Next Article in Journal
Purification, Structural Characterization, and Anti-Inflammatory Effects of a Novel Polysaccharide Isolated from Orostachys fimbriata
Next Article in Special Issue
Evaluating the Conservation State of Naturally Aged Paper with Raman and Luminescence Spectral Mapping: Toward a Non-Destructive Diagnostic Protocol
Previous Article in Journal
Mesoporous Manganese Oxide/Lignin-Derived Carbon for High Performance of Supercapacitor Electrodes
Previous Article in Special Issue
Longitudinal Bottom-Up Proteomics of Serum, Serum Extracellular Vesicles, and Cerebrospinal Fluid Reveals Candidate Biomarkers for Early Detection of Glioblastoma in a Murine Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Intracluster Sulphur Dioxide Oxidation by Sodium Chlorite Anions: A Mass Spectrometric Study

Dipartimento di Chimica e Tecnologie del Farmaco, “Sapienza” University of Rome, 00185 Rome, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(23), 7114; https://doi.org/10.3390/molecules26237114
Submission received: 3 November 2021 / Revised: 21 November 2021 / Accepted: 22 November 2021 / Published: 24 November 2021
(This article belongs to the Special Issue Analytical Chemistry in Italy)

Abstract

:
The reactivity of [NaL·ClO2] cluster anions (L = ClOx; x = 0–3) with sulphur dioxide has been investigated in the gas phase by ion–molecule reaction experiments (IMR) performed in an in-house modified Ion Trap mass spectrometer (IT-MS). The kinetic analysis revealed that SO2 is efficiently oxidised by oxygen-atom (OAT), oxygen-ion (OIT) and double oxygen transfer (DOT) reactions. The main difference from the previously investigated free reactive ClO2 is the occurrence of intracluster OIT and DOT processes, which are mediated by the different ligands of the chlorite anion. This gas-phase study highlights the importance of studying the intrinsic properties of simple reacting species, with the aim of elucidating the elementary steps of complex processes occurring in solution, such as the oxidation of sulphur dioxide.

Graphical Abstract

1. Introduction

Pollution and other environmental issues are typically associated with the atmospheric emissions of exhaust flue gases produced by power plants and industries [1]. Different technologies, collectively known as flue gas cleaning processes, attempt to mitigate the release of greenhouse gases deriving from the burning of coal to generate electrical power [2]. Most efforts in this field are aimed at planning pollutant-control strategies to reduce sulphur dioxide which is referred to as the main precursor of acid rainfalls and atmospheric particulate [3,4,5]. To this end, the European Union established the 2016/2284/UE Regulation that intends to progressively reduce SO2 emissions until 2029 and for the next few years [6].
Among the flue gas desulphurization (FGD) methods, the wet scrubbing system is a low-cost and simple technology based on the reaction between SO2 and an alkaline sorbent, typically limestone [7,8,9]. Although engineers mostly design separate air-cleaning devices for individual gas emission removal, the search for multi-pollutant control systems would reduce the need for large installation areas and operation costs [10]. To this end, sodium chlorite (NaClO2) is one of the most effective reagents for the simultaneous removal of oxides of sulphur (SOx) and nitrogen (NOx) [11]. The addition of this compound to seawater solution has been recently exploited to improve the elimination of SO2 and favour the development of environmentally friendly seawater-based FGD [12].
The strong oxidative properties of NaClO2 allows the conversion of sulphites (SO32−) produced by SO2 absorption to the harmless sulphates (SO42−) that are easily solubilized in water and thus removed [13,14]. Nevertheless, many factors can affect the outcome of the scrubbing process (e.g., pH, temperature, oxidant concentration, oxidant/gas contact time, volumetric gas, liquid flow rates) and the influence of these parameters has to be carefully evaluated in the design of the operating systems [15,16,17]. For instance, solution salinity is known to increase SO2 absorption efficiency, and under alkaline conditions needed for SO2/SO3 conversion, the occurrence of a gas-solid interface reaction between SO2 and NaClO2 gives rise to the formation of Cl·, ClO· and OClO· chlorinated species which may enhance the concomitant NO oxidation in multi-pollutant removal plants [18,19]. On the other side, the above-mentioned factors can contribute to masking the intrinsic reactivity of NaClO2 towards sulphur dioxide preventing the elucidation of the mechanistic details that lead to the oxidation of SO2 and the formation of collateral products.
A successful strategy to avoid solution interfering effects and investigate the chemical processes at a strictly molecular level consists in performing gas-phase studies by mass spectrometry [20,21,22,23,24,25,26]. This technique is one of the most routinely employed for analytical purposes in a plethora of research fields spanning, inter alia, from foods and drugs to biology [27,28,29,30,31,32,33,34] or from geology to atmospheric chemistry [35,36,37,38,39]. Less well known is the use of mass spectrometry in fields such as catalysis, nevertheless, in the last years, mass spectrometry has been widely employed to assess the elementary steps of a chemical transformation by unravelling mechanistic pathways and elucidating the factors which affect the reaction outcome [40,41,42,43,44,45,46]. Accordingly, ion-molecule reaction (IMR) experiments were largely intended to investigate the reactivity of ionic reagents generated at their ground state towards neutral species under single-collision conditions. The gas-phase reaction of free ClO and ClO2 anions towards SO2 has actually provided important information on the intrinsic properties of naked chlorite leading to the oxidation of sulphur dioxide to SO3, SO3·− and SO4·, with the concomitant formation of the chlorinated species ClO, ClO·, and Cl· [47]. These reaction channels, respectively referred to as oxygen-atom (OAT), oxygen-ion (OIT), and double oxygen transfer (DOT), may represent simplified models of large-scale reactions occurring in the atmosphere or involved in the flue-gas desulphurization processes.
In addition, electrospray ionization mass spectrometry has been long-time devoted to the study of salt speciation [48,49,50] showing its capability in controlling the size and charge of cluster ions. As a result, ionic clusters can be considered miniaturized systems to investigate the intrinsic features of matter aggregation phenomena [51,52]. Accordingly, the study of the gas-phase reactions of SO2 with positive and negative carbonate cluster ions contributed to highlighting the major role of the charge in the kinetics of smallest clusters, as well as the different reactivity when charged cluster are ligated to a NaOH molecule [53]. Indeed, a point-charge ligand can generate oriented external electric fields able to change thermodynamics and kinetics of a gas-phase thermal process by controlling the reaction mechanism, efficiencies, and product distribution [54,55,56].
Continuing with our studies focused on the chemistry of sulphur dioxide [57,58,59,60,61,62], here we report on the gas-phase reactivity of negatively charged chlorite cluster ions, [NaL·ClO2] (L = ClOx with x = 0–3), towards SO2 investigated by ion-molecule reaction experiments. In this way, the effect of the ligation of a neutral molecule to ClO2 that changes the ion size and charge distribution of the cluster has been evaluated based on the known reactivity of naked ClO2 species with SO2.

