Ting Zhangab,
Xuefeng Wanga,
Xueli Huanga,
Yinnian Liaoa and
Jinzhu Chen*b
aKey Laboratory of Xinjiang Coal Clean Conversion and Chemical Process, College of Chemistry and Chemical Engineering, Xinjiang University, Urumqi, Xinjiang 830046, PR China
bCAS Key Laboratory of Renewable Energy, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China. E-mail: chenjz@ms.giec.ac.cn; Fax: +86-20-3722-3380; Tel: +86-20-3722-3380
First published on 22nd December 2015
Chemical fixation of carbon dioxide (CO2) to cyclic carbonates was investigated by using bifunctional nucleophile–electrophile catalysts of a metallophthalocyanine–carbon nitride hybrid [MPc/g-C3N4 (M = Co, Cu)] in the absence of any co-catalysts and organic solvents. MPc/g-C3N4 was readily obtained by direct calcination of a mixture of dicyandiamide and metallophthalocyanine under a flowing-nitrogen atmosphere, and the MPc/g-C3N4 prepared at 480 °C (MPc/g-C3N4-480) showed the highest catalytic performance toward the cycloaddition reaction of CO2 to epichlorohydrin (ECH). For bifunctional MPc/g-C3N4, the MPc species function as Lewis acidic centers for ECH activation via electrophilic attack; while, the g-C3N4 moiety, possessing abundant and uncondensed species with the forms of primary amine (NH2) groups and secondary amine (C–NH–C) groups at the edges of graphitic sheets as edge defects, acts as an organic base for CO2 activation through nucleophilic attack. The developed MPc/g-C3N4 is stable and insoluble in any commonly used organic solvents and behaves as heterogeneous catalyst, leading to facile separation and recycling in a CO2 fixation reaction.
For the cycloaddition reaction, a number of catalysts, such as transition metal complexes, organometallic compounds, ionic liquids, alkali metal salts, quaternary ammonium salts and phosphonium salts have been developed so far.1j–s Among them, metallophthalocyanine (MPc) complexes are very attractive, not only because of their bio-inspired chemistry and structural analogy to metalloporphyrin in the active sites of various enzymes which are responsible for peroxide destruction, catalytic oxidation and reduction, and dioxygen transport in nature, but also due to their chemical and thermal stability as well as their facile preparation on a large scale.3 In addition, MPc complexes have been widely used in various material fields, such as liquid crystal, electrochromic and non-linear optical devices, semiconductor and information storage systems.4
In the literature research for MPc-promoted reactions, a series of MPc-based catalysts such as copper phthalocyanine (CuPc),5 MPc complexes (M = Cu2+, Co2+, Ni2+ and Al3+) encapsulated in zeolite-Y,6 and aluminum phthalocyanine complex covalently bonded to MCM-41 silica7 were previously investigated for cycloaddition of CO2 and epoxides to produce cyclic carbonates. Moreover, cobalt phthalocyanine (CoPc) covalently immobilized on cellulose was reported for aerobic oxidation of alkylarenes and alcohols.8 CoPc immobilized on graphene oxide was investigated for photo induced carbon dioxide reduction to methanol.9 In situ synthesized CoPc/TiO2 was explored for photo-induced CO2 reduction to formic acid, methanol, formaldehyde in an aqueous solution using visible light irradiation.10 The methods of encapsulation and covalent immobilization of MPc species realize heterogenization of catalyst, provide easy separation and recycling of the MPc-derived catalyst, promote sufficient interactions between substrate and catalyst, and, most importantly, maintain MPc complex as a catalytically active monomer form by preventing it from self-dimerization.
