Xiangyu
Wu
a,
F. Matthias
Bickelhaupt
*bcd and
Jing
Xie
*a
aKey Laboratory of Cluster Science of Ministry of Education, Beijing Key Laboratory of Photoelectronic/Electrophotonic Conversion Materials, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing, 100081, China. E-mail: jingxie@bit.edu.cn
bDepartment of Chemistry and Pharmaceutical Sciences, AIMMS, Vrije Universiteit Amsterdam, De Boelelaan 1108, 1081 HZ Amsterdam, The Netherlands. E-mail: f.m.bickelhaupt@vu.nl
cInstitute for Molecules and Materials (IMM), Radboud University Nijmegen, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands
dDepartment of Chemical Sciences, University of Johannesburg, Auckland Park, Johannesburg 2006, South Africa
First published on 15th March 2024
We have quantum chemically investigated how microsolvation affects the various E2 and SN2 pathways, their mutual competition, and the α-effect of the model reaction system HOO−(H2O)n + CH3CH2Cl, at the CCSD(T) level. Interestingly, we identify the dual nature of the α-nucleophile HOO− which, upon solvation, is in equilibrium with HO−. This solvent-induced dual appearance gives rise to a rich network of competing reaction channels. Among both nucleophiles, SN2 is always favored over E2, and this preference increases upon increasing microsolvation. Furthermore, we found a pronounced α-effect, not only for SN2 substitution but also for E2 elimination, i.e., HOO− is more reactive than HO− in both cases. Our activation strain and quantitative molecular orbital analyses reveal the physical mechanisms behind the various computed trends. In particular, we demonstrate that two recently proposed criteria, required for solvent-free nucleophiles to display the α-effect, must also be satisfied by microsolvated HOO−(H2O)n nucleophiles.
The peroxide anion HOO− is an interesting nucleophile, because the monohydrated peroxide anion displays dual nucleophile character, where both HOO−(H2O) and (HOOH)(OH−) species are similarly stable.57 Moreover, as HOO−(H2O) reacts with CH3Cl,58 both HOO− and HO− anions are possible attacking nucleophiles, and this was observed in our recent direct dynamic simulation work.36 The introduction of water into the HOO− nucleophile enriches the reaction dynamics for it adds the proton-transfer induced HO−-SN2 pathway; one can expect that if the substrate was ethyl halides, the E2 pathway will emerge and make the dynamics more complicated and interesting.
Besides, HOO− is a typical α-nucleophile, possessing a lone pair of electrons adjacent to the nucleophilic atom. The term α-effect59 has been used to describe the enhanced reactivity of α-nucleophiles compared to that of normal nucleophiles by deviating from the Brønsted-type correlations found for normal nucleophiles.60 There has been extensive discussion on the origin of the α-effect, as well as a controversy about whether the α-effect is controlled by the intrinsic properties of the α-nucleophile or by external solvent effects.61–63 In terms of the intrinsic properties, mechanisms such as ground state destabilization, transition state stabilization, and thermodynamic product stability were proposed to be the origin of the α-effect.63–69 The α-effect has been observed in a variety of SN2 reactions,58,61,70–74 yet fewer studies have addressed its relevance to E2 reactions.75,76 A recent study by Hamlin et al.77 proposed two intrinsic criteria for the α-nucleophile to display the α-effect: (1) a higher energy HOMO and (2) a smaller HOMO lobe and overall amplitude of occupied orbitals on the nucleophilic center compared to the normal nucleophile. These criteria were proposed for solvent-free nucleophiles, and it is intriguing to examine whether they suit microsolvated nucleophiles.