2. Results and Discussion

Oxo-halogenated ions investigated in this work were generated by the negative electrospray ionization of NaClO2 solutions typically yielding a series of singly-charged cluster ions in which NaClO2 is clustered to the ClO2 anion to form aggregates resembling the general formula [(NaClO2)n·ClO2], n varying from 1 to 4 in the m/z range 100–500 (Figure S1). Aggregation phenomena are indeed characteristic of electrosprayed saline compounds [49] and are influenced by the solute concentrations and source parameters [53]. Furthermore, the electric field applied between the capillary and the skimmer plate accounts for the occurrence of electrochemical reactions at the conductive contact-solution interface near the ES emitter [63]. The detection of ClOx (x = 0, 1, 3) anions in addition to the ClO2 parent species suggests the effective occurrence of in-source redox processes. For x = 1 and 3, the corresponding ClO and ClO3 anions do not undergo significant aggregation phenomena. On the contrary, Cl anions promote aggregation with NaClO2 to form [(NaClO2)n·Cl] ions (n = 1–5), and mixed clusters of general formula [NaxClyOz] were also identified as minor species, as shown in the Supplementary Materials (Figure S1). The simplest ClO2 clusters for n = 1 were found at m/z 125 and 157 and respectively attributed to the 35chlorine isotopologue of [NaCl·ClO2] and [NaClO2·ClO2] species. The assignment was based on the distinctive 35/37Cl isotope pattern and on the corresponding collision-induced dissociation (CID) mass spectra. The ion [Na35Cl·35ClO2] at m/z 125 predominantly fragments by losing a Na35Cl neutral counterpart giving rise to the 35ClO2 daughter ion at m/z 67 (Figure 1a). The gas-phase decomposition of the corresponding 35/37Cl isotopomer (m/z 127) predictably leads to the formation of an equal ratio of 35ClO2 and 37ClO2 fragments at m/z 67 and 69, respectively (Figure 1b). The parent ion can be therefore described as a complex of the type [Cl·Na·ClO2] in which both the chloride (Cl) and chlorite (ClO2) anions are coordinated to the sodium cation (Na+). In particular, the chlorite moiety is reasonably consistent with an OClO species rather than with the more stable ClOO isomer, the presence of which can be excluded considering the structure of the precursor salt, NaClO2, and the high energy barrier to the isomerization, calculated to be 51.1 kcal·mol−1, [47] which cannot be overcome by the ions during the ionization process.
Similarly, the CID mass spectrum of the ionic species [Na35ClO2·35ClO2] at m/z 157 shows the only daughter ion 35ClO2 at m/z 67, arising from the loss of a NaClO2 neutral counterpart (Figure 2a). Accordingly, the gas-phase decomposition of the isotopomer at m/z 159 leads to the formation of an equal ratio of 35ClO2 and 37ClO2 fragment ions at m/z 67 and 69 (Figure 2b), accounting for the symmetrical complex [ClO2·Na·ClO2].
Each ionic species described above was in turn isolated into the ion trap and exposed to unreactive gas (He) over long accumulation times. Since no remarkable signal loss occurred, these ions can be considered rather stable gaseous chlorine-based aggregates. When reacted with SO2, they showed a noteworthy reactivity. In the following, the reactivity of selected cluster ions, [Cl·Na·ClO2] and [ClO2·Na·ClO2], will be described in depth, starting from the simplest [Cl·Na·ClO2]. At the occurrence, the formula of the reacting species is written with the sodium cation in the centre, to highlight the reactive anionic units. Similar to the reactions observed with the non-clustered ClO2 ions [47], both [Cl·Na·ClO2] and [ClO2·Na·ClO2] cluster ions promote oxygen-atom transfer (OAT), oxygen-ion transfer (OIT), and double oxygen transfer (DOT) towards SO2. The main difference with the free ClO2 is that when SO2 is oxidised, the oxidised products predominantly remain in the cluster and are not released as free species. Accordingly, the whole mechanistic picture of the reactions between [Cl·Na·ClO2] and [ClO2·Na·ClO2] anions towards SO2 was outlined by identifying direct and consecutive pathways, measuring the rate constants for each reaction channel, and structurally characterizing the ionic products by CID experiments.

2.1. Reactivity of [Cl·Na·ClO2] Cluster Anion

[Cl·Na·ClO2] cluster anions react with SO2 at room temperature giving rise to the products shown in Scheme 1, through a complex series of parallel and consecutive reactions. A kinetic plot showing the time progress of the reaction is displayed in Figure 3. The identity of the ionic products from reactions 1–5 has been probed by collision-induced dissociation as discussed in the following. As reported in Table 1, the reaction of [Cl·Na·ClO2] has a rate constant (kdec) of 2.88 × 10–10 (±30%) cm3 s−1 mol−1 and an efficiency (k/kcoll) of 24.2%.
Although the larger size of [Cl·Na·ClO2] is predictably responsible for the decrease of the overall reaction rate compared to that of naked ClO2 (2.88 vs. 9.10 × 10−10 cm3 s−1 mol−1), the intrinsic reactivity of the two ionic species is comparable, except for small differences in the branching ratios of the three oxygen transfer reactions. For the sake of clarity, the reactivity, OIT, OAT and DOT, is indicated in each reaction channel.
The main reaction of [Cl·Na·ClO2] leads to the formation of the ionic product [Cl·Na·SO3]· at m/z 138 and a ClO· radical species (Equation (1)). The reaction proceeds quickly with a rate constant k1 of 2.13 × 10−10 (±30%) cm3 s−1 mol−1 (Table 2), and a branching ratio of 74.2% (Table 1). The collision-induced dissociation of the product ion at m/z 138 gives rise to the SO3 ion at m/z = 80 (Figure S3) through the loss of a neutral NaCl consistent with a [Cl·Na·SO3]· connectivity, hinting at the occurrence of an intracluster oxidation of SO2 to SO3 through an oxygen ion transfer (OIT) process.
Furthermore, [Cl·Na·SO3]· was found to be unreactive towards SO2 thus confirming the presence of the two notoriously inert Cl and SO3· moieties [47]. The Cl anion only plays a spectator role, whereas the sodium cation is reasonably involved in the coordination of both Cl and SO3· anions.
A minor channel gives rise to [Cl·Na·ClO] at m/z 109 and SO3 (Equation (2)), with a rate constant k2 of 2.51 × 10−11 (±30%) cm3 s−1 mol−1 (Table 2), and a branching ratio of 8.8% (Table 1). The product ion [Cl·Na·ClO] at m/z 109 resembles an aggregate in which a Cl spectator anion and a ClO moiety are both coordinated to sodium cation, as evidenced by its CID mass spectrum. Through this path, SO2 is therefore oxidised to SO3 by an oxygen atom transfer (OAT) reaction. Once formed, [Cl·Na·ClO] displays the distinctive reactivity of the surrounding ClO moiety towards SO2 [47] that consists in a further SO2 to SO3 conversion (Equation (2.1)), through a second OAT process, and in an intracluster reaction giving [Cl·SO3] at m/z = 115 through an OIT process (Equation (2.2)). The rate constants of the two competitive processes are respectively k2.1 = 7.53 × 10−10 (±30%) and k2.2 = 7.43 × 10−11 (±30%) cm3 s−1 mol−1 (Table 2). Not surprisingly, the OAT undergone by [Cl·Na·ClO2] (Equation (2)) is slower with respect to the same process undergone by [Cl·Na·ClO] (Equation (2.1)), reflecting the different reactivity of the free ClO2 and ClO species [47]. The first preferably oxidises SO2 through an OIT process, whereas the OAT is faster in the case of ClO.
Finally, the [Cl·Na·ClO2] parent ion is involved in different reactions collectively responsible for a double oxygen transfer (DOT) to SO2 with the formation of product ions containing a sulphate anion, SO4· (Equations (3)–(5)). The sulphate moiety can be either found as a clustered ion, as in Equations (3) and (4), or it can be a free anion as in Equation (5). In Reactions (3) and (4), upon the oxidation of SO2 to SO4·, a NaCl or Cl· neutral moieties are respectively released. In any case, it is formed through an overall O2 transfer and the DOT processes account for a branching ratio of 17.0% (Table 1).
The ionic product at m/z 154 (Equation (3)) is consistent with a [Cl·Na·SO4] structure according to its fragmentation into SO4· ion at m/z = 96 (Figure S4) and loss of the neutral NaCl. The reaction occurs with a rate constant k3 of 3.06 × 10−11 (±30%) cm3 s−1 mol−1 and represents the main DOT path. Alternatively, the SO4· moiety can remain attached to the Cl· radical, and releasing a NaCl moiety leads to the product ion [Cl·SO4] at m/z 131 with a k4 of 1.47 × 10−11 (±30%) cm3 s−1 mol−1 (Equation (4)). According to the electron affinity values for SO4 (EA = 5.10 eV) and Cl (EA = 3.61 eV) [64], the negative charge of the [Cl·SO4] product ion is mostly located on the SO4 moiety, as confirmed by the dissociation of this cluster into SO4· ion at m/z = 96 (Figure S5).
Finally, SO4· is also generated as a free ion through reaction 5 with a k5 of 3.4 × 10−12 (±30%) cm3 s−1 mol−1. Not surprisingly, the free SO4· ion is the least abundant product formed through the DOT paths. In the clustered species, [Cl·SO4] and [Cl·Na·SO4]·, the negative charge can be more favourably dispersed in larger species.
The comparison with the reactivity of the free ClO2 anion shows that also with the [Cl·Na·ClO2] clustered anions the OIT remains the main reaction channel. When ClO2 was reacted with SO2, the small differences in the electron affinities between ClO (EA of 2.27 eV) and SO3 (EA of 2.06 eV) [64] only resulted in close energies (−24.6 and −25.6 kcal mol−1) calculated for the two alternative exit channels, namely SO3· (+ClO·) and SO3 (+ClO) [47]; therefore, the prevalence of the OIT process was attributed to kinetic factors. Contrarywise, thermochemical factors favoured the OAT reaction over the OIT process in the reactivity of the free ClO anion with SO2 due to the significantly higher electron affinity of Cl (EA = 3.61 eV) with respect to that of SO3 (EA = 2.06 eV). Accordingly, in the reactivity of [Cl·Na·ClO] ion, OAT (Equation (2.1)) prevails over the OIT process (Equation (2.2)) by a ratio of ca. 10/1. The in-depth theoretical analyses performed on the free ClO2 species [47] can also give some insights into the reactivity observed with the clustered chlorite anions. The potential energy surface (PES) of [OClO-SO2] system, was characterised by an early transition state that accounts for the almost barrierless formation of SO3·. In the TS, the negative charge is exclusively located on the preformed SO3 group (1.02 e), that is prone to rapid dissociation into the sulphite radical anion, and that strongly competes with the OAT and DOT processes. The formation of SO3 and SO4· occurs through common intermediates, found on the double well PES, which dissociate reflecting the endothermicity of the two processes. This theoretical analysis is well suited to also explain the reactivity observed for the ligated [Cl·Na·ClO2] cluster ions with SO2, and that of the other ligated species described in the following sections. The NaCl ligand does not affect the outcome of the oxidation reactions. Rather, it seems to have the effect of spreading the charge on the cluster, eventually lowering the reaction rate.
Overall, an increase of DOT and OAT processes at the expense of OIT channel is evidenced for the [Cl·Na·ClO2] cluster ion with respect to the non-clustered ClO2 anion.