Recent research demonstrated possibility of incorporating metalloporphyrin or metallophthalocyanine with sp2-hybridized carbon materials such as graphene via π–π interaction for various catalytic application. For instance, catalyst of graphene–hemin hybrid material was investigated for selective oxidation of pyrogallol, L-arginine and toluene.11 Graphene-iron phthalocyanine (g-FePc) composite was reported as a promising Pt-free catalyst for oxygen reduction reactions.12 Such kind of hybrids generally possess homogeneously distributed metalloporphyrin or metallophthalocyanine species throughout the surface of the graphene. The sp2-hybridized graphene behaves as supports to protect the metalloporphyrin and metallophthalocyanine species from aggregation or destruction in the catalytic reactions, and functions as a π donor to the metalloporphyrin and metallophthalocyanine species through cation–π or π–π interaction between them to modify the catalytic performance of the hybrids.13 In addition, the heterostructures of these hybrids provide a convenient preparation, separation and recycling of the conjugates from the reaction solution.14
In addition to graphene, graphitic carbon nitride (g-C3N4), as an analogue to graphite, possesses two-dimensional stacked structures with π-conjugated planar layers. Moreover, g-C3N4 is a typical solid base material and a series of mesoporous g-C3N4-derived materials were reported for various CO2-related transformations.15 For instance, urea-derived graphitic carbon nitride,16 mesoporous carbon nitride grafted with n-bromobutane,17 and metal halides supported mesoporous carbon nitride18 were investigated for cycloaddition of CO2 to epoxides to afford cyclic carbonates.
In the case of urea-derived graphitic carbon nitride, edge defects, the incompletely-coordinated nitrogen atoms, in the lower crystallinity and polymerization degree of u-g-C3N4 are presumably served as the main active sites in the cycloaddition reaction.16 For mesoporous carbon nitride grafted with n-bromobutane, the catalytic active sites are suggested to be uncondensed amines and Br anions which are obtained from the reaction between n-bromobutane and the N-containing heterocycles of mp-C3N4.17 Finally, a proposed mechanism of metal halides supported mesoporous carbon nitride-promoted cycloaddition reaction of CO2 to propylene oxide generally involves a basic support of mp-C3N4 material to adsorb CO2; whereas the active sites of supported metal halides promote the activation of epoxide and catalyze its transform into carbonates.18
In addition to catalyst, in some cases, co-solvent and co-catalysts play key roles on the cycloaddition reaction. For instance, the presence of co-solvent can facilitate the epoxides diffusion, promote the catalyst dispersion and reduce the reaction system viscosity.1d While, co-catalysts such as 4-dimethylaminopyridine (DMAP), potassium iodide (KI), phenyltrimethylammonium tribromide (PTAT), tetrabutylammonium bromide (TBAB) and tetraphenylphosphonium bromide (TPPB) can function as nucleophilic reagents showing synergetic effect in promoting the cycloaddition reactions.1s,t
Inspired by the above works, in this research, we prepared metallophthalocyanine–carbon nitride hybrids [MPc/g-C3N4 (M = Co, Cu)] as effective and bifunctional catalysts for cycloaddition of CO2 to epichlorohydrin (ECH) to give cyclic 3-chloro-1,2-propylenecarbonate (CPC) under co-catalyst and solvent free conditions (Fig. 1 and 2). In the hybrid MPc/g-C3N4, the MPc species can function as mild Lewis acid site for epoxide activation by coordination of oxygen atom and subsequent polarization of the C–O bond of ECH in the cycloaddition of CO2 to epoxides. While, the g-C3N4 moiety can act as solid base and behave as co-catalyst such as DMAP owing to the presence of a large number of uncondensed terminal amine and amine groups at the edges of the graphitic sheets.19 Moreover, the investigated MPc/g-C3N4 is readily available, air- and moisture-tolerant, recyclable catalyst for the cycloaddition reaction.