In this study, we report a quantum chemical study on the HOO−(H2O)n + CH3CH2Cl reaction (Scheme 1), where n = 0 to 4 is the number of water molecules. We explore the full reaction pathways that, after formation of the initial E2 or SN2 product complexes, lead to the separated products, as shown in Scheme 1. This is in agreement with earlier experimental and simulation studies on closely related microsolvated ion–molecule SN2 reactions, which have shown that the formation of the unsolvated ionic products strongly dominates the formation of the solvated ionic products because of dynamic bottlenecks which make solvent transfer from nucleophiles to leaving groups less likely.45,47,78 The purpose of this study is three-fold, namely, to investigate the effect of solvation on (1) the competing SN2 and E2 reaction pathways, (2) the competing normal HOO−-pathways and the solvent-induced HO−-pathways, and (3) the α-effect on both, the SN2 and E2 reactions, in terms of the nucleophiles' intrinsic properties.
![]() | ||
Scheme 1 SN2 and E2 reaction pathways of HOO−(H2O)n=1–4 + CH3CH2Cl reactions. In pathway c′, one water is the conjugate acid of the nucleophile HO−. |
n | HOO−W-path | HO−PT-path | |||||
---|---|---|---|---|---|---|---|
ΔE‡ | ΔH‡ | ΔG‡ | ΔE‡ | ΔH‡ | ΔG‡ | ||
a Computed at CCSD(T)/aug-cc-pVTZ//MP2/6-311++G(d,p). | |||||||
inv-SN2 | 0 | −14.1 | −13.8 | −3.9 | −14.0 | −14.0 | −5.4 |
1 | −4.6 | −3.8 | 5.2 | −1.4 | −0.4 | 9.5 | |
2 | 0.6 | 1.7 | 13.2 | 2.7 | 3.8 | 16.3 | |
3 | 3.1 | 4.1 | 14.8 | 8.0 | 9.3 | 19.4 | |
4 | 8.0 | 8.3 | 18.2 | 10.0 | 11.0 | 22.4 | |
anti-E2 | 0 | −9.7 | −12.6 | −3.5 | −12.7 | −15.8 | −8.0 |
1 | 2.4 | 0.1 | 7.9 | 3.8 | 1.6 | 9.6 | |
2 | 7.9 | 6.9 | 19.2 | 8.8 | 6.9 | 17.4 | |
3 | 14.0 | 12.5 | 21.6 | 15.2 | 13.7 | 22.0 | |
4 | 17.9 | 16.4 | 25.1 | 18.2 | 16.2 | 26.9 | |
ret-SN2 | 0 | 16.5 | 16.3 | 27.5 | 20.9 | 20.5 | 29.4 |
syn-E2 | 0 | −3.1 | −6.0 | 4.1 | −5.2 | −8.6 | −0.5 |
It has been shown that hydrated peroxide anions, HOO−(H2O)n, tend to abstract a proton from H2O and form more stable species HO−(HOOH)(H2O)n−1 (Fig. S1, ESI†).36,38,57,58 Consequently, the solvent-induced proton-transfer HO−-moiety is a potential nucleophile to compete with the original nucleophile HOO−. Hence, when HOO−(H2O)n reacts with CH3CH2Cl, four pathways are possible: (a) HOO−W-SN2, “W” indicates that the nucleophile is bound with water molecules; (a′) HO−PT-SN2, “PT” indicates that the HO− nucleophile is induced by proton transfer from water molecules; (c) HOO−W-E2; and (c′) HO−PT-E2 (Scheme 1). Note that in the hydrated system, the HOO−W-E2 and HO−PT-E2 pathways generate the same products.
To show the effect of individual solvent molecules, we plotted the potential energy profiles of the HOO−(H2O)n=0,1,2 + CH3CH2Cl reactions in Fig. 2 for both HOO−W-paths (right panel) and HO−PT-paths (left panel). We used 0, 1 and 2 as prefixes to denote the number of solvent molecules when naming the species. The corresponding transition state structures are shown in Fig. 3. The involvement of multiple H2O molecules complicates the structures, so the most stable structures of each species and corresponding energetics were used in the discussion. Information on higher-energy conformational isomers is provided in Fig. S2–S6 (ESI†) for interested readers. Using HO−(HOOH) + CH3CH2Cl as the reference point, as observed, the HOO−W-SN2 reactions (−18.8 kcal mol−1) are more exothermic than HO−PT-SN2 reactions (−11.5 kcal mol−1), where both are more exothermic than E2 reactions (3.1 kcal mol−1). The addition of water molecules to the ion–molecule system stabilizes each species. For the singly- and doubly-hydrated systems, the potential energy profiles of both SN2 and E2 reactions remain double-well shaped. However, due to the differential stabilization effect of water molecules on the reactants and transition states, the barrier heights changed differently for the SN2 and E2 reactions. The energetic values are presented in Table 1 and details will be discussed in the next section.
