DOI:
10.1039/C5RA05544J
(Paper)
RSC Adv., 2015,
5, 54266-54274
Mechanism of fixation of CO2 with an epoxide catalyzed by ZnBr2 and a choline chloride co-catalyst: a DFT study†
Received
28th March 2015
, Accepted 8th June 2015
First published on 8th June 2015
Abstract
To explore the reason for the high activity of the cycloaddition reaction of PO (propylene oxide) with CO2 catalyzed by ZnBr2/CH (choline chloride) co-catalyst, the mechanism has been constructed using a DFT (density functional theory) method. The combination of CH and ZnBr2 will generate three new stable complexes, i.e., [Ch]2[ZnBr2Cl2], Ch+ZnBr2Cl−, and Ch+ZnBrCl2−. The latter two are derived from the dissociation of [Ch]2[ZnBr2Cl2]. The detailed mechanism of a coupling reaction catalyzed by the more stable complex Ch+ZnBrCl2− is explored. It has been elucidated that the attack from the Zn complex and the Br− anion is the major factor in promoting the cleavage of the C–O bond of PO. Finally, the performance of [Ch]2[ZnBr2Cl2] is also investigated, providing less activity, indicating that it should dissociate to gain better catalytic effect.
1 Introduction
The fixation and utilization of carbon dioxide (CO2) have received increasing attention in light of the growing problems of the greenhouse effect and depletion of the fossil fuels.1,2 Emission of CO2 into the atmosphere has already resulted in global warming and climate change, while CO2 also serves as an abundant, cheap, nontoxic C1 raw material. In the past two decades, various pathways of CO2 utilization have been explored. Among them, the synthesis of five-membered cyclic carbonates via the coupling reaction of CO2 with epoxides is one of the most promising methodologies.3 However, the thermodynamic stability and kinetic inertness of CO2 are roadblocks that are in the way of utilizing it as a raw material.4 Exploring efficient catalysts is a promising method to overcome this obstacle. Diverse catalysts have been successfully developed, such as, alkali metal salts,5,6 metal oxides,7,8 transition-metal complexes,9–11 ionic liquids (ILs),12–14 and others.15–17 Most of the catalysts suffer from low activity, harsh reaction conditions, the need of organic co-solvents, or a combination of these.
In recent years, ILs have received much more attention because of their unique properties, including low melting temperature, undetectable vapor pressure, and special solubility for many organic or inorganic compounds, and especially due to their environmental friendliness and feasibility for design.18 Moreover, numerous experiments have proven that the involvement of Lewis acids will dramatically enhance the catalytic activity. Zinc halides are one of the most popular co-catalysts that have appeared in the IL-based catalyst system. This bicomponent catalyst system is robust, air stable, cheap, and free of organic solvent. Some composite catalyst systems have been reported, which have been widely applied in the processes of a regiospecific Fischer indole reaction,19 the protection of carbonyls,20 Diels–Alder reactions,21 and the fixation of CO2.22,23 Zhang et al. firstly investigated the activity of a ZnBr2/CH bicomponent catalyst for the synthesis of cyclic carbonates.24 The turnover frequency (TOF) of the ZnBr2/CH bicomponent catalyst of 494 is higher than that of ZnBr2 of 5 and that of CH of a trace amount, which is similar to other bicomponent catalysts reported in the previous literature.22,25 Although the TOFs are not high enough, the catalytic activity is enhanced by introducing the zinc halide.
To our best knowledge, detailed mechanisms for the coupling of CO2 with PO catalyzed by C(2,4,6)mimCl,26 DMimBr Me2PO4 + ZnBr2,27 TBABr,2,28 TEA(Br–Cl),2 LiBr,29 TBD·HBr,30 KI/glycerol,31 and TBAI/pyrogallol32 have been thoroughly studied. Moreover, the mechanism for the fixation of CO2 with PO mediated by ZnCl2/[BMIM]Cl has also been explored by Xia et al.33 The conclusions can not be totally transferred to the ZnBr2/CH bicomponent system, since the anions of ZnBr2 and CH are different and the Cl− anions are in excess. In a mixture of ZnBr2 and CH with a molar ratio of 1
:
5, lots of Cl− anions are available. What is the most stable configuration of Zn complex? What is the real main catalytic active species for the title reaction? Will the Zn complex exist in the catalytic system as a three-fold coordination complex or a four-fold coordination complex?