Effect of the Ligand

To deeply investigate the role of the NaCl ligand in the reactivity of [Cl·Na·ClO2] ion towards SO2, Cl was first replaced by X anion (X = F, Br, I) to form the corresponding [X·Na·ClO2] reactive species and subsequently Li+ was inserted in place of Na+ to evaluate the role of the cation. Only non-redox-active ligands were used in order to make a comparison with the effect of salinity in solution, where it is known that the increase in ionic strength determines an increase in the absorption efficiency of SO2, [18].
As reported in Table 1, the overall rate constant (kdec) increases with the charge density of X anion, reaching the highest value of 3.75 × 10−10 (±30%) cm3 s−1 mol−1 with the smallest F anion (ion radius = 136 pm) and the lowest value of 1.85 × 10−10 (±30%) cm3 s−1 mol−1 with the largest I anion (ion radius = 216 pm) (Table 1) [65].
Regarding the branching ratios of the three reaction channels, [I·Na·ClO2] and [Br·Na·ClO2] show a reactivity distribution comparable to that of the [Cl·Na·ClO2] ion. On the contrary, an increase of the OIT process at the expense of the OAT and DOT channels was reported for [F·Na·ClO2] cluster species. As a consequence of the high charge density on the fluoride ion, the [F·Na·SO3]· ionic product arising from the OIT process adds an SO2 molecule giving rise to a labile adduct of the type [F·Na·SO3·SO2]· never observed with the other reactant ions. The OAT process increases down the halogen series, whereas the opposite occurs with the OIT process.
Passing to the cation effect, an opposite trend was observed, as the reaction rate decreases by increasing the positive charge density on the metal. The overall rate constant for the [Cl·Li·ClO2] + SO2 reaction is indeed almost three times lower than the corresponding value for the [Cl·Na·ClO2] + SO2 system, highlighting the central role of an external electric field in modelling the reaction kinetics. [54,55,56] Minor effect of the charge density is instead reported for the general reactivity scheme of [Cl·Li·ClO2] anion.
The main role played by the different ligands described above might be due to the spreading the negative charge of ClO2 within the cluster, the effect of which is reflected in the oxidative capacity of ClO2 ion: the higher the charge density of the ligand, the faster the reaction. Passing from F to I, the former forms a tighter ion pair with Na+, making the chlorite anion more available to oxidise sulphur dioxide. A second effect concerns the steric hindrance to the approach of SO2 due to the neutral ligand, whereby ligands of larger dimensions lead to a decrease in the overall reaction rate. An opposite effect is observed with the lithium cation, whose small size increases the interactions with both Cl and ClO2, reducing the oxygen transfer rate of the latter. The effect of the non-redox ligands here tested is different to that played in solution, where an increase in ionic strength (i.e., the salinity) has the effect of increasing the SO2 absorption efficiency by the solutions and therefore the overall efficiency of wet scrubbing processes [18].

2.2. Reactivity of [ClO2·Na·ClO2] Cluster Anion

Passing to [ClO2·Na·ClO2] cluster ion, the reactive channels reported in Scheme 2 have been observed from the reaction with SO2. The identity of the ionic products from Reactions (6)–(11) have been probed by collision−induced dissociation experiments as discussed below.
The reaction of [ClO2·Na·ClO2] with SO2 at 298 K is fast and efficient showing an overall rate constant (kdec) of 7.48 × 10−10 (±30%) cm3 s−1 mol−1 at 298 K. This value is only 0.82 times lower than the corresponding one for the bare ClO2 species, whereas the efficiencies (k/kcoll) of the two processes are similar (66.0% vs. 63.8%, Table 1). The intrinsic reactivity of the [ClO2·Na·ClO2] anion towards SO2 is comparable to that of the [Cl·Na·ClO2] species, as demonstrated by rather close branching ratios for the three reaction pathways (Table 1). Nonetheless, the concomitant presence of two reactive ClO2 moieties in the [ClO2·Na·ClO2] ion gives rise to an intricate reaction picture as shown in the above Scheme 2 and in the kinetic plot of Figure 4.
The main reaction of [ClO2·Na·ClO2] ion at m/z 157 leads to the ionic product at m/z 170, attributed to [ClO2·Na·SO3]·, and a ClO· radical species (Equation (6)). The CID mass spectrum of the ionic product at m/z 170 shows a major dissociation into SO3·, which accounts for a [ClO2·Na·SO3]· structure (Figure S7). As in the case of Cl-clustered species [Cl·Na·ClO2], the main reaction of [ClO2·Na·ClO2] consists of an oxygen ion transfer, resulting in a fast intracluster oxidation of SO2. The rate constant k6 is 6.26 × 10−10 (±30%) cm3 s−1 mol−1 (Table 3), and a branching ratio of 81.8% (Table 1). The intracluster formation of SO3· gives rise to a negatively charged product in which one of the two ClO2 moieties only played a spectator role, whereas the sodium cation is reasonably involved in the coordination of the ClO2 and SO3· anions. However, the presence of a residual ClO2 moiety in the product ion [ClO2·Na·SO3]· is responsible for the consecutive reactivity of this species, which is deeply discussed in the next paragraph (vide infra). The complete reactive scheme of [ClO2·Na·ClO2], integrated with the reactivity of [ClO2·Na·SO3]·, is reported in the Supplementary Materials (Scheme S1), showing the complex and intricate reactivity of an only apparently simple species.
A second path, indeed a minor one, leads to an ionic product at m/z 141 and SO3 (Equation (7)), formed through an OAT from one of the two ClO2 unit to SO2. The branching ratio is only 4.8% (Table 1) and a rate constant k7 of 3.64 × 10−11 (±30%) cm3 s−1 mol−1 (Table 3). The ionic product at m/z 141, resembles an aggregate in which a ClO2 anion and a ClO reactive moiety are both coordinated to sodium cation, although its fragmentation into the ClO3 species at m/z 83 seems to account for a [Cl·Na·ClO3] structure (Figure S8). Nonetheless, the [Cl·Na·ClO3] ion obtained by spraying a NaCl/NaClO3 (1:1) millimolar solution resulted to be not reactive towards SO2 (Figure S9).
Therefore, it seems more likely to attribute the [ClO·Na·ClO2] connectivity to the ion at m/z 141 that rearranges to [Cl·Na·ClO3] upon CID, thus demonstrating the interaction of sodium cation with the ClO2 and ClO anions, rather than Cl and ClO3 species. The presence of two potential reactive units, ClO and ClO2, make the [ClO·Na·ClO2] cluster ion quite reactive. The consecutive OAT process observed (Equation (7.1)) has a rate constant k7.1 of 9.57 × 10−10 (±30%) cm3 s−1 mol−1 (Table 3), which is much higher than k7 relative to the similar OAT process in Equation (7). Again, as for reactions 2 and 2.1, the reason lies in the different reactivity of the free chlorite and hypochlorite anions, the first undergoing faster OIT and the second faster OAT processes. The product ion at m/z 125 corresponds to the [Cl·Na·ClO2] species, as demonstrated by its characteristic fragmentation pattern and the distinctive reactivity discussed in the previous paragraph (Figures S10 and S11).
Four different DOT channels were reported for the [ClO2·Na·ClO2] parent ion. The first three pathways (Equations (8)–(10)) resemble those already described for the [Cl·Na·ClO2] ion (Equations (3)–(5)), as the same product ions at m/z 154, 131, and 96 are respectively detected. The intracluster DOT processes reported in Equations (3) and (8) occurs with similar rate constants k3 and k8, regardless the ligand [NaCl] or [NaClO2] attached to the ClO2 moiety (k3 = 3.06 vs. k8 = 2.51 × 10−11 cm3 s−1 mol−1). A significantly faster formation of the [Cl·SO4] ion at m/z 131 was reported for the [ClO2·Na·ClO2] parent species with respect to [Cl·Na·ClO2] (k9 = 4.17 vs. k4 = 1.47 × 10−11 cm3 s−1 mol−1). Again, the formation of the free [SO4]· product ion represents the lowest DOT process (k10 = 1.34 × 10−11 cm3 s−1 mol−1). In addition, a fourth DOT channel was observed only for the [ClO2·Na·ClO2] parent ion which is worthy of note. In this case, the oxidation of SO2 leads to the product at m/z 119 with a rate constant k11 of 2.22 × 10−11 cm3 s−1 mol−1 (Equation (11)) which subsequently adds a further SO2 molecule with a kadd of 5.74 × 10−11 cm3 s−1 mol−1 (Equation (11.1), Table 3). Although the structure of the ionic species at m/z 119 could not be directly probed owing to its unproductive CID, a possible [Na·SO4] formula can be reasonably supposed. The corresponding ion at m/z 119 was also obtained by electrospraying a solution of Na2SO4 which, once isolated and reacted with SO2, gave a ligated [Na·SOSO2] addition product with a rate constant consistent with kadd, of Equation (11.1), thus confirming the identity of the parent species at m/z 119 (Figure S12). The [Na·SO4] formula accounts for the oxidation of the sulphur atom of sulphur dioxide and the eventual reduction of the chlorine atoms of the [ClO2·Na·ClO2] ion. Both the ClO2 units may be involved in the reaction, in which each ClO2 anion transfers an O·− moiety to SO2 giving rise to an SO42− species and the release of two ClO radicals. This hypothesis was confirmed by replacing one of the two ClO2 anions with the similarly oxygenated, but intrinsically unreactive ClO3 anion to obtain the [ClO3·Na·ClO2] parent ion. When exposed to SO2, this species shows an intrinsic reactivity comparable to that of the [ClO2·Na·ClO2] ion with the only exception of the product at m/z 119 that was not observed, thus highlighting the involvement of both ClO2 anions in the double O· transfer.