![]() | ||
Fig. 3 Structures of transition states of HOO−(H2O)n=0–4 + CH3CH2Cl optimized at MP2/6-311++G(d,p). Bond lengths are in angstrom. |
Regarding the competition between the α-nucleophile HOO−W and the normal nucleophile HO−PT (Fig. 4c and d), the barrier height difference, defined as ΔΔE‡2 = ΔE‡(HO−PT) − ΔE‡(HOO−W), is greater than zero for almost all cases, that is, the α-nucleophile reacts faster in almost all cases. The main trend is that introducing microsolvation, i.e., going from n = 0 to n ≥ 1, enhances the α-effect, i.e., ΔΔE‡2. However, the dependence of the α-effect as a function of introducing more solvent molecules (i.e., along n = 1, 2, 3, 4) is less uniform. For the SN2 paths, the barrier difference ΔΔE‡2(SN2) ranges from 2.0 to 4.9 kcal mol−1, whereas there is a smaller difference between the E2 pathways (Fig. 4d), and ΔΔE‡2(E2) ranges from 0.4 to 1.4 kcal mol−1, indicating that the HO−-E2 path can be strongly competitive to the HOO−-E2 pathway, provided sufficient energy is available to pass the E2 barriers.
In brief, among the four competing pathways, the HOO−W-SN2 path dominates with incremental solvation. In what follows, we seek the reason for barrier height difference upon solvation by analyzing the properties of nucleophiles and transition states.
In line with our previous studies,25,38,55,80 we found herein that microsolvation lowers the energy of the HOMO of both HOO− and HO− systematically upon adding an additional solvent molecule, either H2O or HOOH, by the HOMO–LUMO interaction with the σO–H* (solvent) LUMO (Table S4, ESI†). The HOMO of the microsolvated peroxide anion in HOO−(H2O)2,3,4 (−5.6, −6.0, and −6.6 eV) always remains higher in energy than that of the equivalently microsolvated hydroxide anion in HO−(HOOH)(H2O)1,2,3 (−6.5, −6.9, and −7.3 eV) and HO−(H2O)2,3,4 clusters (−5.7, −6.6, and −7.2 eV; see Table S4, ESI†). As shown previously by Bickelhaupt et al.,25,55 this situation gives rise to a smaller HOMO–LUMO energy gap and a more stabilizing HOMO–LUMO interaction with the LUMO (substrate) (in this work, the substrate is CH3CH2Cl) and, therefore, to a lower barrier for the reactions of microsolvated HOO− than for the corresponding reactions of microsolvated HO−. Indeed, we found a strong correlation of the barrier heights of the SN2 and E2 reactions with the HOMO energy level of the microsolvated peroxide and hydroxide nucleophiles (Fig. 5 and Fig. S8, ESI†). The same holds true for the related quantities of proton affinity (PA) and ethyl cation affinity (ECA), which are measures of the nucleophiles’ ability to bind with a proton and an ethyl cation. The PA and ECA values are defined as the enthalpy change of NuH → Nu− + H+ and CH3CH2Nu → Nu− + CH3CH2+, respectively (see Table S5 for all computed PA and ECA values, ESI†).
With the introduction of solvent molecules, the PA or ECA values of nucleophiles decrease continuously, that is, both types of affinities become weaker. Gratifyingly, the overall barrier heights (ΔE‡) of the SN2 and E2 reactions have, again, a strong linear correlation with the PA or ECA of the various microsolvated peroxide and hydroxide nucleophiles (R2 ≈ 1.00; see Fig. 5). In line with this, there is also a good to reasonable correlation between the height of the barrier and the amount of charge transferred from the nucleophile to the substrate, as computed with natural population analysis (NPA) atomic charges (R2 = 0.85–0.97; see Fig. 5 and Table S7, ESI†).37,83 Thus, a stronger charge transfer goes with a lower barrier.