To elucidate the above questions, we investigated the mechanism for the coupling reaction of CO2 with an epoxide using the M06 method.34 After the computational methods section, the mechanism investigations for the coupling reaction catalyzed by ZnBr2 or CH are firstly discussed, respectively; in the next section, the main catalytic active species are analyzed; subsequently, the role of the composite catalyst is explored and its activity is compared with a single component catalyst, with a conclusion section at the end. It is expected that our work will provide valuable clues for further theoretical and experimental studies for the synthesis of more powerful catalysts.
2 Computational details
The Gaussian 09 program was employed to perform all the electronic calculations.35 All electronic structures of stationary points including the reactants, products, intermediates, and transition states were optimized using the M06 method.34 On the basis of previous experiences,26,30,41 the hydrogen bond is an essential factor for stabilizing the structures. So the noncovalent interactions should be carefully considered. On the other hand, a metal atom is included in the catalytic system. Considering the abovementioned two factors, the M06 functional becomes the final choice because the M06 functional is recommended by Truhlar et al. for application in organometallic and inorganometallic chemistry and for noncovalent interactions.34 Owing to the inclusion of the hydrogen atoms and the Zn atom, the basis set of 6-31+G(d,p) was employed for all atoms to obtain a high ratio of performance to cost, which is also a popular basis set for investigating the IL-based catalytic system and mechanism of other organic reactions.33,36,37 Four criteria are employed in the Gaussian program to judge the completion of optimization. They are the maximum remaining force on one atom of the system as well as the average (RMS, root mean square) force on all atoms, and the maximum displacement, that is, the maximum structural change of one coordinate as well as the average (RMS) change over all structural parameters in the last two iterations. The default RMS force criterion is 3 × 10−4. To ascertain the stability of the Zn complexes, the dissociation energies were refined at the M06/6-311+G(d,p) level based on the M06/6-31+G(d,p) optimized geometries. To identify the nature of the stationary points and to obtain the ZPE (zero-point energy) correction, the vibrational frequency was calculated at the same level. The number of imaginary frequency was employed to identify the minimum or the transition state, i.e., NIMAG = 1 for a saddle point or NIMAG = 0 for a minimum. The MEP (minimum-energy path) was constructed starting from the saddle point geometry and going downhill to the reactant and product using the intrinsic reaction coordinate (IRC) theory.38 Along this energy path, the reaction coordinate s is defined as the distance from the saddle point, with s > 0 referring to the product side. The abovementioned calculations were performed in the gas-phase state. Finally, the critical barrier heights involved in routes 1–5 (ring-opening and ring-closure steps) were corrected at the M06/6-311+G(d,p) level using a self-consistent reaction field (SCRF) method in the polarizable continuum model (PCM)39,40 on the basis of the gas-phase geometries.
3 Results and discussion
3.1 Uncatalyzed reaction
The coupling reaction of CO2 with PO without any catalyst has been substantially investigated in previous studies.26,31,41 It has been identified that the CO2 insertion and C–O bond rupture of PO are completed in a concerted mechanism with a high barrier height. For comparison with the results in this work, the preferential route is calculated at the M06/6-31+G(d,p) level. The corresponding results are plotted in Fig. S1 of the ESI.† At the M06/6-31+G(d,p) level, the barrier height is 55.33 kcal mol−1, which is difficult to overcome.
3.2 ZnBr2/ZnCl2 catalyst or CH catalyst
Before exploring the mechanism of binary catalysts, the coupling reaction of CO2 with PO catalyzed by a single catalyst is investigated to compare the activity and to discover the different features between the catalysts. As described in literature,33 the reaction becomes a stepwise mechanism with two steps. The cleavage of the C–O bond is activated by electrophilic attack from the Zn atom of ZnCl2 on the O atom of PO. The CO2 insertion and ring-closure are completed in the second step, which is the rate-determining step with a barrier height of 54.59 kcal mol−1 at the M06/6-31+G(d,p) level. When the catalyst is changed from ZnCl2 to ZnBr2, the barrier height of the rate-determining step is increased by 5.01 kcal mol−1. The potential energy surface profiles are given in Fig. 1, and the optimized geometries for all involved species are shown in Fig. S2.† Neither ZnCl2 nor ZnBr2 has any reactivity for the coupling reaction of CO2 with PO, which is consistent with the experimental measurements.23,24
 |
| Fig. 1 Potential energy profiles of the route catalyzed by ZnBr2 (–) or ZnCl2 (---). | |
In addition to ZnBr2, the effect of sole CH is also explored to confirm its co-catalytic role. The mechanism of CO2 with PO catalyzed by a single HETEAB (2-hydroxyl-ethyl-triethylammonium bromide) catalyst has been studied in our previous work.42 CH and HETEAB have a similar structure and the same functional group indicating the same mechanism. Different chain lengths and anions between the two catalysts will not modify the mechanism, although they will have a little effect on the barrier height. As depicted in Fig. 2 and Fig. S3,† the most favorable pathway follows a three-step mechanism, that is, ring-opening, CO2 insertion, and ring-closure, with barrier heights of 25.63, 0.76, and 22.71 kcal mol−1, respectively. The barrier height of the rate-determining step (25.63 kcal mol−1) is much lower than that of the ZnBr2-mediated coupling reaction. This is contradictory to the experimental result that the catalytic activity of CH is lower than that of ZnBr2 with trace product yield at 110 °C.24 However, when the temperature is increased to 125 °C the conversion catalyzed by CH is improved to 45.1%.43 The catalytic activity of CH is greatly affected by the temperature, and has great possibility to be better than ZnBr2 at a slightly higher temperature.
