Next Article in Journal
222-Nanometer Far-UVC Exposure Results in DNA Damage and Transcriptional Changes to Mammalian Cells
Next Article in Special Issue
Definition of the Acceptor Substrate Binding Specificity in Plant Xyloglucan Endotransglycosylases Using Computational Chemistry
Previous Article in Journal
Role of the Pangolin in Origin of SARS-CoV-2: An Evolutionary Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Chloronium Cation [(C2H3)2Cl+] and Unsaturated C4-Carbocations with C=C and C≡C Bonds in Their Solid Salts and in Solutions: An H1/C13 NMR and Infrared Spectroscopic Study

by
Evgenii S. Stoyanov
* and
Irina V. Stoyanova
Vorozhtsov Institute of Organic Chemistry, Siberian Branch of Russian Academy of Sciences, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(16), 9111; https://doi.org/10.3390/ijms23169111
Submission received: 6 July 2022 / Revised: 2 August 2022 / Accepted: 10 August 2022 / Published: 14 August 2022
(This article belongs to the Special Issue Structure, Energy and Dynamics of Molecular Interactions)

Abstract

:
Solid salts of the divinyl chloronium (C2H3)2Cl+ cation (I) and unsaturated C4H6Cl+ and C4H7+ carbocations with the highly stable CHB11Hal11 anion (Hal=F, Cl) were obtained for the first time. At 120 °C, the salt of the chloronium cation decomposes, yielding a salt of the C4H5+ cation. This thermally stable (up to 200 °C) carbocation is methyl propargyl, CH≡C-C+-H-CH3 (VI), which, according to quantum chemical calculations, should be energetically much less favorable than other isomers of the C4H7+ cations. Cation VI readily attaches HCl to the formal triple C≡C bond to form the CHCl=CH-C+H-CH3 cation (VII). In infrared spectra of cations I, VI, and VII, frequencies of C=C and C≡C stretches are significantly lower than those predicted by calculations (by 400–500 cm−1). Infrared and 1H/13C magic-angle spinning NMR spectra of solid salts of cations I and VI and high-resolution 1H/13C NMR spectra of VII in solution in SO2ClF were interpreted. On the basis of the spectroscopic data, the charge and electron density distribution in the cations are discussed.

1. Introduction

Simple saturated carbocations (C2–C7) have been studied experimentally in condensed phases [1,2,3,4,5,6,7] and their infrared (IR) spectra have been interpreted [6,7,8]. Protonated arenium cations, starting with C6H7+, have been broadly studied by NMR spectroscopy in liquid superacids and as carborane salts [9,10]. There are more problems with the research on unsaturated nonarenium carbocations, because their salts have not been obtained so far. The simplest and least stable vinyl cation, C2H3+, and isomers of C3H3+ and C3H5+ have been studied experimentally only in vacuum by mass-selected IR spectroscopy [11,12,13]. Their stable isomers have been identified and IR spectra interpreted. Numerous attempts have been made to study the allyl cation, C3H5+, by NMR spectroscopy in liquid superacids at low temperature, and they have failed [14]. The formation of C3H5+ has been proved in a cryogenic superacidic matrix (170 K) by IR spectroscopy [15,16]. With the increasing temperature (230 K), the IR spectrum changes and this phenomenon is attributed to “polymerization”. The main disadvantages of the conventional use of superacids such as FSO3H/SbF5 are the high oxidation potential and reactivity of the Lewis acid used, SbF5, which lead in the case of vinyl cations to rearrangement or decomposition products at temperatures above −100 °C [17,18]. Nevertheless, NMR spectroscopy has revealed the formation (in liquid superacids) of unstabilized C5-dienyl cations with two conjugated C=C bonds at −135 °C [18] and more stable cyclobutenyl and dimethyl-allyl cations [19,20,21]. The vinyl cations stabilized by β-silyl and other electron-donating groups were studied in the 1990s by H.-U. Siehl and coauthors [22,23,24] by NMR spectroscopy in liquid superacids at temperatures below −100 °C and were isolated as carborane salts [25]. Thus, from refs. [15,16,17,18,19,20,21,22,23,24,25], it follows that nonstabilized vinyl cations are stable only at low temperatures (less than −100 °C), and their salts are not isolable in a pure state. Recently, it was shown that this is not the case: carborane salts of the vinyl and allyl types of carbocations C3H5+ and C4H7+, stable at room and elevated temperatures (up to 140–150 °C), were obtained and characterized by X-ray diffraction analysis and IR spectroscopy [26,27,28]. The possibility of the existence of carbocations containing C≡C bonds is discussed on the basis of quantum chemical calculations [29], but they have not been obtained and characterized experimentally.
In this work, we report obtaining pure solid salts—stable at room temperature—of the unsaturated carbocations with double C=C (C4H6Cl+) and triple C≡C bonds (C4H5+) and a so far unknown divinyl chloronium cation, (C2H3)2Cl+. As counterions, carborane anions CHB11Hal11 were chosen, where Hal=F, Cl (hereafter abbreviated as {F11} and {Cl11}, Figure 1) because of their extreme stability and minimal basicity, which promotes the formation of stable salts with highly reactive cations [30]. The cations being studied were characterized by 1H/13C NMR and IR spectroscopy.