%D‡ = %CαCl‡ + %HβCβ‡ |
Indeed, we found a reasonably strong linear correlation of the SN2 and E2 reaction barriers with %D‡ (R2 = 0.81 to 0.98; see Fig. 5). The trend is that the more geometrically distorted the TS, the higher the barrier.
Furthermore, we found that incremental microsolvation makes the transition structures in general more product-like, as elaborated upon in the following. The SN2 pathway and the transition state involve the breaking of the Cα–Cl bond and the formation of the O–Cα bond, whereas the E2 pathway and transition state involve the breaking of both the Cα–Cl bond and the Hβ–Cβ bond and the formation of the peroxide or hydroxide O–Hβ bond (Fig. 3). Table 2 shows that, as the degree of microsolvation increases from 0 to 3, for transition structures of inv-SN2 reactions, the O–Cα bond shortens and the Cα–Cl bond lengthens systematically. In the case of the transition structures of anti-E2 reactions, there are a few irregularities but, by and large, the O–Hβ bond shortens, and both the Cα–Cl bond and the Hβ–Cβ bond lengthen upon going from the unsolvated to the microsolvated situation. In line with these structural characteristics, the Cl leaving group becomes increasingly negatively charged in the transition states of both SN2 and E2 reactions when the degree of microsolvation increases, as reflected by the computed NPA charges (see Table S7, ESI†). For example, the negative charge of the leaving group q(Cl) of the HOO−W SN2 path increases from −0.539 to −0.675 a.u., and the value of the HOO−W E2 path increases from −0.355 to −0.580 a.u. These findings are all consistent with the fact that the transition states become more product-like. Thus, as-stated above, microsolvation shifts the TS to a later, more product-like point along the reaction coordinate.
HOO−W-path | HO−PT-path | |||||
---|---|---|---|---|---|---|
inv-SN2-TS | inv-SN2-TS | |||||
n | r(O–Cα) | r(Cα–Cl) | r(Hβ–Cβ) | r(O–Cα) | r(Cα–Cl) | r(Hβ–Cβ) |
0 | 2.126 | 2.113 | 1.094 | 2.166 | 2.150 | 1.091 |
1 | 2.047 | 2.178 | 1.094 | 2.038 | 2.251 | 1.091 |
2 | 2.015 | 2.204 | 1.094 | 2.019 | 2.266 | 1.092 |
3 | 1.971 | 2.249 | 1.094 | 2.002 | 2.289 | 1.092 |
4 | 1.950 | 2.270 | 1.094 | 2.019 | 2.269 | 1.092 |
anti-E2-TS | anti-E2-TS | |||||||
---|---|---|---|---|---|---|---|---|
n | r(O–Hβ) | r(Cα–Cl) | r(Hβ–Cβ) | r(Cα–Cβ) | r(O–Hβ) | r(Cα–Cl) | r(Hβ–Cβ) | r(Cα–Cβ) |
0 | 1.249 | 1.950 | 1.377 | 1.458 | 1.310 | 1.911 | 1.330 | 1.470 |
1 | 1.171 | 2.045 | 1.464 | 1.433 | 1.146 | 2.032 | 1.508 | 1.435 |
2 | 1.102 | 2.009 | 1.582 | 1.442 | 1.132 | 2.051 | 1.533 | 1.430 |
3 | 1.130 | 2.148 | 1.519 | 1.412 | 1.120 | 2.077 | 1.552 | 1.425 |
4 | 1.114 | 2.164 | 1.546 | 1.411 | 1.159 | 2.054 | 1.500 | 1.428 |
ΔE‡ = ΔEstrain + ΔEint |
As shown in Fig. S9a and S9b (ESI†), the destabilizing strain energy of the E2 path is significantly larger than that of the SN2 path. The reason is the aforementioned larger characteristic distortion associated with the E2 path in which two bonds are breaking (Cα–Cl and Cβ–H) in the substrate as compared to the lesser characteristic distortion associated with the SN2 path along which only one bond (Cα–Cl) is breaking in the substrate.86 The higher activation strain is what makes the E2 barrier higher than the SN2 barrier, and this can only be inverted if the stabilizing nucleophile–substrate interaction is strong enough. As pointed out by Bickelhaupt et al.,86 the E2 pathway goes with a higher TS acidity, i.e., the lower LUMO in the TS, than the SN2 pathway. However, neither HOO− nor HO− are strong enough bases to cause an inversion of barrier heights as determined by the unfavorably high activation strain for E2 reactions involving the CH3CH2Cl substrate. In view of the fact that introducing solvent molecules makes the nucleophile an even poorer electron donor, the E2 barrier becomes higher relative to the SN2 barrier as the degree of solvation increases.
