Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Single solvent molecule effect over SN2 and E2 competition in the hydroperoxide anion reaction with ethyl-iodide

Xiangyu Wua, Chongqin Zhub, Joseph S. Francisco*c and Jing Xie*a
aKey Laboratory of Cluster Science of Ministry of Education, School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 100081, China. E-mail: jingxie@bit.edu.cn
bKey Laboratory of Theoretical and Computational Photochemistry, Ministry of Education, College of Chemistry, Beijing Normal University, Beijing 100875, China
cDepartment of Earth and Environmental Science and Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, USA. E-mail: frjoseph@sas.upenn.edu

Received 12th June 2025 , Accepted 23rd July 2025

First published on 24th July 2025


Abstract

The influence of individual solvent molecules on the dynamics of competing reactions remains largely unexplored for many important chemical systems. Herein, direct dynamics simulations revealed that a single water molecule has multifaceted effects on the reaction between the hydroperoxide anion HOO and C2H5I. The introduction of one water reduced the overall reaction rate and shifted the preference from elimination (E2) to substitution (SN2) reactions because of the differential solvation effect. Increasing the collision energy lowered the overall reactivity but did not change the SN2-to-E2 pathway ratio. Notably, the additional water molecules also induced new competing pathways that used HO as an attacking nucleophile via proton transfer within the nucleophile HOO(H2O); here, both the HO-E2 and HO-SN2 trajectories were observed at small percentages. The occurrence of the HO paths was driven by the extensive proton transfer within the pre-reaction complex well, but was suppressed by the entropy effect and increased barriers. In addition, water molecules complicated the reaction mechanisms, increased the percentage of indirect mechanisms, and affected the dynamic features of proton transfer. As in the solvent-free system, protons were frequently exchanged between the nucleophiles and substrates, whereas in the singly solvated system, proton exchange mainly occurred within the nucleophiles. This work highlights the dynamic role of solvent molecules and may have profound impacts on reaction dynamics, with relevance to organic synthesis and chemistry in biosystems, microdroplets, and aerosols.


Introduction

In addition to providing a liquid-phase reaction environment, solvents have a profound impact on chemical reactions, such as affecting the reaction kinetics through catalysis1–5 or inhibition,6 altering the product branching ratios,7–9 and modifying the reaction dynamics.4,10,11 The influence of the individual solvent molecules on the reaction dynamics has long been a central topic among experimental and theoretical researchers in the fields of atmospheric, physical, and interfacial chemistry.1,3,6,7,9,12–19 In particular, the development of crossed-beam imaging experimental techniques and computational simulation powers enables a detailed investigation of the reaction mechanisms and dynamics.10,20–29 Ion–molecule reactions represent a prevalent category of chemical reactions and can potentially be affected by solvent effects.10,30–32 However, the dynamic and steric effects caused by the individual solvent molecules remain elusive for numerous critical ion-molecule reactions.

The hydroperoxide anion (HOO), which is the conjugate base of hydrogen peroxide (HOOH), is an important species across diverse fields. For example, in atmospheric chemistry, HOO participates in reactions that contribute to acid rain formation and ozone consumption.33 In the cellular biology field, HOO is involved in the oxidative processes linked to ageing and degenerative diseases since it reacts with lipids and proteins.34–37 Industrially, HOO serves as a key oxidant in semiconductor cleaning processes and peroxide bleaching, where it facilitates the breakdown of chromophores to decolorize fabrics.38–40 Additionally, HOO likely participates in hydrocarbon oxidation and combustion processes, such as acetylene combustion; here, it reacts with formaldehyde and formic acid.41,42 Recent advancements in energy storage have also highlighted its importance since shifting the discharge intermediate from superoxide O2 to HOO in aprotic Li–O2 batteries can reduce the byproducts and overpotentials to enhance battery efficiency.43 Due to its multifaceted roles, HOO needs to be comprehensively understood to advance research in atmospheric science, biochemistry, industrial application, and energy technologies.

The hydrated HOO ion has garnered great interest for its unique properties.44–49 Microhydrated HOO ions, HOO(H2O)n, can undergo proton transfer to generate isomers in the form of HO(HOOH)(H2O)n−1, where the latter is lower in energy when the number of water molecules n is less than 6.48 Hence, when hydrated HOO ions interact with a substrate, they can behave as dual nucleophiles, similar to CN, such that either HOO or HO can act as nucleophiles.50–53 Previously, direct dynamics simulations of singly hydrated HOO reacting with CH3Cl and CH3I revealed that the HOO path is dominant and produces CH3OOH, whereas the HO path contributes only a minor fraction of the products and produces CH3OH.32,54 Calculations have shown that increasing the degree of hydration causes the HO path to be even less favourable.55 Enlarging the substrate from methyl halides to ethyl halides introduces a competing reaction to the SN2 pathway, namely, elimination reactions (E2).4,9,56–58 Either HOO or HO can serve as the base for the E2 reaction. Interestingly, when singly hydrated HOO reacts with ethyl halides, the E2 reaction products of the HOO and HO paths are identical (Scheme 1) and have similar barriers.59 Distinguishing these two product channels by experiment is a formidable task. Therefore, the reaction dynamics need to be investigated using computational simulations.


image file: d5sc04298d-s1.tif
Scheme 1 SN2 and E2 reaction pathways of the HOO(H2O) + CH3CH2X reactions.

In this work, we performed direct dynamics simulations on CH3CH2I reacting with HOO and HOO(H2O) in the gas phase to reveal the role of a single water molecule in the reaction dynamics. The key questions we explored are as follows: What is the effect of one water molecule on (1) the reaction kinetics, i.e., the rate constants; (2) the competition between SN2 and E2 reactions, i.e., their ratio and atomic-level mechanism; and (3) the competition between the HOO and HO pathways? (4) What is the role of collision energy? The simulations focus on the microhydrated HOO ions and reveal that the solvent molecules not only affect the competition between the different pathways but also influence the dynamic behaviour of chemical reactions. These findings provide valuable theoretical guidance for the environment, energy, and synthesis chemistry.

