Next Article in Journal
Environment-Assisted Shortcuts to Adiabaticity
Next Article in Special Issue
Thermodynamic Properties of the First-Generation Hybrid Dendrimer with “Carbosilane Core/Phenylene Shell” Structure
Previous Article in Journal
Solenoid Configurations and Gravitational Free Energy of the AdS–Melvin Spacetime
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process

Research Institute of Thermodynamics and Kinetics, Ivanovo State University of Chemistry and Technology, 153000 Ivanovo, Russia
*
Author to whom correspondence should be addressed.
Entropy 2021, 23(11), 1478; https://doi.org/10.3390/e23111478
Submission received: 6 October 2021 / Revised: 26 October 2021 / Accepted: 29 October 2021 / Published: 8 November 2021

Abstract

:
A multi-technique approach based on Knudsen effusion mass spectrometry, gas phase chromatography, mass spectrometry, NMR and IR spectroscopy, thermal analysis, and quantum-chemical calculations was used to study the evaporation of 1-butyl-3-methylimidazolium tetrafluoroborate (BMImBF4). The saturated vapor over BMImBF4 was shown to have a complex composition which consisted of the neutral ion pairs (NIPs) [BMIm+][BF4], imidazole-2-ylidene C8N2H14BF3, 1-methylimidazole C4N2H6, 1-butene C4H8, hydrogen fluoride HF, and boron trifluoride BF3. The vapor composition strongly depends on the evaporation conditions, shifting from congruent evaporation in the form of NIP under Langmuir conditions (open surface) to primary evaporation in the form of decomposition products under equilibrium conditions (Knudsen cell). Decomposition into imidazole-2-ylidene and HF is preferred. The vapor composition of BMImBF4 is temperature-depended as well: the fraction ratio of [BMIm+][BF4] NIPs to decomposition products decreased by about a factor of three in the temperature range from 450 K to 510 K.

1. Introduction

In recent years, ionic liquids (ILs) have become one of the most fast-developing fields of chemistry. This keen interest is due to the unique combination of IL’s properties, which includes the ability to dissolve organic, inorganic and polymer materials together with low vapor pressure at room temperatures [1]. Today, imidazolium-based ILs are the most investigated group of ionic liquids. They are distinguishable from others by quantifiable vapor pressures at 380–500 K and their relatively high decomposition temperatures [2]. This fact gives possibility to the use of these ILs in various heterophase processes such as distillation, chemical gas-phase deposition, etc. Therefore, the investigation of evaporation of the imidazolium-based ILs is particularly important.
It has been experimentally proven that many aprotic ILs evaporate congruently and that their vapors consist of neutral ion pairs (NIPs) [3,4,5,6,7,8]. However, evaporation of a number of aprotic ILs can be accompanied by a partial decomposition of the condensed phase. Some ILs, e.g., those containing chiral center, decompose while heating, and the decomposition processes prevail over evaporation [9]. Decomposition was marked for ILs with high electronegative anions (BF4, PF6, AsF6, SCN, etc.) as well [10,11,12]. The 1-butyl-3-methylimidazolium tetrafluoroborate (BMImBF4) is a striking example of such a compound. As opposed to the alkylimidazolium ILs with NTf2 anion having a simple vapor composition consisting exclusively of NIPs [6,13] at temperatures below the onset of decomposition, the vapor of BMImBF4 contains many other components [14] at quite low temperatures, where any appreciable effects on thermal analysis curves are not evident. However, there is no consensus in the literature about this phenomenon: some works [14,15,16,17,18,19,20,21,22] are concerned with the thermal behavior of BMImBF4, and conclude that the no thermal degradation of IL occurred, while at the same time the other authors [23,24,25,26,27] postulate the thermal decomposition of the investigated compound.
In a recent paper [10], the evaporation mechanism of BMImBF4 was investigated using NMR analysis and mass spectrometry. The authors established the competitive vaporization and thermal decomposition of IL ruled by the sample surface area to volume ratio. The main decomposition route proposed in [10] was the formation of imidazol-2-ylidene (Figure 1a) through an Arduengo carbene [28] by cation-anion interaction. Previously, borane-substituted imidazol-2-ylidene was synthesized in vacuo from BMImBF4 by Taylor et al. [29]. Some researchers [10,29] gave the NMR and mass spectra of pure ylidene and found that its vapor pressure was sufficiently higher than that of IL at the same temperatures. The evaporation of BMImBF4 can also be accompanied by the formation of imidazoles [26,27]. These routes are the formation of 1-methylimidazole, 1-butene, HF, and BF3 (Figure 1b), and 1-butylimidazole, fluoromethane, and BF3 (Figure 1c). The pyrolysis-gas chromatography experiments [26,27] were carried out at much higher temperatures than that of decomposition (873 K and 823 K in refs. [26] and [27] respectively). The authors [26] concluded that the ratio between these two routes is close to 1:1. According to Ohtani et al. [27] the reaction with 1-methylimidazole formation is preferred. Formation of ethylimidazole from similar IL BMImPF6 was observed by KEMS in Ref. [12]. Authors [12] noted the influence of orifice size on vapor composition: the larger the orifice, the larger the contribution of the BMImPF6 vapor species. The investigation of ILs with cyano-functionalized anions [11,30], which are close to the object under study, showed one more degradation route by the intrinsic cyclization (Figure 1d) of the butyl group on the C1 atom.
In a recent paper [31] the kinetic model of maximum operation temperature (MOT) was applied to BMImBF4 IL. This model defines the temperature at which a mass loss of 1%, which can be attributed to thermal decomposition, occurs as a function of variable application time. It was found that MOT decreased exponentially with increasing application time (466 K at 1 h and 348 K at 1 year). The authors [31] also studied the vapor composition of BMImBF4. The decomposition products according to routes b and c (here and further designations as in Figure 1) were found in vapor. Additionally, the formation of imidazole and 1-butene from 1-butylimidazole was suggested and traces of imidazole were found.
Mass spectroscopy [10,14] revealed the partial decomposition of initial IL according to way a and did not find any traces of imidazoles. It should be noted that in a recent paper [12] with the analogous BMImPF6 IL, none of the ylidenes were registered, whereas the ethylimidazole was observed in vapor.
The available thermodynamic studies of BMImBF4 were carried out without any analysis of the gas phase composition [17,18,19]. As a result, the vapor pressures of BMImBF4 obtained in these works disagree by some orders of magnitude. Therefore, the comprehensive analysis of the vapor composition of the mentioned group of ILs is mandatory in thermodynamic investigations.
This work is a multi-technique study of the BMImBF4 evaporation carried out with the use of Knudsen effusion mass spectrometry (KEMS), IR and NMR spectroscopy, thermal analysis, gas-phase chromatography–mass spectrometry (GCMS), and quantum chemical modelling. The main goals are to determine the composition of saturated vapor over BMImBF4 and to clarify the routes of thermal decomposition while heating.