2. Results and Discussion

2.1. The Chloronium Cation

It is formed by a three-stage reaction (see the Experimental Section). The first two stages
H {Cl11} + C2H4Cl2 → C2H4Cl+ {Cl11} + HCl
and
C2H4Cl+ {Cl11} + C2H4Cl2 → (C2H4Cl-Cl+-C2H4Cl) {Cl11}
are similar to those studied for reactions of H{Cl11} with CH2Cl2 [31] or C2H5Cl [32] and should lead to the formation of the salt of the dichloroethylchloronium cation, C2H4Cl-Cl+-C2H4Cl. It decomposes rapidly with the release of two HCl molecules to form mostly expected divinylchloronium (third stage):
(C2H4Cl-Cl+-C2H4Cl) {Cl11} → (C2H3-Cl+-C2H3) {Cl11} + 2HCl.
It will be shown below that in the IR spectra of the obtained salt, characteristic bands of chloronium group C-Cl+-C are observed at 628 and 594 cm−1, thereby unambiguously proving the formation of the chloronium cation.
The optimized structure of the cis-isomer of divinylchloronium at the B3LYP/6-311G++(d,p) level of theory is shown in Scheme 1 (isomer I) with some geometrical parameters (the trans-isomer is almost the same in terms of energy). The presence of two double C=C bonds in cation I does not rule out that under certain conditions, intramolecular cyclization with the formation of isomer II may occur (Scheme 1). The optimization of structure II shows that the C····C distance in the chloronium C–Cl+–C group decreases so much that the C–C bond and the four-membered carbon cycle can form (isomer III).
Energetically, cations II and III are more favorable than I by 31.2 and 11.3 kcal/mol, respectively.
NMR spectra. The magic-angle spinning (MAS) 13C NMR spectrum of the isotope-substituted 13C-(C2H3)2Cl+{Cl11} salt (99 atom% 13C) with high-power 1H decoupling contains four signals (Figure 2). Although the integration of 13C NMR signals is not recommended, we will point out that in this spectrum, the intensities of the signals at (83.2 + 78.6), 60.6 and 43.9 ppm correlate as 2.0:0.95:1.0, respectively. The MAS NMR 13C spectrum registered without 1H decoupling shows that two signals at 83.2 and 78.6 ppm are split with a 1JCH constant of 304 Hz (Figure 2, red), which indicates that they belong to the C atoms of nonequivalent CαH groups (Scheme 1). The other two signals obviously belong to carbon atoms of nonequivalent CβH2 groups. They must have a triplet structure. Nevertheless, due to the weaker influence of the positive charge on them, the spin–spin 13C–1H constant decreases, thereby leading to signal broadening without multiplet resolution. Thus, the chloronium cation contains two nonequivalent CH2CH groups.
The 1H MAS NMR spectrum of the (C2H3)2Cl+{Cl11} salt (with the natural abundance of the 13C isotope) shows the known signal from the CH group of the {Cl11} anion at 3.37 ppm and overlapped signals from the cation (Figure 3). In the 1H MAS NMR spectrum of the isotope-substituted 13C-(C2H3)2Cl+ {Cl11} salt, signals from the cation are broadened due to spin–spin 1H–13C coupling without fine structure resolution. To obtain a high-resolution spectrum, the 12C-(C2H3)2Cl+{Cl11} salt was dissolved in SO2ClF and a 1H NMR spectrum was registered. It turned out to be-time dependent: the HCl signal appears at 1.06 ppm and its intensity increases until the intensity of the CH signal of the {Cl11} anion is reached. This means that chloronium in solution decomposes with the release of one HCl molecule:
(C2H3)2Cl+{Cl11} → C4H5+{Cl11} + HCl.
The 1H NMR spectrum of the freshly prepared solution (recorded within 2–3 min after solution preparation at room temperature) shows strong signals from the chloronium cation and weak signals from HCl and the products of decomposition (Figure 4). Because a solid suspension in a liquid does not have time to completely precipitate within 2–3 min, the resolution of the spectrum is low and the 3JHH decoupling is not observed. The spectrum shows all six signals expected for that of the asymmetric cation (C2H3)2Cl+. They correlate with the signals in the 1H MAS NMR spectrum subjected to separation into six Lorentzian components (Figure 3). Therefore, the 1H NMR spectra of chloronium are similar for a solution and the solid salt, and cation structure is the same in both phases.
The finding that each C2H3 group of the (C2H3)2Cl+ cation yields three H1 signals (Ha, Hb, and Hc) that are known for neutral vinyl chloride with hindered rotation of the CH2 group around the C=C bond [33] means that C2H3 groups contain a double bond. That is, the cation under study is distorted divinylchloronium with two nonequivalent C2H3 groups. Its asymmetry in solid salt (C2H3)2Cl+{Cl11} and in solutions may be due to the emergence of contact ion pairs with cation–anion interaction via the CαH group (Scheme 2). In neutral vinyl chloride, the Cα atom of CH2CαHCl is mostly deshielded because it is mainly affected by the electronegativity of the Cl atom, while atoms of the CH2 group are deshielded less [33,34] (Table 1). The same is observed in cation I: the Cα atom closest to Cl+ is more deshielded than the Cβ atom. There is also an important difference: the screening of H and C atoms in the chloronium cation is greater than that of vinyl chloride, i.e., the charge of the cation contributes to an increase in the screening constant of the C and H atoms.
IR spectra. In Figure 5, IR spectra of the salts of protio and deutero chloronium cation are given. To identify the frequencies of CH, CC, and CCl vibrations, the difference in the spectra between the protio and deutero samples was obtained with such a scaling factor as to fully compensate for the adsorption of the {Cl11} anion. This difference allowed us to determine the bands of the protio cation with positive intensities and those of the deutero cation as negative in the entire frequency range. The frequencies involving the C–H (D) vibrations yielded a common H/D isotopic ratio of 1.326 to 1.367 (Table 2). For the split band at 1238 cm−1, this ratio is 1.024, confirming its assignment to the CC stretch. Two C-Cl stretches of the bridged C-Cl+-C group at 628 and 594 cm−1 (very specific for chloronium cations [31,32]) remained nearly the same, as did another one at 740 cm−1, confirming that it did not involve vibrations from the H atom (can be a ClCC bending vibration).
The IR spectrum of the chloronium cation has two specific features. The first and most important is the absence of a strong C=C stretching band expected at ~1600 cm−1. Instead, it shows a split band of medium intensity at 1238 and 1234 cm−1, which is even lower than that of the t-Bu+ cation (~1290 cm−1). If this is indeed the band of the CC stretch vibration of cation I, then it is slightly stronger than the strengthened single C-C bond in the t-Bu+ cation, but weaker than the aromatic CC bond with one-and-half-bond status.
The second feature is the multiplet structure of the bands of CH2/CD2 stretch vibrations (Figure 5, Table 2); this arrangement is caused by the nonequivalence of the two C2H3 groups of cation I owing to ion pairing. In the salts with carborane anion {F11}, which is much less basic than {Cl11}, the ion pairing should weaken or not form at all. The IR spectrum of the C4H6Cl+{F11} salt contains three CH stretch vibrations of the cation (Figure 6). Their frequencies are strictly proportional to the frequencies of vinyl chloride [35]: νasCH2 3121 cm−1, νsCH2 3030 cm−1 and νCH 3086 cm−1. These data allow their assignment (Figure 6) corresponding to symmetrical cation I. That is, the chloronium cation is uniformly surrounded by {F11} anions. In the region of expected C=C stretching frequencies, a very weak band at 1607 cm−1 is observed. Nevertheless, when the salt is heated to 100 °C and chloronium is completely decomposed with the release of HCl (see Experimental section), a new intense νC=C band of the decomposition product emerges at 1567 cm−1, and a weak band at 1607 cm−1 remains unchanged (Figure 6). This result indicates that the latter band belongs to one of the formed by-products. Thus, the symmetric chloronium cation, just as the asymmetric one, does not show the absorption of the stretching vibration of the C=C double bond in the expected frequency range.
A comparison of the experimental spectrum of chloronium with those calculated for optimized structures of isomers IIII at the B3LYP/6-311G++(d,p) level of theory (Table S3 in SI) does not show good agreement (especially for C=C stretch frequencies), which is generally typical for unsaturated carbocations [26,27].
Nevertheless, it is useful to compare the experimental stretch vibrations of the C–Cl+–C group, which are the most specific indicators of chloronium cations, with those calculated for IIII isomers, given that the νcalcexp ratio must be close to 1.0. For isomer I, empirical frequencies are closest to calculated ones with the νcalcexp ratio of 0.91 (Table 3). For isomers II and III, this ratio is unsatisfactorily low, which is a sufficient reason to exclude them from consideration.