 |
| Fig. 2 Potential energy profile of the most favorable route catalyzed by CH (choline chloride). | |
3.3 Zn-complex
To elucidate the role of ZnBr2/CH in a cycloaddition reaction, the active Zn species that exists in the reaction system is firstly explored. Three possible chlorozincate clusters ZnCl2, ZnCl3−, and ZnCl42− have been considered in Xia’s work.33 The ZnCl3− is the most suitable one to be the actual catalytic component, as it is difficult to decompose into other species. The bromozincate clusters ZnBr2, ZnBr3−, and ZnBr42− are expected to have a similar stability sequence. However, the formations of the Zn complex are more complicated owing to the different anions between ZnBr2 and CH and the excess Cl− anions that exist in the title reaction. Sparked by the testified method used by Xia et al., the possible Zn complexes and their dissociation energies are shown in Table 1. To refine the reaction energy, the single-point energy calculation is performed at the M06/6-311+G(d,p) level on the basis of the M06/6-31+G(d,p) optimized structure. The dissociation of ZnBr2Cl22− into ZnBrCl2− and Br− is exothermic by 50.78 kcal mol−1, suggesting the stability of the ZnBrCl2− anion. Moreover, it is difficult to further dissociate ZnBrCl2− to ZnCl2 and Br− with large endothermic energies.
Table 1 The possible dissociation forms of the Zn complexes and the corresponding dissociation energies (ΔG, kcal mol−1) calculated at the M06/6-31+G(d,p) level and refined at the M06/6-311+G(d,p) level
Dissociation forms |
M06/6-31+G(d,p) |
M06/6-311+G(d,p) |
ZnBr2 + Cl− → ZnBr2Cl− |
−53.38 |
−52.56 |
ZnBr2Cl− + Cl− → ZnBr2Cl22− |
45.32 |
46.50 |
ZnBr2Cl22− → ZnBrCl2− + Br− |
−44.56 |
−50.78 |
ZnBrCl2− → ZnCl2 + Br− |
54.69 |
49.31 |
ChCl + ZnBr2 → Ch+ZnBr2Cl− |
−33.62 |
−28.29 |
Ch+ZnBr2Cl− + ChCl → [Ch]2[ZnBr2Cl2] |
−33.63 |
−21.74 |
[Ch]2[ZnBr2Cl2] → Ch+ZnBrCl2− + ChBr |
36.53 |
21.71 |
Ch+ZnBrCl2− → ZnCl2 + ChBr |
31.46 |
31.66 |
It is in no doubt that the stability of bromo-chlorozincate clusters will be affected by the Ch+ cation. Due to the different anions in CH and ZnBr2, there are more possible situations. Three structures Ch+ZnBr2Cl−, Ch+ZnBrCl2−, and [Ch]2[ZnBr2Cl2] were optimized at the M06/6-31+G(d,p) level and their energies were refined at the M06/6-311+G(d,p) level. The CH and ZnBr2 will form Ch+ZnBr2Cl− spontaneously with a formation energy of −28.29 kcal mol−1. Next, the Ch+ZnBr2Cl− will react with excess CH to produce the more stable [Ch]2[ZnBr2Cl2], which is an exothermic process with a formation energy of −21.74 kcal mol−1. However, the interaction between [Ch]2[ZnBr2Cl2] and PO will make the [Ch]2[ZnBr2Cl2] dissociate into other species. On one hand is that [Ch]2[ZnBr2Cl2] will dissociate into Ch+ZnBr2Cl− and ChCl, which needs 21.74 kcal mol−1 dissociation energy. On the other hand is that it will dissociate into Ch+ZnBrCl2− and ChBr with a lesser dissociation energy of 21.71 kcal mol−1, which is a more favorable route. The major parameters of the optimized geometries for ZnBr2, ZnBr2Cl22−, and [Ch]2[ZnBr2Cl2] are shown in Fig. 3. Due to the interactions between the Ch+ cation and ZnX42−, the Zn–Br bonds are elongated, while the Zn–Cl bonds are shortened. To get a high ratio of computational performance to cost, Ch+, Br−, and ZnBrCl2− will be taken into account as the essential catalysts. In addition, the activity of catalyst [Ch]2[ZnBr2Cl2] is also considered to compare with those of Ch+, Br−, and ZnBrCl2−.