2. Experimental

Thermal analysis of the samples (98% purity, Sigma-Aldrich, St. Louis, MO, USA) was performed on a synchronous thermal analysis instrument Netzsch STA 449 F3 Jupiter (NETZSCH-Gerätebau, Selb, Germany) in the temperature range of 20–500 °C at a speed of 5 °C/min in nitrogen atmosphere. The device has high sensitivity with a resolution of 0.1 μg. In parallel with the data on weight loss, the temperature dependences of the thermal effects were recorded at the resolution of 1 μW.
IR spectroscopic measurements were carried out on a Bruker Tensor 27 (Bruker AXS, Madison, WI, USA) spectrometer with Fourier transform. The operating frequency range was 370–4000 cm−1 with a resolution of 1 cm−1. The instrument makes it possible to obtain both the spectra of the condensed phases and the temperature dependences of the absorption for the gas phase.
NMR spectra 1H, 13C, 11B, 15N in DMSO-d6 at T = 22 °C and T = 70 °C were recorded by a Bruker Avance 500 (Bruker AXS, Madison, US) spectrometer with 5 mm TBI 1H/31P/D-BB z-GRD sensor. A working frequency for 1H was 500.17 MHz, 13C—125.77 MHz, 11B—160.47 MHz, 15N—50.68 MHz. 13C NMR spectra were obtained using broadband proton decoupling (WALTZ 16). 15N chemical shift measurements were made based on two-dimensional HMBC 15N-1H spectra. The solvent signal (DMSO-d6) was used as a reference for 1H, 13C spectra; BF3OEt2 was used for 11B; nitromethane was used for 15N. The signal assignment in spectra of objects under study was performed on a base of literature data [32,33,34,35,36] and NMR-prediction instruments.
A magnetic sector mass spectrometer MI1201 (PO “Electron”, Sumy, Ukraine) coupled with a Knudsen cell was used for vapor analysis. Neutral vapor species were studied using a combined ion source operating in electron-ionization (EI) mode. A detailed description of the apparatus is given elsewhere [37,38,39].
The GCMS experiments were carried out on a Shimadzu GCMS QP2010 Ultra (Shimadzu, Kyoto, Japan). Each sample was analyzed in programmable mode: a column temperature was kept at 100 °C during 5 min, after that the sample was heated with a speed of 5 °C/min up to 250 °C. Two types of columns were used: polar (Agilent DB-17MS capillary column) and non-polar (Zebron ZB-5MS capillary column).

3. Computational Details

The molecular structure of conformers of the neutral ionic pair [C4mim+][BF4] and the cation [C4mim+] has been studied by the density functional theory method (pure B3LYP, B3LYP with D3 version of Grimme’s dispersion [40], long-range-corrected version of B3LYP using the Coulomb-attenuating CAM-B3LYP functional [41], as well as hybrid functional of Truhlar and Zhao M06 [42]) with the use of the Dunning’s correlation consistent triple basis sets cc-pVTZ [43]. All calculations were carried out using the Gaussian 09 package [44]. Comparison of IR spectra of computed molecules with experiment revealed a more accurate description of molecular structure by the CAM-B3LYP functional, which is used in all calculations.
The nine most energetically preferable conformers of the [BMIm+] cation were selected on a basis of conformational analysis. The structures of the BMImBF4 conformers are depicted in Figure S1 (see Supplementary Materials). Each cation conformer can exist together with the [BF4] anion in two forms: “close” (denoted as “a”) and “open” (denoted as “b”). The equilibrium mole fractions of the conformers in the temperature range of 300–500 K found on a basis of their relative free energies calculated by DFT are presented in Figure 2. The conformer 5a due to its lowest energy dominates in vapor at experimental conditions (400–500 K). Its structure is shown in Figure 3.
The IR spectra were modeled in the following manner: the intensities of vibrations of the conformers were multiplied by the corresponding mole fractions and are summarized through observed IR range (40–400 cm−1).

4. Results

4.1. Thermal Analysis

The decomposition temperature Tdec = 685 K was found as the average between the mass loss and DSC data (Figure 4). This value is quite close to that of 679 [31], 697 K [45], 712 K [46], 633–723 K [47] whereas the other literature data are somewhat lower: ~640 K [26], 653 K [48], and 634 K [23]. None of additional effects before decomposition were found.

4.2. NMR Analysis

None of structural changes on heating were revealed (Figure 5). The only effect of temperature change was marked for 11B. It is a broadening of B1 peak due to increasing of anion motion around cation upon heating. The obtained 1H spectra are in agreement with those from [32,34,35,36] and disagree with data from [33]. A possible source of such discrepancy is that, in the latter case, the authors used non-commercial self-synthesized samples and therefore some impurities may not be removed. The theoretical 1H NMR spectra of [BMIm+][BF4] NIP, imidazol-2-ylidene, and bicyclic IL methyl-4C-imidazolium tetrafluoroborate were obtained by quantum chemical calculations at CAM-B3LYP/aug-cc-pVTZ level of theory. The chemical shift scale was calibrated by tetramethylsilane (TMS). The comparison of theoretical (Figure S2, Supplementary Materials) and experimental spectra corresponds to the structure of [BMIm+][BF4] NIP. The comparison of the obtained 1H and 13C spectra with those of imidazole-2-ylidene [10], 1-butylimidazole [49], and 1-methylimidazole [50] indicates the absence of traces of these decomposition products in initial IL.