Computational methods

Direct dynamics simulations were performed at the CAM-B3LYP/6-31+G(d,p)&ECP level of theory using the VENUS/NWChem program.60–62 The 6-31+G(d,p) basis set was used for the H, C, and O atoms, and the LANL2DZdp basis set and effective core potential (ECP) were used for iodine. The choice of this method was based on the accuracy of the structure and reaction enthalpy (ESI Note and Tables S1–S3). The initial distance between CH3CH2I and HOO or HO(HOOH) was 15 Å, and the collision energy (Ecoll) was 0.04 eV (equivalent to 300 K) and 1.0 eV. The vibrational and rotational temperatures of the reactants were 300 K, and the vibrational and rotational energies were sampled for their respective Boltzmann distributions. At an Ecoll of 0.04 eV, the trajectories were calculated at a fixed impact parameter b of 0, 1, 3, 5, 7, 9, 11, 13, and 15 Å; here, the reaction probability was zero at the maximum b under investigation, namely, bmax. At Ecoll = 1.0 eV, the impact parameter b was 0, 1, 3, 5, 7, and 9 Å for the HOO system, and b was 0, 1, 3, 5, and 7 Å for the HO(HOOH) system. For each b, approximately 100–120 trajectories were simulated, resulting in a total of 3178 trajectories. Considering the computational efficiency, the trajectories were integrated by the velocity Verlet method63 with a 1.0 fs time step.

Results and discussion

Overall reactivity

The reactivities of CH3CH2I with HOO and HO(HOOH) at collision energies of 0.04 eV and 1.0 eV were evaluated by plotting the opacity functions Pr(b) versus the impact parameter b for the total reaction and different product channels, i.e., the HOO-SN2, HO-SN2, HOO-E2, and HO-E2 pathways (Fig. 1). In comparison, increasing the collision energy decreases both the maximum impact parameter bmax and the reaction probability at each b, i.e., Pr(b), which together lower the overall reactivity. For example, for the unsolvated HOO + CH3CH2I reaction, the value of bmax decreases from 15.0 Å at an Ecoll of 0.04 eV to 9 Å at 1.0 eV; Pr(b) at b = 1 Å decreases from 0.89 at 0.04 eV to 0.23 at 1.0 eV. Evaluating the total integral cross sections (ICSs) σr via the formula image file: d5sc04298d-t1.tif provides an σr value of 452.8 ± 7.6 Å2 at 0.04 eV and a much smaller value of 17.7 ± 1.9 Å2 at 1.0 eV. The corresponding reaction rate constants of k(Ecoll, Tv, Tr) = v(Ecoll)σ(Ecoll, Tv, Tr) are (23.8 ± 0.4) and (4.7 ± 0.5) × 10−10 cm3 mol−1 s−1. These results indicate that increasing Ecoll from 0.04 to 1.0 eV reduces the rate constant by a factor of 4.
image file: d5sc04298d-f1.tif
Fig. 1 Opacity functions Pr(b) of the different pathways for CH3CH2I reacting with (a and c) HOO and (b and d) HO(HOOH) at collision energies of 0.04 eV (top) and 1.0 eV (bottom).

Adding one water molecule to the nucleophile barely affects the value of bmax but lowers Pr(b). For example, at b = 1.0 Å and Ecoll = 0.04 eV, the total Pr(b) decreases from 0.89 for the unsolvated system to 0.49 for the singly solvated system. As a result, this leads to a decrease in the cross-section and reaction rate constant. For the singly solvated HO(HOOH) + CH3CH2I system, σr is 209.2 ± 8.0 Å2 at 0.04 eV and 7.5 ± 1.0 Å2 at 1.0 eV, and the respective rate constants k are (9.3 ± 0.4) and (1.7 ± 0.2) × 10−10 cm3 mol−1 s−1. Taken together, adding one water to the system decreases the rate constant by factors of 1.6 and 1.8 at Ecoll of 0.04 and 1.0 eV, respectively.

Product channels

The SN2 and E2 reactions compete with each other during the reaction process. For the unsolvated system, both the HOO-SN2 and the HOO-E2 pathways occur throughout the range of b from 0 to bmax. Notably, the E2 pathway remains dominant over the SN2 pathway at each b (Fig. 1a and c). The branching ratio derived from the ICSs for E2[thin space (1/6-em)]:[thin space (1/6-em)]SN2 is 60.4[thin space (1/6-em)]:[thin space (1/6-em)]39.6 at 0.04 eV and 62.7[thin space (1/6-em)]:[thin space (1/6-em)]37.3 at 1.0 eV (Fig. 2a and c). These results show that increasing Ecoll slightly enhances the advantage of E2 over SN2 reactions, even though the overall reactivity is reduced.
image file: d5sc04298d-f2.tif
Fig. 2 Branching ratios of the HOO-SN2 (blue), HOO-E2 (red), HO-SN2 (green) and HO-E2 (orange) pathways for CH3CH2I reacting with (a and c) HOO and (b and d) HO(HOOH) at collision energies of 0.04 eV (top) and 1.0 eV (bottom).

The addition of one water molecule to the system introduces significant changes in the products and their ratios, in addition to affecting the reactivity. At an Ecoll of 0.04 eV, in addition to the normal HOO-SN2 and E2 pathways, additional pathways are observed owing to proton transfer within the nucleophile. One is the HO-SN2 pathway; here, HO behaves as the attacking nucleophile and generates CH3OH + H2O2 + I, and this reaction occurs at b values of 1, 3, 5, 7, 9, and 11 Å. The other is the HO-E2 pathway, which generates the same products as normal HOO-E2, i.e., C2H4 + H2O2 + H2O + I, but the attacking base is HO. This reaction occurs at b values of 0, 1, 3, 7, 9 and 13 Å. However, both pathways are quite rare, where the HO-SN2 and HO-E2 paths account for 2.2% and 2.6%, respectively, of the overall products. At an Ecoll of 1.0 eV, only the HO-E2 pathway is observed, with a minor contribution of 6.8%. Nevertheless, the observation of these HO-attacking trajectories highlights the unique role of water molecules in shaping the reaction dynamics.32,54

Furthermore, the solvated products, i.e., I(H2O) and I(HOOH), are observed in both the SN2 and E2 reaction trajectories. Although the calculated reaction energies are more negative for the solvated products than for the separated products (Fig. S1), the unsolvated product I ions dominate over the solvated I(solvent) ions. In comparison, the reaction energy of the CH3OOH + I(H2O) products is −47.8 kcal mol−1, and this value is 11.4 kcal mol−1 lower than that of the CH3OOH + I + H2O products; the reaction energy of the C2H4 + I(H2O) + H2O2 products is −22.7 kcal mol−1, and this value is 11.4 kcal mol−1 lower than that of the C2H4 + I + H2O + H2O2 products. At an Ecoll of 0.04 eV, the simulated product anion ratio I[thin space (1/6-em)]:[thin space (1/6-em)]I(solvent) is approximately 97[thin space (1/6-em)]:[thin space (1/6-em)]3, highlighting that solvated pathways are suppressed because of the dynamic nature of the reactions in the gas phase; these results are consistent with previous experimental and theoretical works.8,22,30,64 When the collision energy is increased to 1.0 eV, no solvated ions are found. Halogen-bonded complexes, either [C2H5⋯I⋯OOH] or [C2H5⋯I⋯OOH](H2O), are observed in both solvent-free and singly solvated systems and are less than 1%.