2.2. The C4-Carbocations

Salts of the chloronium cation decompose at temperatures above 100 °C with the release of HCl and the formation of carbocation C4H5+ (see the Experimental part). Three noncyclic isomers—IV, V, and VI—with the lowest energy are possible (the energy of V and VI exceeds that of IV by 5.8 and 8.00 kcal/mol, respectively) [29]. Their optimized structures at the B3LYP/6-311G++(d,p) level of theory are depicted in Scheme 3 with CC distances. Let us consider with which isomer the NMR and IR spectra of the C4H5+ cation are the most consistent.
NMR spectra. The 1H MAS NMR spectrum of the isotope-substituted 13C-C4H5+{Cl11} salt is given in Figure 7. It features a signal from the (12C)-H atom of the anion at 3.37 ppm and three signals from the cation. The intensity of the strongest signal at 2.55 ppm exceeds that of the anion by 3.3-fold, indicating that it belongs to the CH3 group of the cation. Two other signals of equal intensity at 7.44 and 9.35 ppm obviously belong to two CH groups of the cation. This result clearly supports the formation of isomer VI. The signal at 7.44 ppm is a doublet with a 1JCH coupling constant of 184 Hz. Other cation signals should also be unresolved doublets because all H atoms are bonded to one 13C atom.
The 1H NMR spectrum was separated into Lorentzian components, taking into account the fact that the cation signals are doublets, and the anion signal is a singlet (Figure 7). A well-resolved doublet at 7.44 ppm may belong to the H atom of the H-(C≡C) group, and the signal at 9.35 ppm should be attributed to the H atom of the CH[CH3] group, because the broadening of the doublet components due to the spin–spin interaction with hydrogen atoms of the methyl group impairs their resolution.
13C MAS NMR analysis with high-power 1H decoupling yields four signals of this sample (Figure 8). The 13C NMR spectrum registered without 1H decoupling does not result in the manifestation of a fine structure owing to 13C-1H spin–spin interactions. The signal with a chemical shift of 36 ppm is typical for the CH3 group in carbocations. To interpret the remaining signals, it is desirable to see their hyperfine structure. For this, it is necessary to record the NMR spectra of salt solutions, if possible.
The salt of cation VI was dissolved in SO2ClF. Gaseous SO2ClF was liquefied in an NMR tube at −15 °C using a stream of gaseous SO2ClF passed through a desiccant P2O5. The 1H NMR spectrum of the solvent showed a weak broad signal at 8.90 ppm from H+(H2O)n, and a second one at 1.06 ppm from molecular HCl, which is formed upon the decomposition of SO2ClF with water. That is, the solvent contained HCl and H+(H2O)nCl as impurities. The solution was obtained by the anaerobic condensation of SO2ClF in an NMR tube with the 13C-C4H5+{Cl11} salt at the bottom; the tube was sealed and shaken until a saturated solution was obtained. Its 1H NMR spectrum contained weak solute signals along with a weak H+(H2O)n signal that broadened and shifted to 7.1 ppm, indicating a decrease in water protonation. At the same time, the HCl signal almost disappeared. All of these data mean that HCl joined the cation.
The 1H NMR spectrum of the solute shows a known singlet from the CH group of the anion (3.37 ppm) and four signals from the cation, which are split into doublets due to the 1H−13C spin–spin interaction (Figure 9). The integrals of the cation signals strictly conform to the 1:1:1:3 ratio. Thus, this carbocation is not the same as that in the original solid salt. It has one CH3 group (at 2.69 ppm) and three CH groups. Obviously, upon salt dissolution, the HCl molecule joins the formal triple CC bond of cation VI to form the C4H6Cl+ cation VII (Scheme 4).
In the 1H NMR spectrum of cation VII, the signal of the CH3 group at 2.69 ppm is a doublet with 1JH-C = 130 Hz. Each doublet component is split into a quartet because of a spin–spin interaction of 1H with nucleus Hα, Cα and Cβ with 2JHCα3JHHα3JHCβ ≈ 7 Hz. Two other signals at 7.04 ppm and 8.42 ppm with constants 1JH-C = 168 and 158 Hz, respectively, are characteristic of the CH groups with sp2 hybridization of the carbon atom and belong to the middle groups, CαH and CβH. The doublet at 8.42 ppm has a quintet-like structure due to the spin–spin interaction of the Hα atom with the protons of the CH3 and CHβ groups and the C(H3), Cβ and Cγ carbon atoms. If all of the constants of these interactions are approximately the same and equal to 7 Hz, then the splitting will be quintet. The doublet signal at 7.04 ppm with 1JH-C = 168 Hz belongs to the Hβ atom. Its multiplicity is poorly resolved due to the spin–spin interaction of the Hβ atom with more differentiated atoms of groups CαH and CγH. Finally, the signal at 9.58 ppm with 1JCH 183 Hz does not have ultrafine structure and belongs to the terminal CγH(Cl) group. Its increased 1JCγH compared to 1JCβH (168 Hz) can be explained by the high electronegativity of the Cl atom bound to Cγ.
The 13C NMR spectrum of the same solution recorded with broad-band decoupling 13C {1H} shows four signals (Figure 10, black). The signal at 20.8 ppm is a doublet with 1JCC = 39 Hz. It belongs to the sp3 C atom of the methyl group interacting with the Cα atom’s spin with sp2 hybridization. The components of this doublet are also split into doublets owing to a long-range interaction with Cβ, with 2JCCβ = 7 Hz. The signal at 129 ppm is a triplet belonging to the Cβ atom bound to Cα and Cγ atoms with 1JC-C = 60 Hz, typical of the JCsp2-Csp2 constants. The signal at 186 ppm is a superposition of two doublets with 1JC-C = 60 Hz. It belongs to the Cα atom bound to significantly different atoms, C(H3) and Cβ. Finally, the doublet signal at 203 ppm with the constant 1JC-C = 60 Hz belongs to the Cγ atom of the CγHCl terminal group.
The monoresonance 13C NMR spectrum of this solution is in agreement with the signal assignment proposed above (Figure 10, blue). The signal of the CH3 group is split into a triplet with 1JC-H = 130 Hz (1JC-C = 39 Hz); the Cβ signal is split into two triplets with 1JC-H = 168 Hz (1JC-C = 60 Hz); two doublet Cα signals are doubled with 1JC-H = 158 Hz (1JC-C = 60 and 39 Hz) and the signal from the Cγ atom shows a split doublet with 1JC-H = 183 Hz and 1JC-C = 60 Hz. All data from the 13C NMR spectrum of cation VII are summarized in Table 4.
To test whether there was a similarity between the chemical shifts of the 13C signals of cations VII and VI, the interpreted signals from VII were compared with those from VI. An unexpectedly good correlation was obtained (Figure 11), which implies that the shielding of 13C nuclei in carbon skeletons Ijms 23 09111 i001 of both cations was similar. Based on this similarity, it is possible to interpret the 13C signals of cation VI (Table 5).
IR spectra. The formation of the C4H5+ cation from chloronium at 120 °C results in the disappearance of the bands of the most characteristic stretch vibrations of the C-Cl+-C moiety of chloronium, and in the appearance of a new spectrum with a highly intense band at 1563 cm−1, typical of the stretching vibrations of the multiple CC bond of unsaturated hydrocarbons (Figure 12). The sequential recording of IR spectra with increasing temperature does not yield absorption bands of any intermediates, except for the spectrum of the final product, which increases in intensity. This means that if intermediate products are formed, they are unstable and quickly transform into the final product. It is expected that the decomposition of chloronium produces isomer IV, which, according to quantum chemical calculations, has the lowest energy among all the open-chain isomers, IVVI [29]. Nevertheless, according to NMR spectra, cation VI is formed. Therefore, if isomer IV comes into being at the first stage, then it quickly transitions into isomer VI (Scheme 5).
Stronger heating of the C4H5+{Cl11} salt to 200 °C does not induce any changes in the IR spectrum of carbocation VI, indicating its thermal stability. The IR spectra of C4H5+ are similar between its salts with {F11} and {Cl11} anions (Table S4 in SI), which means that its structure is not affected by any change in the basicity of the anionic environment. Because the absorption pattern of the {F11} anion overlaps with a larger part of the C4H5+ spectrum, as compared to that of {Cl11}, we discuss the spectrum of C4H5+{Cl11} salt.
In Figure 12, the IR spectra of salts C4H5+{Cl11} and C4D5+{Cl11} are depicted. Their difference, obtained via the complete compensation of the absorption of the {Cl11} anion, allows detecting all bands of the protio-cation with positive intensity and those of the deutero-cation with negative intensity. Pairs of bands with the H/D isotopic ratio of 1.34–1.36 belong to CH vibrations (Table 6). For the intense band at 1563 cm−1 and the middle one at 1447 cm−1 (protio-sample), this ratio is ~1.036 and 1.057, respectively, which means that they belong to the stretching vibrations of CC bonds. Their frequencies approach the double and one-and-a-half bond status, respectively, and the formal structure of the cation VI is more correctly represented as shown in Scheme 6. On the other hand, the CC stretch frequencies of cation VI can be considered as asymmetric (1563 cm−1) and symmetric (1447 cm−1) CCC vibrations. Then, structure VI can be described as resonance between two structures VIa and VIb (Scheme 6).
The IR spectrum of unsaturated cation VII was found to have low intensity because its salt was obtained in a small amount by the evaporation of the solvent from its solution in SO2ClF. Nevertheless, the most important frequencies are detectable with high confidence (Figure S2 in SI). The IR spectrum of VII does not contain an intense band of C=C stretching above 1600 cm−1, which is present in the other studied isomers of the C4H6Cl+ cation, VIII (at 1680 cm−1) and IX (at 1710 cm−1), with π-electron density concentrated mainly on the C=C bond [28] (Scheme 7). Instead, its IR spectrum shows two bands of CC stretches at 1440 and 1262 cm−1, whose low frequencies are more consistent with symmetrical and asymmetrical vibrations, respectively, of allyl moiety Ijms 23 09111 i002. The frequency of the C-Cl stretch vibration is seen at 636 cm−1 (Figure S2 in SI), which is much lower than that of isomers VIII (at 768 cm−1) and IX (at 733 cm−1) [28] and matches the frequency of asymmetric vibration of the bridged C–Cl+–C group of the stablest dimethylchloronium cation, which decomposes at elevated temperature [32]. Therefore, the C-Cl bond is weak, and with an increase in temperature, cation VII can lose the HCl molecule.