In the Brønsted-type I correlation, one plots reaction barriers (ΔH‡) against the proton affinity (PA) of HOO−(H2O)n and HO−(H2O)n for n = 0 to 4. We found that there is a good correlation between ΔH‡ and the PA of the normal nucleophiles HO−(H2O)n for both SN2 (Fig. 6a) and E2 (Fig. 6b) reactions. Thus, nucleophiles with larger PA values have lower barriers and, therefore, a higher reactivity. All the points of HOO−(H2O)n deviate downward from this correlation line, with deviation values ΔΔH‡ = ΔH‡(HO− Brønsted path PA) − ΔΔH‡(HOO−) of more than 5 kcal mol−1, revealing the existence of the α-effect for microhydrated HOO− anions. The barriers of E2 reactions are more sensitive to the PA, as reflected by the larger slope of the correlation lines.
In the Brønsted-type II correlation, at each degree of hydration n, one plots the ΔH‡ of a set of other reference normal nucleophiles, including HO−, H2N− and HS−, against the PA, and compares this with the barrier for the HOO− anion and its PA. As shown in Fig. 7, when n = 0 to 3, the barriers of HOO−(H2O)n show a downward shift from the Brønsted-type II ΔH‡versus PA correlation line of the normal nucleophiles for both SN2 and E2 reactions. This downward shift becomes less obvious for E2 when n = 1–3. This is consistent with the report of Hamlin et al.,77 where the unsolvated nucleophiles were considered reacting with ethyl halides. To our knowledge, no such correlation has been plotted for microsolvated nucleophiles reacting with ethyl halides, nevertheless, examples with methyl halides exist.58,74,88–90 Experimental studies by Bierbaum's group58,74 and computational investigations by Ren's group88,90 suggested that the α-effect exists for HOO−(H2O) and HOO−(CH3OH) reacting with methyl chloride. Herein, we expand the exploration of this phenomenon to ethyl halides covering both SN2 and E2 reactions. In brief, the Brønsted-type II analyses also suggest that HOO−(H2O)n nucleophiles display the α-effect as compared with their normal nucleophile counterparts.
The mechanism behind the α-effect in SN2 reactions of unsolvated nucleophiles has been recently studied by Hamlin et al.77 who identified two criteria an α-nucleophile needs to fulfill in order to show the α-effect: (1) a small HOMO lobe and overall reduction of occupied amplitude on the nucleophilic center, in order to reduce the repulsive occupied–occupied orbital overlap between the nucleophile and the substrate and (2) a sufficiently high HOMO energy level, in order to still engage in a strong HOMOnucleophile–LUMOsubstrate orbital interaction with the substrate. Herein, we examine whether the microsolvated HOO−(H2O)n nucleophiles satisfy the two criteria. First, as shown in Fig. S10 (ESI†), the key occupied orbitals of nucleophiles, the HOMO lobes of HOO−(H2O)n, are smaller than those of HO−(H2O)n and HO−(HOOH)(H2O)n−1 for nearly all degrees of microsolvation. This is further confirmed by the fact that the total negative charge on the nucleophilic center is significantly lower in the α-nucleophiles than in the corresponding normal nucleophile (see Fig. S1, ESI†). Second, if a DFT method is used, the HOMO levels of HOO−(H2O)n are consistently at higher energy than those of HO−(H2O)n and HO−(HOOH)(H2O)n−1. So, both criteria77 are indeed satisfied also in the case of microsolvated α-nucleophiles HOO−(H2O)n. Accordingly, the microsolvated HOO−(H2O)n anion shall display an α-effect, consistent with the result given by Brønsted-type correlations.