Moreover, the additional water molecules shift the preference between the SN2 and E2 pathways compared with the solvent-free case (Fig. 2b and d). For the singly solvated system, at an Ecoll of 0.04 eV, the HOO-SN2 path has a higher Pr(b) than the HOO-E2 path from b of 0 to 5 Å, and they have similar Pr(b) values at b values of 7 to 13 Å. At an Ecoll of 1.0 eV, the HOO-SN2 path has a higher Pr(b) at b values of 1 and 5 Å. The resulting branching ratios of E2[thin space (1/6-em)]:[thin space (1/6-em)]SN2 are approximately 46[thin space (1/6-em)]:[thin space (1/6-em)]54 under both collision energies. Considering only the HOO pathways, i.e., excluding the HO pathways, SN2 is also preferred over E2. The phenomenon that solvent molecules promote a portion of the SN2 path has also been observed in F(CH3OH) + CH3CH2Br reactions.4,5 Why does a single water molecule shift the preference between the SN2 and E2 pathways in the studied system? To answer this question, we analysed the potential energy profiles, the atomistic mechanisms and the dynamics of water molecules in the following section.

Potential energy profiles

The potential energy profile of each SN2 and E2 pathway (Fig. 3) displays a double-well shape, where a central transition state (TS) connects a shallow well of pre-reaction complex (RC) and a deep well of the postreaction complex (PC). All SN2 and E2 reactions are highly exoergic, and the SN2 products are more exoergic.
image file: d5sc04298d-f3.tif
Fig. 3 Potential energy profile of CH3CH2I reacting with HOO and HO(HOOH). (a) refers to the backside attack HOO-SN2 pathway (blue); (b) refers to the backside attack HO-SN2 pathway (green); (c) refers to the antisymmetric HOO-E2 pathway (red); (c’) refers to the syn-symmetric HOO-E2 pathway (pink); (d) refers to the antisymmetric HO-E2 pathway (orange). Energy values (in kcal mol−1) excluding zero-point energy corrections are in normal text, and the values of Gibbs free energy at 298.15 K are in parentheses.

For the unsolvated system, both the back-side SN2 pathway (denoted as a) and the anti-E2 pathway (denoted as c) share the same pre-reaction complex (denoted as 0aRC), with a relative energy of −18.3 kcal mol−1 with respect to reactants HOO + C2H5I. The relative energies of the transition states of the SN2 pathway (0aTS, −17.9 kcal mol−1) and the anti-E2 pathway (0cTS, −17.7 kcal mol−1) exceed that of 0aRC by less than 1 kcal mol−1. These results indicate that overcoming both barriers is easy. When the E2 reaction proceeds via a syn-E2 mechanism, the HOO group abstracts the Hβ atom from the same side of leaving group I, and its transition state (0c′TS) is much greater in energy, with a value of −7.4 kcal mol−1 with respect to the reactants. The halogen complex [C2H5⋯I⋯OOH] (−19.0 kcal mol−1) has a similar energy as 0aRC.

One water molecule is added to the HOO anion to form HO(HOOH), and HOO(H2O) is 29.1 kcal mol−1 exoergic. Starting from HO(HOOH) + C2H5I, the reaction energies of the HOO-SN2, HO-SN2, and E2 pathways are −36.4, −30.0, and −11.3 kcal mol−1, respectively. The HOO-SN2 and HOO-E2 pathways share the same RC, i.e., 1aRC, and the HO-SN2 (denoted as b) and HO-E2 pathways (denoted as d) share the same RC, denoted as 1bRC. The nomenclatures of TS and PC use the a, b, c, and d annotations. The pre-reaction complex for the HO path (1bRC, −14.3 kcal mol−1) is slightly more stable than that for the HOO path (1aRC, −13.0 kcal mol−1), as observed for the reaction with CH3Cl/CH3I.32,54 Nevertheless, the transition states of the HO paths are higher in energy than that for the HOO-path (either the SN2 path or the E2 path). Arranged in ascending order of transition state energies, the order of pathways is HOO-SN2 < HO-SN2 < HOO-E2(anti) < HO-E2(anti). The energies are −9.4, −7.1, −6.2, and −4.0 kcal mol−1 for 1aTS, 1bTS, 1cTS, and 1dTS, respectively. When the free energies are compared, the transition state of HO-SN2 has the highest free energy among these four TSs. As expected, the syn-E2 path has an even higher transition state, 1c′TS, whose relative energy is 2.7 kcal mol−1 in electronic energy and 9.6 kcal mol−1 in free energy.

Assuming a Boltzmann distribution of these four pathways and according to the calculated free energies of the TSs, the thermal ratio at 300 K of HOO-SN2[thin space (1/6-em)]:[thin space (1/6-em)]HOO-E2 is 19.5[thin space (1/6-em)]:[thin space (1/6-em)]80.5 for the unsolvated system, and the thermal ratio of HOO-SN2[thin space (1/6-em)]:[thin space (1/6-em)]HO-SN2[thin space (1/6-em)]:[thin space (1/6-em)]HOO-E2[thin space (1/6-em)]:[thin space (1/6-em)]HO-E2 is 87.9[thin space (1/6-em)]:[thin space (1/6-em)]0.1[thin space (1/6-em)]:[thin space (1/6-em)]11.7[thin space (1/6-em)]:[thin space (1/6-em)]0.3 for the singly solvated system. At an Ecoll of 0.04 eV, corresponding to a temperature of 300 K, the ratios given by direct dynamic simulations are 39.6[thin space (1/6-em)]:[thin space (1/6-em)]60.4 and 51.8[thin space (1/6-em)]:[thin space (1/6-em)]2.2[thin space (1/6-em)]:[thin space (1/6-em)]43.4[thin space (1/6-em)]:[thin space (1/6-em)]2.6. The thermal distribution predictions (based on stationary point calculations) and dynamic simulation results are consistent in that the HOO-E2 path is dominant under unsolvated conditions, whereas the HOO-SN2 path is dominant under singly solvated conditions. Notably, the simulations provide a much higher percentage of the HOO-E2 path, differing by 20–30%, and a slightly higher percentage of the HO paths. These results highlighted the dynamic characteristics of the reaction process. The dynamic preference of the E2 product channel has also been observed in the F(CH3OH) + C2H5Br reaction,4,5 where the dominant pathway given by simulation becomes the E2 path instead of the SN2 path, which was predicted by the thermal distribution. The atomistic mechanisms discussed in the next section provide a detailed picture of the dynamic characteristics.

