound of about 21.0 kcal mol−1, which is the contribution to the n = 1 alkylogue. To obtain a more precise number, the values of Ind(n) were plotted against r(S–O−) of the geometry-optimized alkylogues of HSO4−, as shown in Fig. 5. The dependent variable used was r(S–O−) because the stabilization of the negative charge is expected to diminish as the distance between O− and the HOSO2 group increases. A nonlinear least-squares fit of these data allowed for the extrapolation of Ind(n) back to n = 0 to yield 51.0 ± 6.4 kcal mol−1 as the contribution by inductive/field effects toward the deprotonation-enthalpy enhancement of sulfuric acid over ethanol.
The general function that was used for the nonlinear least-squares fit is provided in eqn (2).
| Ind(n) = [Ae−br(S–O−)] + C/([r(S–O−)]d) | (2) |
Here, A, b, C and d are fitting parameters. The first term on the right side of eqn (2) is an exponential decay, which is expected for inductive effects.53,54 The second term on the right side of the equation is an inverse-power relationship, which is the general form expected for through-space intermolecular interactions.55 Least-squares fitting of this function to the Ind(n) data yields: A = 69.09 ± 2.36; b = 0.35 ± 0.01; C = 13.40 ± 4.73; d = 0.86 ± 0.22. The resulting equation reproduces the Ind(n) data with an average discrepancy of 0.02 kcal mol−1. To extrapolate Ind(n) to n = 0, the value of 1.47 Å was substituted for r(S–O−), according to Table 1, to yield 51.0 ± 6.4 kcal mol−1 as the contribution by inductive/field effects toward the deprotonation-enthalpy enhancement of sulfuric acid over ethanol. The 6.4 kcal mol−1 uncertainty reflects the standard deviations of A, b, C and d obtained from the nonlinear fit.
The difference in the calculated deprotonation enthalpies between sulfuric acid and ethanol, 66.4 kcal mol−1, is taken to be the sum of the inductive/field effect and the delocalization effect from direct orbital interactions. Thus, the delocalization effect is worth approximately 15.4 kcal mol−1.
The difference between the electron delocalization energies in H2SO4 and HSO4− is 40.7 kcal mol−1, but the corresponding difference in C2H5OH and C2H5O− is also significant, at 22.5 kcal mol−1. Thus, overall the electron delocalization effect contributes 18.2 kcal mol−1 to the deprotonation-enthalpy enhancement of sulfuric acid compared with ethanol. This value is close to the value of 15.4 kcal mol−1 indirectly derived from the alkylogue extrapolation method. These two very different approaches, therefore, consistently suggest that the intramolecular electron delocalization effect makes an important contribution to the strong acidity of sulfuric acid.
Footnote |
| † Electronic supplementary information (ESI) available: Cartesian coordinates, absolute energies, thermally corrected enthalpies, and number of imaginary frequencies for each calculated structure; complete citation for ref. 42. See DOI: 10.1039/c4cp04110k |
| This journal is © the Owner Societies 2015 |