 |
| Fig. 3 Optimized geometries for ZnBr2, ZnBr2Cl22−, and [Ch]2[ZnBr2Cl2]. | |
3.4 Ch+, Br−, and ZnBrCl2− catalysts
According to experience obtained from the literature and our previous work,33,41,44 a three-step mechanism is more favorable than a two-step mechanism. Neither activating the C atom of CO2 nor activating the O atom of CO2 is easy to complete because of the stability of CO2. Thus, only a three-step mechanism will be considered in the following study. In the three-step mechanism, the oxygen atom and carbon atom of the epoxide are firstly activated by the Lewis acid and halide anion, respectively; subsequently, the CO2 reacts with the activated ring-opening substrate; finally, the cyclic carbonate is formed by subsequent intramolecular cyclization and the catalyst is regenerated.33,44 The oxygen atom can be motivated by the Ch+, ZnBrCl2−, or both of them. Four possible pathways are located on the basis of the different species employed to activate PO. Both the substituted and non-substituted carbon atom will be attacked, so another corresponding four routes are easily found by extension. For clarity, only the most favorable route of each kind will be discussed in detail.
Route 1.
Since the Lewis acid can activate the oxygen atom of PO, it is easy to conject that more acidity is helpful to promote the ring-opening. Both ZnBrCl2− and Ch+ are employed to attack the O atom of PO, which is also the mechanism proposed by Zhang et al.24 The optimized geometries for the first transition state are plotted in Fig. 4 and the potential energy profiles are shown in Fig. 5. The other structures are shown in Fig. S4.† The ChBr, ZnBrCl2−, and PO firstly form a trimolecule complex 1-a with a relative energy of −64.84 kcal mol−1. Hydrogen bonds are formed to stabilize the PO and ZnBrCl2− with the distances of O1–H⋯Cl, C1–H⋯O2, and C2–H⋯O2 being 2.32, 2.20, and 2.58 Å, respectively. Taking the complex 1-a as a starting point, the nucleophilic attack of the Br− anion on the C atom of PO occurs associated with the synergetic attack from both the Zn atom of ZnBrCl2− and the H atom of the hydroxyl group on the O atom of PO via TS1-1. Overcoming the barrier height of 30.31 kcal mol−1, the intermediate 1-b is generated. In 1-b, not only ZnBrCl2− but also the –OH group of Ch+ have interactions with the O atom of PO. Next, the CO2 is introduced into the reaction to form intermediate 1-c by electrostatic attraction between the O2 atom and the C3 atom of CO2. Subsequently, the attack of CO2 on PO results in the formation of another intermediate 1-d by surmounting a barrier height of 6.09 kcal mol−1. Through the steric adjustment of the configuration, 1-d is converted into configuration 1-e which is a little more stable. In the following step, complex 1-e will convert into product-like intermediate 1-f via TS1-3. The imaginary vibration of TS1-3 corresponds to the shrinking of the C4–O3 bond, which results in the formation of PC. Finally, the catalysts are easily released and the catalytic cycle is completed. The ring-opening step is the rate-determining step in route 1 with a barrier height of 30.31 kcal mol−1, which is lower than the routes catalyzed by single catalyst ZnBr2 but higher than CH. However, the barrier height is still not easy enough to surmount at 110 °C. Are there any other favorable routes or competitive routes? Since the ring-opening step promoted by both Ch+ and ZnBrCl2− is not good enough, could one of them have a better performance? So the possibility of the O atom of PO being attacked by a single species, Ch+ or ZnBrCl2−, is explored. Three other routes are discovered and the potential energy profiles are also exhibited in Fig. 5. All of them are a three-step mechanism with ring-opening, CO2 insertion, and ring-closure. The main difference is the ring-opening step, thus, only the first step is discussed in the following routes. The optimized geometries of the first transition states are shown in Fig. 4 and the other structures are shown in Fig. S4.†
 |
| Fig. 4 Optimized geometries for the first transition states included in routes 1(TS1-1), 2(TS2-1), 3(TS3-1), and 4(TS4-1) of the cycloaddition reaction catalyzed by Ch+, Br−, and ZnBrCl2−. | |
 |
| Fig. 5 Potential energy profiles of routes 1–3 (---) and 4 (–) for the cycloaddition reaction catalyzed by Ch+, Br−, and ZnBrCl2−. TSm-n (n = 1, 2, or 3) denotes the first, second or third transition state involved in route m, and m-n (n = a–f) denotes the intermediate in route m. | |
Route 2.