Atomistic mechanisms

Different types of mechanisms were observed by tracking the animations of each reactive trajectory. The distribution of the mechanisms is shown in Fig. 4. The mechanisms are classified into direct and indirect mechanisms, depending on whether intermediates are formed during the trajectory.20,65 For the unsolvated system, the reaction mechanism is primarily direct. At an Ecoll of 0.04 eV, the direct HOO-SN2 mechanism accounts for 24.2%, and the direct HOO-E2 mechanism accounts for 40.4%; at an Ecoll of 1.0 eV, the corresponding values are 34.2% and 57.8%, respectively. As expected, a higher collision energy leads to more direct trajectories. For both the HOO-SN2 and E2 pathways, both direct stripping (DS) and direct rebound (DR) mechanisms were observed, where the former is more common. The features of these mechanisms have been described in previous works,20,65,66 and representative animations are presented in the ESI Movies. The addition of one water molecule increases the percentage of indirect mechanisms. For the singly hydrated system, at an Ecoll of 0.04 eV, the indirect HOO-SN2 mechanism accounts for 40.5%, and the indirect HOO-E2 mechanism accounts for 29.8%; at an Ecoll of 1.0 eV, the corresponding values are 22.2% and 21.8%, respectively. These values are much higher than the corresponding values from the unsolvated system, which are 15.4, 20.0, 3.1, and 4.9% (Fig. 4). The distributions of the product I ion velocity scattering angle are shown in Fig. S2, where the indirect mechanism shows isotropic scattering.
image file: d5sc04298d-f4.tif
Fig. 4 Distributions of the reaction mechanisms for the SN2 and E2 pathways for CH3CH2I reacting with (a and b) HOO and (c and d) HO(HOOH) at collision energies of 0.04 and 1.0 eV. Note: For unsolvated systems, the 0aRC mechanism involves α-elimination and Hα-exchange mechanisms.

For both unsolvated and singly solvated systems, the observed indirect mechanisms include the formation of various intermediates and roundabout (Ra) mechanisms, where the latter contribute less than 3%. Previous reports of roundabout mechanisms of E2 reactions were limited to F nucleophiles;5,56,67,68 this work extends the roundabout mechanism to HOO and HO(HOOH) reactions.

For both the SN2 and E2 pathways, the dominant indirect mechanism is the formation of the HOO⋯C2H5I complex (0aRC) for the unsolvated system at both Ecoll values and the singly solvated system at an Ecoll of 1.0 eV. The dominant indirect mechanism becomes the formation of the (H2O)HOO⋯C2H5I complex (1aRC) for the singly solvated system at an Ecoll of 0.04 eV. In addition, the mechanism of the singly solvated system was complicated by the participation of multiple intermediates, including 0aRC, 1aRC, 1bRC, and 1aPC. To demonstrate the indirect mechanisms, we provide snapshots of representative trajectories in Fig. 5. Fig. 5a shows an indirect HOO-SN2 trajectory that is trapped in the pre-reaction complex 1aRC well for approximately 4 ps before the products are formed. Fig. 5c displays an indirect HOO-E2 trajectory that is trapped in the 1aRC well for ∼9 ps. Then, the HOO(H2O) group abstracts the Hβ atom from the opposite side of leaving group I; specifically the anti-E2 mechanism occurs, and E2 products are formed. The lifetimes of pre-reaction complexes 1aRC were analysed by plotting ln[N(t)/N(0)] versus time (t) for the trajectories, where N(0) refers to the number of trajectories that experienced the formation of RCs and N(t) refers to the number of trajectories whose lifetimes of RCs are no less than t (Fig. S3). The resulting dissociation rate constant of 1aRC is 0.13 ps−1 for the SN2 path and 0.16 ps−1 for the E2 path. Consequently, the corresponding half-lives of 1aRC are 5.33 and 4.33 ps; these results indicate that the reaction system was trapped in the pre-reaction complex well.


image file: d5sc04298d-f5.tif
Fig. 5 Snapshots of the representative trajectories showing the indirect mechanism of the HOO(H2O) + CH3CH2I reaction. The distance between the Cα atom and I atom r(Cα-I) (purple), the distance between the Cα atom and the attacking O atom r(Cα–O) (green), and the distance between Hβ and the attacking O atom r(Hβ–O) (orange) are shown as a function of time. (a) HOO-SN2 pathway, (b) HO-SN2 pathway, (c) HOO-E2 pathway, (d) HO-E2 pathway.

In addition to the anti-E2 mechanism, we also observed trajectories that follow the syn-E2 mechanism (Fig. S4b). Owing to the higher transition states (Fig. 3) and greater steric hindrance than those of the anti-E2 paths, the syn-E2 trajectories constitute only a small portion of all E2 trajectories. For the unsolvated system, the fraction of syn-E2 trajectories is approximately 3% and 8% at Ecoll values of 0.04 and 1.0 eV; for the singly solvated system, the corresponding fractions are 0.9% and 0.