4.3. KEMS

Mass spectrometric experiments were performed in the temperature range of 424–514 K, much below the decomposition temperature (685 K) found by the thermal analysis. The background subtracted mass spectrum recorded at 472 K and the energy of ionizing electrons of 40 eV is shown in Figure 6. In contrast with alkylimidazoilum ILs with NTf2 anion [6,13] the obtained mass spectrum has some prominent features. First, the parent cation with m/z = 139 has very low relative intensity; second, the lightweight fragment ions have high intensities with dominating ion with m/z = 96; third, the ions with the higher mass than that of the parent cation were also registered (m/z = 158, 187).
The temperature dependencies of ion currents in the form ln(IT) vs. 1/T and the ionization efficiency curves were measured for the most intense ions (Figure 7). The ion appearance energies (AE) obtained from the ionization efficiency curves by a linear extrapolation method together with the slopes of the temperature dependencies are listed in Table 1. The energy scale was calibrated using the background signal of HI+ (IE(HI) = 10.38 eV [51]).
The temporal dependencies of the ion currents for ions with m/z = 82, 96, 137, 139, and 187 were measured during 36 h at T = 480 K and are shown in Figure 8. The ion currents for m/z = 96, 137, and 187 increase in time, the ion current for m/z = 82 is practically time-independent, while the ion current for m/z = 139 decreases in time.
The mass spectrum obtained in this work considerably differs from that in [14]. In the latter work the major peak in mass spectrum was m/z = 139, with co-dominating m/z = 82. Despite the ion with m/z = 158 was registered, none of the heavier ions were found. It should be mentioned that the evaporation in [14] was carried out from the open surface in Langmuir conditions while in our work the evaporation was performed in Knudsen conditions.
To clarify the influence of evaporation conditions on vapor composition, the additional experiments on BMImBF4 evaporation from the open Knudsen cell (intermediate between Knudsen and Langmuir conditions) and from the entirely open surface of IL (Langmuir conditions) were carried out. The recorded mass spectra for the selected peaks are given in Table 2. The mass spectrum from the open surface is very close to that obtained in [14]. The tendency of increasing of the intensity of the parent cation (m/z = 139) is observed in the effusion cell—open cell–open surface series, while the intensity of ions with m/z = 82, 96, 137 decreases in the same series. The ion with m/z = 187 was found only in Knudsen conditions. In work [12] the same effect of the mass spectrum dependence from the area of the effusion orifice was reported for the BMImPF6. The changes in the mass spectra at different evaporation conditions for BMImBF4 and BMImPF6 are shown in Figure 9. One can see that the behavior of relative intensities is the same for both ILs and strongly depends on a ratio of effusion area to evaporation area.

4.4. IR-Spectroscopy

The IR-spectra (Figure 10) were recorded for the initial IL, the residue after the mass spectrometric experiment, and the distillate collected from the surface of the collimator located in front of the effusion orifice. The analysis of the spectra revealed the absence of any substantial changes in the condensed phase in the Knudsen cell during mass spectrometric experiments, when the sample was heated up to 514 K. However, the IR-spectrum of the distillate had some distinctive features in 800–1000 cm−1 region. All obtained spectra were identical to those in [17]. To attribute registered peaks in the spectra a quantum chemical modelling of the vibrational spectrum (CAM-B3LYP/cc-pVTZ level of theory) was performed. All theoretical spectra were calculated at 500 K as a combination of those for conformers (18 conformers in the case of BMImBF4; 9 for ylidene; two for bicyclic IL) taking into account their mole fractions (Figure 11 and Figure 12).
Comparison of the theoretical and experimental spectra showed that the initial IL as well as the residue after the KEMS experiment consisted of BMImBF4 whereas the distillate contains imidazole-2-ylidene along with BMImBF4.

4.5. GCMS

The GCMS mass spectra and the ion profiles in BMImBF4 chromatograms with a polar and nonpolar chromatograph column are shown in Figure 11. The chromatogram profile on the nonpolar column recorded for the pure undiluted IL was rather broadened (Figure 11d). The most likely explanation is an interaction of BMImBF4 with the column material. The addition of ethanol into the sample results in a peak narrowing (Figure 11f). The only peak on the chromatogram pointed out the presence of a single nonpolar compound in vapor.
The chromatogram recorded on the polar column (Figure 11a) had the only peak as well (despite of a slight splitting both peaks have the same mass spectra). Its mass spectrum is characterized by the dominating ion with m/z = 82 and the sufficiently lower intensity peaks with m/z = 159, 110, 55, and 42.