2.3. Comparative Discussion of the IR and NMR Spectroscopic Data

2.3.1. The Chloronium Cation

Even though, according to calculations, cyclic cation III is energetically more favorable than open-chain I by 31.2 kcal/mol, it is I that forms in the solid salt and exists in solutions. A specific feature of cation I is its asymmetry in salts with {Cl11} anions, owing to the ion pairing with the anion. In SO2ClF solutions, cation I retains its asymmetry due to the formation of contact ion pairs. In a salt with the least basic {F11} anions, the IR spectrum of the cation corresponds to a symmetrical one, as predicted by quantum chemical calculations for a naked cation I in vacuum. Such a high sensitivity of the cation to the influence of the environment indicates its high polarizability.
The asymmetry of cation I in ion pairs arises due to its interaction with the anion, not through its Cl atom, but rather through the CαH group, as illustrated in Scheme 2. In other words, the positive charge is more substantial on the CαH group.
The asymmetry of the cation is manifested mostly in the vibrations of its CH stretches rather than in those of its CC stretches, which differ only by 5 cm−1. A comparison of its IR spectrum with the interpreted spectrum of neutral vinyl chloride [35] (without C-Cl and CCCl vibrations) allows the assignment of its most characteristic frequencies (Table S6 in SI)), and shows a good correlation for all frequencies with the exception of that for C=C stretching (Figure 13), which deviates downward from the correlation line by 340 cm−1. The same discrepancy is observed when we compare the experimental and calculated C=C stretching frequencies of cation I. The calculated νC=C of 1602 cm−1 (Table S3 in SI after scaling by a factor of 0.9674) is even slightly greater than that of neutral vinyl chloride and exceeds the experimental one for I by 360 cm−1. Put another way, the positive charge of cation I has the strongest effect on its C=C bond, and this phenomenon is poorly described by quantum chemical calculations. The nature of the low frequency of C=C stretching vibrations and its high sensitivity to the presence of electron-donating substituents in the cation are discussed in the [28].
Cation I also possesses unusual features in NMR spectra: the impact of the positive charge on its C2H3- groups increases the screening constants of the C atoms while weakening the C=C bond, compared to those of neutral C2H3Cl. Thus, a positive charge reduces the π-electron density between carbon atoms, bringing its nature closer to a single σ-bond (the screening of C atoms increases).
A comparison of cation I with its saturated analog, diethylchloronium (C2H5)2Cl+ [32], suggests that frequencies of symmetric and asymmetric stretching vibrations of the chloronium C-Cl+-C group in I are higher by 80 and 83 cm−1, respectively. This is a large difference. It means that the C-Cl+-C group attracts electron density from the C2H5- group, thereby imparting a weak π-character to C-Cl bonds and reducing the π-character of the C=C bond. Thus, the C2H3 group under the influence of a positive charge becomes an effective electron donor, attracting a positive charge.

2.3.2. The C4-Carbocations

When chloronium decomposes at >120 °C, the most stable isomer, IV, forms first (Scheme 5). Nonetheless, if this happens, then IV quickly converts to VI, which, according to quantum chemical calculations at the B3LYP/6-311G++(d,p) or MP2(full)/6-311G(d,p) levels of theory, is energetically less favorable by 8.0 or 7.6 kcal/mol [29], respectively. The calculations also predict that the transition of IV to VI takes place in two stages, with an activation energy of 39 kcal/mol [29], which should prevent the formation of VI. Nevertheless, the formation of isomer VI was revealed to be the most preferable.
Isomer VI in carborane salts can be in a homogeneous and inhomogeneous (with predominant interaction with one counterion) anionic environment. Both states, differing in νC=C frequencies, have been found for vinyl-type carbocations [26,27,28]. On the other hand, isomer VI in salts with {Cl11} and with the least basic {F11} anion shows similar frequencies of C≡C stretching, pointing to its homogeneous anionic environment.
The IR spectrum of isomer VI has two features related to the vibration frequencies of C≡C and C-H stretches. The first one is νC≡C at 1563 cm−1, which is 73 cm−1 higher than that of the C=C double bond (1490 cm−1) of cations (CH3)2C=C+H or CH3C+=CH [26,27], and this is expected. Nonetheless, the calculated C≡C stretch (at B3LYP/6-311G++(d,p) level of theory, with a scaling of 0.967) is 2020 cm−1, which significantly exceeds the experimental one by 450 cm−1. A similar discrepancy between the calculated and experimental values of the C=C stretching frequencies equal to 362–246 cm−1 has been documented for vinyl carbocations, CH3-CH=CH+ and CH3-C+=CH2, respectively [26,27].
The second feature of the IR spectra of VI is the reduced frequency of stretching vibrations of its CH3 group by ~120 cm−1 relative to those of neutral alkanes. This is possible if the CH3 group is involved in hyperconjugation with the partially empty 2pz orbital of the Cα atom. The decrease in frequency is not very large compared to that observed for t-butyl+ and i-propyl+ alkane carbocations (~160 cm−1), which have a well-pronounced hyperconjugation effect [6]. This observation clearly indicates that in isomer VI, the triple C≡C bond is partially delocalized (Scheme 6) with the predominant contribution of resonance structure VIa, because a weak contribution of resonance structure VIb promotes the weak filling of the 2pz orbital of the Cα atom and the hyperconjugation effect is partially attenuated.
The calculated IR spectrum of isomer VI revealed that its CH3 group should not be disturbed by hyperconjugation (Table S5 in SI). Similarly, the calculated frequency of the CγH group at 3137 cm−1 is typical for the H-C≡C group of neutral molecules and exceeds the empirical one (at 3058 cm−1) by ~80 cm−1. Therefore, the application of quantum chemical calculations to the interpretation of IR spectroscopic data should be carried out with great care. The same conclusion was drawn earlier in a study on vinyl-type carbocations [26,27].
The obtained NMR spectra allow us to discern some pattern in their changes for alkanes, alkenes, and unsaturated carbocations. A strongly deshielded 13C NMR signal (>150 ppm relative to TMS) can be confidently assigned to a positively charged carbon atom. The largest chemical shift has been documented for the central carbon atom of the tert-butyl cation at 335.2 ppm [36]. In unsaturated carbocations, the C=C double bond is often delocalized, forming allyl moiety Ijms 23 09111 i003. Chemical shifts of its C atoms usually vary depending on the substituents attached to it: from 170 to 190 ppm for C1,3, and 170–150 ppm for C2 atoms, respectively [37], thus pointing to greater charge localization on atoms C1 and C3. If the double bond is not delocalized, as for example, in the Ijms 23 09111 i004 moiety of the 1-cyclopropylcyclopropyli-dene methyl cation, then the positive charge is located mainly on the C2 atom with a chemical shift of 234 ppm and less on atoms C1 (51.7 ppm) and C3 (21.2 ppm) [38]. In vinyl-type carbocations R′C+=CR″2 stabilized by electron-donating groups R′ and R″, such as Ijms 23 09111 i005, the positive charge on the C=C bond is greatly reduced due to the combined influence of t-Bu and the two β-silyl substituents directly attached to it [39]. In the C13 NMR spectra, the signal from the Cα atom is observed in a weak field at 202.7 ppm and from the Cβ atom at 75.5 ppm, that is, the charge is mainly localized on the Cα atom.
Vinyl cation VII, formally studied by us, has a double C=C bond in the Ijms 23 09111 i006 moiety, whose carbon atoms possess the chemical shifts Cγ (203 ppm), Cβ (129 ppm) and Cα (183 ppm). They do not match a triatomic moiety with one CC double bond but rather correspond to slightly asymmetric allyl moiety Ijms 23 09111 i007. Therefore, it is more correct to represent structure VII as shown in Scheme 8.
Cation VI formally contains a triple CC bond. According to the IR spectra, it is delocalized to form the Ijms 23 09111 i008 moiety (Scheme 6). The chemical shifts of its C atoms (230 (Cγ), 150 (Cβ) and 215 ppm (Cα)) are consistent with those of the Ijms 23 09111 i009 part of vinyl cation VII (Figure 11). This means that electron density distributions—across the CC bonds of the two moieties, which differ in formal double and triple CC bonds—are of the same type, with a lower positive charge on the Cα atom than on atoms Cβ and Cγ. The observed participation of H atoms of the CH3 group in hyperconjugation with the partially empty 2pz orbital of the Cα atom indicates its sp2 hybridization. That is, formally, the Cα atom must carry a considerable positive charge.