In the abovementioned anti-E2 and syn-E2 trajectories, the nucleophile first abstracts Hβ from the substrate CH3CH2I. For the unsolvated system at an Ecoll of 1.0 eV, the E2 trajectories may be initiated by a nucleophile abstracting a Hα atom. As shown in Fig. S5b, HOO initially abstracts an Hα atom from the CH3CH2I substrate to form HOOH, followed by the C–I bond breaking, and then the substrate becomes CH3CH. Thereafter, the substrate goes through an Hβ-transfer transition state (EαTS) to generate CH2[double bond, length as m-dash]CH2. This mechanism is called α-elimination (Eα) in previous simulations of F + CH3CH2Br reactions,5 and this name is adopted here. The barrier of EαTS is as high as 7.9 kcal mol−1 (Fig. S6); thus, this mechanism is highly unfavourable and is observed only at an Ecoll of 1.0 eV and accounts for 2.2% of the HOO-E2 path.

In addition, the trajectories that experience the formation of 0aRC sometimes involve hydrogen exchange (HE) between the nucleophile HOO and substrates, as observed in both the SN2 and E2 pathways (Fig. S7b and S5c). Similar to the first step of the Eα mechanism, the nucleophile initially abstracts an Hα atom from the CH3CH2I substrate. The newly formed HOOH moiety transfers an H atom to Cα simultaneously or after a period, and then, the system reforms the CH3CH2I and HOO parts. This H-atom can be either the same Hα or the H that originally belonged to the nucleophile HOO. For the latter case, a transition state was located with a high barrier of 9.5 kcal mol−1. Proton exchange may occur several times within a trajectory. These multiple types of hydrogen exchange trajectories are presented in Fig. S5 and S7. Overall, these trajectories account for 0.36% (2.17%) of the SN2 reactions and 0.63% (0.12%) of the E2 reactions at an Ecoll of 0.04 eV (1.0 eV). Notably, the Hα-abstraction may generate CH3CHI + HOOH products. The product channel is 48.6 endothermic and was not observed during our simulation. However, a trace amount of CH3CHCl was observed in previous simulations of F/HO + CH3CH2Cl.69,70

Interestingly, these α-elimination and Hα-exchange mechanisms were not observed in the singly hydrated system. This can be understood by comparing the proton affinities (PAs). The PA of HOO is 375.9 kcal mol−1, and the PA of HO(HOOH) is 357.0 kcal mol; thus, the abstraction of either Hα or Hβ from CH3CH2I by HO(HOOH) is more difficult than that by HOO. Instead, the proton transfer between the nucleophiles and substrates is completely suppressed by proton transfer within the singly solvated nucleophile HO(HOOH). Representative trajectory snapshots are shown in Fig. S8 and S9.

For the water-induced HO-SN2 and HO-E2 pathways, Fig. 5b and d show that these trajectories also experience a long time (approximately 6–7 ps) within the pre-reaction complex well; here, the nucleophile group strongly interconverts between HOO(H2O) and HO(HOOH). The singly solvated pre-reaction complexes (H2O)HOO⋯C2H5I (1aRC) and (HOOH)HO⋯C2H5I (1bRC) are close in energy, and the latter is approximately 1 kcal mol−1 more stable. However, tracking the trajectories within the well indicates that the system prefers to stay in the configuration of (H2O)HOO⋯C2H5I. For example, for a representative HO-SN2 trajectory (Fig. 5b) and considering the time within the pre-reaction complex well, the system adopts the (H2O)HOO⋯C2H5I configuration 69% of the time, and this value is 73% for a representative HO-E2 trajectory (Fig. 5d). The more stable (HOOH)HO⋯C2H5I configuration only accounts for 31% and 27%, respectively. A detailed analysis is provided in Fig. S10. This occurs because the symmetric structure of (HOOH)HO in 1bRC can easily be disturbed during the dynamic reaction process, namely, the entropy effect. Hence, the lower probability of the HO configurations within the well and the higher barrier both lead to the suppression of the HO paths, both SN2 and E2. Nevertheless, the simulated HO paths have a slightly greater portion than those predicted by the thermal distribution at 300 K; these results indicate the dynamic effects of proton transfer during the reaction processes.

Finally, we investigated the reason for the suppression of the thermodynamically favoured solvated product I(solvent) by the I product from the simulation. To understand this, we plotted the time at which the H2O molecule departed from the system against the formation time of CH3CH2OOH for the HOO-SN2 pathway and against the formation time of CH2[double bond, length as m-dash]CH2 for the HOO-E2 pathway (Fig. 6). At Ecoll = 0.04 eV, the formation of CH3CH2OOH or CH2[double bond, length as m-dash]CH2 simultaneously occurred with the departure of H2O, mostly within 20 ps. As Ecoll increased to 1.0 eV, the majority of SN2 reactions occurred within 5 ps, whereas the H2O molecules were removed much earlier, mainly before 2.5 ps. Similarly, the E2 reactions occurred within ∼4 ps, and the H2O molecules were removed before 2.0 ps. With a time width of ±250 fs centred on the unit slope line, the percentages of the trajectories for SN2 and E2 were 65.2% and 71.5%, respectively, at Ecoll = 0.04 eV, and the corresponding values decreased to 46.0% and 41.4%, respectively, at Ecoll = 1.0 eV. The above analysis, together with the mechanism analysis, indicated that a large portion of the collision energy and/or the reaction exothermicity were transferred to the relative translational energy of the solvent and the remaining parts. Consequently, solvated products were rarely observed.


image file: d5sc04298d-f6.tif
Fig. 6 Water-leaving time as a function of CH3CH2OOH (left) and CH2[double bond, length as m-dash]CH2 (right) formation time for the HOO(H2O) + CH3CH2I reaction at 0.04 (top) and 1.0 eV (bottom) collision energies.

Conclusions

In this work, the effect of a single water molecule on the reaction between the hydroperoxide anion and ethyl iodide was investigated by direct dynamics simulation. The introduction of one water molecule had multiple effects. First, it lowered the overall reactivity; here, the rate constant was reduced by 61% and 63% in comparison to that of the solvent-free reaction, under the considered collision energies of 0.04 and 1.0 eV. Second, it changed the preference between the competing SN2 and E2 reactions. The SN2-to-E2 ratio was approximately 40[thin space (1/6-em)]:[thin space (1/6-em)]60 for the solvent-free system, and this ratio changed to approximately 54[thin space (1/6-em)]:[thin space (1/6-em)]46 for the single-solvation system under both Ecoll values. A single solvent shifted the preference from E2 to SN2 because of the differential solvation effect that led to a higher barrier for the E2 path than for the SN2 path. However, the dynamical simulations resulted in more E2 trajectories than what was predicted by the potential energy profiles. Increasing Ecoll did not affect the SN2-to-E2 ratio but lowered the overall reactivity by 82%.