5. Discussion

To determine the vapor composition of BMImBF4 a proper mass spectrum interpretation should be carried out. Molecular precursors of the ions were defined based on the data of the KEMS and GCMS experiments.
The identification of the molecular precursors of the ions was performed on the basis of two principles: (1) the ions from the same molecule usually show close slopes of the temperature dependencies of ion currents, and (2) AE of fragment ions increases with decreasing their masses. In addition, for experiments on the Knudsen/Langmuir evaporation, the following statement is true: the ratio of ion currents from the same molecule does not depend on the evaporation conditions.
The parent cation with m/z = 139 has AE (12.4 ± 0.5 eV), which is consistent with the value from [14] (12.8 ± 0.4 eV). This AE value is considerably higher than those obtained for the parent cations of prototypical ILs BMImNTf2 (9.3 ± 0.3 eV) [12] and EMImNTf2 (8.9 ± 0.2 eV) [52] and at the same time it is closer to the AE obtained for the similar BMImPF6 IL (11.3 ± 0.5 eV) [12]. This fact points out another nature of electron ionization of such class of ILs caused by the stronger cation-anion interaction. As a result, the ion with m/z = 158 corresponding to BMImF+ has appeared. The scheme of this ion formation suggested in [14] includes the intramolecular rearrangement during ionization. AE (BMImF+/BMImBF4) = 11.7 ± 0.5 eV is lower than that of the parent cation. The same situation was observed for origination of BMImF+ from BMImPF6 (AE (BMImF+/BMImPF6) = 11.1 ± 0.3 eV) [12]. The slopes of temperature dependency of ion current for the ions with m/z = 158 and m/z = 139 are very close (Table 1) indicating the origination of the ion with m/z = 158 directly from NIP. This fact is additionally confirmed by GCMS data on a polar column where the signal with m/z = 158 was detected. An indirect proof of the origin of the ion with m/z = 158 from NIP is the constant ratio 158/139 observed with the different sizes of effusion orifice in the experiment with BMImPF6 [12].
The ion with m/z = 49 is BF2+ having AE = 16.9 ± 0.5 eV close to AE (BF2+/BF3) = 16 ± 1 eV [53] indicating BF3 as a possible molecular precursor of this ion. At the same time, the slope of the temperature dependence for an ion with m/z = 49 is similar to that for ions with m/z = 139 and 158 (Table 1). This indicates the second source of origin of the BF2+ ion from [BMIm+][BF4] NIP.
The ions with m/z = 82, 96, 137, and 187 have similar slopes of the temperature dependencies of ion currents. Their AE values increases in the m/z series 137-96-82 corresponding to the abovementioned assignment rule. An exception from this rule is the ion with m/z = 187 having the highest appearance energy. However, this circumstance can be explained assuming the different ionization process for this ion. Let us demonstrate it basing on the AE values of BF3. The bond between boron and fluorine is quite strong, even to detach one fluorine atom from boron trifluoride it needs the high energy AE (BF2+/BF3) = 16 ± 1 eV [53]. Therefore, the relatively high AE value of the ion with m/z = 187 can be explained by the nature of B-F interaction. A similar situation is observed [54] upon ionization of the B2F4 molecule with AE (BF+/B2F4) < AE (BF2+/B2F4) < AE (B2F3+/B2F4) due to the different processes for the BF+, BF2+, and B2F3+ ions formation including different co-products. The data of KEMS, GCMS and Knudsen/Langmuir evaporation experiments confirmed a single source of the ions with m/z = 96, 137, and 187. Comparison of the mass spectra obtained for IL by GCMS and KEMS, and by DIMS [10] for imidazole-2-ylidene showed their qualitative similarity (Table 3). Small quantitative differences can be explained by some additional contribution into these signals from NIP in our work and a possible difference in the ion source constructions as well as different types of mass analyzers. Therefore, one can conclude that imidazole-2-ylidene is the main molecular precursor of these ions. Ylidene has no ionic bond and its ionization mechanism is close to those of inorganic polyhalides (see [55]) where the intensity of the molecular ion is less (or absent at all) than that of the first dissociative ion. That is why there is no molecular ion with m/z = 206, but the ion with m/z = 187 is present in the mass spectrum.
The assignment of the ion with m/z = 82 is rather complicated. Analysis of the temperature dependencies of ion currents and the ionization efficiency curves allows us to assign it to the same neutral precursor as for the ions with m/z = 96, 137, and 187, i.e., to imidazole-2-ylidene. However, according to GCMS data, the ion with m/z = 82 is present in the mass spectrum on both the polar and nonpolar columns. The previous studies [3,4] of ILs with the NTf2 anion show that the ion with m/z = 82 is common for alkylimidazolium ILs. This ion can also be formed from 1H-imidazoles via routes b and c (Figure 1). Analysis of possible evaporation routes according to the schemes depicted in Figure 1b,c was performed on the basis of data from the NIST Mass Spectrometry Data Center [56]. A list of the main peaks in mass spectra for various decomposition processes [56] is given in Table 4. According to route c (Figure 1) the intensity of the ion with m/z = 34 corresponding to CH3F should be very strong; the same is expected for the ion with m/z = 97—the fragment from 1-butylimidazole. However, the KEMS data don’t support this assumption. All main peaks corresponding to route b were found in the mass spectrum. The presence of ions with m/z = 41 and m/z = 42 in our mass spectrum and their absence in mass spectrum of pure ylidene [10] pointed out the possibility of evaporation of IL by way b. Hence the ion with m/z = 82 has at least three sources: ylidene, 1-methylimidazole, and BMImBF4. The most significant results supporting this conclusion were obtained in an isothermal evaporation experiment (Figure 8). The intensity of the ion with m/z = 82 was almost time-independent, while the intensities of ions with m/z = 96, 137, 187 increased in time considerably. However, the growth in the intensity of these ions was accompanied by a rapid decreasing of the parent cation signal (m/z = 139). Therefore, one can assume that at the initial stage the ion with m/z = 82 originates from both NIP and decomposition products, but at the end of the isothermal evaporation experiment the decomposition products are the main molecular precursors of this ion.
The formation of bicyclic IL is not confirmed, because only one peak was registered in the GCMS experiment with a polar column and there were no traces of cyclization in NMR spectra.
To summarize, the vapor over BMImBF4 consists of NIPs and decomposition products according to routes a and b. Most of the previous investigations [17,18,19] used TGA and IR-spectroscopy to control the condensed phase of BMImBF4 and postulated the absence of any significant decomposition. No traces of these products was found from TGA and IR data on the condensed phase in our work. This discrepancy can be explained as follows. Analysis of the reported in literature [57,58,59,60,61] vapor pressures of potential dissociation products of IL under study (Figure 12) shows that they (with the exception of ylidene) are several orders of magnitude higher than those of prototypical IL BMImNTf2 [8]. Therefore, at experimental temperatures, these lightweight products cannot accumulate inside the effusion cell and they rapidly evaporate. That is why the IR-spectrum of the residue in an effusion cell is almost identical to that of the initial IL, whereas in the distillate collected from the cold parts of vacuum chamber the peaks attributed to imidazole-2-ylidene increased. The assessed vapor pressure of imidazole-2-ylidene is about one order of magnitude higher than that of IL. It leads to an amount of these vapor species becomes the higher the nearer are the evaporation conditions to equilibrium (closed system).
The differences in vapor composition under Knudsen and Langmuir conditions, as demonstrated in [12], can be explained by the kinetically hindered decomposition of IL. In a closed system (Knudsen cell) the evaporation flux is in equilibrium with the reverse flux from the cell walls. The highly volatile decomposition products accumulate inside the effusion cell and their pressure become considerable. The reverse flux is absent under Langmuir conditions, leading to the decrease in the pressure of the decomposition products, which is limited by the hindered decomposition speed.

6. Conclusions

The evaporation of BMImBF4 IL is characterized by a complex vapor composition which leads to the appearance of atypical ions in its EI mass spectrum at much lower temperatures (424–514 K) than decomposition temperatures obtained by the TGA (685 K) method. Combined analysis of the KEMS, GCMS, NMR, and IR-spectroscopy data together with a thermal analysis and quantum chemical modelling reveal three competing routes of BMImBF4 evaporation: (1) congruent in the form of NIPs; (2) with decomposition in the form of imidazole-2-ylidene and HF; and (3) with decomposition in the form of 1-methylimidazole, 1-butene, HF, and BF3. Two other possible routes of decomposition of BMImBF4 in the form of bicyclic IL and H2 as well as 1-butylimidazole, CH3F, and BF3 are found to be negligible. Quantitative analysis of the vapor composition and vaporization thermodynamics will be given in future papers.
The vapor composition of BMImBF4 strongly depends on the evaporation conditions. Under equilibrium conditions (Knudsen cell), decomposition products prevail in vapor, while under Langmuir conditions (open surface), evaporation in the form of NIP is preferred. Vapor composition is temperature-dependent as well: the amount of [BMIm+][BF4] NIPs relative to that of the decomposition products decreases by about a factor of three in the temperature range from 450 K to 510 K. The main reason for this specific evaporation of BMImBF4 is a high reactivity of the C1 atom in the imidazole ring, together with the high electronegativity of the anion. Similar peculiarities were observed for BMImPF6 evaporation and can be expected for all alkylimidazolium ILs with anions like BF4, PF6, AsF6, SCN, etc.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/e23111478/s1, Figure S1: Structures of BMImBF4 conformers, Figure S2. Experimental and theoretically predicted NMR spectra.