Our above analyses show that the rate-accelerating α-effect in microsolvated E2 and microsolvated SN2 reactions goes hand-in-hand with a rise in orbital energy of the π-antibonding HOMO and the reduced amplitude of density on the nucleophilic center. The mechanism behind this is that the α-nucleophile has, due to its reduced density on the nucleophilic center, less steric (Pauli) repulsion with the substrate than the normal nucleophile and, therefore, a more stabilizing overall interaction;77 this difference in Pauli repulsion does not occur for the proton affinity because the proton has no occupied orbitals. However, if the ethyl cation affinity (ECA) instead of the PA is used as the Brønsted-correlation parameter, the α-effect is diminished; that is, the downward deviation of the barriers for the α-nucleophiles from the barrier versus the ECA Brønsted-type correlation is significantly reduced because the Pauli-reduction lowering in the case of α-nucleophiles relative to normal nucleophiles now happens not only in the interaction with the substrate in the TS of the reaction but also in the interaction with the carbon acid CH3CH2+ which defines the ECA. We had found this previously for SN2 reactions of unsolvated α-nucleophiles.77
Herein, we have been able to extend this finding to microsolvated nucleophiles and to E2 reactions. Thus, the barriers versus ECA correlations were constructed (Fig. S12d, ESI†), and the degree of downward deviation is indeed greatly reduced. In fact, as shown in Fig. 6c and d, the ΔH‡ values of both the HOO−W-path (α-nucleophile) and the HO−W-path (normal nucleophile) have a good linear relationship with the ECA. This phenomenon of reduced deviation is also observed when the HY−(H2O)0–3 is used as the reference, where Y = O, S, and HN (Fig. S15, ESI†), i.e. type II correlation. Altogether, our computed Brønsted-type correlations reveal that the microhydrated HOO−(H2O)0–4 nucleophiles exhibit the α-effect in both SN2 and E2 reactions.
The SN2 path dominates the E2 path in our model systems. Adding water molecules further enhances the dominance of the SN2 reaction. This is so for both nucleophiles, HOO−(H2O)0–4 and HO−(HOOH)(H2O)0–3. Thus, the E2 barrier rises further above the SN2 barrier with each additional water molecule. This trend emerges from the combination of two factors: (i) the SN2 mechanism is associated with a smaller characteristic distortion and thus less activation strain ΔEstrain than the E2 mechanism; (ii) therefore, as the nucleophile–substrate interaction ΔEint is weakened due to microsolvation, the barrier for the E2 path rises faster than that for SN2 and the latter pathway becomes more dominant.
In the SN2 substitution, the initial HOO− nucleophile is clearly more reactive than the associated solvent-induced HO− nucleophile. But, in the E2 elimination, the difference in reactivity is significantly smaller, with HOO− still being somewhat more reactive. Thus, we found that the HOO−(H2O)0–4 nucleophiles display the α-effect in both the SN2 and, to a lesser extent, also in the E2 reaction. We show that the microsolvated α-nucleophiles satisfy the earlier proposed criteria for the occurrence of the α-effect, namely, a higher-energy HOMO and less occupied amplitude on the nucleophilic center, as compared to the corresponding normal nucleophile.
Our present work provides a unified description and rationalization of the reaction potential energy surface (PES) and kinetic and structural phenomena determined by this PES. A next leap forward that we envisage is the exploration of the complex dynamics taking place on this mechanistically rich multi-mechanistic PES of the HOO−(H2O)0–4 + CH3CH2Cl reactions.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cp00671b |
This journal is © the Owner Societies 2024 |