Third, the extra water molecules facilitated new pathways via proton transfer between HOO(H2O) and HO(HOOH), thus leading to the HO-SN2 and HO-E2 paths, in addition to the traditional HOO-SN2 and HOO-E2 paths, where the latter two paths were dominant. The percentages of HOO-SN2, HOO-E2, HO-SN2, and HO-E2 paths were 51.8%, 43.4%, 2.2%, and 2.6%, respectively, at an Ecoll of 0.04 eV and 54.6%, 38.6%, 0, and 6.8%, respectively, at an Ecoll of 1.0 eV. Extensive proton transfer within the pre-reaction complex well drove the occurrence of the HO paths. However, the lower probability of HO configurations that caused by the entropy effect and the relatively higher barrier led to their low percentages. We anticipate that the solvent molecule inducing new pathways is not an isolated case, especially for protic solvents such as H2O, NH3, and CH3OH. However, such studies are quite limited so far. Singly-solvated anions like OH(H2O), NH2(NH3), and CH3O(CH3OH) experience extensive proton exchange, but generate the same type of anions. Anions like F(H2O), HO(NH3), HOO(NH3), and CH3O(H2O) can generate new nucleophiles through proton transfer, but they are usually higher in energy. Their dynamics remain to be explored in the future.

Fourth, the introduction of one water molecule complicated the reaction mechanisms, increased the percentage of indirect mechanism, and affected the dynamical feature of proton transfer. For the solvent-free system, proton transfer occurred between nucleophiles and substrates; for example, α-elimination and Hα-exchange mechanisms were observed. In contrast, for the singly solvated system, proton transfer mainly occurred within the nucleophile, i.e., the configurations shifted between (H2O)HOO⋯C2H5I and (HOOH)HO⋯C2H5I. Finally, solvated products were observed for the singly solvated reactions, but at a very low percentage of 3%. This occurred because a large portion of the collision energy and/or the reaction exothermicity were transferred to the translation of the products.

To conclude, this work revealed that for the HOO + C2H5I reaction, a single water molecule affected more than the reactivity and competition between the SN2 and E2 reactions. It also induced new competing pathways that used HO as an attacking nucleophile via proton transfer and complicated the reaction mechanisms. These dynamic roles of the individual solvent molecules observed in this work could reveal new facets of reaction dynamics relevant to organic synthesis, biochemistry,71,72 microdroplet chemistry,73–76 and chemistry in aerosols.77–80

Data availability

Data are in ESI.

Author contributions

J. X. designed the research. X. Wu performed the computations, analyzed the data, prepared the figures and tables, and wrote the draft of the paper. C. Z. edited the paper. J. X. and J. S. F. supervised the research and edited the paper.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Key Research and Development Program of China (2024YFA1509602), the National Natural Science Foundation of China (no. 22273004), and the Teli Fellowship and the Innovation Foundation (no. 2021CX01026) from the Beijing Institute of Technology, China.

References

  1. E. Vöhringer-Martinez, B. Hansmann, H. Hernandez, J. S. Francisco, J. Troe and B. Abel, Science, 2007, 315, 497–501 CrossRef PubMed.
  2. G. I. Almerindo and J. R. Pliego, Chem. Phys. Lett., 2006, 423, 459–462 CrossRef CAS.
  3. X. Liu, J. Xie, J. Zhang, L. Yang and W. L. Hase, J. Phys. Chem. Lett., 2017, 8, 1885–1892 CrossRef CAS PubMed.
  4. X. Liu, J. Zhang, L. Yang and W. L. Hase, J. Am. Chem. Soc., 2018, 140, 10995–11005 CrossRef CAS PubMed.
  5. H. Wang, X. Liu, S. Zhao, G. Fu, W. Zhen, L. Yang and J. Zhang, Precis. Chem., 2024, 2, 40–48 CrossRef CAS PubMed.
  6. D. K. Bohme and G. I. Mackay, J. Am. Chem. Soc., 1981, 103, 978–979 CrossRef CAS.
  7. F. M. Bickelhaupt, E. J. Baerends and N. M. M. Nibbering, Chem.–Eur. J., 1996, 2, 196–207 CrossRef CAS.
  8. N. Eyet, J. J. Melko, S. G. Ard and A. A. Viggiano, Int. J. Mass Spectrom., 2015, 378, 54–58 CrossRef CAS.
  9. T. Hansen, J. C. Roozee, F. M. Bickelhaupt and T. A. Hamlin, J. Org. Chem., 2022, 87, 1805–1813 CrossRef CAS PubMed.
  10. R. Otto, J. Brox, S. Trippel, M. Stei, T. Best and R. Wester, Nat. Chem., 2012, 4, 534–538 CrossRef CAS PubMed.
  11. L. Yang, X. Liu, J. Zhang and J. Xie, Phys. Chem. Chem. Phys., 2017, 19, 9992–9999 RSC.
  12. K. Doi, E. Togano, S. S. Xantheas, R. Nakanishi, T. Nagata, T. Ebata and Y. Inokuchi, Angew. Chem., Int. Ed., 2013, 52, 4380–4383 CrossRef CAS PubMed.
  13. C. Zhu, X. C. Zeng, J. S. Francisco and I. Gladich, J. Am. Chem. Soc., 2020, 142, 5574–5582 CrossRef CAS PubMed.
  14. B. Tang, Q. Bai, Y.-G. Fang, J. S. Francisco, C. Zhu and W.-H. Fang, J. Am. Chem. Soc., 2024, 146, 21742–21751 CrossRef PubMed.
  15. Y.-G. Fang, L. Wei, J. S. Francisco, C. Zhu and W.-H. Fang, J. Am. Chem. Soc., 2024, 146, 21052–21060 CrossRef CAS PubMed.
  16. Z. Wan, C. Zhu and J. S. Francisco, J. Am. Chem. Soc., 2023, 145, 17478–17484 CrossRef CAS PubMed.
  17. Z. Wan, Y. Fang, Z. Liu, J. S. Francisco and C. Zhu, J. Am. Chem. Soc., 2023, 145, 944–952 CrossRef CAS PubMed.
  18. G. T. Dunning, D. R. Glowacki, T. J. Preston, S. J. Greaves, G. M. Greetham, I. P. Clark, M. Towrie, J. N. Harvey and A. J. Orr-Ewing, Science, 2015, 347, 530–533 CrossRef CAS PubMed.
  19. F. Xie, D. S. Tikhonov and M. Schnell, Science, 2024, 384, 1435–1440 CrossRef CAS PubMed.
  20. J. Xie, R. Otto, J. Mikosch, J. Zhang, R. Wester and W. L. Hase, Acc. Chem. Res., 2014, 47, 2960–2969 CrossRef CAS PubMed.
  21. R. Otto, J. Xie, J. Brox, S. Trippel, M. Stei, T. Best, M. R. Siebert, W. L. Hase and R. Wester, Faraday Discuss., 2012, 157, 41–57 RSC.
  22. R. Otto, J. Brox, S. Trippel, M. Stei, T. Best and R. Wester, J. Phys. Chem. A, 2013, 117, 8139–8144 CrossRef CAS PubMed.
  23. B. Bastian, T. Michaelsen, L. Li, M. Ončák, J. Meyer, D. H. Zhang and R. Wester, J. Phys. Chem. A, 2020, 124, 1929–1939 CrossRef CAS PubMed.
  24. B. Bastian, T. Michaelsen, M. Ončák, J. Meyer and R. Wester, Int. J. Mass Spectrom., 2021, 462, 116526 CrossRef CAS.
  25. X. Lu, L. Li, X. Zhang, B. Fu, X. Xu and D. H. Zhang, J. Phys. Chem. Lett., 2022, 13, 5253–5259 CrossRef CAS PubMed.
  26. X. Lu, C. Shang, L. Li, R. Chen, B. Fu, X. Xu and D. H. Zhang, Nat. Commun., 2022, 13, 4427 CrossRef CAS PubMed.
  27. D. A. Tasi and G. Czakó, Chem. Sci., 2021, 12, 14369–14375 RSC.
  28. Y. Feng, L. Zhou, Q. Wan, S. Lin and H. Guo, Chem. Sci., 2018, 9, 5890–5896 RSC.
  29. R. Yin, B. Jiang and H. Guo, ACS Catal., 2022, 12, 6486–6494 CrossRef CAS.
  30. J. Xie, R. Otto, R. Wester and W. L. Hase, J. Chem. Phys., 2015, 142, 244308 CrossRef PubMed.
  31. J. Xie, X. Ma, J. Zhang, P. M. Hierl, A. A. Viggiano and W. L. Hase, Int. J. Mass Spectrom., 2017, 418, 122–129 CrossRef CAS.
  32. C. Zhao, X. Ma, X. Wu, D. L. Thomsen, V. M. Bierbaum and J. Xie, J. Phys. Chem. Lett., 2021, 12, 7134–7139 CrossRef CAS PubMed.
  33. M. A. Vincent, I. J. Palmer, I. H. Hillier and E. Akhmatskaya, J. Am. Chem. Soc., 1998, 120, 3431–3439 CrossRef CAS.
  34. J. A. Cotruvo, Jr. and J. Stubbe, Biochemistry, 2010, 49, 1297–1309 CrossRef PubMed.
  35. L. Q. Tuan, H. Umakoshi, T. Shimanouchi and R. Kuboi, Enzyme Microb. Technol., 2009, 44, 101–106 CrossRef CAS.
  36. M. A. Babizhayev, K. S. Vishnyakova and Y. E. Yegorov, Fundam. Clin. Pharmacol., 2011, 25, 139–162 CrossRef CAS PubMed.
  37. L. Qiao, Y. Lu, B. Liu and H. H. Girault, J. Am. Chem. Soc., 2011, 133, 19823–19831 CrossRef CAS PubMed.
  38. H. Wang, Z. Song, W. Liu and H. Kong, Microelectron. Eng., 2011, 88, 1010–1015 CrossRef CAS.
  39. S. Siddiqui, M. Keswani, B. Brooks, A. Fuerst and S. Raghavan, Microelectron. Eng., 2013, 102, 68–73 CrossRef CAS.
  40. S. H. Zeronian and M. K. Inglesby, Cellulose, 1995, 2, 265–272 CrossRef CAS.
  41. S. W. Benson and P. S. Nangia, Acc. Chem. Res., 1979, 12, 223–228 CrossRef CAS.
  42. A. B. Fialkov, Prog. Energy Combust. Sci., 1997, 23, 399–528 CrossRef CAS.
  43. Y. Qiao, S. Wu, J. Yi, Y. Sun, S. Guo, S. Yang, P. He and H. Zhou, Angew. Chem., Int. Ed., 2017, 56, 4960–4964 CrossRef CAS PubMed.
  44. Z. Ma, D. Anick and M. E. Tuckerman, J. Phys. Chem. B, 2014, 118, 7937–7945 CrossRef CAS PubMed.
  45. D. L. Thomsen, J. N. Reece, C. M. Nichols, S. Hammerum and V. M. Bierbaum, J. Am. Chem. Soc., 2013, 135, 15508–15514 CrossRef CAS PubMed.
  46. D. L. Thomsen, C. M. Nichols, J. N. Reece, S. Hammerum and V. M. Bierbaum, J. Am. Soc. Mass Spectrom., 2014, 25, 159–168 CrossRef CAS PubMed.
  47. D. L. Thomsen, J. N. Reece, C. M. Nichols, S. Hammerum and V. M. Bierbaum, J. Phys. Chem. A, 2014, 118, 8060–8066 CrossRef CAS PubMed.
  48. D. J. Anick, J. Phys. Chem. A, 2011, 115, 6327–6338 CrossRef CAS PubMed.
  49. F. Yu, J. Chem. Phys., 2018, 148, 014302 CrossRef PubMed.
  50. Z. Kerekes, D. A. Tasi and G. Czakó, J. Phys. Chem. A, 2022, 126, 889–900 CrossRef CAS PubMed.
  51. A. Gutal and M. Paranjothy, Phys. Chem. Chem. Phys., 2023, 25, 15015–15022 RSC.
  52. X. Liu, S. Tian, B. Pang, H. Li and Y. Wu, Phys. Chem. Chem. Phys., 2023, 25, 14812–14821 RSC.
  53. X. Liu, W. Guo, H. Feng, B. Pang and Y. Wu, J. Phys. Chem. A, 2023, 127, 7373–7382 CrossRef CAS PubMed.
  54. X. Wu, C. Zhao, S. Zhang and J. Xie, J. Phys. Chem. A, 2024, 128, 2393–2398 CrossRef CAS PubMed.
  55. X. Wu, C. Zhao and J. Xie, ChemPhysChem, 2022, 23, e202200285 CrossRef CAS PubMed.
  56. E. Carrascosa, J. Meyer, J. Zhang, M. Stei, T. Michaelsen, W. L. Hase, L. Yang and R. Wester, Nat. Commun., 2017, 8, 25 CrossRef PubMed.
  57. L. Yang, J. Zhang, J. Xie, X. Ma, L. Zhang, C. Zhao and W. L. Hase, J. Phys. Chem. A, 2017, 121, 1078–1085 CrossRef CAS PubMed.
  58. X. Wu, S. Zhang and J. Xie, Phys. Chem. Chem. Phys., 2022, 24, 12993–13005 RSC.
  59. X. Wu, F. M. Bickelhaupt and J. Xie, Phys. Chem. Chem. Phys., 2024, 26, 11320–11330 RSC.
  60. W. L. Hase, R. J. Duchovic, X. Hu, A. Komornicki, K. F. Lim, D. H. Lu, G. H. Peslherbe, K. N. Swamy, S. R. V. Linde, A. J. C. Varandas, H. Wang and R. J. Wolf, Quantum Chem. Program Exch. Bull, 1996, 16, 671 Search PubMed.
  61. M. Valiev, E. J. Bylaska, N. Govind, K. Kowalski, T. P. Straatsma, H. J. J. Van Dam, D. Wang, J. Nieplocha, E. Apra, T. L. Windus and W. A. de Jong, Comput. Phys. Commun., 2010, 181, 1477–1489 CrossRef CAS.
  62. U. Lourderaj, R. Sun, S. C. Kohale, G. L. Barnes, W. A. de Jong, T. L. Windus and W. L. Hase, Comput. Phys. Commun., 2014, 185, 1074–1080 CrossRef CAS.
  63. L. Verlet, Phys. Rev., 1967, 159, 98–103 CrossRef CAS.
  64. J. Zhang, L. Yang, J. Xie and W. L. Hase, J. Phys. Chem. Lett., 2016, 7, 660–665 CrossRef CAS PubMed.
  65. J. Xie and W. L. Hase, Science, 2016, 352, 32–33 CrossRef CAS PubMed.
  66. J. Zhang, J. Mikosch, S. Trippel, R. Otto, M. Weidemüller, R. Wester and W. L. Hase, J. Phys. Chem. Lett., 2010, 1, 2747–2752 CrossRef CAS.
  67. S. Zhao, H. Wang, G. Fu, W. Zhen, M. Liu, L. Yang and J. Zhang, Precis. Chem., 2023, 1, 507–515 CrossRef CAS.
  68. X. Wu, F. Ying, H. Wang, L. Yang, J. Zhang and J. Xie, ACS Phys. Chem. Au, 2024, 4, 581–592 CrossRef CAS PubMed.
  69. J. Meyer, V. Tajti, E. Carrascosa, T. Győri, M. Stei, T. Michaelsen, B. Bastian, G. Czakó and R. Wester, Nat. Chem., 2021, 13, 977–981 CrossRef CAS PubMed.
  70. A. B. Nacsa, C. Tokaji and G. Czakó, Faraday Discuss., 2024, 251, 604–621 RSC.
  71. D. R. B. Brittain, R. Pandey, K. Kumari, P. Sharma, G. Pandey, R. Lal, M. L. Coote, J. G. Oakeshott and C. J. Jackson, Chem. Commun., 2011, 47, 976–978 RSC.
  72. D. O'Hagan and J. W. Schmidberger, Nat. Prod. Rep., 2010, 27, 900–918 RSC.
  73. X. Yan, R. M. Bain and R. G. Cooks, Angew. Chem., Int. Ed., 2016, 55, 12960–12972 CrossRef CAS PubMed.
  74. S. Banerjee, E. Gnanamani, X. Yan and R. N. Zare, Analyst, 2017, 142, 1399–1402 RSC.
  75. K. Li, K. Gong, J. Liu, L. Ohnoutek, J. Ao, Y. Liu, X. Chen, G. Xu, X. Ruan, H. Cheng, J. Han, G. Sui, M. Ji, V. K. Valev and L. Zhang, Cell Rep. Phys. Sci., 2022, 3, 100917 CrossRef CAS.
  76. S. Jin, H. Chen, X. Yuan, D. Xing, R. Wang, L. Zhao, D. Zhang, C. Gong, C. Zhu, X. Gao, Y. Chen and X. Zhang, JACS Au, 2023, 3, 1563–1571 CrossRef CAS PubMed.
  77. J. Li, N. T. Tsona and L. Du, Phys. Chem. Chem. Phys., 2018, 20, 10650–10659 RSC.
  78. N. Tsona Tchinda, L. Du, L. Liu and X. Zhang, Atmos. Chem. Phys., 2022, 22, 1951–1963 CrossRef.
  79. Y. Liu, Q. Ge, T. Wang, R. Zhang, K. Li, K. Gong, L. Xie, W. Wang, L. Wang, W. You, X. Ruan, Z. Shi, J. Han, R. Wang, H. Fu, J. Chen, C. K. Chan and L. Zhang, Chem, 2024, 10, 330–351 CAS.
  80. J. Liu, Y. Zhao, X. Lian, D. Li, X. Zhang, J. Chen, B. Deng, X. Lan and Y. Shao, Molecules, 2024, 29, 4033 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: Direct rebound mechanism of the HOO-SN2 path (Video S1); direct stripping mechanism of the HOO-SN2 path (Video S2); direct rebound mechanism of the HOO-E2 path (Video S3); direct stripping mechanism of the HOO-E2 path (Video S4); indirect mechanism, roundabout mechanism of the HOO-SN2 path (Video S5); indirect mechanism of the H-exchange of the HOO-SN2 path (Video S6); indirect mechanism of the HOO-anti-E2 path (Video S7); indirect mechanism of the HOO-syn-E2 path (Video S8); indirect mechanism of the α-elimination of the HOO-E2 path (Video S9); indirect mechanism of the HO-SN2 path (Video S10); indirect mechanism of the HO-E2 path (Video S11). See DOI: https://doi.org/10.1039/d5sc04298d

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.