Author Contributions

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

Funding

This research was funded by Russian Science Foundation, grant number 21-73-00041, and Ministry of Science and Higher Education of Russia, grant number 075-15-2021-671.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study is available in article.

Acknowledgments

This work was supported by Russian Science Foundation under grant No. 21-73-00041. The study was carried out using the resources of the Center for Shared Use of Scientific Equipment of the ISUCT (with the support of the Ministry of Science and Higher Education of Russia, grant No. 075-15-2021-671).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wasserscheid, P.; Welton, T. (Eds.) Ionic Liquids in Synthesis; Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2002; p. 355. [Google Scholar]
  2. Earle, M.J.; Esperança, J.M.S.S.; Gilea, M.A.; Lopes, J.N.C.; Rebelo, L.P.N.; Magee, J.W.; Seddon, K.R.; Widegren, J.A. The distillation and volatility of ionic liquids. Nature 2006, 439, 831–834. [Google Scholar] [CrossRef]
  3. Armstrong, J.P.; Hurst, C.; Jones, R.G.; Licence, P.; Lovelock, K.R.J.; Satterley, C.J.; Villar-Garcia, I.J. Vapourisation of ionic liquids. Phys. Chem. Chem. Phys. 2007, 9, 982. [Google Scholar] [CrossRef] [PubMed]
  4. Leal, J.P.; Esperança, J.M.S.S.; da Piedade, M.E.M.; Lopes, J.N.C.; Rebelo, L.P.N.; Seddon, K.R. The nature of ionic liquids in the gas phase. J. Phys. Chem. A 2007, 111, 6176–6182. [Google Scholar] [CrossRef] [PubMed]
  5. Zaitsau, D.H.; Kabo, G.J.; Strechan, A.A.; Paulechka, Y.U.; Tschersich, A.; Verevkin, S.P.; Heintz, A. Experimental vapor pressures of 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imides and a correlation scheme for estimation of vaporization enthalpies of ionic liquids. J. Phys. Chem. A 2006, 110, 7303. [Google Scholar] [CrossRef] [PubMed]
  6. Chilingarov, N.S.; Medvedev, A.A.; Deyko, G.S.; Kustov, L.M.; Chernikova, E.A.; Glukhov, L.M.; Markov, V.Y.; Ioffe, I.N.; Senyavin, V.M.; Polyakova, M.V.; et al. Mass spectrometric studies of 1-ethyl-3-methylimidazolium and 1-propyl-2,3-dimethylimidazolium bis(trifluoromethyl)-sulfonylimides. Rapid Commun. Mass Spectrom. 2015, 29, 1227–1232. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, C.; Luo, H.; Li, H.; Dai, S. Direct UV-spectroscopic measurement of selected ionic-liquid vapors. Phys. Chem. Chem. Phys. 2010, 12, 7246–7250. [Google Scholar] [CrossRef] [PubMed]
  8. Brunetti, B.; Ciccioli, A.; Gigli, G.; Lapi, A.; Misceo, N.; Tanzi, L.; Ciprioti, S.V. Vaporization of the prototypical ionic liquid BMImNTf2 under equilibrium conditions: A multitechnique study. Phys. Chem. Chem. Phys. 2014, 16, 15653–15661. [Google Scholar] [CrossRef] [PubMed]
  9. Dunaev, A.M.; Motalov, V.B.; Kudin, L.S.; Sergeev, D.N.; Akopyan, A.V. Decomposition temperatures, IR and mass spectra of some chiral ionic liquids based on alkylimidazolium. Russ. Chem. J. 2017, 61, 35–41. [Google Scholar]
  10. Clarke, C.J.; Puttick, S.; Sanderson, T.J.; Taylor, A.W.; Bourne, R.A.; Lovelock, K.R.J.; Licence, P. Thermal stability of dialkylimidazolium tetrafluoroborate and hexafluorophosphate ionic liquids: Ex situ bulk heating to complement in situ mass spectrometry. Phys. Chem. Chem. Phys. 2018, 20, 16786–16800. [Google Scholar] [CrossRef]
  11. Chambreau, S.D.; Schenk, A.C.; Sheppard, A.J.; Yandek, G.R.; Vaghjiani, G.L.; Maciejewski, J.; Koh, C.J.; Golan, A.; Leone, S.R. Thermal Decomposition Mechanisms of Alkylimidazolium Ionic Liquids with Cyano-Functionalized Anions. J. Phys. Chem. A 2014, 118, 11119–11132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Volpe, V.; Brunetti, B.; Gigli, G.; Lapi, A.; Ciprioti, S.V.; Ciccioli, A. Toward the Elucidation of the Competing Role of Evaporation and Thermal Decomposition in Ionic Liquids: A Multitechnique Study of the Vaporization Behavior of 1-Butyl-3-methylimidazolium Hexafluorophosphate under Effusion Conditions. J. Phys. Chem. B 2017, 121, 10382–10393. [Google Scholar] [CrossRef] [PubMed]
  13. Dunaev, A.M.; Motalov, V.B.; Kudin, L.S.; Butman, M.F. Molecular and Ionic Composition of Saturated Vapor over EMImNTf2 Ionic Liquid. J. Mol. Liq. 2016, 219, 599–601. [Google Scholar] [CrossRef]
  14. Deyko, A.; Lovelock, K.R.J.; Licence, P.; Jones, R.G. The vapour of imidazolium-based ionic liquids: A mass spectrometry study. Phys. Chem. Chem. Phys. 2011, 13, 16841–16850. [Google Scholar] [CrossRef] [PubMed]
  15. Meine, N.; Benedito, F.; Rinaldi, R. Thermal stability of ionic liquids assessed by potentiometric titration. Green Chem. 2010, 12, 1711–1714. [Google Scholar] [CrossRef]
  16. Swiderski, K.; McLean, A.; Gordon, C.M.; Vaughana, D.H. Estimates of internal energies of vaporisation of some room temperature ionic liquids. Chem. Commun. 2004, 19, 2178–2179. [Google Scholar] [CrossRef]
  17. Zaitsau, D.H.; Yermalayeu, A.V.; Schubert, T.J.S.; Verevkin, S.P. Alkyl-imidazolium tetrafluoroborates: Vapor pressure, thermodynamics of vaporization, and enthalpies of formation. Mol. Liq. 2017, 242, 951–957. [Google Scholar] [CrossRef]