We speculate that PO is only activated by the hydroxyl group of Ch+. As shown in Fig. 4, the ring-opening of PO in route 2 is motivated by the cooperation of electrophilic attack from the hydroxyl group and nucleophilic attack from the Br− anion via TS2-1. The ZnBrCl2− is coordinated to the O atom of the hydroxyl group, which will make the acidity of the hydroxyl group greater. Thus, the ring-opening step is promoted, and the energy barrier is reduced to 23.42 kcal mol−1. Moreover, there is a proton transfer process in route 2, which is a common appearance for hydroxyl-functionalized ILs.41 The rate-determining step is the ring-closure step for route 2 with a barrier height of 25.03 kcal mol−1. So the ring-opening and ring-closure are two competitive steps.
Route 3.
In route 2, Ch+ is taken as the vital catalytic species and the co-catalyst ZnBrCl2− is utilized to stabilize the Ch+. Thus, it is easy to reverse them. The ZnBrCl2− is coordinated to the O atom of PO to promote the ring-opening. While the Ch+ plays two roles: one is that the atoms H1, H2, and H3 of Ch+ are exerted to stabilize the ZnBrCl2−via hydrogen bonds; the other is that the interaction between Ch+ and ZnBrCl2− will make the ZnBrCl2− a better electrophilic species. Similar to route 2, the ring-opening and ring-closure steps for route 3 are also competitive steps with the closer barrier heights of 23.47 and 24.63 kcal mol−1.
Route 4.
Except that the H atoms from the CH3 group will form the hydrogen interaction with ZnBrCl2−, the –OH group may play the same role with a better result. The O atom of PO is attacked by the ZnBrCl2− and the H atom in the –OH group is coordinated to the Cl1 atom of ZnBrCl2−via a hydrogen bond to form route 4. The barrier heights of the ring-opening, CO2 insertion, and ring-closure steps are 17.73, 5.76, and 14.75 kcal mol−1, respectively. Compared with routes 1–3, route 4 is the most favorable one with the lowest barrier height. To obtain more reliable results, the key barrier heights were corrected at the M06/6-311+G(d,p) level using a PCM model. The corresponding results are listed in Table 2. The barrier heights mostly decreased, especially for the ring-opening step of route 1. However, the most favorable route is still route 4, which is the same as in the above conclusion.
Table 2 The barrier heights (ΔE, kcal mol−1) for the first and third elementary steps involved in routes 1–4
Route |
M06/6-31+G(d,p) |
M06/6-311+G(d,p) (PCM) |
Ring-opening |
Ring-closure |
Ring-opening |
Ring-closure |
1 |
30.31 |
10.38 |
19.49 |
15.67 |
2 |
23.42 |
25.03 |
21.97 |
22.25 |
3 |
23.47 |
24.63 |
17.36 |
20.60 |
4 |
17.73 |
14.75 |
15.55 |
18.44 |
We hope to find valuable clues that are helpful for proposing mechanisms in future and for designing new catalytic systems with better activity. A combination of ZnBrCl2− and Ch+ has more acidity than just one of them, which looks to be a better electrophilic reagent. However, the bulk structure hinders the close interaction between catalyst and PO. The distance of Zn–O(PO) is 2.05 Å in TS1-1 (route 1). The same distance in TS4-1 is 2.01 Å. Shorter distances between catalyst and substrate will give a stronger interaction that is helpful for improving the catalytic performance. As compared with Ch+, ZnBrCl2− is a better electrophilic reagent, so the most favorable route is where ZnBrCl2− is employed as the main catalyst and Ch+ is utilized to stabilize it. Designing ILs with more acidity will be helpful for improving the positive charge of the Zn center of ZnBrCl2−via the weak interactions between them. Consequently, it is profitable to improve the catalytic ability of ZnBrCl2−.