3. Methods and Materials

The carborane acids, H(CHB11Hal11) with Hal=F, Cl, were prepared as described previously [40,41]. They were purified by sublimation at 150–160 °C under pressure of 10−5 Torr on cold Si windows in a specially designed IR cell. The formed thin translucent layer yielded an intense IR spectrum. If this film is slightly wetted with a drop of 1,2-dichloroethane (DCE) and is allowed to dry quickly, the following reaction proceeds with the release of HCl:
H{Hal11} + 2 C2H4Cl2 → (C2H3-Cl+-C2H3){Hal11} + 3HCl.
The amount of the formed HCl was determined as follows: the weighed portion of H {Cl11} was placed on the bottom of the IR cell and wetted with a drop of DCE. The quantity of released HCl was determined by measuring the intensity of its absorption at 2821 cm−1 with subsequent conversion to millimoles using a calibration curve. The determined molar ratio HCl/H {Cl11} is equal to 3 (Table S1 and Figure S1 in SI). Chemical analysis of the content of C and H in the salts of chloronium confirmed its chemical composition (Table S2 in SI). When d4-DCE is used, reaction (1) proceeds with the formation of (C2D3-Cl+-C2D3) {Cl11} and the release of both HCl and DCl (detected in IR spectra). Thus, reaction (1) proceeds via two steps:
H{Cl11} + C2D4Cl2 → C2D4Cl{Cl11} + HCl,
and
C2D4Cl{Cl11} + C2D4Cl2 → (C2D3-Cl+-C2D3){Cl11} + 2 DCl.
After completion of reaction (1), the product film formed on the surface of the Si windows has a yellow color due to the presence of impurities. These are easily removed by washing the film with cold dichloromethane (DCM), because they are readily soluble, whereas the main product dissolves slowly. The remaining film turns white and its IR spectrum confirms the removal of a small amount of impurities.
At ≥100 °C the chloronium cation decomposes into an unsaturated carbocation with the release of HCl, in accordance with Equation (3)
(C2H3-Cl+-C2H3){Cl11} → C4H5+{Cl11} + HCl,
in a way similar to the mechanism of diethyl-chloronium decomposition into the t-Bu+ cation [32]. The thermal decomposition of deuterated divinyl-chloronium proceeds similarly, with the formation of the salt of the C4D5+ cation and the release of DCl.
The salts of 13C-substituted (C2H3)2Cl+ and C4H5+ cations were obtained via reactions (1)–(3) using 13C-dichloroethane (99 atom% 13C) from Sigma-Aldrich (Taufkirchen, Germany).
Solutions of carbocation salts in SO2ClF for NMR measurements were prepared under anaerobic conditions. A solid salt was placed at the bottom of an NMR tube. Through a capillary, gaseous SO2ClF was injected into it. At −15 °C, it was condensed into a liquid. The tube was sealed and shaken until a saturated solution was obtained. 1H and 13C NMR spectra of the solutions were recorded on a Bruker DRX 500 spectrometer (Heidelberg, Germany). Chemical shifts in the 1H NMR spectra were measured relative to an internal standard, the (C)H signal of the anion CHB11Cl11, which is equal to 3.376 ppm in a CH2Cl2 solution referenced to TMS. Chemical shifts in the 13C NMR spectra were referenced to TMS as an external standard.
The preparation of samples for magic-angle spinning (MAS) NMR measurements is described in detail in the Supporting Information (SI).
The MAS NMR measurements were performed on a Bruker Avance-400 NMR spectrometer, Germany, (9.4 T) equipped with a broad-band double resonance MAS NMR probe (4 mm rotor diameter). The samples in the sealed glass ampoules were tightly inserted into MAS rotors and spun at 5–9 kHz. 13C MAS NMR spectra were recorded at ambient temperature at the resonance frequency of 100.63 MHz. The length of the 90° pulse was 6.5 µs, and 2000–10,000 scans were accumulated for a spectrum with a repetition delay of 10 s. To eliminate the effect of J-coupling between 13C and 1H nuclei, high-power proton decoupling was applied during the period of signal acquisition at the power level equivalent of a 90° 1H pulse of 5.0 µs length. 1H MAS NMR spectra were acquired at the frequency of 400.13 MHz by means of Hahn-echo pulse sequence (π/2–τπτ–acquisition), where delay τ is equal to one rotor period, and the length of the π/2 excitation pulse is 5.0 μs. The 1H MAS NMR spectrum was accumulated after 16 scans with the repetition delay of 60 s, ensuring quantitative conditions. Chemical shifts in the 13C NMR spectra were referenced to TMS as an external standard and in the 1H NMR spectra relative to an internal standard: the (C)H signal of anion CHB11Cl11, which is equal to 3.376 ppm in the CH2Cl2 solution referenced to TMS.
IR spectra were obtained on a Shimadzu IRAffinity-1S spectrometer (Japan) housed inside a glovebox in the 4000–400 cm−1 frequency range in transmittance and attenuated total reflectance mode (ATR). The spectra were manipulated using the GRAMMS/A1 (7.00) software from Thermo Scientific.
All the quantum chemical calculations were performed at the B3LYP/6-311G++(d,p) level plus the D3 dispersion correction energy term [42] with an ultrafine integration grid within the framework of the Gaussian’09 package [43].