  18. Liang, R.; Yang, M.; Zhou, Q. Thermal Stability, Equilibrium Vapor Pressure and Standard Enthalpy of Vaporization of 1-Butyl,3-methylimidazoliumTetrafluoroborate. Acta Phys.-Chim. Sin. 2010, 26, 1468–1472. [Google Scholar] [CrossRef]
  19. Krannich, M.; Heym, F.; Jess, A. Characterization of Six Hygroscopic Ionic Liquids with Regard to Their Suitability for Gas Dehydration: Density, Viscosity, Thermal and Oxidative Stability, Vapor Pressure, Diffusion Coefficient, and Activity Coefficient of Water. J. Chem. Eng. Data 2016, 61, 1162–1176. [Google Scholar] [CrossRef]
  20. Deyko, A.; Lovelock, K.R.J.; Corfield, J.-A.; Taylor, A.W.; Gooden, P.N.; Villar-Garcia, I.J.; Licence, P.; Jones, R.G.; Krasovskiy, V.G.; Chernikova, E.A.; et al. Measuring and predicting ΔvapH298 values of ionic liquids. Phys. Chem. Chem. Phys. 2009, 11, 8544–8555. [Google Scholar] [CrossRef]
  21. Lovelock, K.R.J.; Deyko, A.; Licence, P.; Jones, R.G. Vaporisation of an ionic liquid near room temperature. Phys. Chem. Chem. Phys. 2010, 12, 8893–8901. [Google Scholar] [CrossRef] [PubMed]
  22. Deyko, A.; Hessey, S.G.; Licence, P.; Chernikova, E.A.; Krasovskiy, V.G.; Kustov, L.M.; Jones, R.G. The enthalpies of vaporisation of ionic liquids: New measurements and predictions. Phys. Chem. Chem. Phys. 2012, 14, 3181–3193. [Google Scholar] [CrossRef]
  23. Fredlake, C.P.; Crosthwaite, J.M.; Hert, D.G.; Aki, S.N.V.K.; Brennecke, J.F. Thermophysical Properties of Imidazolium-Based Ionic Liquids. J. Chem. Eng. Data 2004, 49, 954–964. [Google Scholar] [CrossRef]
  24. Ngo, H.L.; LeCompte, K.; Hargens, L.; McEwen, A.B. Thermal properties of imidazolium ionic liquids. Thermochim. Acta 2000, 357, 97–102. [Google Scholar] [CrossRef]
  25. Huddleston, J.G.; Visser, A.E.; Reichert, W.M.; Willauer, H.D.; Broker, G.A.; Rogers, R.D. Characterization and comparison of hydrophilic and hydrophobic room temperature ionic liquids incorporating the imidazolium cation. Green Chem. 2001, 3, 156–164. [Google Scholar] [CrossRef]
  26. Eapen, T.; Deepthi, T.; Kunduchi, P.V.; Benny, K.G. Mechanistic outlook on thermal degradation of 1,3-dialkyl imidazolium ionic liquids and organoclays. RSC Adv. 2016, 6, 9421–9428. [Google Scholar] [CrossRef]
  27. Ohtani, H.; Ishimura, S.; Kumai, M. Thermal decomposition behaviors of imidazolium-type ionic liquids studied by pyrolysis-gas chromatography. Anal. Sci. 2008, 24, 1335–1340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Arduengo, A.J.; Harlow, R.L.; Kline, M. A stable crystalline carbine. J. Am. Chem. Soc. 1991, 113, 361–363. [Google Scholar] [CrossRef]
  29. Taylor, A.W.; Lovelock, K.R.J.; Jones, R.G.; Licence, P. Borane-substituted imidazol-2-ylidenes: Syntheses in vacuo. Dalton Trans. 2011, 40, 1463–1470. [Google Scholar] [CrossRef]
  30. Kan, H.; Tsengy, M.; Chu, Y. Bicyclic imidazolium-based ionic liquids: Synthesis and characterization. Tetrahedron 2007, 63, 1644–1653. [Google Scholar] [CrossRef]
  31. Knorr, M.; Icker, M.; Efimova, A.; Schmidt, P. Reactivity of Ionic Liquids: Studies on Thermal Decomposition Behavior of 1-Butyl-3-methylimidazolium Tetrafluoroborate. Thermochim. Acta 2020, 694, 178786. [Google Scholar] [CrossRef]
  32. Chaudhary, G.R.; Bansal, S.; Mehta, S.K.; Ahluwalia, A.S. Thermophysical and Spectroscopic Studies of Pure 1-Butyl-3-methylimidazolium Tetrafluoroborate and Its Aqueous Mixtures. J. Solut. Chem. 2014, 43, 340–359. [Google Scholar] [CrossRef]
  33. Cha, S.; Ao, M.; Sung, W.; Moon, B.; Ahlström, B.; Johansson, P.; Ouchi, Y.; Kim, D. Structures of ionic liquid-water mixtures investigated by IR and NMR spectroscopy. Phys. Chem. Chem. Phys. 2014, 16, 9591–9601. [Google Scholar] [CrossRef]
  34. Zheng, Y.-Z.; Wang, N.-N.; Luo, J.-J.; Zhou, Y.; Yu, Z.-W. Hydrogen-bonding interactions between [BMIM][BF4] and acetonitrile. Phys. Chem. Chem. Phys. 2013, 15, 18055–18064. [Google Scholar] [CrossRef] [PubMed]
  35. Dharaskar, S.A.; Wasewar, K.L.; Varma, M.N.; Shende, D.Z.; Yoo, C. Synthesis, characterization and application of 1-butyl-3-methylimidazolium tetrafluoroborate for extractive desulfurization of liquid fuel. Arab. J. Chem. 2016, 9, 578–587. [Google Scholar] [CrossRef] [Green Version]
  36. Li, Y.; Hu, Y.; Chen, G.; Wang, Z.; Jin, X. Rapid proton diffusion in hydroxyl functionalized imidazolium ionic liquids. Sci. China Ser. B Chem. 2017, 60, 734–739. [Google Scholar] [CrossRef]
  37. Dunaev, A.M.; Motalov, V.B.; Kudin, L.S. A High-Temperature Mass-Spectrometric Method for Determination of the Electron Work Function of Ionic Crystals: Lanthanum, Cerium, and Praseodymium Triiodides. Russ. J. Gen. Chem. 2017, 87, 632–638. [Google Scholar] [CrossRef]
  38. Dunaev, A.M.; Kryuchkov, A.S.; Kudin, L.S.; Butman, M.F. Automatic complex for high temperature investigation on basis of mass spectrometer MI1201. Izv. Vyssh. Uchebn. Zaved. Khim. Khim. Tekhnol. 2011, 54, 73–77. (In Russian) [Google Scholar]