3.5 Ch+, Br−, ZnBrCl2−, and Ch+ catalysts
Considering that there are abundant catalysts in the real system, they are omitted in the above calculation. How will they affect the reaction barrier heights? To elucidate the puzzle, the coupling reaction of CO2 with PO catalyzed by Ch+, Br−, ZnBrCl2−, and Ch+ following the mechanisms of routes 3 and 4 is investigated at the M06/6-31+G(d,p) level. Then, the barrier heights are corrected at the M06/6-311+G(d,p) level using the PCM model. The optimized geometries for the intermediates and transition states are shown in Fig. S5.† As shown in Table 3, the barrier heights are varied, however, the barrier heights of route 3′ are higher than that of route 4′. The favorable route determined using the catalysts Ch+, Br−, and ZnBrCl2− could be expected to be reliable.
Table 3 The barrier heights (ΔE, kcal mol−1) for the first and third elementary steps involved in routes 3, 4, 3′, 4′, and 5 calculated at the M06/6-311+G(d,p) (PCM)//M06/6-31+G(d,p) level
Route |
The essential catalyst |
Ring-opening |
Ring-closure |
3 |
Ch+, Br−, and ZnBrCl2− |
17.36 |
20.60 |
3′ |
Ch+, Br−, ZnBrCl2−, and Ch+ |
16.06 |
20.86 |
4 |
Ch+, Br−, and ZnBrCl2− |
15.55 |
18.44 |
4′ |
Ch+, Br−, ZnBrCl2−, and Ch+ |
12.11 |
16.89 |
5 |
[Ch]2[ZnBr2Cl2] |
27.06 |
12.47 |
3.6 [Ch]2[ZnBr2Cl2] catalyst
The formation of [Ch]2[ZnBr2Cl2] is exothermic indicating a spontaneous process. There is a possibility that it does not dissociate. Up to now, no theoretical study has been reported for the mechanism of the coupling reaction of CO2 with PO catalyzed by [Ch]2[ZnBr2Cl2]. On the basis of experience obtained from Section 3.4 and other literature,44,45 the most favorable route is that the ZnBr2Cl22− is coordinated to the O atom of PO as the electrophilic species and two Ch+ cations are utilized to stabilize the ZnBr2Cl22−. The excess Cl− anions that exist in the reaction system will jointly assist the ring-opening by attacking the C atom of PO. According to this conjecture, a route with a similar mechanism to route 4 should be located. The difference is that the ring-opening of PO is activated by a tetra-coordinated Zn complex not a tri-coordinated one. It is unfortunate that one Zn–Cl bond is elongated in the process of forming a transition state. As a result, the ring-opening step is still activated by a tri-coordinated Zn complex rather than a tetra-coordinated Zn complex.
Alternatively, the ring-opening can be promoted by the synergetic effects of the tetra-coordinated Zn complex and the IL. As shown in Fig. 6 and Fig. S6,† the O is attacked by the –OH group of the IL, and simultaneously the C atom of PO is motivated by the Br atom of ZnBr2Cl22− leading to an energy-rich substrate. The two following steps are the CO2 insertion and the ring-closure. The ring-opening step is the rate-determining step with a barrier height of 25.34 kcal mol−1, which is higher than that of route 4. As shown in Table 3, the same situation occurs after the barrier heights are corrected at the M06/6-311+G(d,p) (PCM) level. So the catalytic components Ch+, Br−, and ZnBrCl2− will facilitate the reaction much easier.
 |
| Fig. 6 Potential energy profiles of route 4 (–) for the cycloaddition reaction catalyzed by Ch+, Br−, and ZnBrCl2− and route 5 (---) catalyzed by [Ch]2[ZnBr2Cl2]. | |
4 Conclusion
The mechanism of the coupling reaction of CO2 with PO catalyzed by ZnBr2/CH is studied using a DFT method. The theoretical study clearly indicates the reason why the cycloaddition of CO2 with PO will be more favorable in the presence of ZnBr2/CH compared with a single catalyst of ZnBr2 or CH. It has been identified that using solely ZnBr2 almost has no catalytic effect on the cycloaddition reaction with a much higher barrier height. While single component catalyst CH presented moderate catalytic activity for the rate-determining step. In the composite catalyst, the real catalytic components are Ch+, ZnBrCl2−, and Br− rather than [Ch]2[ZnBr2Cl2]. The ring-opening is promoted by the electrostatic interaction between the Zn atom of ZnBrCl2− and the O atom of PO, and simultaneous nucleophilic attack from the Br− anion of ChBr. In the whole catalytic process, Ch+ is employed to stabilize the Zn complex and Br− anion via hydrogen bonds, although Ch+ does not directly have an interaction with PO. Moreover, the interaction between the –OH group of Ch+ and the Zn complex will make the latter become a stronger electrophilic species, which will facilitate the rupture of the C–O bond of PO. These cooperative effects reduce the barrier height. The calculated results elucidate the mechanism of composite catalyst ZnBr2/CH, which is not consistent with the experimental assumption.