4. Conclusions

Pure salts of the divinyl-chloronium cation, (C2H3)2Cl+, methylated propargyl (VI), and C4-allyl type VII carbocations with extremely stable and weakly basic carborane anions, CHB11Cl11 and CHB11F11, were obtained for the first time. They dissolved in DCM without interaction with the solvent but transformed into other carbocations. Therefore, they were studied in a solid phase by IR and 1H/13C MAS NMR spectroscopy. Only cation VII was stable in solutions of its salt in SO2ClF, which made it possible to obtain its detailed NMR spectra with hyperfine structure.
These cations are thermally stable at room and elevated temperatures (up to 100–110 °C for chloronium or ~200 °C for the methyl-propargyl carbocation). We reported the same earlier for vinyl-type carbocations [26,27,28], C3H5+ and C4H7+, thereby refuting the widespread belief that unsaturated (nonaromatic) carbocations are unstable at temperatures higher than −100 °C. The high stability of unsaturated carbocations is facilitated by their formal double and triple CC bonds, which contribute to the distribution of electron density over CC bonds and to the dispersion of a positive charge over the molecule.
The composition of the divinyl-chloronium cation is determined by the conditions of its preparation, and the structure is unambiguously proven by MAS NMR and IR spectroscopy. Its formal double C=C bonds have a bond status of less than one-and-half with low π-electron density between C-C atoms. At the same time, the electronic shielding of its C atoms is substantial, which is unexpected. The charge of the cation is concentrated mainly on the C2H3 groups. For this reason, it forms a contact ion pair with an anion through the CH group of one of the C2H3 groups, and this process considerably lowers its symmetry. All these features of the chloronium cation are indicative of its high polarizability.
In the C4H6Cl+ cation, the electron density distribution over CC bonds and the nature of the double C=C bond are very sensitive to the position of the Cl atom in the carbon chain, Ijms 23 09111 i010. When the Cl atom is attached to the Cα/Cβ, atom, isomers Ijms 23 09111 i011 and Ijms 23 09111 i012 arise, with well-defined and close-to-equivalent C=C double bonds. If the Cl atom is attached to the Cγ atom, then allyl-type isomer VII is formed with considerable alignment of the CC bonds in the Ijms 23 09111 i013 moiety, with the charge dispersed over it. This phenomenon drives strong change in the CC stretch frequencies.
In propargyl carbocation VI, π-electron density of its formally triple C≡C bond is distributed to the neighboring C-C bond, thereby upgrading the neighboring bond to one-and-a-half-bond status and demoting itself to two-and-a-half-bond status. The asymmetry of the Ijms 23 09111 i014 moiety is evidenced by the higher frequency of the Ijms 23 09111 i015 stretch at 1563 cm−1 than that of the double C=C bond in vinyl cations (1490 cm−1) [26,27]. Despite the asymmetry, the positive charge is effectively distributed over both CC bonds: the involvement of the hydrogen atoms of the CH3 group in hyperconjugation with the 2pz orbital of the Cα atom means that it has sp2 hybridization and a major positive charge is concentrated on it. Meanwhile, it follows from the MAS NMR spectra that a comparable charge is also localized on the Cγ atom.
Our important finding is that the charged formal triple C≡C bond in carbocation VI can accept an HCl molecule, giving rise to allyl cation VII with the Cl atom on a formal double C=C bond.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23169111/s1. References [35,44] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, E.S.S.; methodology, E.S.S.; validation, E.S.S. and I.V.S.; formal analysis, I.V.S.; investigation, E.S.S. and I.V.S.; resources, I.V.S.; data curation, E.S.S. and I.V.S.; writing—original draft preparation, E.S.S. and I.V.S.; writing—review and editing, E.S.S.; visualization, E.S.S. and I.V.S.; supervision, E.S.S.; project administration, E.S.S.; funding acquisition, E.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation (state registration No. 1021052806375-6-1.4.3).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Victor I. Mamatyuk for a useful discussion of the NMR spectra, Viktor Yu. Kovalskii for providing the quantum chemical calculations and Sergey S. Arsumanov for assistance with the recording of the MAS NMR spectra. The authors are grateful to the Multi-Access Chemical Research Center of the Siberian Branch of the Russian Academy of Sciences for spectral and analytical measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Olah, G.A.; Baker, E.B.; Evans, J.C.; Tolgyesi, W.S.; McIntyre, J.S.; Bastien, I.J. Stable Carbonium Ions. V.1a Alkylcarbonium Hexafluoroantimonates. J. Am. Chem. Soc. 1964, 86, 1360–1373. [Google Scholar] [CrossRef]
  2. Olah, G.A.; Bollinger, J.M.; Cupas, C.A.; Lukas, J. Stable carbonium ions. XXXIV. 1-Methylcyclopentyl cation. J. Am. Chem. Soc. 1967, 89, 2692–2694. [Google Scholar] [CrossRef]
  3. Olah, G.A.; DeMember, J.R.; Commeyras, A.; Bribes, J.L. Stable carbonium ions. LXXXV. Laser Raman and infrared spectroscopic study of alkylcarbonium ions. J. Am. Chem. Soc. 1971, 93, 459–463. [Google Scholar] [CrossRef]
  4. Olah, G.A. Stable Carbonium Ions in Solution. Science 1970, 168, 1298–1311. [Google Scholar] [CrossRef] [PubMed]
  5. Brouwer, D.M.; Hogeveen, H. Electrophilic Substitutions at Alkanes and in Alkylcarbonium Ions. Prog. Phys. Org. Chem. 1972, 9, 179–240. [Google Scholar]
  6. Stoyanov, E.S.; Gomes, G.P. tert-Butyl Carbocation in Condensed Phases: Stabilization via Hyperconjugation, Polarization, and Hydrogen Bonding. J. Phys. Chem. A 2015, 119, 8619–8629. [Google Scholar] [CrossRef]
  7. Stoyanov, E.S.; Nizovtsev, A.S. Stabilization of carbocations CH3+, C2H5+, i-C3H7+, tert-Bu+, and cyclo-pentyl+ in solid phases: Experimental data versus calculations. Phys. Chem. Chem. Phys. 2017, 19, 7270–7279. [Google Scholar] [CrossRef]
  8. Stoyanov, E.S. Stabilization of Saturated Carbocations in Condensed Phases. J. Phys. Chem. A 2017, 121, 9638–9644. [Google Scholar] [CrossRef]
  9. Koptyug, V.A. Contemporary Problems in Carbonium Ion Chemistry III: Arenium Ions-Structure and Reactivity; Rees, C., Ed.; Springer: Berlin/Heidelberg, Germany, 1984; pp. 1–227. [Google Scholar]
  10. Reed, C.A.; Kim, K.-C.; Stoyanov, E.S.; Stasko, D.; Tham, F.S.; Mueller, L.J.; Boyd, P.D.W. Isolating Benzenium Ion Salts. J. Am. Chem. Soc. 2003, 125, 1796–1804. [Google Scholar] [CrossRef]
  11. Duncan, M.A. Infrared Laser Spectroscopy of Mass-Selected Carbocations. J. Phys. Chem. A 2012, 116, 11477–11491. [Google Scholar]
  12. Ricks, A.M.; Douberly, G.E.; Schleyer, P.V.R.; Duncan, M.A. Infrared spectroscopy of gas phase C3H3+ ions: The cyclopropenyl and propargyl cations. J. Chem. Phys. 2010, 132, 051101. [Google Scholar] [CrossRef] [PubMed]
  13. Douberly, G.E.; Risks, A.M.; Schleyer, P.V.R.; Duncan, M.A. Infrared spectroscopy of gas phase C3H5+: The allyl and 2-propenyl cations. J. Chem. Phys. 2008, 128, 021102. [Google Scholar] [CrossRef] [PubMed]
  14. Vogel, P. Carbocation Chemistry; Elsevier: Amsterdam, The Netherlands, 1985; p. 173. [Google Scholar]
  15. Buzek, P.; Schleyer, P.R.; Vančik, H.; Mihalic, Z.; Gauss, J. Generation of the Parent Allyl Cation in a Superacid Cryogenic Matrix. Angew. Chem. Int. Ed. 1994, 33, 448–451. [Google Scholar] [CrossRef]
  16. Mišić, V.; Piech, K.; Bally, T. Carbocations Generated under Stable Conditions by Ionization of Matrix-Isolated Radicals: The Allyl and Benzyl Cations. J. Am. Chem. Soc. 2013, 135, 8625–8631. [Google Scholar] [CrossRef]
  17. Siehl, H.-U.; Müller, T.; Gauss, J.; Buzek, P.; Schleyer, P.V.R. Cyclopropylcyclopropylidenemethyl Cation: A Unique Stabilized Vinyl Cation Characterized by NMR Spectroscopy and Quantum Chemical ab Initio Calculations. J. Am. Chem. Soc. 1994, 116, 6384–6387. [Google Scholar] [CrossRef]
  18. Siehl, H.-U.; Müller, T.; Gauss, J. NMR Spectroscopic and quantum chemical characterization of the (E)- and (Z)-isomers of the penta-1,3-dienyl-2-cation. J. Phys. Org. Chem. 2003, 16, 577–581. [Google Scholar] [CrossRef]
  19. Olah, G.A.; Staral, J.S.; Liang., G. Novel aromatic systems. I. Homocyclopropenyl cation, the simplest 2.pi. homoaromatic system. J. Am. Chem. Soc. 1974, 96, 6233–6235. [Google Scholar] [CrossRef]