  39. Sergeev, D.N.; Dunaev, A.M.; Ivanov, D.A.; Golovkina, Y.A.; Gusev, G.I. Automatization of mass spectrometer for the obtaining of ionization efficiency functions. Prib. Tekhnika Eksperimenta 2014, 1, 139–140. (In Russian) [Google Scholar] [CrossRef]
  40. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  41. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  42. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  43. Dunning, J. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  44. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.02; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  45. van Valkenburg, M.E.; Vaughn, R.L.; Williams, M.; Wilkes, J.S. Thermochemistry of ionic liquid heat-transfer fluids. Thermochim. Acta 2005, 425, 181–188. [Google Scholar] [CrossRef]
  46. Schreiner, C. Synthese und Charakterisierung neuer Ionischer Flüssigkeiten auf der Basis gemischter Fluoroborat-Anionen. Ph.D. Thesis, University of Regensburg, Regensburg, Germany, 2009. [Google Scholar]
  47. Holbrey, J.D.; Seddon, K.R. The phase behaviour of 1-alkyl-3-methylimidazolium tetrafluoroborates; ionic liquids and ionic liquid crystals. Chem. Soc. Dalton Trans. 1999, 2133–2140. [Google Scholar] [CrossRef]
  48. Erdmenger, T.; Vitz, J.; Wiesbrock, F.; Schubert, U.S. Influence of different branched alkyl side chains on the properties of imidazolium-based ionic liquids. J. Mater. Chem. 2008, 18, 5267–5273. [Google Scholar] [CrossRef]
  49. Hog, M.; Schneider, M.; Studer, G.; Bäuerle, M.; Föhrenbacher, S.A.; Scherer, H.; Krossing, I. An Investigation of the Symmetric and Asymmetric Cleavage Products in the System Aluminum Trihalide/1-Butylimidazole. Chem. A Eur. J. 2017, 23, 11054–11066. [Google Scholar] [CrossRef] [PubMed]
  50. Pachler, K.G.R.; Pachter, R.; Wessels, P.L. Carbon-13proton coupling constants in N-substituted imidazoles. A 13C NMR study and MO calculations. Magn. Reson. Chem. 1981, 17, 278–284. [Google Scholar] [CrossRef]
  51. Kimura, K.; Katsumata, S.; Achiba, Y.; Yamazaki, T.; Iwata, S. Ionization energies, Ab initio assignments, and valence electronic structure for 200 molecules. In Handbook of HeI Photoelectron Spectra of Fundamental Organic Compounds; Japan Scientific Societies Press: Tokyo, Japan, 1981; p. 268. [Google Scholar]
  52. Strasser, D.; Goulay, F.; Kelkar, M.S.; Maginn, E.J.; Leone, S.R. Photoelectron Spectrum of Isolated Ion-Pairs in Ionic Liquid Vapor. J. Phys. Chem. A 2007, 111, 3191–3195. [Google Scholar] [CrossRef]
  53. Farber, M.; Srivastava, R.D.; Moyer, J.W. Mass spectrometric determination of the thermodynamics of potassium hydroxide and minor potassium-containing species required in magnetohydrodynamic power systems. J. Chem. Thermodyn. 1982, 14, 1103–1113. [Google Scholar] [CrossRef]
  54. Dibeler, V.H.; Liston, S.K. Mass-spectrometric study of photoionization. XII. Boron trifluoride and diboron tetrafluoride. Inorg. Chem. 1968, 7, 1742–1746. [Google Scholar] [CrossRef]
  55. Pogrebnoi, A.M.; Kudin, L.S.; Motalov, V.B.; Goryushkin, V.F. Vapor species over cerium and samarium trichlorides, enthalpies of formation of (LnCl3)n molecules and Cl−(LnCl3)n ions. Rapid Commun. Mass Spectrom. 2001, 15, 1662–1671. [Google Scholar] [CrossRef] [PubMed]
  56. Acree, W.E., Jr.; Chickos, J.S. NIST Chemistry WebBook, NIST Standard Reference Database Number 69; Linstrom, P.J., Mallard, W.G., Eds.; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1998; p. 20899. [Google Scholar] [CrossRef]
  57. Stull, D.R. Vapor Pressure of Pure Substances. Organic and Inorganic Compounds. Ind. Eng. Chem. 1947, 39, 517–540. [Google Scholar] [CrossRef]
  58. Sheft, I.; Perkins, A.J.; Hyman, H.H. Anhydrous Hydrogen Fluoride: Vapor Pressure and Liquid Density. J. Inorg. Nucl. Chem. 1973, 35, 3677–3680. [Google Scholar] [CrossRef]
  59. Michels, A.; Wassenaar, T. Vapour pressure of methylfluoride. Physica 1948, 14, 104–110. [Google Scholar] [CrossRef]
  60. Coffin, C.C.; Maass, O. The Preparation and Physical Properties of α-,β- and γ-Butylene and Normal and Isobutane. J. Am. Chem. Soc. 1928, 50, 1427–1437. [Google Scholar] [CrossRef]
  61. Jiménez, P.; Roux, M.V.; Turrión, C. Thermochemical properties of N-heterocyclic compounds IV. Enthalpies of combustion, vapour pressures and enthalpies of sublimation, and enthalpies of formation of 2-methylimidazole and 2-ethylimidazole. J. Chem. Thermodyn. 1992, 24, 1145–1149. [Google Scholar] [CrossRef]
Figure 1. Schemes of BMImBF4 decomposition with formation of imidazole-2-ylidene and HF (a); 1-methylimidazole, 1-butene, HF, and BF3 (b); 1-butylimidazole, fluoromethane, and BF3 (c); H2 and bicyclic IL (d).
Figure 1. Schemes of BMImBF4 decomposition with formation of imidazole-2-ylidene and HF (a); 1-methylimidazole, 1-butene, HF, and BF3 (b); 1-butylimidazole, fluoromethane, and BF3 (c); H2 and bicyclic IL (d).
Entropy 23 01478 g001
Figure 2. Computed vapor composition of the BMImBF4 conformers at 300–500 K (the structures corresponding to designations see in Supplementary Materials, Figure S1).
Figure 2. Computed vapor composition of the BMImBF4 conformers at 300–500 K (the structures corresponding to designations see in Supplementary Materials, Figure S1).