2.3. Reactivity of [SO3·Na·ClO2]· Cluster Anion

To better investigate the consecutive reactivity of the product ion at m/z 170, arising from the [ClO2·Na·ClO2] parent species through Equation (6), the putative [ClO2·Na·SO3]· ion was isolated from the sequence 157 to 170 (MS2-isolated) and separately reacted with SO2. The reactivity observed is illustrated in Scheme 3.
The overall reaction shows a rate constant (kdec) of 3.74 × 10−10 (±30%) cm3 s−1 mol−1 and an efficiency of 33.0% at 298 K (Table 1). These values agree with those reported for the other [X·Na·ClO2] parent species analysed and having a unique ClO2 reactive moiety (Table 1). Regarding, instead, the distribution of the three reaction paths, an even more pronounced increase of the DOT channels, accounting for a total amount of 22.7%, was observed (Table 1). The time progress of the reaction is described by the kinetic plot in Figure 5 and the rate constants of each pathway are reported in Table 4.
The intracluster OIT process (Equation (12)) proceeds quickly, showing a k12 of 2.63 × 10−10 (±30%) cm3 s−1 mol−1 (Table 4) and leading to the formation of an ion at m/z 183. Unfortunately, the CID mass spectrum of this species does not allow to distinguish between a [SO3·Na·SO3] or a [Na·SOSO2] structure (Figure S14), the latter already observed as a product of the DOT process involving the [ClO2·Na·ClO2] parent ion. Nonetheless, based on the reactivity of naked ClO2 and knowing that the SO3· moiety is notoriously unreactive with SO2, it is reasonable to suppose a [SO3·Na·SO3] general formula for this species.
The [ClO2·Na·SO3]· parent ion is also involved in an OAT reaction proceeding with a k13 of 2.60 × 10−11 (±30%) cm3 s−1 mol−1 and giving rise to an ion at m/z 154 that is consistent with a [ClO·Na·SO3]· structure. The consecutive OAT reactivity of this species leading to the ion at m/z 138 (Equation (13.1); k13.1 = 1.09 × 10−10 (±30%) cm3 s−1 mol−1, Table 4, Figure S11) accounts for the presence of the surrounding reactive ClO moiety in [ClO·Na·SO3]· (m/z = 154). When MS3-isolated into the ion trap by the sequence 170 to 154 and exposed to SO2, the ionic species at m/z 154 is only partially reactive towards this neutral gas. A portion of the ionic population at m/z 154 survives over time, hinting at the concomitant presence of the unreactive [Cl·Na·SO4]· species together with the [ClO·Na·SO3]· isobaric ion that is consumed in the consecutive reaction (Figure S15). The [Cl·Na·SO4]· species can reasonably arise from a direct intracluster DOT channel (Equation (13b)), as previously observed in analogous processes involving the [Cl·Na·ClO2] and [ClO2·Na·ClO2] parent ions (Equations (3) and (8)). As a result, the O2 transfer from ClO2 to SO2 triggers the release of a neutral SO3 moiety according to the electron affinity values of the species involved in the reaction [27,28,29,30,31,32,33,34,65]. Unfortunately, it was not possible to independently measure k13b, which is therefore included with that of the OAT transfer, k13. As a consequence, the branching ratio of the OAT might be slightly overestimated, at the expense of that for the DOT process, which could therefore be underestimated. Two other DOT pathways reported in Equations (14) and (15) were also previously observed for the [Cl·Na·ClO2] and [ClO2·Na·ClO2] parent clusters (Equations (4) and (9), Equations (5) and (10)). All these pathways show rather similar formation rate constants of the 10−11 cm3 s−1 mol−1 order of magnitude (Table 4).
[ClO2·Na·SO3]· also reacts with SO2, leading to the [Na·SO4] product ion at m/z 119 (Equation (16)). The reaction, showing a k16 of 4.57 × 10−11 cm3 s−1 mol−1, proceeds with an intracluster O2 transfer. Such unusual reactivity probably involves both the ClO2 anion that triggers a classic O2 transfer and the SO3· moiety that may be responsible for an electron transfer, giving rise to an SO42− anion through a concerted rearrangement. As previously reported (Equation (11.1)), the consecutive addition of an SO2 molecule to the [Na·SO4] product ion is observed, thus confirming the identity of the ion at m/z 119.
Finally, as to the reactivity of higher species such as [(NaClO2)n·ClO2], only the rate constant relative to cluster with n = 2 has been measured (Table 1), which does not appear to be affected by the number of additional NaClO2 units compared to [ClO2·Na·ClO2]. However, it was not possible to evaluate the branching ratio of the OIT, OAT and DOT processes of [(NaClO2)2·ClO2], due to the low intensity signals relative to parent cluster ions, and to the complex array of peaks resulting from the reaction with SO2.

3. Materials and Methods

Mass spectrometric experiments were carried out on an LTQ-XL linear ion-trap mass spectrometer (Thermo Fisher Scientific) that was in-house modified to perform ion-molecule reactions (IMR) [53]. Water-acetonitrile (1:1) solutions of NaClO2 at millimolar concentrations were injected into the ESI (electrospray ionization) source of the instrument at a flow rate of 5 μL min−1 via the on-board syringe pump and using nitrogen as sheath and auxiliary gas (flow rate = 11 and 2 arbitrary units respectively, a. u. ~0.37 L min−1). Other [ML·ClO2] cluster anions (L = F, Br, I, ClO3; M = Li, Na) investigated in this work were obtained from millimolar solutions of 1:1 ML and NaClO2 salts dissolved in water-acetonitrile (1:1). To generate chlorite cluster ions and optimize the ion transmission, spray voltage was tuned in the 1.8–3.2 kV range, whereas the capillary temperature was set at 275 °C. The distribution of the ionic aggregates strictly depends on the capillary and tube lens voltage. Hence, these parameters were in turn optimized to increase the signal intensity of the parent ion under investigation.
Once generated, reagent ions were transferred to the vacuum region of the trap, mass-to-charge isolated and reacted with sulphur dioxide. Each reaction product was then mass selected by a further step of isolation, that is typical of MSn experiments performed by Ion Trap mass spectrometers, and the consecutive reactivity of these species was probed towards SO2 to unravel a complete reaction picture. Furthermore, the ionic reactants and products were structurally characterized by collision-induced dissociation (CID) experiments performed by increasing the energy of mass-selected ions in the presence of helium as collision gas (pressure of ca. 3 × 10−3 Torr). Depending on the species of interest, normalized collision energies ranging between 20% and 40% were typically applied with an activation time of 30 ms. Ions were isolated with a window of 1 m/z, and the Q value was optimized to ensure stable trapping fields for all the ionic species under investigation.
Sulphur dioxide was introduced into the trap through a deactivated fused silica capillary that enters the vacuum chamber from a 6.25 mm hole placed in the backside of the mass spectrometer. The pressure of the neutral gas was kept constant by a metering valve and measured by a Granville-Phillips Series 370 Stabil Ion Vacuum Gauge. Owing to the position of the Pirani gauge, the actual sulphur dioxide pressure was estimated after calibration of the pressure reading [66]. Typical pressures of SO2 ranged between 1.1 × 10−7 and 8.0 × 10−7 Torr, the uncertainty was estimated to be ± 30%. The signals of the ionic reactant and products were monitored over time as a function of the neutral concentration and for each reaction time an average of 10 scan acquisitions was recorded. The normalized collision energy was set to zero and the activation Q value was optimized to ensure stable trapping fields for all the ions. Xcalibur 2.0.6 software was used to acquire all the displayed mass spectra.
All the reactions can be regarded as pseudo-first-order processes due to the excess of neutral gas relative to the reactant ion in the trap. DynaFit4 software package [67] was used to perform nonlinear least squares regression to simultaneously fit reactant and products concentration versus time. Experimental data from the kinetic analyses were fitted to a mathematical model consistent with the postulated reaction mechanism. To check the validity of the kinetic schemes, the obtained unimolecular rate constants were used to simulate the time progress of the reactions using the kinetic simulation function contained in DynaFit4. Bimolecular rate constants k (cm3 molecule−1 s−1) were obtained dividing the pseudo-first-order constants (s−1) by the concentration of neutral reagent gas. The branching ratios between the three channels (OIT, OAT, DOT) were calculated from the constants of formation of the primary direct products for each reactive species. The reaction efficiency was calculated as the ratio of the bimolecular rate constant k to the collision rate constant (kcoll), according to the average dipole orientation (ADO) theory [68]. To ensure the accuracy of the k values, approximately 15 independent measurements for each precursor ion were performed on different days over a sevenfold neutral pressure range. The standard deviation in the fitting parameters of the kinetic modelling used is usually evaluated between 10–20%, whereas the uncertainty attached to the measurement of the neutral pressure is typically evaluated ±30%.