  20. Mayr, H.; Olah, G.A. Stable carbocations. 202. Ring closure reactions of allyl to cyclopropylcarbinyl cations. J. Am. Chem. Soc. 1977, 99, 510–513. [Google Scholar] [CrossRef]
  21. Olah, G.A.; Spear, R.J. Stable carbocations. CLXXX. Carbon-13 and proton nuclear magnetic resonance spectroscopic study of phenyl-, methyl-, and cyclopropyl-substituted alkenyl (allyl) cations. Further studies of the trend of charge distribution and the relative delocalization afforded by phenyl, methyl, and cyclopropyl. J. Am. Chem. Soc. 1975, 97, 1539–1546. [Google Scholar]
  22. Siehl, H.-U. Dicoordinated Carbocations; Rappoport, Z., Stang, P., Eds.; Wiley: Chichester, UK; New York, NY, USA, 1997; pp. 189–236. [Google Scholar]
  23. Siehl, H.-U. Stable Carbocation Chemistry; Prakash, G.K.S., Schleyer, P.V.R., Eds.; Wiley: New York, NY, USA, 1997; pp. 165–196. [Google Scholar]
  24. Siehl, H.-U.; Kaufmann, F.-P.; Apeloig, Y.; Braude, V.; Danovich, D.; Berndt, A.; Stamatis, N. The First Persistent β-Silyl-Substituted Vinyl Cation. Angew. Chem. 1991, 30, 1479–1482. [Google Scholar] [CrossRef]
  25. Müller, T.; Margraf, D.; Syha, Y.; Nasiri, H.R.; Kaiser, C.; Maier, R.; Boltre, B.; Juhasz, M.; Reed, C.A. Recent Developments in Carbocation and Onium Ion Chemistry; Kenneth, K.L., Ed.; American Chemical Society: Chicago, IL, USA, 2007; Chapter 3; pp. 51–67. [Google Scholar]
  26. Stoyanov, E.S.; Bagryanskaya, I.Y.; Stoyanova, I.V. Unsaturated vinyl-type carbocation [(CH3)2C=CH]+ in its carborane salts. ACS Omega 2021, 6, 15834–15843. [Google Scholar] [CrossRef] [PubMed]
  27. Stoyanov, E.S.; Bagryanskaya, I.Y.; Stoyanova, I.V. Isomers of the Allyl Carbocation C3H5+ in Solid Salts: Infrared Spectra and Structures. ACS Omega 2021, 6, 23691–23699. [Google Scholar] [CrossRef] [PubMed]
  28. Stoyanov, E.S.; Bagryanskaya, I.Y.; Stoyanova, I.V. An IR-spectroscopic and X-ray-structural study of vinyl-type carbocations in their carborane salts. ACS Omega 2022, 7, 27560–27572. [Google Scholar] [CrossRef]
  29. Cunje, A.; Rodriquez, C.F.; Lien, M.H.; Hopkinson, A.C. The C4H5+ Potential Energy Surface. Structure, Relative Energies, and Enthalpies of Formation of Isomers of C4H5+. J. Org. Chem. 1996, 61, 5212–5220. [Google Scholar] [CrossRef]
  30. Reed, C. Carborane acids. New “strong yet gentle” acids for organic and inorganic chemistry. Chem. Commun. 2005, 13, 1669–1677. [Google Scholar] [CrossRef]
  31. Stoyanov, E.S. Chemical Properties of Dialkyl Halonium Ions (R2Hal+) and Their Neutral Analogues, Methyl Carboranes, CH3−(CHB11Hal11), Where Hal = F., Cl. J. Phys. Chem. A 2017, 121, 2918–2923. [Google Scholar] [CrossRef] [PubMed]
  32. Stoyanov, E.S.; Stoyanova, I.V.; Tham, F.S.; Reed, C.A. Dialkyl Chloronium Ions. J. Am. Chem. Soc. 2010, 132, 4062–4063. [Google Scholar] [CrossRef] [PubMed]
  33. Whipple, E.B.; Steward, W.E.; Reddy, G.S.; Goldstein, J.H. NMR Spectra of Vinyl Chloride and the Chloroethylenes. J. Chem. Phys. 1961, 34, 2136–2138. [Google Scholar] [CrossRef]
  34. Breitmair, E.; Volter, W. Carbon-13 NMR Spectroscopy, 2nd ed.; Gordon & Breach Science: New York, NY, USA, 1978; p. 149. [Google Scholar]
  35. Narita, S.; Ichinohe, S.; Enomoto, S. Infrared Spectra of Vinyl Chloride and Vinyl Chloride-d3. J. Chem. Phys. 1959, 31, 1151–1157. [Google Scholar] [CrossRef]
  36. Olah, G.A.; Donovan, D.J. Stable carbocations. 208. Carbon-13 nuclear magnetic resonance spectroscopic study of alkyl cations. The constancy of carbon-13 nuclear magnetic resonance methyl substituent effects and their application in the study of equilibrating carbocations and the mechanism of some rearrangements. J. Am. Chem. Soc. 1977, 99, 5026–5037. [Google Scholar]
  37. Olah, G.A.; Staral, J.S.; Spear, R.J.; Liang, G. Novel aromatic systems. II. Cyclobutenyl cations and the question of their homoaromaticity. Preparation and study of the homocyclopropenium ion, the simplest homoaromatic system. J. Am. Chem. Soc. 1975, 97, 5489–5497. [Google Scholar] [CrossRef]
  38. Stanton, J.F.; Gauss, J.; Siehl, H.-U. CCSD(T) calculation of NMR chemical shifts: Consistency of calculated and measured 13C chemical shifts in the 1-cyclopropylcyclopropylidenemethyl cation. Chem. Phys. Lett. 1996, 262, 183–186. [Google Scholar] [CrossRef]
  39. Muller, T.; Juhasz, M.; Reed, C.A. The X-ray Structure of a Vinyl Cation. Angew. Chem. Int. Ed. 2004, 43, 1543–1546. [Google Scholar] [CrossRef] [PubMed]
  40. Juhasz, M.; Hoffmann, S.; Stoyanov, E.S.; Kim, K.; Reed, C.A. The strongest isolable acid. Angew. Chem. Int. Ed. 2004, 43, 5352–5355. [Google Scholar] [CrossRef] [PubMed]
  41. Nava, M.; Stoyanova, I.V.; Cummings, S.; Stoyanov, E.S.; Reed, C.A. The Strongest Brønsted Acid: Protonation of Alkanes by H(CHB11F11) at Room Temperature. Angew. Chem. Int. Ed. 2014, 53, 1131–1134. [Google Scholar] [CrossRef]
  42. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  43. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  44. Kesharwani, M.K.; Brauer, B.; Martin, J.M.L. Frequency and Zero-Point Vibrational Energy Scale Factors for Double-Hybrid Density Functionals (and Other Selected Methods): Can Anharmonic Force Fields Be Avoided? J. Phys. Chem. A 2015, 119, 1701–1714. [Google Scholar] [CrossRef]
Figure 1. Icosahedral carborane anion CHB11Hal11 (Hal=F, Cl).
Figure 1. Icosahedral carborane anion CHB11Hal11 (Hal=F, Cl).
Ijms 23 09111 g001
Scheme 1. The optimized structures of isomers of chloronium cation C4H6Cl+.
Scheme 1. The optimized structures of isomers of chloronium cation C4H6Cl+.
Ijms 23 09111 sch001
Figure 2. The 13C MAS NMR spectrum of the isotope-substituted 13C-(C2H3)2Cl+{Cl11} salt (99% 13C) with (black) and without (red) high-power 1H decoupling. Signals of the one vinyl group differing from another are marked with asterisks.
Figure 2. The 13C MAS NMR spectrum of the isotope-substituted 13C-(C2H3)2Cl+{Cl11} salt (99% 13C) with (black) and without (red) high-power 1H decoupling. Signals of the one vinyl group differing from another are marked with asterisks.
Ijms 23 09111 g002
Figure 3. 1H MAS NMR spectrum of the (C2H3)2Cl+{Cl11} salt.
Figure 3. 1H MAS NMR spectrum of the (C2H3)2Cl+{Cl11} salt.
Ijms 23 09111 g003
Figure 4. The 1H NMR spectrum of the (C2H3)2Cl+{Cl11} salt dissolved in SO2ClF. The signals from the vinyl group that more strongly bonded with the Cl+ atom are labeled with an asterisk. The signals from the decomposition products that grow with time are marked with the “♣” sign. Designations Ha, Hb and Hc are given in the Scheme 2.
Figure 4. The 1H NMR spectrum of the (C2H3)2Cl+{Cl11} salt dissolved in SO2ClF. The signals from the vinyl group that more strongly bonded with the Cl+ atom are labeled with an asterisk. The signals from the decomposition products that grow with time are marked with the “♣” sign. Designations Ha, Hb and Hc are given in the Scheme 2.
Ijms 23 09111 g004
Scheme 2. Schematic representation of the structure of cation I with magnetically nonequivalent H atoms a and b involved in ion pairing with an anion.
Scheme 2. Schematic representation of the structure of cation I with magnetically nonequivalent H atoms a and b involved in ion pairing with an anion.
Ijms 23 09111 sch002
Figure 5. IR spectra in transmittance of the (C2H3-Cl+-C2H3) {Cl11} salt (red) and its deuterated analog (blue). Asterisks indicate bands of the {Cl11} anion (two bands of CH stretching of the anion of variable intensity are observed due to the partial involvement of the CH group in H bonding). For a better examination of the bands of CH stretching vibrations, the difference between the protio and deutero spectra is shown (dashed black curve). The intensities of both spectra are reduced to unit anion intensity.
Figure 5. IR spectra in transmittance of the (C2H3-Cl+-C2H3) {Cl11} salt (red) and its deuterated analog (blue). Asterisks indicate bands of the {Cl11} anion (two bands of CH stretching of the anion of variable intensity are observed due to the partial involvement of the CH group in H bonding). For a better examination of the bands of CH stretching vibrations, the difference between the protio and deutero spectra is shown (dashed black curve). The intensities of both spectra are reduced to unit anion intensity.