Entropy 23 01478 g002
Figure 3. Structure of conformer 5a.
Figure 3. Structure of conformer 5a.
Entropy 23 01478 g003
Figure 4. DTA results for BMImBF4.
Figure 4. DTA results for BMImBF4.
Entropy 23 01478 g004
Figure 5. NMR spectra of BMImBF4 at 295–343 K: (a) 1H; (b) 15N; (c) 13C; (d) 11B. Atom numbering is in accordance with Figure 3.
Figure 5. NMR spectra of BMImBF4 at 295–343 K: (a) 1H; (b) 15N; (c) 13C; (d) 11B. Atom numbering is in accordance with Figure 3.
Entropy 23 01478 g005
Figure 6. EI background subtracted mass spectrum of BMImBF4 at 472 K.
Figure 6. EI background subtracted mass spectrum of BMImBF4 at 472 K.
Entropy 23 01478 g006
Figure 7. Temperature dependencies of ion currents (a) and ionization efficiency curves (b). The intensities of ions with m/z = 139 and 187 are scaled by a factor 0.25; m/z = 49 by 0.14, and m/z = 158 by 0.11 to be more illustrative.
Figure 7. Temperature dependencies of ion currents (a) and ionization efficiency curves (b). The intensities of ions with m/z = 139 and 187 are scaled by a factor 0.25; m/z = 49 by 0.14, and m/z = 158 by 0.11 to be more illustrative.
Entropy 23 01478 g007
Figure 8. Time dependencies of ion currents at 480 K.
Figure 8. Time dependencies of ion currents at 480 K.
Entropy 23 01478 g008
Figure 9. Intensity ratio of ion currents of BMImBF4 (a) and BMImPF6 [12] (b).
Figure 9. Intensity ratio of ion currents of BMImBF4 (a) and BMImPF6 [12] (b).
Entropy 23 01478 g009
Figure 10. Experimental IR-spectra of condensed phase of BMImBF4 and theoretical IR-spectra of BMImBF4, imidazole-2-ylidene, and bicyclic IL.
Figure 10. Experimental IR-spectra of condensed phase of BMImBF4 and theoretical IR-spectra of BMImBF4, imidazole-2-ylidene, and bicyclic IL.
Entropy 23 01478 g010
Figure 11. GCMS mass spectra and ion profiles on chromatograms of BMImBF4 with polar (a,b) and nonpolar chromatograph column without (c,d) and with (e,f) addition of ethanol.
Figure 11. GCMS mass spectra and ion profiles on chromatograms of BMImBF4 with polar (a,b) and nonpolar chromatograph column without (c,d) and with (e,f) addition of ethanol.
Entropy 23 01478 g011
Figure 12. Temperature dependencies of vapor pressures of some decomposition products of BMImBF4. Temperature dependence of BMImNTf2 is shown for comparison. Assessed pressure of ylidene was marked as a circle.
Figure 12. Temperature dependencies of vapor pressures of some decomposition products of BMImBF4. Temperature dependence of BMImNTf2 is shown for comparison. Assessed pressure of ylidene was marked as a circle.
Entropy 23 01478 g012
Table 1. Ion appearance energies AE and coefficients a of the ln(I∙T) = −1000∙a/T + b dependency at T = 488 K.
Table 1. Ion appearance energies AE and coefficients a of the ln(I∙T) = −1000∙a/T + b dependency at T = 488 K.
m/zAE, eVaAssigned Ion
4916.9 ± 0.512.970 ± 0.274BF2+
8213.0 ± 0.517.486 ± 0.364MIm+
9611.3 ± 0.517.365 ± 0.442MMIm+
13711.0 ± 0.517.327 ± 0.455C8H13N2+
13912.4 ± 0.514.249 ± 0.390BMIm+
15811.7 ± 0.513.445 ± 0.363BMImF+
18713.8 ± 0.517.968 ± 0.510BMImBF2+
A standard uncertainty is given with the “±” sign.
Table 2. Mass spectra of BMImBF4 recorded at different evaporation conditions.
Table 2. Mass spectra of BMImBF4 recorded at different evaporation conditions.
T, Km/z
8296137139158187
Effusion cell487250455951001868
Open cell47015728463100**-
Open surface471525221100**-
Open surface *50144231510013-
*—reproduced as well as possible from Figure 3 in [14]; **—not measured.
Table 3. The relative intensities of four main peaks in mass spectra of BMImBF4 and imidazole-2-ylidene.
Table 3. The relative intensities of four main peaks in mass spectra of BMImBF4 and imidazole-2-ylidene.
CompoundT, Km/z
8296137187
GCMS (nonpolar column)BMImBF4523341001819
KEMS (effusion cell)BMImBF4487551002115
DIMS [10]Imidazole-2-ylidene323281002764
Table 4. List of the main peaks in mass spectra [56] according to different decomposition processes (for a, b, c, and d designations see Figure 1).
Table 4. List of the main peaks in mass spectra [56] according to different decomposition processes (for a, b, c, and d designations see Figure 1).
ProcessThe Main Peaks with m/z
BMImBF4(s) = BMImBF4(g)82, 139, 158
a82, 96, 137, 187
b20, 28, 41, 42, 49, 54, 56, 81, 82
c33, 34, 49, 81, 82, 97
d137
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dunaev, A.M.; Motalov, V.B.; Kudin, L.S. The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process. Entropy 2021, 23, 1478. https://doi.org/10.3390/e23111478

AMA Style

Dunaev AM, Motalov VB, Kudin LS. The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process. Entropy. 2021; 23(11):1478. https://doi.org/10.3390/e23111478

Chicago/Turabian Style

Dunaev, Anatoliy M., Vladimir B. Motalov, and Lev S. Kudin. 2021. "The Composition of Saturated Vapor over 1-Butyl-3-methylimidazolium Tetrafluoroborate Ionic Liquid: A Multi-Technique Study of the Vaporization Process" Entropy 23, no. 11: 1478. https://doi.org/10.3390/e23111478

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