4. Conclusions

Mass spectrometry has been used to elucidate the gas-phase reactivity of [NaL·ClO2] cluster anions (L = ClOx with x = 0–3) with sulphur dioxide. These charged species can be taken as simplified models of large-scale reactions occurring in solution or in the flue-gas desulphurization processes, which are accomplished by sodium chlorite solutions.
The kinetic analysis has shown that SO2 was efficiently oxidised by oxygen atom transfer (OAT), oxygen ion transfer (OIT), and double oxygen transfer (DOT) respective to SO3, SO3· and SO4·. In the case of OIT and DOT processes, an intracluster reaction was observed, by which the oxidised ionic forms of SO2, namely SO3· and SO4·, remain within the cluster and are not released as a free species. The results here reported show that when ClO2 is ligated to a non−redox active molecule, the complexation leads to a moderate reduction in the rate of oxidation processes, without substantially influencing the branching ratio. This effect contrasts, but not surprisingly, with what is observed in solution, where dissolved salts increase the SO2 capture by increasing ionic strength of the solutions. In the gas phase, the direct and strong interaction of the chlorite anion with the ligand is detrimental to the reaction rate. However, the effect of redox active ligands, metallic or metal-free, could be quite different, as suggested by the reactivity observed with [ClO2·Na·ClO2], in which the second reactive ClO2 moiety succeeds in increasing the rate of the oxidation. Therefore, the ligation with a redox active group, different from the chlorite one, could succeed in tuning the oxidation processes.

Supplementary Materials

The following are available online. Figure S1: Full-scan mass spectrum of a NaClO2 salt solution; Figure S2: Ion-molecule reaction between isolated [Cl·Na·ClO2] cluster ion and SO2; Figure S3: CID mass spectrum of [Cl·Na·SO3]· ion at m/z 138; Figure S4: CID mass spectrum of [Cl·Na·SO4]· ion at m/z 154; Figure S5: CID mass spectrum of [Cl·SO4] ion at m/z 131; Figure S6: ion-molecule reaction between isolated [ClO2·Na·ClO2] cluster ion at m/z 157 SO2; Figure S7: CID mass spectrum of [ClO2·Na·SO3]· product ion at m/z 170; Figure S8: CID mass spectrum of (a) [ClO·Na·ClO2] product ion at m/z and (b) [Cl·Na·ClO3] standard ion at m/z 141; Figure S9: mass spectrum of the ion-molecule reaction of (a) [ClO·Na·ClO2] ion at m/z and (b) [Cl·Na·ClO3] standard ion at m/z 141 towards SO2; Figure S10: mass spectrum of the ion-molecule reaction between [Cl·Na·ClO2] consecutive product ion at m/z 125, MSn-isolated from the reaction sequence m/z 157 to m/z 141 to m/z 125 and SO2; Figure S11: magnified plot of the kinetic reported in Figure 4; Figure S12: mass spectrum of the ion-molecule reaction of (a) [Na·SO4] product ion at m/z 119, MSn-isolated, and (b) [Na·SO4] standard ion at m/z 119 towards SO2; Figure S13: mass spectrum of the ion-molecule reaction between [ClO2·Na·SO3] product ion at m/z 170, MSn-isolated from the reaction sequence m/z 157 to m/z 170 and SO2; Figure S14: CID mass spectrum of product ion at m/z 183; Figure S15: ion-molecule reaction of (a) a mixed ionic population at m/z 154, MSn-isolated from the reaction sequence m/z 157 to m/z 170 to m/z 159 and (b) [Cl·Na·SO4]− ion at m/z 154, MSn-isolated from the reaction of [Cl·Na·ClO2]− reactant ion towards SO2; Figure S16: mass spectrum of the ion-molecule reaction between [Na·SO4] product ion at m/z 119 and SO2; Scheme S1: Complete reaction sequences of [ClO2·Na·ClO2] ion at m/z 157 and of its product ion [ClO2·Na·SO3]· ion at m/z 170.

Author Contributions

Conceptualization, A.T. and C.S.; methodology, A.T.; validation, A.T., F.P. and C.S.; Funding acquisition, G.d.P.; investigation C.S.; data curation, C.S.; Supervision G.d.P.; Visualization C.S.; writing—original draft preparation, A.T. and C.S.; writing—review and editing A.T., C.S., F.P. and G.d.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Sapienza Rome University “Progetti di Ateneo” (Projects n. RM11816428291DFF, RM11715C81A54060, AR21916B746C06BB). C.S. thanks the Dipartimento di Chimica e Tecnologie del Farmaco, Sapienza Rome University, for a post-doc position within the project Dipartimenti di Eccellenza-L. 232/2016.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples are not available from the authors.