Ijms 23 09111 g005
Figure 6. The ATR IR spectrum of the (C2H3-Cl+-C2H3) {F11} salt (red) and the same reduced by a factor of 5 along the vertical axis to show the absorption of C-F vibrations of the anion (dashed curve). The spectrum of the salt after heating at 100 °C for the C=C stretch region is highlighted in blue. Asterisks denote bands of the {F11} anion.
Figure 6. The ATR IR spectrum of the (C2H3-Cl+-C2H3) {F11} salt (red) and the same reduced by a factor of 5 along the vertical axis to show the absorption of C-F vibrations of the anion (dashed curve). The spectrum of the salt after heating at 100 °C for the C=C stretch region is highlighted in blue. Asterisks denote bands of the {F11} anion.
Ijms 23 09111 g006
Scheme 3. Optimized structures of lowest-in-energy open-chain isomers of the C4H5+ cation.
Scheme 3. Optimized structures of lowest-in-energy open-chain isomers of the C4H5+ cation.
Ijms 23 09111 sch003
Figure 7. The 1H MAS NMR spectrum of the C4H5+{Cl11} salt. The components of each split signal are shown in one color.
Figure 7. The 1H MAS NMR spectrum of the C4H5+{Cl11} salt. The components of each split signal are shown in one color.
Ijms 23 09111 g007
Figure 8. 13C MAS NMR analysis with high-power 1H decoupling of the 13C-C4H5+ {Cl11} salt. Each signal and the corresponding spinning sidebands, marked with asterisk, have the same color.
Figure 8. 13C MAS NMR analysis with high-power 1H decoupling of the 13C-C4H5+ {Cl11} salt. Each signal and the corresponding spinning sidebands, marked with asterisk, have the same color.
Ijms 23 09111 g008
Scheme 4. The joining of HCl to the cation (VI) with the formation of cation (VII).
Scheme 4. The joining of HCl to the cation (VI) with the formation of cation (VII).
Ijms 23 09111 sch004
Figure 9. The 1H NMR spectrum of the C4H6Cl+{Cl11} solution in SO2ClF. The broad signal at ~7 ppm belongs to protonated water.
Figure 9. The 1H NMR spectrum of the C4H6Cl+{Cl11} solution in SO2ClF. The broad signal at ~7 ppm belongs to protonated water.
Ijms 23 09111 g009
Figure 10. 13C NMR analysis of the C4H6Cl+ {Cl11} solution in SO2ClF, performed either with (black) or without (blue) high-power 1H decoupling.
Figure 10. 13C NMR analysis of the C4H6Cl+ {Cl11} solution in SO2ClF, performed either with (black) or without (blue) high-power 1H decoupling.
Ijms 23 09111 g010
Figure 11. Correlation between chemical shifts of interpreted signals of 13C NMR of the cation VII and those of the cation VI.
Figure 11. Correlation between chemical shifts of interpreted signals of 13C NMR of the cation VII and those of the cation VI.
Ijms 23 09111 g011
Figure 12. IR spectra in transmittance of the protium-(red) and deuterium-(blue) C4H5+ carbocation in their salts with the {Cl11} anion in the frequency range of CH stretches (a) and below 1600 cm−1 (b). The intensities of both spectra are reduced to unit anion intensity.
Figure 12. IR spectra in transmittance of the protium-(red) and deuterium-(blue) C4H5+ carbocation in their salts with the {Cl11} anion in the frequency range of CH stretches (a) and below 1600 cm−1 (b). The intensities of both spectra are reduced to unit anion intensity.
Ijms 23 09111 g012
Scheme 5. Proposed formation of the C4H5+ cation from the chloronium cation.
Scheme 5. Proposed formation of the C4H5+ cation from the chloronium cation.
Ijms 23 09111 sch005
Scheme 6. The structure of carbocation (VI), represented by two resonance structures (VIa) and (VIb).
Scheme 6. The structure of carbocation (VI), represented by two resonance structures (VIa) and (VIb).
Ijms 23 09111 sch006
Scheme 7. The structure of two vinyl-type isomers of the C4H6Cl+ cation.
Scheme 7. The structure of two vinyl-type isomers of the C4H6Cl+ cation.
Ijms 23 09111 sch007
Figure 13. Correlation between the empirical frequencies of the C2H3 group of the C2H3Cl molecule and those of the (C2H3)2Cl+ cation.
Figure 13. Correlation between the empirical frequencies of the C2H3 group of the C2H3Cl molecule and those of the (C2H3)2Cl+ cation.
Ijms 23 09111 g013
Scheme 8. Structure of isomer VII according to NMR and IR spectroscopy.
Scheme 8. Structure of isomer VII according to NMR and IR spectroscopy.
Ijms 23 09111 sch008
Table 1. 1H/13C NMR chemical shifts (ppm) of the chloronium cation in the (C2H3)2Cl+{Cl11} salt in comparison with those of vinyl chloride (the signals marked with an asterisk belong to the second C2H3 group, which differs from the first one).
Table 1. 1H/13C NMR chemical shifts (ppm) of the chloronium cation in the (C2H3)2Cl+{Cl11} salt in comparison with those of vinyl chloride (the signals marked with an asterisk belong to the second C2H3 group, which differs from the first one).
AtomCation ICH2CHCl [33,34]AtomCation ICH2CHCl
[33,34]
SolidSolutionSolid
Ha5.58
5.14 *
4.87
4.49 *
6.13Cα83.2
78.6 *
124.9
Hb4.774.11
3.97 *
5.23Cβ60.6
43.9 *
116
Hc2.73
2.22 *
2.91
2.49 *
5.39
Table 2. Comparison of the IR spectra of cations C4H6Cl+ and C4D6Cl+ in their salts with the {Cl11} anion.
Table 2. Comparison of the IR spectra of cations C4H6Cl+ and C4D6Cl+ in their salts with the {Cl11} anion.
SampleC-H StretchesC-H BendingCC StretchCH BendingCCCl BendingCCl+C Stretches
protio3055 30442987 2974 2961 29381418 *132412691238914740628 594
deuterio2303 22952244 2218 2191 2171****9601208694748635 582
ratio1.327 1.3261.331 1.341 1.351 1.353--1.3231.0241.3200.9890.989 1.020
* Average of three components; ** overlapped with the anion.
Table 3. A comparison of the stretch vibrations of the C–Cl+–C group of the chloronium cation under study with those calculated for isomers IIII at the B3LYP/6-311G++(d,p) level of theory.
Table 3. A comparison of the stretch vibrations of the C–Cl+–C group of the chloronium cation under study with those calculated for isomers IIII at the B3LYP/6-311G++(d,p) level of theory.
IsomerStretch Vibrations of C-Cl+-C Group, cm−1
(calc a/exp)
C4H6Cl+ (exp)628594
I (calc)568 (0.91)538 (0.91)
II (calc)494 (0.79)385 (0.65)
III (calc)478 (0.76)460 (0.77)
a without scaling.
Table 4. NMR data a for cation VII in a SO2ClF solution.
Table 4. NMR data a for cation VII in a SO2ClF solution.
NucleusCH3CαHCβHCγH (Cl)
1H δ2.698.427.049.58
1JCH130158168183
3JHH7b7b
13C δ20.8183.6129.2203.6
1JCC39606060
a Chemical shifts are in parts per million, coupling constants are in hertz. b not determined.
Table 5. Chemical shifts δ * in the MAS NMR spectra of isomer VI in its C4H5+{Cl11} salt.
Table 5. Chemical shifts δ * in the MAS NMR spectra of isomer VI in its C4H5+{Cl11} salt.
NucleusCH3CαHCβCγH (1JCH in Hz)
1H2.259.35-7.44 (184)
13C36215150231
* in ppm, for 1H vs. internal standard, the H(C) signal of the anion, equal to 3.37 ppm in a dichloromethane solution.
Table 6. IR spectra of C4H5+ and C4D5+ cations (isomers VI) in their salts with the {Cl11} anion.
Table 6. IR spectra of C4H5+ and C4D5+ cations (isomers VI) in their salts with the {Cl11} anion.
C4H5+C4D5+H/D RatioAssignment *
305822841.339νCαH (CαD) and
303922511.350νCβH (CβD)
294221291.381νasCH3 (CD3)
2891--2νCC overtone
287120771.380νasCH3 (CD3)
282920551.377νsCH3 (CD3)
15631521
1497
1.028
1.044
νC≡C
144713681.057νCC
13279971.341δCH (CD)
13039711.341δCH (CD)
9647191.354δCH (CD)
* ν is the stretching mode; δ is the deformation mode.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Stoyanov, E.S.; Stoyanova, I.V. The Chloronium Cation [(C2H3)2Cl+] and Unsaturated C4-Carbocations with C=C and C≡C Bonds in Their Solid Salts and in Solutions: An H1/C13 NMR and Infrared Spectroscopic Study. Int. J. Mol. Sci. 2022, 23, 9111. https://doi.org/10.3390/ijms23169111

AMA Style

Stoyanov ES, Stoyanova IV. The Chloronium Cation [(C2H3)2Cl+] and Unsaturated C4-Carbocations with C=C and C≡C Bonds in Their Solid Salts and in Solutions: An H1/C13 NMR and Infrared Spectroscopic Study. International Journal of Molecular Sciences. 2022; 23(16):9111. https://doi.org/10.3390/ijms23169111

Chicago/Turabian Style

Stoyanov, Evgenii S., and Irina V. Stoyanova. 2022. "The Chloronium Cation [(C2H3)2Cl+] and Unsaturated C4-Carbocations with C=C and C≡C Bonds in Their Solid Salts and in Solutions: An H1/C13 NMR and Infrared Spectroscopic Study" International Journal of Molecular Sciences 23, no. 16: 9111. https://doi.org/10.3390/ijms23169111

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop