Next Article in Journal
Enhancement of Light Olefins Selectivity Over N-Doped Fischer-Tropsch Synthesis Catalyst Supported on Activated Carbon Pretreated with KMnO4
Previous Article in Journal
Iridium-Catalyzed Transfer Hydrogenation of Ketones and Aldehydes Using Glucose as a Sustainable Hydrogen Donor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergy between Ionic Liquids and CuCl2 in Gas–Liquid Phase Reactions of Acetylene Hydrochlorination

1
Industrial Catalysis Insttute, Laboratory Breeding Base of Green Chemistry-Synthesis Technology, Zhejiang University of Technology, Hangzhou 310014, China
2
College of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310014, China
3
College of Environment, Zhejiang University of Technology, Hangzhou 310014, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(6), 504; https://doi.org/10.3390/catal9060504
Submission received: 9 May 2019 / Revised: 28 May 2019 / Accepted: 30 May 2019 / Published: 31 May 2019
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)

Abstract

:
We studied the acetylene hydrochlorination in gas–liquid phase reactions using ionic liquids (IL) as the reaction media and CuCl2 as the catalyst. The Cu-IL catalyst showed strong synergy between the IL and the Cu(II) active catalytic species. For [PrMIm]Cl, the Cu-IL catalyst exhibited significant enhancement of the catalytic activity in comparison with the CuCl2 catalyst supported on activated carbon and the IL alone as the catalyst. We have also performed DFT calculations of the reaction process, which provides a good explanation of our experimental results and for the synergetic effect. Our result suggests that ILs may be used to improve the activity of other metallic catalysts for the hydrochlorination reaction of acetylene.

1. Introduction

The conventional catalyst HgCl2 for acetylene hydrochlorination in the PVC industry has to be replaced by mercury-free catalysts worldwide due to its pollution to the environment. Early in 1984, Hutchings uncovered a correlation between the standard electrode potential of various metals and their catalytic activity, and proposed that gold is a highly active catalyst for acetylene hydrochlorination [1]. Since then, a lot of focus havs been concentrated on noble metals including Au, Pt, and Pd as catalysts for their high standard electrode potentials [2,3,4]. However, these noble metal based catalysts are very costly, and still suffer from issues of loss of activity [5]. In recent years, searching for alternative non-noble metal catalysts with low costs while still maintaining high catalytic activity has attracted a lot of interest and taken a lot of research effort.
Among the non-noble metal catalysts that have been studied in the literature are Cu, Sn, Mn, Zn, Co, and Bi [6,7,8,9,10,11]. These metals often have relatively lower standard electrode potentials when compared with noble metals. Therefore, a mono-metallic catalyst of these metals usually does not have very high catalytic activity. In order to improve the catalytic activity, bimetallic and trimetallic catalysts have been prepared and investigated in the literature [11,12,13] in the hope that the synergy between multicomponents of metals may lead to much enhanced activity. The non-noble metals may act as an assistive to a noble metal catalyst, or they can cooperate with other non-noble metals to form bimetallic or trimetallic catalysts [14,15].
Recently, the extraordinary properties of ionic liquids (ILs), such as low vapor pressure and high salvation strength, have caught a lot of attention [16,17,18]. There have been many reports of significant promotion of catalytic activity by ILs [18]. Furthermore, ILs alone have been used as a mercury-free catalyst as well. Cao and coworkers [19] studied the influence of some ionic liquids on the performance of mercury-free catalysts and found that [BMIm]Cl could increase the acetylene conversion rate up to 68%. In addition, they also studied its effect on the performance of other metal chlorides that act as the main active species of the catalysts. Except for SnCl4, they found significant enhancement of the acetylene conversion rate for HgCl2, MnCl4, H2[PtCl6], and H[AuCl4]. Zhao et al. prepared supported IL phase (SILP) stabilized Au(III) catalyst for acetylene hydrochlorination, without [20] and with [21] the additive of CuCl2, and reported an outstanding catalytic performance. Zhang et al. [22] recently reported that the imidazolium type of IL, [BMIm]BF4, may substantially increase the performance of a Ru catalyst supported on activated carbon (AC), despite it being not a good catalyst on its own. Lately, Cao et al. [23] investigated the performance of imidazolium ILs as mercury-free catalysts for acetylene hydrochlorination. They found that [BMIm]Cl exhibited good catalytic performance. Strong promotional effect of ILs on Pd catalyst using the SILP technique has also been reported lately [24]. Thus far, existing research demonstrates that ILs may promote the performance of a catalyst. Nonetheless, there are still no systematic study of such enhancement effect, such as extent of the enhancement, decisive factors, and underlying mechanism. For example, Cao et al. [19] did not provide comparison against the performance of the metal chloride catalysts under the same preparation and reaction conditions but without the ILs.
On the other hand, studies reveal that nitrogen doping to the activated carbon support may significantly enhance the catalytic activity [25,26,27,28]. This has been attributed to the modification effect of the nitrogen-containing species to the surface chemistry of the AC support. It also demonstrates that these nitrogen-containing species may serve as active sites for catalytic reactions [29,30,31,32]. Based on these studies, here we chose imidazolium based ILs, which contain two N atoms per molecule, as our main reaction media. At the same time, we prepared the catalyst by adding non-noble metal chlorides, including CuCl2, FeCl3, SnCl2, ZnCl2, BiCl2, and BiCl2. These metal chlorides have been studied as an additive in preparing bimetallic catalysts with Au (or Pt and Pd), which may possess higher activity than the mono-metallic Au catalyst [7,15]. Our result showed that CuCl2 has the best catalytic performance. When combined with ILs, CuCl2 and the N sites of the imidazolium ring act together to increase the adsorption of HCl and C2H2, and thus facilitate the hydrochlorination reaction of acetylene. In this paper, we will focus our study on the synergetic effect of ILs and metals, and find out the best combination as a catalyst for acetylene hydrochlorination, and try to reveal the underlying mechanism.
Our experimental result demonstrates that [PrMIm]Cl and CuCl2 can cooperate and exhibit a good synergy in catalyzing acetylene hydrochlorination reactions. The synergetic catalyst exhibits much better performance than the CuCl2/AC catalyst and [PrMIm]Cl alone as a catalyst. To understand this result, we have performed a mechanistic study using DFT calculations.
The authors of [23] proposed a model for the reaction process of [BMIm]Cl alone as a mercury-free catalyst for acetylene hydrochlorination. However, there have been no calculations for [PrMIm]Cl as a catalyst, nor for the catalyst CuCl2-[PrMIm]Cl. For this reason, we have performed quantum chemistry calculations using the DFT theory. Our calculations provide a good explanation for our experimental result and the synergy between the IL and CuCl2.

2. Results

In this section, we will first present the result of the evaluation experiments, followed by characterizations of the catalysts. Then we will show our result of DFT calculations, to reveal a mechanism for the enhancement effect of the ILs.

2.1. Experimental Results

We analyzed the thermal stability of the ILs using thermal gravimetric analysis (TGA). The TGA curves for [PrMIm]Cl and [BMIm]Cl are shown in Supplementary Figure S1. The thermal decomposition products were determined by infrared FT-IR measurements. The FT-IR data indicate that the onset of mass loss is about 195 °C and 150 °C for [BMIm]Cl and [PrMIm]Cl, respectively, far below the thermal decomposition temperature Td. Therefore, for [BMIm]Cl and [PrMIm]Cl, the ideal reaction temperature should be kept below 195 °C and 150 °C, respectively, in order to maintain long term stability.
The TGA result for other ILs in our experiments was similar, with rapid thermal decomposition occurring around the thermal decomposition temperature Td (see Table 2 below in the Materials section). As for the small amount of decomposition products, there are thus far no conclusive data to tell whether such decomposition will continue or stop at this temperature. The TGA analysis cannot provide such information, since the temperature was raised typically at a rate of 10 °C /min, and thus only provides short-term data.
Shown in Figure 1 are the FT-IR spectra of (a) the IL [PrMIm]Cl and (b) the catalyst [PrMIm]Cl-CuCl2, respectively. The broad peak around 3400 cm−1 was from the stretching modes of various O–H bonds from water contamination. The band around 3000 cm−1 mainly came from the C–H bonding in the IL. In addition, there should also be broad vibrational modes associated the N–H bonds around 3200 cm−1 [33]. In comparison, we find that the absorption peak at 1458.6 cm−1 was shifted up to 1462.7 cm−1, and the peak at 3073.3 cm−1 was shifted to 3076.5 cm−1, while the peak at 1636.3 cm−1 was shifted down to 1634.4 cm−1. Shown in Figure 2 is the calculated adsorption structure of [CuCl4]2− on the cation [PrMIm]+. Based on this structure, we calculated the IR absorption spectrum. After careful comparison and review of the animation for the vibration modes of the calculated spectrum, we found that the 1458.6 cm−1 and 1462.7 cm−1 modes involved rather complex collective vibrations of stretching and twisting of the C–C and C–N bonds on the imidazolium ring and the C–H bonds on the propyl chain. The 3073.3 cm−1 and 3076.5 cm−1 peaks mainly involved the stretching modes of the C–H bonds on the methyl group and the propyl chain. The downshifted vibrations at 1636.3 cm−1 and 1634.4 cm−1 mainly came from the stretching of the C=C double bond on the ring that was farther away from the [CuCl4]2− complex, as shown in Figure 2. Evidently, the shift in vibrational frequencies was caused by the adsorption and interaction of the Cu complex near the imidazolium ring.
In Figure 3, we present (a) the measured C2H2 conversion rate XA and (b) the vinyl chloride monomer (VCM) selectivity SVCM as a function of on-stream reaction time, for the evaluation experiments of the acetylene hydrochlorination reaction using the CuCl2-[PrMIm]Cl catalyst at different temperatures. Here we used our typical reaction conditions: CuCl2 concentration = 0.065 mol L−1, C2H2 gas hourly space velocity (GHSV) = 370 h−1, and the reactant feed ratio VHCl:VC2H2 = 1.2. There was an initial incubation period for the catalyst, and then the conversion rate reached a maximum. The data in Figure 3a indicate that at the reaction temperature of 140 °C, the maximum C2H2 conversion rate reached up to 39%, with a VCM selectivity of about 96%. The maximum conversion rate started to decrease at higher and lower temperatures. The maximum C2H2 conversion rate and the corresponding VCM selectivity are plotted as a function of temperature in Supplementary Figure S2. From both Figure 3 and Figure S2, we concluded that for a given CuCl2 concentration and the C2H2 GHSV, the best reaction temperature for catalyst CuCl2-[PrMIm]Cl was about 140 °C. It exhibited high stability; the conversion rate dropped by only 4% after 12 h on stream.
To understand why the initial incubation period was necessary, we did XPS analysis of the fresh CuCl2-[PrMIm]Cl catalyst as prepared before the reaction. The 2p core level XPS spectrum of Cu is shown in Figure 4. The spectrum was analyzed using the XPSPeak 4.1 software, and the C 1s core level peak at 284.8 eV was used to calibrate the binding energy, to get rid of possible charge effect. We obtained the binding energy of 932.24 eV for the Cu 2p3/2 electrons. Comparison with typical data for various valence states of Cu in the XPS handbook [34] indicates that this binding energy is close to that of CuCl and Cu0, while the binding energy of CuCl2 is 935.0 ±0.5 eV. Among various compounds containing Cu2+, only for CuS and Cu(OAc)2, the binding energy of the Cu 2p3/2 may be compatible to Figure 4. Clearly, they were not present in our samples. Furthermore, a typical Cu2+ peak is expected to have strong companion peaks, whereas we only found relatively rather weak companion peaks in the figure at 959.9 eV and 941.6 eV, with a relative signal strength that is consistent with Cu+ or Cu0. Therefore, Figure 4 suggests that CuCl2 in [PrMIm]Cl was reduced to CuCl. The introduction of HCl could re-oxidize CuCl into CuCl2 [35]. Indeed, the oxidization capability of HCl on metallic catalytic species in acetylene hydrochlorination have been known for a long time [2,5]. Similar reduction of the cationic catalytic species has been known essentially for all other metal catalysts for acetylene hydrochlorination as well, including the highly active Au/C catalyst. Therefore, it is a common phenomenon that an initial incubation period for the catalytic activity is needed [36]. Hutchings and coworkers proposed that the catalytically active Au3+ species could be reduced to Au0 by the carbon support during the catalyst preparation process, while other factors also had significant influences [37]. While it is well known that C2H2 can reduce the cationic catalytic species and cause loss of catalytic activity [21], we suspect that the IL [PrMIm]Cl itself might also become reductive during the catalyst preparation process, in the absence of the oxidative HCl.
We repeated the above experiment but for other ILs with anion Cl. Among these ILs, the CuCl2-[BMIm]Cl combination had an optimal reaction temperature of about 160 °C. Nonetheless, the difference between 140 °C and 160 °C was small. We obtained an initial maximum acetylene conversion rate of about 32%. It had a VCM selectivity of about 96%, close to that of the CuCl2-[PrMIm]Cl combination. Other CuCl2-IL combinations exhibited relatively lower catalytic activity.
In Table 1, we list the maximum C2H2 conversion rate and corresponding VCM selectivity for the CuCl2 catalyst when different ILs were used as the reaction media. In order to facilitate the comparison, we chose the same reaction conditions including temperature 140 °C and the same C2H2 GHSV.
Shown in Figure 5 is the bar chart of the data in Table 1. It tells that among all ILs studied, [PrMIm]Cl had the highest acetylene conversion rate. In terms of VCM selectivity, the difference between these ILs was not as dramatic. In fact, the main difference between these different ILs lies in the length of the alkyl chains, as one can tell from their structures as listed in Supplementary Table S1. From left to right in Figure 5, namely, from EMIm to OMIm, the names of the cations can be represented uniformly by CnMIm, with n = 2, 3, 4, 6, 8. The allyl group in AMIm contains a C=C double bond, HEMIm contains an extra hydroxyl group –OH, and BMMIm contains an extra methyl group. These detailed structural differences have a strong influence on the adsorption of the Cu complexes and the formation of intermediate transition states, and have thus changed the catalytic activity of the corresponding CuCl2-IL catalysts.
We also studied the catalytic performance of different metal chlorides as catalysts using the ILs as the reaction media. The result for acetylene conversion rate and the VCM selectivity using [PrMIm]Cl as the reaction media is shown in Table 2 and the bar chart of Figure 6. From Table 2 and Figure 6, it is easily seen that except CuCl2 for which the C2H2 conversion rate was nearly 40%, the catalytic activities of all other metal chlorides were low. Overall, the trend for the activity exhibits a positive correlation with the standard electrode potential of the metal ions, in agreement with [1]. We also tried [PrMIm]Cl alone as the catalyst, but its activity turned out to be very low.
Combining Figure 5 and Figure 6, we concluded that the best combination was CuCl2-[PrMIm]Cl. Therefore, we shall focus our study on this combination.
Next, we studied the effect of the concentration of CuCl2 on the performance of the CuCl2-[PrMIm]Cl catalyst. Shown in Figure 7 are the conversion rate and selectivity as a function of CuCl2 concentration. The result reveals that, in a large range of the CuCl2 concentration, the dependence of the catalytic activity on the CuCl2 concentration is very weak. It is clear that the conversion was limited by mass transfer as the C2H2 conversion rate saturates. In principle, in the very low concentration kinetic regime, the conversion rate should be roughly proportional to the concentration. As the concentration increases, the conversion rate tends to saturate. The data in Figure 7 indicate that the kinetic regime of proportionality was lower than 0.025 mol L−1, which is the lowest concentration studied in the experiment. Indeed, we see a slight increase of the C2H2 conversion rate as a function of CuCl2 concentration.
For comparison, we also performed a control experiment using [PrMIm]Cl alone as the catalyst, and the result demonstrated that the acetylene conversion rate was negligibly small, as compared to the combination. In addition, we performed a similar reaction using AC supported CuCl2 as the catalyst, with the same amount of Cu as for the lowest CuCl2 concentration (0.025 mol L−1) case in Figure 7. After 2 h incubation time, the acetylene conversion rate reached its maximum of 18.5%, under the same reaction conditions. This was substantially lower than the maximum conversion rate, 38%, of the CuCl2-[PrMIm]Cl catalyst at [CuCl2] = 0.025 mol L−1, which is also close to the maximum rate 39% at [CuCl2] = 0.065 mol L−1. It should be noted, however, due to the presence of the AC support, the experiment with the Cu/AC catalyst cannot be viewed as a good control. Nonetheless, it might be reasonable to assume that, without the AC support, the performance of CuCl2 alone would be poorer. It should be emphasized, however, that since the kinetic regime is below the lowest CuCl2 concentration in Figure 7, here we mostly compare the saturation values of the acetylene conversion rate with and without the IL. One should be cautious when comparing the reaction rate on a molar basis.
Then we turned to the effect of the GHSV of the reactant gases. Shown in Figure 8 are the acetylene conversion rate and the VCM selectivity of the CuCl2-[PrMIm]Cl catalyst as a function of GHSV of C2H2. Our result indicates that the VCM selectivity remained unchanged at the level of about 96% as the GHSV increased. However, the C2H2 conversion rate has dropped significantly, from 50% to 40%. As one may expect, this more of less reflects that fact that, for a fixed amount of catalyst, increase in GHSV may cause part of C2H2 to be removed from the reaction chamber before it has a chance to react. Nonetheless, we note that the dependence of the conversion rate on GHSV was not as strong as we had expected.
It should be mentioned that the systematic variations of the catalytic activity with various parameters in Figure 3, Figure 7 and Figure 8 indicate that our experimental results are fully reproducible.

2.2. Mechanistic Study Using DFT

Now we present our result from DFT calculations. Unless mentioned explicitly, all calculations were done with DMol3.
First, we calculated the catalytic reaction process assuming [PrMIm]Cl as the catalyst without the participation of CuCl2. To determine where the most likely adsorption site is located, we calculated the LUMO of the cation [PrMIm]+, which is shown in Supplementary Figure S3. It can be seen that, near the two N atoms, there exist the largest unoccupied p orbitals, which help to adsorb HCl or C2H2 molecules. Therefore, the active adsorption sites are in the neighborhood of the N atoms on the imidazolium ring, as expected.
Then we proceeded to determine the structural relation between the anion Cl and the cation [PrMIm]+ inside the IL. According to [23], Cl- needs to be in the neighborhood of [PrMIm]+ and form strong chemical adsorption. Possible structures are shown in Figure 9. The structure in Figure 9b has a slightly lower energy than Figure 9a (by 2.27 kcal/mol), and is thus more stable. However, we emphasize that these are the (meta-)stable structures of a single pair of anion and cation. It does not necessarily mean that the anions and cations in the bulk IL will keep such structures. To determine which case is right for the bulk IL, we calculated the IR spectra for both [PrMIm]+ alone and the [PrMIm]-Cl structures shown Figure 9b. The calculated IR spectra are shown in Figure 10. The main difference between these two structures is that there appear to be three very strong vibrational modes for [PrMIm]-Cl at 982.7 cm−1, 2763.3 cm−1 and 3042.7 cm−1. The IR spectrum for the structure in Figure 9a is roughly the same, with the strongest 2763.3 cm−1 peak downshifted to 2759.73 cm−1. The animations of the vibrational modes show that these modes are associated with the strong binding between the anion Cl and the cation [PrMIm]+. The frequencies of other vibrational modes are also modified slightly by Cl. These computed IR spectra are to be compared with those measured experimentally, shown in Figure 1a. We find that these three very strong modes are clearly absent in the measured spectrum, while other modes with relatively lower intensity can be found in the FT-IR data. This suggests that the Cl anions in the bulk IL are in the unbound state, rather than tightly adsorbed to any specific cations. This picture is in fact consistent with its ionic liquid nature. Therefore, our further calculations were all based on the cation [PrMIm]+ without the presence of Cl.
According to our DFT calculations, the possible reaction path when [PrMIm]Cl alone is used as the catalyst is shown in Figure 11, following the direction of the arrows. The adsorption (Ads) of the reactants (R) C2H2 on the IL contributes a fairly small decrease to the system energy. This may have to do with the symmetric, non-dipolar charge distribution of C2H2. There is a rather high energy barrier for the intermediate transition state (TS). As the VCM forms, the system energy decreases by the amount of the binding energy, followed by the desorption (Des) of VCM. Finally, one obtains the product (P) VCM. The system energy change is small as well for the Des step, suggesting that the adsorption of VCM on the IL is also weak. The system energy E and the Gibbs free energy G at each stage are shown in Figure 12, as the red and blue curves, respectively. The data for E and G are listed in Table 3. It can be seen that the barriers for the energy and the Gibbs free energy are as high as 27.28 kcal/mol and 29.57 kcal/mol, respectively. Therefore, the reaction following this path is rather slow. It should be noted that it is the barrier in the Gibbs free energy that determines the reaction rate, even though the energy barrier is often used in the discussion.
It should also be pointed out that, since the reaction takes place in the liquid phase, the free energy should not contain the pV term, namely, we have G = E − ST. Both the zero point energy (ZPE) and finite temperature (140 °C) corrections are considered in the energy and free energy calculations.
To verify that we have found the right transition state, we did an intrinsic reaction coordinate (IRC) scan. The obtained energy curve is shown in Figure 13 as a function of IRC. The forward evolution leads to the Des state, while the reverse evolution leads back to the Ads state (plus HCl). This confirms that we have indeed found the correct transition state. Figure 13 also shows that the forward reaction has a much lower energy barrier than the backward reaction. This is because acetylene hydrochlorination is an exothermic reaction with a high reaction heat.
Besides the interaction between the ILs and HCl as well as C2H2, we studied how CuCl2 participates in the catalysis. It is known that in a Cl rich environment, CuCl2 often exists in the form of [CuCl4]2− and [Cu2Cl6]2−. An isolated [CuCl4]2− usually has a (meta)stable square planar or flattened tetrahedral structure. Careful studies [38,39] show that the structure of [CuCl4]2- exhibits a continuous distribution, with the most probable angle of about 136° between two Cu–Cl bonds. There is also a fairly large probability for this angle to be 90°. The former corresponds to a flattened tetrahedron, while the latter to a planar square. The reason that regular tetrahedron structure is unstable is that there exists very strong Jahn-Teller distortion in this system, which leads to strong structural deformation [39]. Thus far, there is still no consensus about the most probable angle. However, it is certain that due to the existence of all different possible angles, there are structural phase transitions at certain temperatures. Similarly, an isolated [Cu2Cl6]2− complex may also exist in planar structure or flattened edge-sharing double-tetrahedral structure. In both cases, the planar structure has a higher energy, and is thus a meta-stable state. For example, for [CuCl4]2−, the energy of the square planar structure is higher than that of flattened tetrahedral structure by 11.60 kcal/mol. Shown in Supplementary Figure S4 are the stable structures of (a) [CuCl4]2− and (b) [Cu2Cl6]2−, corresponding to flattened tetrahedral and flattened edge-sharing double tetrahedral structures, respectively.
Next, we calculated the catalytic reactions near the Cu(II) site of the CuCl2 catalyst. In the [PrMIm]Cl reaction media, since CuCl2 appears in the form of [CuCl4]2− and [Cu2Cl6]2−, they will be attracted to the neighborhood of the N atoms of the cationic [PrMIm]+ complex.
The reaction steps and the energy levels for each step are shown in Figure 14. The red (E) and blue (G) lines represent the energy and Gibbs free energy as catalyzed by [Cu2Cl6]2−, while the black (E) and green (G) lines are for the energy and Gibbs free energy levels as catalyzed by [CuCl4]2−. All E and G levels with respect to the reactants are also listed in Table 4.
Now there are two intermediate states and two transition states. For the case of [Cu2Cl6]2−, the structures of each step are shown in Figure 15. As can be seen from the Figure, [Cu2Cl6]2− is adsorbed to the neighborhood of the N atoms of [PrMIm]Cl. Then a C2H2 molecule is adsorbed to [Cu2Cl6]2− (Ads), with an adsorption energy of −10.50 kcal/mol. Next, C2H2 and [Cu2Cl6]2− react and form an intermediate state IM1, for which the C atoms of C2H2 replace two of the Cl atoms of [Cu2Cl6]2−, and the two Cl atoms form covalent bonds with the C atoms. The energy of IM1 is lower than the reactant by 17.62 kcal/mol. After the transition state TS1 (with energy 0.71 kcal/mol), one of the two Cu–C bonds and one of the two C–Cl valence bonds break apart, and the detached Cl atom forms a covalent bond with Cu, and thus the intermediate state IM2. The IM2 state has the lowest energy in the entire reaction path, −47.6206 kcal/mol. Then the system passes through the transition state TS2, and the other Cu–C bond breaks apart, while an HCl molecule comes to the neighborhood. The HCl molecule combines with the broken C2H2Cl, contributes an H atom, and forms C2H3Cl, while the leftover Cl atom joins the Cu2Cl5 complex, and recovers Cu2Cl6. In this Des stage, the system energy is −40.34 kcal/mol, relative to the reactants. The two transition states have fairly high energy barriers. Especially for TS1, the breaking of the Cu–C bond leaves two free atoms such that the system energy is substantially higher than that of IM1, by 18.33 kcal/mol. Note that the energy of the Des stage is slightly higher than the product. The energy barrier from IM1 to TS1 is substantially higher than that from IM2 to TS2, and therefore is the determining step for the reaction rate. If one considers Gibbs free energy, then the energy levels will differ slightly, and the reaction barrier would become 14.81 kcal/mol.
The molecular structures of each reaction step for the case of [CuCl4]2− are shown in Figure 16. The process is very similar to the case of [Cu2Cl6]2− above. First, the [CuCl4]2− complex is adsorbed to the N atoms of [PrMIm]Cl, followed by adsorption of a C2H2 molecule to [CuCl4]2− (Ads), with an adsorption energy of −6.58 kcal/mol. Then C2H2 reacts with [CuCl4]2− and form the intermediate state IM1, for which the C atoms of C2H2 replace two Cl atoms of [CuCl4]2- and form two covalent Cu–C bonds, and these two Cl atoms now form covalent C–Cl bonds with the C atoms of C2H2. IM1 has an energy of −10.30 kcal/mol. Next, via the transition state TS1, one of the Cu–C bonds and one of the C–Cl bonds break apart, and form a Cu–Cl bond, and thus an intermediate state IM2, which has the lowest energy of the entire reaction path, −44.50 kcal/mol. Now via the transition state TS2, the other Cu–C bond breaks apart, while an HCl molecule approaches. The H atom of HCl detaches and combines with the C2H2Cl complex, yielding C2H3Cl, while the Cl atom groups with CuCl3 and recovers CuCl4. In the Des step, various covalent bonds relax to their equilibrium lengths so that stable C2H3Cl appears, with a system energy of −39.24 kcal/mol. The two transition states have rather high energy barriers. In particular, after the Cu–C bond breaks, the TS1 state has a dramatically elevated energy level with respect to IM1. The barrier from IM1 to TS1 is 16.67 kcal/mol, higher than that from IM2 to TS2, and thus is the decisive step for the reaction rate. Taking into account of the entropy at 140 °C, the reaction barrier in terms of Gibbs free energy is reduced down to 11.14 kcal/mol, implying that high temperature makes the forward reaction easier.
For the two forms of Cu(II) complexes, [PrMIm]Cl not only provides Cl, but also participates in the adsorption of the reactants so that it acts synergistically with CuCl2 in the catalysis. As can be seen from Figure 15 and Figure 16, the shape of [PrMIm]+ deviates significantly from its free form due to attractions from the reactants and the Cu complexes. Nonetheless, we do not consider [PrMIm]Cl as an active catalytic ingredient in Figure 15 and Figure 16, as the catalytic reactions mainly take place near the Cu complexes.
Comparing the catalytic reaction processes under the two forms of Cu complexes, one can see that the [CuCl4]2− case has an overall lower energy. Furthermore, the rate determining Gibbs free energy barrier, 11.14 kcal/mol, is significantly lower than that for the [Cu2Cl6]2− case, 14.81 kcal/mol. Therefore, one can conclude that, when using [PrMIm]Cl as the reaction media, the largest contribution of CuCl2 to the catalytic reaction comes from the form of [CuCl4]2−. Nonetheless, the contribution in the form of [Cu2Cl6]2− is not completely negligible, as the energy barrier difference between them is not huge (about ΔG = 4.5 kBT). On the other hand, Figure 12 and Figure 14 represent catalytic reactions using [PrMIm]Cl and CuCl2 as the catalyst, respectively. From the height of the barriers, one concludes that the contribution from [PrMIm]Cl as a catalyst to the acetylene hydrochlorination reaction rate is negligible, since the reaction rate R has an exponential dependence on the barrier height ΔG, namely,
R   T   e Δ G / k B T .
For T = 413 K, we have kBT/molecule = 0.8207 kcal/mol. Based on this relationship, the contribution from [Cu2Cl6]2− is about 1.14% of that of [CuCl4]2−, while the contribution of [PrMIm]Cl as a catalyst, with the barrier ΔG = 27.28 kcal/mol, is 2.88 × 10−9 times that of [CuCl4]2−, which is completely negligible.
In summary, our DFT calculations show that the main catalysis takes place near the Cu(II) active site. In the environment of IL, Cu(II) often exists in the form of [CuCl4]2− and [Cu2Cl6]2− complexes, and are adsorbed to the cation of the IL. At the same time, the IL facilitates the reaction by attracting C2H2 and HCl to the neighborhood of the Cu complex. While both Cu complexes can catalyze the acetylene hydrochlorination reaction, [CuCl4]2− dominates the contribution. Without CuCl2, the IL alone exhibits very low catalytic activity, as is consistent with our experiment.

3. Discussion

Our experimental result for the activity of the CuCl2-IL catalysts should be compared to that in the literature. We notice that Cao et al. [19] obtained a C2H2 conversion rate of 68% at a relatively low GHSV for the catalyst CuCl2-[BMIm]Cl, whereas we obtained the best conversion rate of about 39% for CuCl2-[PrMIm]Cl but at a much higher GHSV = 370 h−1. For CuCl2-[BMIm]Cl, our best conversion rate at this high GHSV was about 32%. However, they also reported a fairly rapid drop of the conversion rate with increasing GHSV, while we saw a fairly slow change with GHSV. As indicated in their Figure 6, the conversion rate for different temperatures and CuCl2 concentrations dropped to about 37–40% when the GHSV was increased to 90 h−1. It seems that if their GHSV had increased further, their conversion rate would have dropped more rapidly to a level below our present result.
Cao’s group also studied the effect of different anions on the catalytic activity [23], and found that among ILs with the same cation, catalysts prepared with the one with the anion Cl had the best catalytic performance. This was also confirmed by our preliminary result. Nevertheless, as mentioned earlier, we avoided using ILs with non-Cl anions altogether since they would inevitably introduce unwanted by-products, such as C2H3Br, etc.
Our calculated reaction process and reaction mechanism should also be compared with existing literature. The theoretical calculations of Ren et al. showed that under room temperature, the hydrogen bond between C2H2 and the ILs is critical in the solubility of C2H2 in the ILs [40]. The cations of the ILs may interact strongly with C2H2 via π–π bonding. The length of the alkyl chains does not strongly affect this π–π bonding. The high solubility helps C2H2 to approach the catalyst more closely so as to facilitate the catalytic reaction. Our results of the synergetic effect of the ILs are consistent with the calculated high solubility of C2H2 in the ILs.
Regarding the reaction mechanism of [BMIm]Cl alone as the catalyst, [23] did quantitative computations. They thought that at room temperature, the following dynamical balances exist
HCl + Cl ⇌ [HCl2]
HCl + [HCl2] ⇌ [H2Cl3]
Cl + [H2Cl3] ⇌ 2[HCl2]
and they proposed a reaction mechanism. From their results, the formation and opening of hydrogen bond is the key to the catalytic reaction. First of all, the Cl anion forms a hydrogen bond with the C(2)-H group on the imidazolium ring so that a six element ring forms, with 1 Cl, 1 N, 2 C, and 2 H atoms. This ring is relatively small in that the C2H2 molecule cannot interact strongly with it. When a HCl molecule joins, it may form a [HCl2] radical with the previous Cl, so that the six-element ring is enlarged to be an eight-element ring. In result, the hydrogen bond between the previous Cl- and the C(2)-H group is weakened while the ring grows, so that a C2H2 molecule may be inserted to form an intermediate state. In particular, the C2H2 may form a hydrogen bond with the [HCl2] radical, while it interacts with the imidazolium ring via π–π bonding. After charge transfer and bond reorganization, a vinyl chloride molecule is formed.
Our calculated process with [PrMIm]Cl as the catalyst should be compared with their proposed mechanism. However, we find that in our case of [PrMIm]Cl, only the six-element ring was involved, and our transition state was confirmed to connect the Ads step with the Des step via an IRC scan, as shown in Figure 13. Nonetheless, we do acknowledge that there might be other reaction paths, connecting the reactants with the products. Indeed, this can be seen from the fact that, for a molecular structure as complicated as [PrMIm]Cl-CuCl2, there are multiple meta-stable adsorption structures with slightly different adsorption energies.
Regarding the reactions catalyzed by either [CuCl4]2− or [Cu2Cl6]2−, the intermediate IM1 state involves formation of two Cu–C bonds, one of which has to be broken later. This IM1 state has a rather low energy, which is thus expected to have a fairly long lifetime. We could not identify a process that bypasses the IM1 step and evolves from the Ads step to IM2 directly via the transition state TS1. Should such a process exist, the reaction energy barrier would become very low, and the reaction would become superfast.
Thus far, we are not aware of theoretical calculations about the CuCl2-IL catalysts for acetylene hydrochlorination that can be compared to our current study. Our calculations show that, without CuCl2, the reaction rate is essentially negligible. This is different from the findings of [23], but consistent with our own experimental result. The C2H2 conversion rate for the regular CuCl2/AC catalyst from our control experiment under our reaction conditions was 18.5%, much lower than our current result. These results serve as a manifestation of the synergetic effect of CuCl2 and [PrMIm]Cl in catalyzing the reaction of acetylene hydrochlorination. Microscopically, the Cl rich environment of the IL leads to the formation of anionic [CuCl4]2− and [Cu2Cl6]2− complexes, which are attracted to the cations of the IL. At the same time, the high solubility of C2H2 (as well as HCl) in the IL greatly increases the chances for the reactants as well as the Cu(II) complexes to get close to each other and complete the reaction cycles.
It should be mentioned that the IL calculated in [23] was [BMIm]Cl, rather than [PrMIm]Cl. However, due to the relatively lower catalytic activity of the [BMIm]Cl-CuCl2 combination, we did not do calculations for the case where [BMIm]Cl acts as the catalyst alone. The different choices of ILs might provide reconciliation between our findings and [23] regarding the catalytic activity of the IL as a catalyst for acetylene hydrochlorination.
For the future, we look to include more metal chlorides and other compounds in our calculations, and have more thorough comparisons between experiment and theory, in hopes to find better metal–IL combinations for which the metal–IL synergy works to yield an excellent catalyst for acetylene hydrochlorination.

4. Materials and Methods

Major imidazolium type of ILs (purity > 99%) include [BMIm]Cl, [BMIm]Br, [PrMIm]Cl, [PrMIm]Br, [EMIm]Cl, [HMIm]Cl, [BMIm]BF4, [EMIm]BF4, [BMIm]PF6, and [EMIm]PF6, etc., which differ mainly in the length of the alkyl chains and the anions. They were purchased from Lanzhou Greenchem ILS, LICP, CAS. (Lanzhou, China), and used as received. To avoid complications in the VCM selectivity, only ILs with Cl anions were used. The chemical formulae, molecular weights, and structures of the ILs with Cl and different alkyl chains can be found in Supplementary Table S1. The metal chlorides were purchased from Alladin Regents (Shanghai) Co. Ltd. High purity reactants C2H2 (>99.9%) and HCl (>99.9%) were purchased from Hangzhou Jingong Special Gases, Ltd. and Shanghai Chengkung Gas Industry, Ltd., and used without further purification.
Our gas–liquid phase reaction apparatus for acetylene hydrochlorination is schematically shown in Supplementary Figure S5. To remove residue impurity gases including water and air, the temperature of the reaction chamber was raised gradually to 120 °C and N2 was let into the reaction path for 20 min. Then we turned on the HCl flow, with a flow rate of 0.5 L/h, to activate the catalyst for 1 h. Next, the temperature of the reaction chamber was raised to the reaction condition, 140 °C, and we turned on the flow of C2H2, with a volume feed ratio of VHCl:VC2H2 = 1.2. The flow rate was controlled by the mass flow controller (MFC). Before entering the reaction chamber, C2H2 was passed through the K2Cr2O2 and H2SO4 bottles to remove impurity gas components. The typical GHSV for C2H2 was set to be 370 h−1. The temperature of reaction chamber was controlled by the oil bath, whose temperature was measured by the thermal couple (TC). After reaction, the product and residual reactant gases passed though the absorbing bottle, to remove HCl, before they enter the gas chromatography analyzer.
The ILs were pre-processed before the experiments began: we took 15 ml IL in a small glass container, and placed in an oven at 80 °C for 20 min, so that it melted into liquid completely. Then we added the metal chloride MClx that had been dried to remove water. After stirring actively in an ultrasonic bath for 10 min, we obtained a homogeneous liquid of IL with MClx dissolved. In Supplementary Figure S6, we show the IL [PrMIm]Cl in (a) solid and (b) liquid states and (c) as a solution of CuCl2, exhibiting sand-yellow, light yellow, and green colors, respectively. Then the catalyst was put in the reaction chamber that was placed in an oil bath, as shown in Figure S5.
In order to determine the appropriate reaction temperature, we surveyed the thermodynamic transition temperatures and the thermal decomposition temperature Td, as shown in Table 5. Here listed are the melting temperature Tm, freezing temperature Tf, glass transition temperature Tg, and Td. Due to the big size mismatch between the cation and the anion in an IL, and that the cations are normally very large, most ILs exhibit glassy behavior; there are no sharp transition temperatures, unlike that of a good crystalline solid, and the melting and freezing temperatures often differ. For this reason, we find different temperatures from different sources. Nevertheless, except for [BMMIm]Cl, the rest ILs have a Tm below 100 °C. Note that the freezing temperature may be substantially lower than the melting point, implying that the ILs may exhibit supercritical behavior; a melted IL will not freeze until a much lower temperature. For [PrMIm]Cl, one can melt it at 80 °C, then it will remain a liquid at room temperature. More importantly, these ILs are thermally stable above about 250 °C, well above our chosen reaction temperature, 140 °C. On the other hand, we notice that the actual onset temperature for thermal decomposition is substantially lower than Td, therefore, to maintain long term stability, we chose a reaction temperature that was substantially below Td as well. At our reaction temperature, the ILs remained in perfect stable liquid state.
The best reaction temperature differs for different ILs. We prepared a series of catalysts of different CuCl2–IL combinations, and carried out the evaluation experiment under different parameters, including temperature, CuCl2 concentration, GHSV of C2H2, and on-stream reaction time, etc. We measured the conversion rate of C2H2, selectivity of VCM, and stability of the catalyst, etc., in order to determine the best IL. In comparison, we also performed control studies with CuCl2 and ILs alone as separate catalysts, to ascertain the promotional effect of the ILs to CuCl2. For CuCl2 without an IL, we used AC as the support.

Methods of DFT Calculations

Our DFT calculations were performed mainly using the quantum chemistry computation software package DMol3 [44], including the optimization of the molecular structures and the computation of the transitional states. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional [45] and all-electron double numerical basis set with polarized function (DNP) have been employed [46]. The convergence tolerance of energy, maximum force and maximum displacement are 1.0 × 10−5 Ha, 2.0 × 10−3 Ha/Å and 5.0 × 10−3 Å (1Ha = 27.21 eV, equivalent to 627.5095 kcal/mol) for geometry optimization. The complete LST/QST search protocol, the SCF tolerance of 1.0 × 10−6 Ha and smearing of 0.005 Ha are set for transition states. The Grimme method for DFT-D correction is considered for all calculations. Each atom in the storage models is allowed to relax to the minimum in the enthalpy without any constraints.
In addition, we also use Gaussian 16 [47] to do independent calculations for some simple molecular optimization and reactions. We employed the widely used B3LYP functional [48,49,50,51] with the 6-31G(D) basis set [52,53], which includes the diffusive function corrections and has been known to give reliable results. To ensure better accuracy [54], we chose the unrestricted method for the B3LYP calculations. We use the ‘tight convergence criteria’ set by the software, namely, maximum force 1.5 × 10−5 Ha/Å, root-mean-square (RMS) force 1.0 × 10−5 Ha/Å, maximum displacement 6.0 × 10−5 Å, and RMS displacement 4.0 × 10−5 Å. In addition, we use quadratically convergent self-consistent field algorithm (QC SCF) or XQC SCF to speed up the convergence. The errors for our energy calculations are usually (0.01~1.0) × 10−5 Ha.

5. Conclusions

We have studied the catalytic performance of CuCl2 among other metal chlorides as a catalyst for gas–liquid phase reactions of acetylene hydrochlorination, using ILs as the reaction media. Among various imidazolium-based ILs with different alkyl chains, we find that when [PrMIm]Cl was the reaction media, the metal catalyst exhibited the highest catalytic activity. Under the conditions T = 140 °C, C2H2 GHSV = 370 h−1, and reactant gas feed volume ratio VHCl:VC2H2 = 1.2, the [PrMIm]Cl-CuCl2 combination exhibits the highest acetylene conversion rate, up to 39%, much higher than the sum of that of the CuCl2/AC catalyst and of the IL alone as the catalyst. Our DFT calculation of the reaction process for the [PrMIm]Cl-CuCl2 catalyst show that CuCl2 are adsorbed in the neighborhood of the N atoms of [PrMIm]Cl, and the main catalytic reactions occur at the Cu(II) active sites. Without the participation of Cu2+, reactions at the N sites of the ILs have a rather high energy barrier and thus this contribution is negligible to the total reaction rate. As for the reactions near the Cu(II) active sites, among the two most common ionic complexes, [CuCl4]2− and [Cu2Cl6]2−, the former has the lowest energy barrier, and hence the highest reaction rate. Since the difference in barrier height is not very big, the contribution of the latter is small but should not be simply neglected.
Our result revealed strong synergetic effects between Cu and the ILs in the catalytic reaction of acetylene hydrochlorination. In an IL which contains Cl, it is easy for CuCl2 to attract extra Cl to form stable [CuCl4]2− and [Cu2Cl6]2− complexes, which are easily adsorbed to the cation, especially the N atoms of the IL. At the same time, C2H2 has a high solubility in the imidazolium type of ILs. This helps to bring C2H2 and HCl close to each other and to the Cu(II) complexes. On the other hand, the Cl- anion of HCl and that of CuCl2 as well as the Cl of the ILs are identical quantum particles. This facilitates the formation of intermediate states and transition states, so as to lower the reaction barrier and expedite the reaction of HCl and C2H2 near the Cu(II) sites to form C2H3Cl.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/6/504/s1. Figure S1: TGA curves of the IL [PrMIm]Cl (black thick line) and [BMIm]Cl (red thin line), Figure S2: The highest C2H2 conversion rate XA and the VCM selectivity SVCM as a function of reaction temperature T, for the catalyst CuCl2-[PrMIm]Cl. Reaction conditions: CuCl2 concentration = 0.065 mol L−1, C2H2 GHSV = 370 h−1, feed volume ratio VHCl:VC2H2 = 1.2, Figure S3: (a) LUMO orbital and (b) chemical structure of the [PrMIm]+ cation, as calculated using Gaussian 16. Here the dark gray, light gray and blue balls represent the C, H and Cl atoms, respectively. And the red and green colors represent (different signs of) the LUMO orbitals, Figure S4: Main structure for CuCl2 to appear in the IL [Prmim]Cl. Panel (a) is the flattened tetrahedron for the anionic [CuCl4]2− complex, and panel (b) is the edge-sharing double tetrahedron structure for the anionic [Cu2Cl6]2− complex, Figure S5: Schematic diagram of the apparatus for gas-liquid phase reactions using ILs as the solvent and reaction media. After passing through the MFC, C2H2 and HCl enter the reaction chamber, which is placed in an oil bath. The residual reactant and product gases then pass through the absorbing bottle, Figure S6: Physical appearance of (a) solid and (b) liquid [PrMIm]Cl and (c) its solution of CuCl2 (placed in a glass, bird view), exhibiting colors of sand-yellow, light yellow, and green, respectively, Table S1: Structures and molecular weights of some common imidazolium-based ILs that were used in the experiments.

Author Contributions

Conceptualization, Y.Y. (Yi Yu), J.Z., and X.L.; Methodology, Y.Y. (Yi Yu) and B.W.; Validation, J.Z and X.L.; Formal analysis, Y.Y. (Yi Yu); Computation, Y.Y. (Yi Yu); Data curation, Y.Y. (Yi Yu), Y.Y. (Yuxue Yue), B.W., and H.H.; Writing—original draft preparation, Y.Y. (Yi Yu); Writing—review and editing, Y.Y. (Yi Yu), J.Z., and X.L.; Supervision, X.L.; Project administration, J.Z.; Funding acquisition, X.L.

Funding

This work was supported by the National Natural Science Foundation of China (NSFC; grant nos. 21606199 and 21476207) and the Science and Technology Department of Zhejiang Province, China (Projects for Applications of Technology, grant no. 2017C37004).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hutchings, G.J. Vapor phase hydroehlorination of acetylene: Correlation of catalytic activity of supported metal chloride catalysts. J. Catal. 1985, 96, 292–295. [Google Scholar] [CrossRef]
  2. Johnston, P.; Carthey, N.; Hutchings, G.J. Discovery, Development, and Commercialization of Gold Catalysts for Acetylene Hydrochlorination. J. Am. Chem. Soc. 2015, 137, 14548–14557. [Google Scholar] [CrossRef] [PubMed]
  3. Malta, G.; Freakley, S.J.; Kondrata, S.A.; Hutchings, G.J. Acetylene hydrochlorination using Au/carbon: a journey towards single site catalysis. Chem. Commun. 2017, 53, 11733–11746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Xu, H.; Luo, G.H. Green production of PVC from laboratory to industrialization: State-of-the-art review of heterogeneous non-mercury catalysts for acetylene hydrochlorination. J. Ind. Eng. Chem. 2018, 65, 13–25. [Google Scholar] [CrossRef]
  5. Nkosi, B.; Coville, N.J.; Hutchings, G.J. Vapour phase hydrochlorination of acetylene with group VIII and IB metal chloride catalysts. Appl. Catal. 1988, 43, 33–39. [Google Scholar] [CrossRef]
  6. Wieiaut, J.P.; Vandalfsen, J. Undefined. Rec. Truv. Chim. 1932, 51, 636. [Google Scholar]
  7. Wang, S.J.; Shen, B.X.; Song, Q.L. Kinetics of acetylene hydrochlorination over bimetallic Au–Cu/C catalyst. Catal. Lett. 2010, 134, 102–109. [Google Scholar] [CrossRef]
  8. Wang, L.; Shen, B.X.; Zhao, J.G.; Bi, X.T. Trimetallic Au-Cu-K/AC for acetylene hydrochlorination. Can. J. Chem. Eng. 2017, 95, 1069–1075. [Google Scholar] [CrossRef]
  9. Zhang, H.Y.; Dai, B.; Wang, X.G.; Xu, L.L.; Zhu, M.Y. Hydrochlorination of acetylene to vinyl chloride monomer over bimetallic Au–La/SAC catalysts. J. Ind. Eng. Chem. 2012, 18, 49–54. [Google Scholar] [CrossRef]
  10. Zhang, H.Y.; Li, W.; Jin, Y.H.; Sheng, W.; Hu, M.C.; Wang, X.Q.; Zhang, J.L. Ru-Co(III)-Cu(II)/SAC catalyst for acetylene hydrochlorination. Appl. Catal. B Environ. 2016, 189, 56–64. [Google Scholar] [CrossRef]
  11. Zhao, J.; Xu, J.T.; Xu, J.H.; Ni, J.; Zhang, T.T; Xu, X.L.; Li, X.N. Activated-carbon-supported Gold–Cesium(I) as highly effective catalysts for hydrochlorination of acetylene to vinyl chloride. ChemPlusChem 2015, 80, 196–201. [Google Scholar] [CrossRef]
  12. Zhao, J.; Zhang, T.T.; Di, X.X.; Xu, J.T.; Gu, S.C.; Zhang, Q.F.; Ni, J.; Li, X.N. Activated carbon supported ternary Gold–Cesium(I)–Indium(III) catalyst for the hydrochlorination of acetylene. Catal. Sci. Technol. 2015, 5, 4973–4984. [Google Scholar] [CrossRef]
  13. Zhai, Y.Y.; Zhao, J.; Di, X.X.; Di, S.X.; Wang, B.L.; Yue, Y.X.; Sheng, G.F; Lai, H.X.; Guo, L.L.; Wang, H.; et al. Carbon-supported perovskite-like CsCuCl3 nanoparticles: a highly active and cost-effective heterogeneous catalyst for the hydrochlorination of acetylene to vinyl chloride. Catal. Sci. Technol. 2018, 8, 2901–2908. [Google Scholar] [CrossRef]
  14. Zhao, J.; Gu, S.C.; Xu, X.L.; Zhang, T.T.; Di, X.X.; Pan, Z.Y.; Li, X.N. Promotional effect of copper(II) on an activated carbon supported low content bimetallic Gold–Cesium(I) catalyst in acetylene hydrochlorination. RSC Adv. 2015, 5, 101427–101436. [Google Scholar] [CrossRef]
  15. Wang, B.L.; Lai, H.X.; Yue, Y.X.; Sheng, G.F; Deng, Y.Q.; He, H.H.; Guo, L.L.; Zhao, J.; Li, X.N. Zeolite supported ionic liquid catalysts for the hydrochlorination of acetylene. Catalysts 2018, 8, 351. [Google Scholar] [CrossRef]
  16. Welton, T. Room temperature ionic liquids: Solvents for synthesis and catalysis. Chem. Rev. 1999, 99, 2071–2084. [Google Scholar] [CrossRef]
  17. Wasserscheid, P.; Welton, T. Ionic Liquids in Synthesis; Wiley-VCH Verlag: Stuttgart, Germany, 2002. [Google Scholar]
  18. Vekariya, R.L. A review of ionic liquids: Applications towards catalytic organic transformations. J. Mol. Liq. 2017, 227, 44–60. [Google Scholar] [CrossRef]
  19. Qin, G.; Song, Y.H.; Jin, R.; Shi, J.; Yu, Z.Y.; Cao, S.K. Gas–liquid acetylene hydrochlorination under nonmercuric catalysis using ionic liquids as reaction media. Green Chem. 2011, 13, 1495–1498. [Google Scholar] [CrossRef]
  20. Zhao, J.; Gu, S.C.; Xu, X.L.; Zhang, T.T.; Yu, Y.; Di, X.X.; Ni, J.; Pan, Z.Y.; Li, X.N. Supported ionic-liquid-phase-stabilized Au(III) catalyst for acetylene hydrochlorination. Catal. Sci. Technol. 2016, 6, 3263–3270. [Google Scholar] [CrossRef]
  21. Zhao, J.; Yu, Y.; Xu, X.L.; Di, S.X.; Wang, B.L.; Xu, H.; Ni, J.; Guo, L.L.; Pan, Z.Y.; Li, X.N. Stabilizing Au(III) in supported-ionic-liquid-phase (SILP) catalyst using CuCl2 via a redox mechanism. Appl. Catal. B 2017, 206, 175–183. [Google Scholar] [CrossRef]
  22. Li, Y.; Dong, Y.Z.; Li, W.; Han, Y.; Zhang, J.L. Improvement of imidazolium-based ionic liquids on the activity of ruthenium catalyst for acetylene hydrochlorination. Mol. Catal. 2017, 443, 220–227. [Google Scholar] [CrossRef]
  23. Zhou, X.F.; Xu, S.G.; Liu, Y.L.; Cao, S.K. Mechanistic study on metal-free acetylene hydrochlorination catalyzed by imidazolium-based ionic liquids. Mol. Catal. 2018, 461, 73–79. [Google Scholar] [CrossRef]
  24. Zhao, J.; Yue, Y.X.; Sheng, G.F.; Wang, B.L.; Lai, H.X.; Di, S.X.; Zhai, Y.Y.; Guo, L.L.; Li, X.N. Supported ionic liquid-palladium catalyst for the highly effective hydrochlorination of acetylene. Chem. Eng. J. 2019, 360, 38–46. [Google Scholar] [CrossRef]
  25. Xu, J.H.; Zhao, J.; Xu, J.T.; Zhang, T.T.; Li, X.N.; Di, X.X.; Ni, J.; Wang, J.G.; Cen, J. Influence of surface chemistry of activated carbon on the activity of gold/activated carbon catalyst in acetylene hydrochlorination. Ind. Eng. Chem. Res. 2014, 53, 14272–14281. [Google Scholar] [CrossRef]
  26. Zhao, J.; Zhang, T.T.; Di, X.X.; Xu, J.T.; Xu, J.H.; Feng, F.; Ni, J.; Li, X.N. Nitrogen-modified activated carbon supported bimetallic Gold–Cesium(I) as highly active and stable catalyst for the hydrochlorination of acetylene. RSC Adv. 2015, 5, 6925–6931. [Google Scholar] [CrossRef]
  27. Zhao, J.; Xu, J.T.; Xu, J.H.; Zhang, T.T.; Di, X.X.; Ni, J.; Li, X.N. Enhancement of Au/AC acetylene hydrochlorination catalyst activity and stability via nitrogen-modified activated carbon support. Chem. Eng. J. 2015, 262, 1152–1160. [Google Scholar] [CrossRef]
  28. He, H.H.; Zhao, J.; Wang, B.L.; Yue, Y.X.; Sheng, G.F.; Wang, Q.T.; Yu, L.; Hu, Z.-T.; Li, X.N. Highly active AuCu-based catalysts for acetylene hydrochlorination prepared using organic aqua regia. Materials 2019, 12, 1310. [Google Scholar] [CrossRef]
  29. Di, X.X.; Zhao, J.; Yu, Y.; Xu, X.L.; Gu, S.C.; He, H.H.; Zhang, T.T.; Li, X.N. One-pot synthesis of nitrogen and sulfur co-doped activated carbon supported AuCl3 as efficient catalysts for acetylene hydrochlorination. Chin. Chem. Lett. 2016, 27, 1567–1571. [Google Scholar] [CrossRef]
  30. Zhao, J.; Wang, B.L.; Yue, Y.X.; Sheng, G.F.; Lai, H.X.; Wang, S.S.; Yu, L.; Zhang, Q.F.; Feng, F.; Hu, Z.T.; et al. Nitrogen- and phosphorus-codoped carbon-based catalyst for acetylene hydrochlorination. J. Catal. 2019, 373, 240–249. [Google Scholar] [CrossRef]
  31. Zhao, J.; Wang, B.L.; Xu, X.L.; Yu, Y.; Di, S.X; Xu, H.; Zhai, Y.Y.; He, H.H.; Guo, L.L.; Pan, Z.Y.; et al. Alternative solvent to aqua regia to activate Au/AC catalysts for the hydrochlorination of acetylene. J. Catal. 2017, 350, 149–158. [Google Scholar] [CrossRef]
  32. Zhao, J.; Wang, B.L.; Yue, Y.X.; Di, S.X; Zhai, Y.Y.; He, H.H.; Sheng, G.F.; Lai, H.X.; Zhu, Y.H.; Guo, L.L.; et al. Towards a greener approach for the preparation of highly active Gold/Carbon catalyst for the hydrochlorination of ethyne. J. Catal. 2018, 365, 153–162. [Google Scholar] [CrossRef]
  33. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part B: Applications in Coordination, Organometallic, and Bioinorganic Chemistry, 6th ed.; Wiley: Hoboken, NJ, USA, 2009. [Google Scholar]
  34. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Hangbook of X-ray Photoelectron Spectroscopy; Chastain, J., King, R.C., Jr., Eds.; Physical Electronics Inc.: Eden Prairie, MN, USA, 1995. [Google Scholar]
  35. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 1997. [Google Scholar]
  36. Malta, G.; Kondrat, S.A.; Freakley, S.J.; Davies, C.J.; Dawson, S.; Liu, X.; Lu, L.; Dymkowski, K.; Fernandez-Alonso, F.; Mukhopadhyay, S.; et al. Deactivation of a single-site Gold-on-Carbon acetylene hydrochlorination catalyst: An X-ray absorption and inelastic neutron scattering study. ACS Catal. 2018, 8, 8493–8505. [Google Scholar] [CrossRef]
  37. Conte, M.; Davies, C.J.; Morgan, D.J.; Carley, A.F.; Johnston, P.; Hutchings, G.J. Characterization of Au3+ species in Au/C catalysts for the hydrochlorination reaction of acetylene. Catal. Lett. 2014, 144, 1–8. [Google Scholar] [CrossRef]
  38. Rakitin, Y.V.; Kalinnikova, V.T.; Khodasevichb, S.G.; Novotortsev, V.M. Structures of the complex ions [CuCl4]2− and [CuCl5]3−. Russ. J. Coord. Chem. 2007, 33, 891–895. [Google Scholar] [CrossRef]
  39. Kelley, A.; Nalla, S.; Bond, M.R. The square-planar to flattened-tetrahedral CuX42−(X = Cl, Br) structural phase transition in 1,2,6-trimethylpyridinium salts. Acta Cryst. 2015, B71, 48–60. [Google Scholar] [CrossRef]
  40. Zhao, X.; Xing, H.; Yang, Q.; Li, R.; Su, B.; Bao, Z.; Yang, Y.; Ren, Q. Differential solubility of ethylene and acetylene in room-temperature ionic liquids: A theoretical study. J. Phys. Chem. B 2012, 116, 3944–3953. [Google Scholar] [CrossRef]
  41. Zhang, S.J.; Lu, X.M.; Zhou, Q.; Li, X.H.; Zhang, X.P.; Li, S.C. Ionic Liquids: Physicochemical Properties; Elsevier: Amsterdam, The Netherlands, 2009. [Google Scholar]
  42. Haynes, W.M.; Lide, D.R.; Bruno, T.J. CRC Handbook of Chemistry and Physics, 97th ed.; CRC Press, Taylor & Francis Group: Boca Raton, FL, USA, 2017. [Google Scholar]
  43. Plechkova, N.V.; Seddon, K.R. Ionic liquids completely UnCOILed–Critical expert overviews; Wiley: Hoboken, NJ, USA, 2015. [Google Scholar]
  44. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  45. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  46. Delley, B. An all-electron numerical method for solving the local density functional for polyatomic molecules. J. Chem. Phys. 1990, 92, 508–517. [Google Scholar] [CrossRef]
  47. Frisch, M.J. Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  48. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  49. Lee, C.; Yang, W.T.; Parr, R.G. Development of the Colle–Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef]
  50. Stephens, P.J. Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields: A comparison of local, nonlocal, and hybrid density functional. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  51. Kim, K.; Jordan, K.D. Comparison of density functional and MP2 calculations on the water monomer and dimer. J. Phys. Chem. 1994, 98, 10089–10094. [Google Scholar] [CrossRef]
  52. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular-orbital studies of organic-molecules. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  53. Hariharan, P.C.; Pople, J.A. Accuracy of AH equilibrium geometries by single determinant molecular-orbital theory. Mol. Phys. 1974, 27, 209–214. [Google Scholar] [CrossRef]
  54. Foresman, J.B.; Frisch, A. Exploring Chemistry with Electronic Structure Methods, 2nd ed.; Gaussian, Inc.: Pittsburgh, PA, USA, 1996. [Google Scholar]
Figure 1. FT-IR spectra of (a) [PrMIm]Cl and (b) the catalyst CuCl2-[PrMIm]Cl.
Figure 1. FT-IR spectra of (a) [PrMIm]Cl and (b) the catalyst CuCl2-[PrMIm]Cl.
Catalysts 09 00504 g001
Figure 2. The structure of the [CuCl4]2− complex adsorbed on to the cation [PrMIm]+, as calculated using the computational chemistry software, Gaussian 16. Here the blue, the smallest light grey, dark grey, brick red in the upper right corner, and green balls represent N, H, C, Cu, and Cl atoms, respectively.
Figure 2. The structure of the [CuCl4]2− complex adsorbed on to the cation [PrMIm]+, as calculated using the computational chemistry software, Gaussian 16. Here the blue, the smallest light grey, dark grey, brick red in the upper right corner, and green balls represent N, H, C, Cu, and Cl atoms, respectively.
Catalysts 09 00504 g002
Figure 3. (a) C2H2 conversion rate XA and (b) VCM selectivity SVCM of the catalyst CuCl2-[PrMIm]Cl as a function of reaction time at different temperatures. Reaction conditions: Concentration of CuCl2 = 0.065 mol·L−1, C2H2 GHSV = 370 h−1, feed volume ratio VHCl:VC2H2 = 1.2.
Figure 3. (a) C2H2 conversion rate XA and (b) VCM selectivity SVCM of the catalyst CuCl2-[PrMIm]Cl as a function of reaction time at different temperatures. Reaction conditions: Concentration of CuCl2 = 0.065 mol·L−1, C2H2 GHSV = 370 h−1, feed volume ratio VHCl:VC2H2 = 1.2.
Catalysts 09 00504 g003
Figure 4. XPS spectrum of Cu 2p core electrons for the fresh as-prepared catalyst [PrMIm]Cl-CuCl2. The binding energy of the observed 2p3/2 peak is 932.24 eV.
Figure 4. XPS spectrum of Cu 2p core electrons for the fresh as-prepared catalyst [PrMIm]Cl-CuCl2. The binding energy of the observed 2p3/2 peak is 932.24 eV.
Catalysts 09 00504 g004
Figure 5. Acetylene conversion rate XA and VCM selectivity SVCM of the CuCl2 catalyst, using IL [M]Cl with different cations [M]+ as the reaction media. Reaction temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, and [CuCl2] = 0.065 mol L−1.
Figure 5. Acetylene conversion rate XA and VCM selectivity SVCM of the CuCl2 catalyst, using IL [M]Cl with different cations [M]+ as the reaction media. Reaction temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, and [CuCl2] = 0.065 mol L−1.
Catalysts 09 00504 g005
Figure 6. Acetylene conversion rate and VCM selectivity of different metal chlorides, using [PrMIm]Cl as the reaction medium. Reaction temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, and concentration of metal chloride 0.10 mol L−1.
Figure 6. Acetylene conversion rate and VCM selectivity of different metal chlorides, using [PrMIm]Cl as the reaction medium. Reaction temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, and concentration of metal chloride 0.10 mol L−1.
Catalysts 09 00504 g006
Figure 7. Acetylene conversion rate and VCM selectivity of CuCl2-[PrMIm]Cl, as a function of the concentration of CuCl2. Reaction conditions: temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2.
Figure 7. Acetylene conversion rate and VCM selectivity of CuCl2-[PrMIm]Cl, as a function of the concentration of CuCl2. Reaction conditions: temperature 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2.
Catalysts 09 00504 g007
Figure 8. Acetylene conversion rate and VCM selectivity of CuCl2-[PrMIm]Cl, as a function of C2H2 GHSV. Reaction conditions: temperature 140 °C, CuCl2 concentration 0.065 mol L−1, VHCl:VC2H2 = 1.2.
Figure 8. Acetylene conversion rate and VCM selectivity of CuCl2-[PrMIm]Cl, as a function of C2H2 GHSV. Reaction conditions: temperature 140 °C, CuCl2 concentration 0.065 mol L−1, VHCl:VC2H2 = 1.2.
Catalysts 09 00504 g008
Figure 9. Possible structural relationship between the anion and the cation of the IL [PrMIm]Cl. (a) A 6-element or (b) 7-element ring forms with the Cl anion, C, N, and H atoms. Also labeled in the left panel is the charge distribution. Here the blue, smallest light grey, dark gray, and green balls represents N, H, C, and Cl atoms, respectively. Unlike the imidazolium ring, the 6- and 7-element rings are not as stable, since the dashed lines here only indicate moderately strong Coulomb attractions.
Figure 9. Possible structural relationship between the anion and the cation of the IL [PrMIm]Cl. (a) A 6-element or (b) 7-element ring forms with the Cl anion, C, N, and H atoms. Also labeled in the left panel is the charge distribution. Here the blue, smallest light grey, dark gray, and green balls represents N, H, C, and Cl atoms, respectively. Unlike the imidazolium ring, the 6- and 7-element rings are not as stable, since the dashed lines here only indicate moderately strong Coulomb attractions.
Catalysts 09 00504 g009
Figure 10. Infrared spectrum of the [PrMIm]+cation and the [PrMIm]-Cl structure shown in Figure 9b, as calculated via DFT. Very strong vibration modes appear in the lower panel, at 982.7 cm−1, 2763.3 cm−1 and 3042.7 cm−1. The animation of the calculated vibrations shows that they are related to the Cl anion. Other vibration frequencies have also showed small shift due to the influence of Cl.
Figure 10. Infrared spectrum of the [PrMIm]+cation and the [PrMIm]-Cl structure shown in Figure 9b, as calculated via DFT. Very strong vibration modes appear in the lower panel, at 982.7 cm−1, 2763.3 cm−1 and 3042.7 cm−1. The animation of the calculated vibrations shows that they are related to the Cl anion. Other vibration frequencies have also showed small shift due to the influence of Cl.
Catalysts 09 00504 g010
Figure 11. Reaction path of acetylene hydrochlorination, using [PrMIm]Cl alone as the catalyst. Shown here is a process participated by the [PrMIm]+ cation. The blue, light grey, dark grey, and green balls represent N, H, C, and Cl atoms, respectively.
Figure 11. Reaction path of acetylene hydrochlorination, using [PrMIm]Cl alone as the catalyst. Shown here is a process participated by the [PrMIm]+ cation. The blue, light grey, dark grey, and green balls represent N, H, C, and Cl atoms, respectively.
Catalysts 09 00504 g011
Figure 12. Step-by-step reaction process, in which the reactants C2H2 and HCl are adsorbed and react near the active N sites of [PrMIm]Cl. Here [PrMIm]+ serves as the catalyst. The red and blue lines represent the change in energy E and in the Gibbs free energy G, respectively.
Figure 12. Step-by-step reaction process, in which the reactants C2H2 and HCl are adsorbed and react near the active N sites of [PrMIm]Cl. Here [PrMIm]+ serves as the catalyst. The red and blue lines represent the change in energy E and in the Gibbs free energy G, respectively.
Catalysts 09 00504 g012
Figure 13. Energy curve from the IRC scan of the TS state in Figure 12, calculated using Gaussian 16. Starting from the saddle point of the TS state, the forward and backward scan lead to the Des and Ads (plus HCl) states, respectively. The forward reaction has a lower energy barrier.
Figure 13. Energy curve from the IRC scan of the TS state in Figure 12, calculated using Gaussian 16. Starting from the saddle point of the TS state, the forward and backward scan lead to the Des and Ads (plus HCl) states, respectively. The forward reaction has a lower energy barrier.
Catalysts 09 00504 g013
Figure 14. Step-by-step reaction process, in which the reactants C2H2 and HCl are adsorbed and react near the Cu(II) active site. The red and blue lines represent the cases with the participation of [Cu2Cl6]2− while the black and green lines represent the cases catalyzed by [CuCl4]2−.
Figure 14. Step-by-step reaction process, in which the reactants C2H2 and HCl are adsorbed and react near the Cu(II) active site. The red and blue lines represent the cases with the participation of [Cu2Cl6]2− while the black and green lines represent the cases catalyzed by [CuCl4]2−.
Catalysts 09 00504 g014
Figure 15. Reaction path of acetylene hydrochlorination near the Cu(II) atom, as catalyzed jointly by the catalyst CuCl2 and the reaction media [PrMIm]Cl. Shown is the process in which the anion [Cu2Cl6]2− participates in the catalysis. Here Cu exists in the form of [Cu2Cl6]2−, and is adsorbed near the N atoms of [PrMIm]Cl. Here, brick red spheres denote Cu atoms, and the color coding for other atoms is the same as Figure 11.
Figure 15. Reaction path of acetylene hydrochlorination near the Cu(II) atom, as catalyzed jointly by the catalyst CuCl2 and the reaction media [PrMIm]Cl. Shown is the process in which the anion [Cu2Cl6]2− participates in the catalysis. Here Cu exists in the form of [Cu2Cl6]2−, and is adsorbed near the N atoms of [PrMIm]Cl. Here, brick red spheres denote Cu atoms, and the color coding for other atoms is the same as Figure 11.
Catalysts 09 00504 g015
Figure 16. Reaction path of acetylene hydrochlorination near the Cu(II) atom, as catalyzed jointly by the catalyst CuCl2 and the reaction media [PrMIm]Cl. Shown is the process in which the anion [CuCl4]2− participates in the catalysis. Here Cu exists in the form of [CuCl4]2−, and is adsorbed near the N atoms of [PrMIm]Cl. The color coding is the same as Figure 15.
Figure 16. Reaction path of acetylene hydrochlorination near the Cu(II) atom, as catalyzed jointly by the catalyst CuCl2 and the reaction media [PrMIm]Cl. Shown is the process in which the anion [CuCl4]2− participates in the catalysis. Here Cu exists in the form of [CuCl4]2−, and is adsorbed near the N atoms of [PrMIm]Cl. The color coding is the same as Figure 15.
Catalysts 09 00504 g016
Table 1. Catalytic performance of CuCl2, using IL [M]Cl with different cations [M]+ as the reaction media
Table 1. Catalytic performance of CuCl2, using IL [M]Cl with different cations [M]+ as the reaction media
MEMImPrMImBMImHMImOMImAMImHEMImBMMIm
XA (%)2239322415181612
SVCM (%)9396969490938885
Reaction conditions: 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, concentration of CuCl2 = 0.065 mol L−1.
Table 2. Catalytic performance of different metal chlorides, using [PrMIm]Cl as the reaction medium.
Table 2. Catalytic performance of different metal chlorides, using [PrMIm]Cl as the reaction medium.
MClxCuCl2FeCl3ZnCl2SnCl2NiCl2BiCl3
XA (%)39191610159
SVCM (%)96.59596949293
Reaction conditions: 140 °C, C2H2 GHSV = 370 h−1, VHCl:VC2H2 = 1.2, [CuCl2] = 0.10 mol L−1.
Table 3. Energy levels of the reaction steps shown in Figure 12 (kcal/mol).
Table 3. Energy levels of the reaction steps shown in Figure 12 (kcal/mol).
EnergyRAdsTSDesP
[PrMIm]+: E0−4.5442225.02324−43.40721−41.09202
[PrMIm]+: G0−2.3614624.91584−34.7227−33.65141
Table 4. Energy levels of the reaction steps shown in Figure 14 (kcal/mol).
Table 4. Energy levels of the reaction steps shown in Figure 14 (kcal/mol).
EnergyRAdsIM1TS1IM2TS2DesP
[Cu2Cl6]2−: E0−10.4962−17.6158−0.7144−47.6207−36.6548−40.3414−41.0920
[Cu2Cl6]2−: G0−9.2365−14.1023−0.7096−40.6574−30.2568−33.3256−33.6514
[CuCl4]2−: E0−6.5772−10.30164.3732−44.4982−35.6521−39.2439−41.0920
[CuCl4]2−: G0−5.4169−46.92164.2157−38.1267−29.3651−32.9651−33.6514
Table 5. Thermodynamic transition temperatures of common imidazolium-based ILs and those used in the experiment.
Table 5. Thermodynamic transition temperatures of common imidazolium-based ILs and those used in the experiment.
ILTm (°C)Tf (°C)Tg (°C)Td (°C)
[BMIm]Cl41 ~ 70-−76 ~ −69254
[PrMIm]Cl60, 62−140-281 ~ 282
[EMIm]Cl77 ~ 9073-281 ~ 285
[HMIm]Cl−85 , −75-−75 253
[OMIm]Cl12, 12.3-−87 ~ −77243
[AMIm]Cl17, 49–51--273
[HEMIm]Cl60.8 ± 2, 86−111--
[BMMIm]Cl79, 99,100---
Note: The melting tempertures were mainly from the ChemSpider website at http://www.chemspider.com, and the handbook of ionic liquids [41,42,43].

Share and Cite

MDPI and ACS Style

Yu, Y.; Yue, Y.; Wang, B.; He, H.; Hu, Z.-T.; Zhao, J.; Li, X. Synergy between Ionic Liquids and CuCl2 in Gas–Liquid Phase Reactions of Acetylene Hydrochlorination. Catalysts 2019, 9, 504. https://doi.org/10.3390/catal9060504

AMA Style

Yu Y, Yue Y, Wang B, He H, Hu Z-T, Zhao J, Li X. Synergy between Ionic Liquids and CuCl2 in Gas–Liquid Phase Reactions of Acetylene Hydrochlorination. Catalysts. 2019; 9(6):504. https://doi.org/10.3390/catal9060504

Chicago/Turabian Style

Yu, Yi, Yuxue Yue, Bolin Wang, Haihua He, Zhong-Ting Hu, Jia Zhao, and Xiaonian Li. 2019. "Synergy between Ionic Liquids and CuCl2 in Gas–Liquid Phase Reactions of Acetylene Hydrochlorination" Catalysts 9, no. 6: 504. https://doi.org/10.3390/catal9060504

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