Acknowledgements
We thank the State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Wuhan Institute of Physics and Mathematics, Chinese Academy of Sciences for providing computational resources. This work was financially supported by the National Natural Science Foundation of China (21376063, 21476061), Program for He’nan Innovative Research Team in University (15IRTSTHN005), Natural Science Foundation of He’nan Province of China (134300510008, 144300510032, 142300410120), Science Foundation of Henan Province (14A150034), and Foundation for University Key Teachers from the He’nan Educational Committee.
References
- A. Decortes, A. M. Castilla and A. W. Kleij, Angew. Chem., Int. Ed., 2010, 49, 9822–9837 CrossRef CAS PubMed.
- J. Q. Wang, K. Dong, W. G. Cheng, J. Sun and S. J. Zhang, Catal. Sci. Technol., 2012, 2, 1480–1484 CAS.
- M. North, R. Pasquale and C. Young, Green Chem., 2010, 12, 1514–1539 RSC.
-
United Nations Development Programme, United Nations Department of Economic, Social Affairs, and World Energy Council, World Energy Assessment: Energy and the Challenge of Sustainability, United Nations Development Programme, New York, 2000 Search PubMed.
- J. W. Huang and M. Shi, J. Org. Chem., 2003, 68, 6705–6709 CrossRef CAS PubMed.
- J. L. Song, Z. F. Zhang, B. X. Han, S. Q. Hu, W. J. Li and Y. Xie, Green Chem., 2008, 10, 1337–1341 RSC.
- T. Yano, H. Matsui, T. Koike, H. Ishiguro, H. Fujihara, M. Yoshihara and T. Maeshima, Chem. Commun., 1997, 1129–1130 RSC.
- K. Yamaguchi, K. Ebitani, T. Yoshida, H. Yoshida and K. Kaneda, J. Am. Chem. Soc., 1999, 121, 4526–4527 CrossRef CAS.
- X. B. Lu, B. Liang, Y. J. Zhang, Y. Z. Tian, Y. M. Wang, C. X. Bai, H. Wang and R. Zhang, J. Am. Chem. Soc., 2004, 126, 3732–3733 CrossRef CAS PubMed.
- F. Jutz, J. D. Grunwaldt and A. Baiker, J. Mol. Catal. A: Chem., 2008, 279, 94–103 CrossRef CAS.
- D. S. Bai, H. W. Jing, Q. Liu, Q. Q. Zhu and X. F. Zhao, Catal. Commun., 2009, 11, 155–157 CrossRef CAS.
- Y. Y. Song, H. W. Jing, B. Li and D. S. Bai, Chem.–Eur. J., 2011, 17, 8731–8738 CrossRef CAS PubMed.
- Q. He, J. W. O’Brien, K. A. Kitselman, L. E. Tompkins, G. C. T. Curtis and F. M. Kerton, Catal. Sci. Technol., 2014, 4, 1513–1528 CAS.
- B. H. Xu, J. Q. Wang, J. Sun, Y. Huang, J. P. Zhang, X. P. Zhang and S. J. Zhang, Green Chem., 2015, 17, 108–122 RSC.
- M. North and R. Pasquale, Angew. Chem., Int. Ed., 2009, 48, 2946–2948 CrossRef CAS PubMed.
- A. Berkessel and M. Brandenburg, Org. Lett., 2006, 8, 4401–4404 CrossRef CAS PubMed.
- H. S. Kim, J. J. Kim, S. D. Lee, M. S. Lah, D. Moon and H. G. Jang, Chem.–Eur. J., 2003, 9, 678–686 CrossRef CAS PubMed.
- J. P. Hallett and T. Welton, Chem. Rev., 2011, 111, 3508–3576 CrossRef CAS PubMed.
- R. C. Morales, V. Tambyrajah, P. R. Jenkins, D. L. Davies and A. P. Abbott, Chem. Commun., 2004, 158–159 RSC.
- Z. Y. Duan, Y. L. Gu and Y. Q. Deng, Catal. Commun., 2006, 7, 651–656 CrossRef CAS.
- A. P. Abbott, G. Capper, D. L. Davies, R. H. Rasheed and V. Tambyrajah, Green Chem., 2002, 4, 24–26 RSC.
- H. S. Kim, J. J. Kim, H. Kim and H. G. Jang, J. Catal., 2003, 220, 44–46 CrossRef CAS.
- J. Sun, L. Wang, S. J. Zhang, Z. X. Li, X. P. Zhang, W. B. Dai and R. Mori, J. Mol. Catal. A: Chem., 2006, 256, 295–300 CrossRef CAS.
- W. G. Cheng, Z. Z. Fu, J. Q. Wang, J. Sun and S. J. Zhang, Synth. Commun., 2012, 42, 2564–2573 CrossRef CAS.
- F. W. Li, L. F. Xiao, C. G. Xia and B. Hu, Tetrahedron Lett., 2004, 45, 8307–8310 CrossRef CAS.
- H. Sun and D. J. Zhang, J. Phys. Chem. A, 2007, 111, 8036–8043 CrossRef CAS PubMed.
- J. K. Lee, Y. J. Kim, Y. S. Choi, H. Lee, J. S. Lee, J. Hong, E. K. Jeong, H. S. Kim and M. Cheong, Appl. Catal., B, 2012, 111–112, 621–627 CrossRef CAS.
- J. Q. Wang, J. Sun, W. G. Cheng, K. Dong, X. P. Zhang and S. J. Zhang, Phys. Chem. Chem. Phys., 2012, 14, 11021–11026 RSC.
- Y. Ren, C. H. Guo, J. F. Jia and H. S. Wu, J. Phys. Chem. A, 2011, 115, 2258–2267 CrossRef CAS PubMed.
- S. Foltran, R. Mereau and T. Tassaing, Catal. Sci. Technol., 2014, 4, 1585–1597 CAS.
- J. Ma, J. L. Liu, Z. F. Zhang and B. X. Han, Green Chem., 2012, 14, 2410–2420 RSC.
- C. J. Whiteoak, A. Nova, F. Maseras and A. W. Kleij, ChemSusChem, 2012, 5, 2032–2038 CrossRef CAS PubMed.
- F. Wang, C. Z. Xu, Z. Li, C. G. Xia and J. Chen, J. Mol. Catal. A: Chem., 2014, 385, 133–140 CrossRef CAS.
- Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, 215–241 CrossRef CAS.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision A.02, Gaussian Inc., Wallingford, CT, 2009 Search PubMed.
- T. Clark, J. Chandrasekhar, G. W. Spitznagel and P. V. R. Schleyer, J. Comput. Chem., 1983, 4, 294–301 CrossRef CAS.
- I. Skoric, F. Pavosevic, M. Vazdar, Z. Marinic, M. Sindler-Kulyk, M. Eckert-Maksic and D. Margetic, Org. Biomol. Chem., 2011, 9, 6771–6778 CAS.
- K. Fukui, J. Phys. Chem., 1970, 74, 4161–4163 CrossRef CAS.
- S. Miertus, E. Scrocco and J. Tomasi, Chem. Phys., 1981, 55, 117–129 CrossRef CAS.
- S. Miertus and J. Tomasi, Chem. Phys., 1982, 65, 239–241 CrossRef CAS.
- L. Wang, X. F. Jin, P. Li, J. L. Zhang, H. Y. He and S. J. Zhang, Ind. Eng. Chem. Res., 2014, 53, 8426–8435 CrossRef CAS.
- L. Wang, P. Li, X. F. Jin, J. L. Zhang, H. Y. He and S. J. Zhang, J. CO2 Util., 2015, 10, 113–119 CrossRef CAS.
- J. Sun, S. J. Zhang, W. G. Cheng and J. Y. Ren, Tetrahedron Lett., 2008, 49, 3588–3591 CrossRef CAS.
- F. Castro-Gómez, G. Salassa, A. W. Kleij and C. Bo, Chem.–Eur. J., 2013, 19, 6289–6298 CrossRef PubMed.
- R. C. Luo, X. T. Zhou, W. Y. Zhang, Z. X. Liang, J. Jiang and H. B. Ji, Green Chem., 2014, 16, 4179–4189 RSC.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra05544j |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.