References

  1. Lima, R.; Bachmann, R.T. Pollutant emissions from modern incinerators. Int. J. Environ. Pollut. 2002, 18, 336. [Google Scholar] [CrossRef]
  2. Vehlow, J. Air pollution control systems in WtE units: An overview. Waste Manag. 2015, 37, 58–74. [Google Scholar] [CrossRef]
  3. Muralikrishna, I.V.; Manickam, V. Environmental Management; Elsevier Inc.: Amsterdam, The Netherlands, 2017. [Google Scholar]
  4. Seinfeld, J.H.; Pandis, S.N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 3rd ed.; John Wiley and Sons: New York, NY, USA, 2016. [Google Scholar]
  5. Dupart, Y.; King, S.M.; Nekat, B.; Nowak, A.; Wiedensohler, A.; Herrmann, H.; David, G.; Thomas, B.; Miffre, A.; Rairoux, P.; et al. Mineral dust photochemistry induces nucleation events in the presence of SO2. Proc. Natl. Acad. Sci. USA 2012, 109, 20842–20847. [Google Scholar] [CrossRef] [Green Version]
  6. Directive (EU) 2016/2284 of the European Parliament and of the Council of 14 December 2016 on the Reduction of National Emissions of Certain Atmospheric Pollutants, Amending Directive 2003/35/EC and Repealing Directive 2001/81/EC.
  7. Woodard & Curran, Inc. Industrial Waste Treatment Handbook, 2nd ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2006. [Google Scholar]
  8. Hu, G.; Dam-Johansen, K.; Wedel, S.; Peterhansen, J. Review of the direct sulfation reaction of limestone. Prog. Energy Combust. Sci. 2006, 32, 386–407. [Google Scholar] [CrossRef]
  9. Koralegedara, N.H.; Pinto, P.X.; Dionysiou, D.; Al-Abed, S.R. Recent advances in flue gas desulphurization gypsum processes and applications—A review. J. Environ. Manag. 2019, 251, 109572. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, Y.; Guo, T.; Chen, Z. Experimental study on simultaneous desulfurization and denitrification from flue gas with composite absorbent. Environ. Prog. Sustain. Energy 2010, 30, 216–220. [Google Scholar] [CrossRef]
  11. Adewuyi, Y.G.; He, X.; Shaw, H.; Lolertpihop, W. Lolertpihop, Simultaneous absorption and oxidation of NO and SO2 by aqueous solutions of sodium chlorite. Chem. Eng. Commun. 1999, 174, 21–51. [Google Scholar] [CrossRef]
  12. Flagiello, D.; Di Natale, F.; Erto, A.; Lancia, A. Wet oxidation scrubbing (WOS) for flue-gas desulphurization using sodium chlorite seawater solutions. Fuel 2020, 277, 118055. [Google Scholar] [CrossRef]
  13. Chu, H.; Chien, T.-W.; Twu, B.-W. The absorption kinetics of NO in NaClO2/NaOH solutions. J. Hazard. Mater. 2001, B84, 241–252. [Google Scholar] [CrossRef]
  14. Chu, H.; Chien, T.-W.; Twu, B.-W. Simultaneous absorption of SO2 and no in a stirred tank reactor with NaClO2/NaOH solutions. Water Air Soil Pollut. 2003, 143, 337–350. [Google Scholar] [CrossRef]
  15. Pourmohammadbagher, A.; Jamshidi, E.; Ale-Ebrahim, H.; Dabir, S. Study on simultaneous removal of NOx and SO2 with NaClO2 in a novel swirl wet system. Ind. Eng. Chem. Res. 2011, 50, 8278–8284. [Google Scholar] [CrossRef]
  16. Chien, T.-W.; Chu, H. Removal of SO2 and NO from flue gas by wet scrubbing using an aqueous NaClO2 solution. J. Hazard. Mater. 2000, 80, 43–57. [Google Scholar] [CrossRef]
  17. Krzyżyńka, R.; Zhao, Y.; Hutson, N. Absorption of NOx, SO2, and mercury in a simulated additive-enhanced wet flue gas desulphurization scrubber. Polish J. Environ. Stud. 2010, 19, 1255–1262. [Google Scholar]
  18. Andreasen, A.; Mayer, S. Use of Seawater Scrubbing for SO2 Removal from Marine Engine Exhaust Gas. Energy Fuels 2007, 21, 3274–3279. [Google Scholar] [CrossRef]
  19. Byun, Y.; Hamilton, I.P.; Tu, X.; Shin, D.N. Formation of chlorinated species through reaction of SO2 with NaClO2 powder and their role in the oxidation of NO and Hg0. Environ. Sci. Pollut. Res. 2014, 21, 8052–8058. [Google Scholar] [CrossRef]
  20. Coelho, F.; Eberlin, M.N. The bridge connecting gas-phase and solution chemistries. Angew. Chem. Int. Ed. 2011, 50, 5261–5263. [Google Scholar] [CrossRef]
  21. Zhu, W.; Yuan, Y.; Zhou, P.; Zeng, L.; Wang, H.; Tang, L.; Guo, B.; Chen, B. The expanding role of electrospray ionization mass spectrometry for probing reactive intermediates in solution. Molecules 2012, 17, 11507–11537. [Google Scholar] [CrossRef]
  22. Fabris, D. Mass spectrometric approaches for the investigation of dynamic processes in condensed phase. Mass Spectrom. Rev. 2004, 24, 30–54. [Google Scholar] [CrossRef]
  23. Troiani, A.; Rosi, M.; Garzoli, S.; Salvitti, C.; de Petris, G. Iron-Promoted C-C Bond Formation in the Gas Phase. Angew. Chem. Int. Ed. 2015, 54, 14359–14362. [Google Scholar] [CrossRef]
  24. Troiani, A.; De Petris, G.; Pepi, F.; Garzoli, S.; Salvitti, C.; Rosi, M.; Ricci, A. Base-Assisted Conversion of Protonated D -Fructose to 5-HMF: Searching for Gas-Phase Green Models. ChemistryOpen 2019, 8, 1190–1198. [Google Scholar] [CrossRef] [Green Version]
  25. Vikse, K.L.; Ahmadi, Z.; Manning, C.C.; Harrington, D.A.; McIndoe, J.S. Powerful insight into catalytic mechanisms through simultaneous monitoring of reactants, products, and intermediates. Angew. Chem. Int. Ed. 2011, 50, 8304–8306. [Google Scholar] [CrossRef] [PubMed]
  26. Gronert, S. Mass spectrometric studies of organic ion/molecule reactions. Chem. Rev. 2001, 101, 329–360. [Google Scholar] [CrossRef] [PubMed]
  27. Domínguez, I.; Frenich, A.G.; Romero-González, R. Mass spectrometry approaches to ensure food safety. Anal. Methods 2020, 12, 1148–1162. [Google Scholar] [CrossRef]
  28. Wang, X.; Wang, S.; Cai, Z. The latest developments and applications of mass spectrometry in food-safety and quality analysis. Trends Anal. Chem. 2013, 52, 170–185. [Google Scholar] [CrossRef]
  29. Cimino, P.; Troiani, A.; Pepi, F.; Garzoli, S.; Salvitti, C.; Di Rienzo, B.; Barone, V.; Ricci, A. From ascorbic acid to furan derivatives: The gas phase acid catalyzed degradation of vitamin C. Phys. Chem. Chem. Phys. 2018, 20, 17132–17140. [Google Scholar] [CrossRef]
  30. Geoghegan, K.F.; Kelly, M.A. Biochemical applications of mass spectrometry in pharmaceutical drug discovery. Mass Spectrom. Rev. 2005, 24, 347–366. [Google Scholar] [CrossRef]
  31. Loos, G.; Van Schepdael, A.; Cabooter, D. Quantitative mass spectrometry methods for pharmaceutical analysis. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20150366. [Google Scholar] [CrossRef] [Green Version]
  32. Forte, G.; Chiarotto, I.; Giannicchi, I.; Loreto, M.A.; Martinelli, A.; Micci, R.; Pepi, F.; Rossi, S.; Salvitti, C.; Stringaro, A.; et al. Characterization of naproxen–polymer conjugates for drug-delivery. J. Biomater. Sci. Polym. Ed. 2015, 27, 69–85. [Google Scholar] [CrossRef]
  33. Vestal, M.L. The Future of Biological Mass Spectrometry. J. Am. Soc. Mass Spectrom. 2011, 22, 953–959. [Google Scholar] [CrossRef] [Green Version]
  34. Finehout, E.J.; Lee, K.H. An introduction to mass spectrometry applications in biological research. Biochem. Mol. Biol. Educ. 2004, 32, 93–100. [Google Scholar] [CrossRef]
  35. de Laeter, J.R. Mass spectrometry and geochronology. Mass Spectrom. Rev. 1998, 17, 97–125. [Google Scholar] [CrossRef]
  36. Petrie, S.; Bohme, D.K. Ions in space. Mass Spectrom. Rev. 2006, 26, 258–280. [Google Scholar] [CrossRef] [PubMed]
  37. Cacace, F.; Cipollini, R.; DE Petris, G.; Troiani, A. The impervious route to the elusive HOOO anion. Int. J. Mass Spectrom. 2003, 228, 717–722. [Google Scholar] [CrossRef]
  38. Cacace, F.; de Petris, G.; Pepi, F.; Troiani, A. Direct experimental evidence for the H2O+O2 charge transfer complex: Crucial support to atmospheric photonucleation theory. Angew. Chem. Int. Ed. 2000, 39, 367–369. [Google Scholar] [CrossRef]
  39. Cacace, F.; de Petris, G.; Pepi, F.; Rosi, M.; Troiani, A. Ionization of ozone/chlorofluorocarbon mixtures in atmospheric gases: Formation and remarkable dissociation of [CHXYO3]+ complexes (X = H, Cl, F; Y= Cl, F). Chem. Eur. J. 2000, 6, 2572–2581. [Google Scholar] [CrossRef]
  40. Johnson, G.; Tyo, E.C.; Castleman, A.W. Cluster reactivity experiments: Employing mass spectrometry to investigate the molecular level details of catalytic oxidation reactions. Proc. Natl. Acad. Sci. USA 2008, 105, 18108–18113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Johnson, G.E.; Mitrić, R.; Bonačić-Koutecký, V.; Castleman, A. Clusters as model systems for investigating nanoscale oxidation catalysis. Chem. Phys. Lett. 2009, 475, 1–9. [Google Scholar] [CrossRef]
  42. Schlangen, M.; Schwarz, H. Effects of ligands, cluster size, and charge state in gas-phase catalysis: A happy marriage of experimental and computational studies. Catal. Lett. 2012, 142, 1265–1278. [Google Scholar] [CrossRef] [Green Version]
  43. Dietl, D.-C.N.; Troiani, A.; Schlangen, M.; Ursini, O.; Angelini, G.; Apeloig, Y.; de Petris, G.; Schwarz, H. Mechanistic Aspects of Gas-Phase Hydrogen-Atom Transfer from Methane to [CO].+ and [SiO].+: Why Do They Differ? Chem. Eur. J. 2013, 19, 6662–6669. [Google Scholar] [CrossRef]
  44. Schlangen, M.; Schwarz, H. Probing elementary steps of nickel-mediated bond activation in gas-phase reactions: Ligand- and cluster-size effects. J. Catal. 2011, 284, 126–137. [Google Scholar] [CrossRef]
  45. Vikse, K.L.; McIndoe, J.S. Mechanistic insights from mass spectrometry: Examination of the elementary steps of catalytic reactions in the gas phase. Pure Appl. Chem. 2015, 87, 361–377. [Google Scholar] [CrossRef] [Green Version]
  46. Vikse, K.L.; Ahmadi, Z.; McIndoe, J.S. The application of electrospray ionization mass spectrometry to homogeneous catalysis. Coord. Chem. Rev. 2014, 279, 96–114. [Google Scholar] [CrossRef]
  47. Troiani, A.; Rosi, M.; Salvitti, C.; de Petris, G. The oxidation of sulfur dioxide by single and double oxygen transfer paths. ChemPhysChem 2014, 15, 2723–2731. [Google Scholar] [CrossRef]
  48. Stewart, I.I. Electrospray mass spectrometry: A tool for elemental speciation. Spectrochim. Acta Part B 1999, 54, 1649–1695. [Google Scholar] [CrossRef]
  49. Hao, C.; March, R.E. Electrospray ionization tandem mass spectrometric study of salt cluster ions: Part 2 Salts of polyatomic acid groups and of multivalent metals. J. Mass Spectrom. 2001, 36, 509–521. [Google Scholar] [CrossRef] [PubMed]
  50. Eberlin, M.N. Electrospray ionization mass spectrometry: A major tool to investigate reaction mechanisms in both solutions and the gas phase. Eur. J. Mass Spectrom. 2007, 13, 19–28. [Google Scholar] [CrossRef]
  51. Castleman, A.W., Jr.; Keesee, R.G. Gas-phase clusters: Spanning the states of matter. Science 1988, 241, 36–42. [Google Scholar] [CrossRef]
  52. Bondybey, V.E.; Beyer, M.K. How many molecules make a solution? Int. Rev. Phys. Chem. 2002, 21, 277–306. [Google Scholar] [CrossRef]
  53. Troiani, A.; Salvitti, C.; de Petris, G. Gas-phase reactivity of carbonate ions with sulphur dioxide: An experimental study of clusters reactions. J. Am. Soc. Mass Spectrom. 2019, 30, 1964–1972. [Google Scholar] [CrossRef]
  54. Shaik, S.; Mandal, D.; Ramanan, S.S.D.M.R. Oriented electric fields as future smart reagents in chemistry. Nat. Chem. 2016, 8, 1091–1098. [Google Scholar] [CrossRef]
  55. Yue, L.; Li, J.; Zhou, S.; Sun, X.; Schlangen, M.; Shaik, S.; Schwarz, H. Control of product distribution and mechanism by ligation electric field in the thermal activation of methane. Angew. Chem. Int. Ed. 2017, 56, 10219–10223. [Google Scholar] [CrossRef]
  56. Geng, C.; Li, J.; Schlangen, M.; Shaik, S.; Sun, X.; Wang, N.; Weiske, T.; Yue, L.; Zhou, S.; Schwarz, H. Oriented external electric fields as mimics for probing the role of metal ions and ligands in the thermal gas-phase activation of methane. Dalton Trans. 2018, 47, 15271–15277. [Google Scholar] [CrossRef]
  57. de Petris, G.; Cartoni, A.; Troiani, A.; Angelini, G.; Ursini, O. Ursini, Water activation by SO2+ ions: An effective source of OH radicals. PhysChemChemPhys 2009, 11, 9976–9978. [Google Scholar] [CrossRef]
  58. de Petris, G.; Troiani, A.; Rosi, M.; Angelini, G.; Ursini, O. Selective activation of C-Cl and C-F bonds by SO+ radical cations: An experimental and computational study. ChemPlusChem 2013, 78, 1065–1072. [Google Scholar] [CrossRef]
  59. Troiani, A.; Rosi, M.; Garzoli, S.; Salvitti, C.; de Petris, G. Vanadium hydroxide cluster ions in the gas phase: Bond-forming reactions of doubly-charged negative ions by SO2-promoted V-O activation. Chem. Eur. J. 2017, 23, 11752–11756. [Google Scholar] [CrossRef] [PubMed]
  60. Troiani, A.; Rosi, M.; Garzoli, S.; Salvitti, C.; de Petris, G. Sulphur dioxide cooperation in hydrolysis reactions of vanadium oxide and hydroxide cluster dianions. New J. Chem. 2018, 42, 4008–4016. [Google Scholar] [CrossRef]
  61. Troiani, A.; Rosi, M.; Garzoli, S.; Salvitti, C.; de Petris, G. Effective redox reactions by chromium oxide anions: Sulfur dioxide oxidation in the gas phase. Int. J. Mass Spectrom. 2018, 436, 18–22. [Google Scholar] [CrossRef]
  62. Salvitti, C.; Rosi, M.; Pepi, F.; Troiani, A.; de Petris, G. Reactivity of transition metal dioxide anions MO2 (M = Co, Ni, Cu, Zn) with sulfur dioxide in the gas phase: An experimental and theoretical study. Chem. Phys. Lett. 2021, 776, 138555. [Google Scholar] [CrossRef]
  63. E Bartmess, J.; Georgiadis, R. Empirical methods for determination of ionization gauge relative sensitivities for different gases. Vacuum 1983, 33, 149–153. [Google Scholar] [CrossRef]
  64. Kuzmic, P. Program DYNAFIT for the Analysis of Enzyme Kinetic Data: Application to HIV Proteinase. Anal. Biochem. 1996, 237, 260–273. [Google Scholar] [CrossRef]
  65. Bowers, M.T.; Su, T. Interactions between Ions and Molecules; Plenum Press: New York, NY, USA, 1975. [Google Scholar]
  66. Van Berkel, G.J.; Kertesz, V. Using the Electrochemistry of the Electrospray Ion Source. Anal. Chem. 2007, 79, 5510–5520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Linstrom, P.J.; Mallard, W.G. (Eds.) NIST Chemistry WebBook; NIST Standard Reference Database Number 69; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1997. Available online: http://webbook.nist.gov (accessed on 20 October 2021).
  68. Tro, N.J. Chemistry, International Edition; Pearson Education Inc.: Upper Saddle River, NJ, USA, 2016; pp. 341–343. [Google Scholar]
Figure 1. ESI-(-) CID mass spectra of the isolated (a) [Na35Cl·35ClO2] at m/z 125 and (b) [Na35Cl·37ClO2] and [Na37Cl·35ClO2] at m/z 127.
Figure 1. ESI-(-) CID mass spectra of the isolated (a) [Na35Cl·35ClO2] at m/z 125 and (b) [Na35Cl·37ClO2] and [Na37Cl·35ClO2] at m/z 127.
Molecules 26 07114 g001
Figure 2. ESI-(-) CID mass spectra of the isolated (a) [Na35ClO2·35ClO2] at m/z 157 and (b) [Na35ClO2·37ClO2] and [Na37ClO2·35ClO2] at m/z 159.
Figure 2. ESI-(-) CID mass spectra of the isolated (a) [Na35ClO2·35ClO2] at m/z 157 and (b) [Na35ClO2·37ClO2] and [Na37ClO2·35ClO2] at m/z 159.
Molecules 26 07114 g002
Scheme 1. Reactivity scheme [Cl·Na·ClO2] ions (m/z = 125) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Scheme 1. Reactivity scheme [Cl·Na·ClO2] ions (m/z = 125) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Molecules 26 07114 sch001
Figure 3. Kinetic plot and best fit lines of the reaction of isolated [NaCl·ClO2] ions (m/z = 125) with SO2. P SO2 = 5.26 × 10−7 Torr. [Cl·Na·ClO2] (m/z 125) (R2 = 0.9991); [Cl·Na·SO3]· (m/z 138) (R2 = 0.9986); + [Cl·Na·ClO] (m/z 109) (R2 = 0.9770); [Cl·Na·SO4]· (m/z 154) (R2 = 0.9915); × [Cl·SO4] (m/z 131) (R2 = 0.9868); [SO4]· (m/z 96) (R2 = 0.9846); [Cl·Na·Cl] (m/z 93) (R2 = 0.9946); [Cl·SO3] (m/z 115) (R2 = 0.9946).
Figure 3. Kinetic plot and best fit lines of the reaction of isolated [NaCl·ClO2] ions (m/z = 125) with SO2. P SO2 = 5.26 × 10−7 Torr. [Cl·Na·ClO2] (m/z 125) (R2 = 0.9991); [Cl·Na·SO3]· (m/z 138) (R2 = 0.9986); + [Cl·Na·ClO] (m/z 109) (R2 = 0.9770); [Cl·Na·SO4]· (m/z 154) (R2 = 0.9915); × [Cl·SO4] (m/z 131) (R2 = 0.9868); [SO4]· (m/z 96) (R2 = 0.9846); [Cl·Na·Cl] (m/z 93) (R2 = 0.9946); [Cl·SO3] (m/z 115) (R2 = 0.9946).
Molecules 26 07114 g003
Scheme 2. Reactivity scheme [ClO2·Na·ClO2] ions (m/z = 157) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Scheme 2. Reactivity scheme [ClO2·Na·ClO2] ions (m/z = 157) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Molecules 26 07114 sch002
Figure 4. Kinetic plot and best fit lines of the reaction of isolated [NaClO2·ClO2] ions (m/z = 157) with SO2. P SO2 = 3.30 × 10−7 Torr. [ClO2·Na·ClO2] (m/z 157) (R2 = 0.9991); [ClO2·Na·SO3]· (m/z 170) (R2 = 0.9986); [SO3·Na·SO3] (m/z 183) (R2 = 0.9998); × [ClO2·Na·ClO] (m/z 141) (R2 = 0.9888); [Cl·Na·ClO2] (m/z 125) (R2 = 0.9965); + [Cl·Na·SO3]·− (m/z 138) (R2 = 0.9967); [Cl·Na·SO4]·− (m/z 154) (R2 = 0.9912); [SO4]·− (m/z 96) (R2 = 0.9946); [Cl·SO4] (m/z 131) (R2 = 0.9992); [Na·SO4] (m/z 119) (R2 = 0.9971).
Figure 4. Kinetic plot and best fit lines of the reaction of isolated [NaClO2·ClO2] ions (m/z = 157) with SO2. P SO2 = 3.30 × 10−7 Torr. [ClO2·Na·ClO2] (m/z 157) (R2 = 0.9991); [ClO2·Na·SO3]· (m/z 170) (R2 = 0.9986); [SO3·Na·SO3] (m/z 183) (R2 = 0.9998); × [ClO2·Na·ClO] (m/z 141) (R2 = 0.9888); [Cl·Na·ClO2] (m/z 125) (R2 = 0.9965); + [Cl·Na·SO3]·− (m/z 138) (R2 = 0.9967); [Cl·Na·SO4]·− (m/z 154) (R2 = 0.9912); [SO4]·− (m/z 96) (R2 = 0.9946); [Cl·SO4] (m/z 131) (R2 = 0.9992); [Na·SO4] (m/z 119) (R2 = 0.9971).
Molecules 26 07114 g004
Scheme 3. Reactivity scheme [ClO2·Na·SO3]· ions (m/z = 170) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Scheme 3. Reactivity scheme [ClO2·Na·SO3]· ions (m/z = 170) with SO2. Oxygen transfer channels (OIT, OAT, DOT) and bimolecular rate constants are indicated for each reaction.
Molecules 26 07114 sch003
Figure 5. Kinetic plot and best fit lines of the reaction of isolated [ClO2·Na·SO3]· ions (m/z = 170) with SO2. P SO2 = 2.60∙× 10−7 Torr. [ClO2·Na·SO3]· (m/z 170) (R2 = 0.9991); [SO3·Na·SO3]/[Na·SO4·SO2] (m/z 183) (R2 = 0.9986); [ClO·Na·SO3]· / [Cl·Na·SO4]· (m/z 154) (R2 = 0.9686); [Cl·Na·SO3]· (m/z 138) (R2 = 0.9880); × [Cl·SO4] (m/z 131) (R2 = 0.9950); [Na·SO4] (m/z 119) (R2 = 0.9973); [SO4]· (m/z 96) (R2 = 0.9946).
Figure 5. Kinetic plot and best fit lines of the reaction of isolated [ClO2·Na·SO3]· ions (m/z = 170) with SO2. P SO2 = 2.60∙× 10−7 Torr. [ClO2·Na·SO3]· (m/z 170) (R2 = 0.9991); [SO3·Na·SO3]/[Na·SO4·SO2] (m/z 183) (R2 = 0.9986); [ClO·Na·SO3]· / [Cl·Na·SO4]· (m/z 154) (R2 = 0.9686); [Cl·Na·SO3]· (m/z 138) (R2 = 0.9880); × [Cl·SO4] (m/z 131) (R2 = 0.9950); [Na·SO4] (m/z 119) (R2 = 0.9973); [SO4]· (m/z 96) (R2 = 0.9946).
Molecules 26 07114 g005
Table 1. Rate constants (cm3 s−1 mol−1), branching ratios (%Σ) and efficiencies (k/kcoll) for the reactions of [L·M·ClO2] (L = Cl, F, Br, I, ClO2, SO3; M = Na, Li) anions with SO2. The reaction of ClO2 with SO2 is also reported for comparative purposes. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer.
Table 1. Rate constants (cm3 s−1 mol−1), branching ratios (%Σ) and efficiencies (k/kcoll) for the reactions of [L·M·ClO2] (L = Cl, F, Br, I, ClO2, SO3; M = Na, Li) anions with SO2. The reaction of ClO2 with SO2 is also reported for comparative purposes. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer.
Reactionkdec (×10−10) §k/kcoll (%)OITOATDOT
ClO2 + SO2 [47]9.1063.886.54.29.3
[Cl·Na·ClO2] + SO22.8824.274.28.817.0
[F·Na·ClO2] + SO23.7530.889.23.77.1
[Br·Na·ClO2] + SO22.1218.773.410.116.5
[I·Na·ClO2] + SO21.8516.868.913.717.4
[Cl·Li·ClO2] + SO21.8114.584.77.97.4
[ClO2·Na·ClO2] + SO27.4866.081.84.813.4
[SO3·Na·ClO2] + SO23.7433.070.46.922.7
[(NaClO2)2·ClO2] + SO27.3968.1---
§ ±30%.
Table 2. Rate constants (cm3 mol−1 s−1) for the reactions of the [Cl·Na·ClO2] anion with SO2. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer.
Table 2. Rate constants (cm3 mol−1 s−1) for the reactions of the [Cl·Na·ClO2] anion with SO2. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer.
Reactivityk × 10−10 cm3 mol−1 s−1 §
OITk1k2.2
2.130.743
OATk2k2.1
0.2517.53
DOTk3k4k5
0.3060.1470.034
§ ±30%.
Table 3. Rate constants (cm3 mol−1 s−1) for the reaction: [ClO2·Na·ClO2] + SO2. Only the most relevant rate constants are reported. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer; add: addition reaction.
Table 3. Rate constants (cm3 mol−1 s−1) for the reaction: [ClO2·Na·ClO2] + SO2. Only the most relevant rate constants are reported. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer; add: addition reaction.
Reactivityk × 10−10 cm3 mol−1 s−1 §
OITk6
6.26
OATk7k7.1
0.3649.57
DOTk8k9k10k11
0.2510.4170.1340.222
 SO2 addkSO2
0.574
§ ±30%.
Table 4. Rate constants (cm3 mol−1 s−1) for the reaction [ClO2·Na·SO3]· + SO2. Only the most relevant rate constants are reported. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer; add: addition reaction.
Table 4. Rate constants (cm3 mol−1 s−1) for the reaction [ClO2·Na·SO3]· + SO2. Only the most relevant rate constants are reported. OIT: oxygen ion transfer; OAT: oxygen atom transfer; DOT: double oxygen transfer; add: addition reaction.
Reactivityk × 10−10 cm3 mol−1 s−1 §
OITk12
2.63
OATk13k13.1
0.2601.09
DOTk14k15k16
0.1670.2220.457
SO2 addkadd
0.494
§ ±30%.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Salvitti, C.; Pepi, F.; Troiani, A.; de Petris, G. Intracluster Sulphur Dioxide Oxidation by Sodium Chlorite Anions: A Mass Spectrometric Study. Molecules 2021, 26, 7114. https://doi.org/10.3390/molecules26237114

AMA Style

Salvitti C, Pepi F, Troiani A, de Petris G. Intracluster Sulphur Dioxide Oxidation by Sodium Chlorite Anions: A Mass Spectrometric Study. Molecules. 2021; 26(23):7114. https://doi.org/10.3390/molecules26237114

Chicago/Turabian Style

Salvitti, Chiara, Federico Pepi, Anna Troiani, and Giulia de Petris. 2021. "Intracluster Sulphur Dioxide Oxidation by Sodium Chlorite Anions: A Mass Spectrometric Study" Molecules 26, no. 23: 7114. https://doi.org/10.3390/molecules26237114

Article Metrics

Back to TopTop