FT-IR spectra of g-C3N4-480, CoPc, and CoPc/g-C3N4-480 with various CoPc contents were compared in Fig. 3a. The FT-IR spectra of g-C3N4-480 show several absorption bands in the region of 1240 to 1650 cm−1 corresponding to the typical stretching vibration of CN heterocycles.16,20 The characteristic peak at 803 cm−1 is assigned to the triazine units (C6N7), suggesting the formation of typical structure of g-C3N4.20b,21 The broad peaks at around 3000–3400 cm−1 were indicative of primary (terminal –NH2 groups) and secondary (NH) amines linked to the edges of graphitic CN sheets and physically adsorbed water (H–O–H) molecules.18,20b,22
The FT-IR spectra of CoPc/g-C3N4-480 with various CoPc contents and g-C3N4-480 closely resemble each other. As shown in Fig. 3a, with increasing content of CoPc in CoPc/g-C3N4-480, the characteristic absorption peaks of CoPc at 732 cm−1 assigned as the Pc ring vibration become stronger,10 while that of the –NH2 and –NH vibration between 3000 and 3400 cm−1 become weaker, confirming the incorporation of CoPc species in CoPc/g-C3N4-480 and decrease in polymerization degree of g-C3N4 in CoPc/g-C3N4-480.
Fig. 3b shows CoPc/g-C3N4 samples with various terminal calcination temperatures. The multiple absorption bands of the CN heterocycles in the region of 1240–1650 cm−1 become broader with the increasing temperature, while the broad bands located at 3000–3400 cm−1 become weaker, further confirming the increase in polymerization degree. FT-IR analysis of CuPc/g-C3N4 with various CuPc contents and terminal calcination temperatures reveal similar tendency with that of CoPc/g-C3N4 samples (Fig. S1, ESI†).
XRD patterns of g-C3N4-480, CoPc, and CoPc/g-C3N4-480 with various CoPc contents are shown in Fig. 4a. In the case of g-C3N4-480, the strongest peak at 27.41° is a characteristic interlayer stacking peak of aromatic systems, indexed for graphitic materials as the (002) peak (Fig. 4a). While, the small angle peak at 13.08°, corresponding to interplanar distance, is indexed as (100), which is associated with interlayer stacking.23 For the composite CoPc/g-C3N4-480, the intensity of characteristic diffraction peaks significantly decreased with increasing CoPc contents if compared with pristine g-C3N4-480, suggesting a slightly distorted structure of nitride pores after the increasing incorporation of CoPc species (Fig. 4a).24 Notably, the intensity of diffraction peaks corresponding to CoPc species at 7.04, 9.26, 23.84, 33.08 and 40.43 (Fig. 4a, JCPDS file no. 44-1994) were observed increasing with CoPc contents in CoPc/g-C3N4-480 samples. In the case of CoPc/g-C3N4 samples obtained with various terminal calcinations temperatures, the XRD patterns show the same peak positions (Fig. 4b). However, these patterns generally display a slight broadening width and a decreasing intensity of the overall peaks following higher calcination temperatures. XRD analysis of CuPc/g-C3N4 with various CuPc contents and terminal calcinations temperatures reveal similar tendency with that of CoPc/g-C3N4 sample (Fig. S2, ESI†).
The TEM images of both CoPc/g-C3N4 and CuPc/g-C3N4 showed the typical layered platelet-like morphology of g-C3N4, indicating the MPc (M = Co, Cu) species were found to be uniformly distributed through g-C3N4 support in MPc/g-C3N4 without any agglomeration (Fig. 5).
The thermal stabilities of CoPc/g-C3N4-480 (0.66) and CuPc/g-C3N4-480 (0.80) were investigated by thermogravimetric analysis to confirm the stability of the hybrid materials. The curves of weight loss versus temperature for CoPc/g-C3N4-480 and CuPc/g-C3N4-480 were plotted in Fig. S3 (ESI†). The significant weight loss of hybrid samples, which corresponds to sample degradation, starts at about 615 °C for CoPc/g-C3N4-480 and 670 °C for CuPc/g-C3N4-480, respectively.
The surface compositions of CoPc/g-C3N4 and CuPc/g-C3N4 were further investigated with XPS. Peaks corresponding to carbon, nitrogen, cobalt for CoPc/g-C3N4 and copper for CuPc/g-C3N4 were presented in the survey scan (Fig. 6a). Fig. 6b shows the N(1s) XPS spectrum of CoPc/g-C3N4 and CuPc/g-C3N4. The N(1s) line of CoPc/g-C3N4 was deconvoluted into three superimposed peaks at 398.7, 400.4 and 401.7 eV corresponding to the aromatic nitrogen bonded to two carbon (CN–C) in triazine or heptazine rings,16 the sp2 hybridized nitrogen connected to three atoms (C–N(–C)–C or C–N(–H)–C),25 and the sp3 hybridized terminal nitrogen (such as N–H2 and N–O) of the heptazine rings.26 In the case of the N(1s) XPS of CuPc/g-C3N4, two peaks at 398.4 and 400.4 eV was evidently observed as depicted in Fig. 6b.
![]() | ||
Fig. 6 (a) XPS scan survey and (b) N(1s) XPS spectra for CoPc/g-C3N4-480 (0.66) and CuPc/g-C3N4-480 (0.80); (c) Co(2p) XPS spectra of CoPc/g-C3N4; and (d) Cu(2p) XPS spectra of CuPc/g-C3N4. |
As shown in Fig. 6c, the Co(2p) XPS spectrum of CoPc/g-C3N4 shows two main peaks located at 796.0 and 780.8 eV corresponding to Co(2p1/2) and Co(2p3/2), respectively, indicating the presence of Co(II)-centered CoPc species in CoPc/g-C3N4.27 In the case of CuPc/g-C3N4, the presence of Cu(II) species was supported by the appearance of two main peaks at 954.6 and 934.7 eV assigned to Cu(2p1/2) and Cu(2p3/2), respectively, as shown in Fig. 6d, which is closer to the Cu(II) position of a Cu(II)Pc.28 The Co(2p) XPS analysis of CoPc/g-C3N4 and Cu(2p) XPS analysis of CuPc/g-C3N4 therefore confirmed successful formation of MPc/g-C3N4 composites.
Fig. 8a shows the influence of reaction temperature and time on the CoPc/g-C3N4-promoted coupling reaction of CO2 and ECH without using co-catalyst and solvent. CPC yield generally increased with reaction time at all reaction temperatures investigated. Reaction temperature dramatically promoted the cycloaddition reaction and CPC yield significantly increased from 56.2% at 110 °C to 100% at 130 °C after 28 h. Notably, an induction period around 0–12 h was observed when the cycloaddition reaction was performed at low reaction temperature of 110 °C. In addition, the induction period decreased with reaction temperature. ECH has chloromethyl group which might undergo some reactions with the amino group of g-C3N4, giving off the Cl− anion. Thus, Cl− anion is suggested to function as a nucleophile for the epoxide ring opening, which might be related to the observed induction period. To rule out this possibility, KCl was added to the cycloaddition reaction at 110 °C and the induction period disappeared with increased CPC yield. However, KCl shows very limited promotion effect on the CPC yield after the induction period. While, co-catalyst KI can significant enhance CPC yield under the investigated conditions, suggesting the presence of the synergistic effect between CoPc/g-C3N4 and I− anion. Moreover, several other epoxides having no chloromethyl group, such as propylene oxide and styrene oxide were examined and reduced yields of the corresponding cyclic carbonates were observed (Table 1). The effect of initial CO2 pressure on the cycloaddition reaction revealed that the increase in CO2 pressure from 1.0 to 3.0 MPa promoted the CPC yield from 70.3% to 97.6% (Fig. 8b).
The influence of catalyst loading amounts on the cycloaddition reaction revealed that the CPC yield steeply increased with the catalyst loading levels, with a significant rise to CPC yield of 97.6% for a 0.33 mol% [CoPc] species relative to ECH (Fig. 9). Further increase of the CoPc/g-C3N4 loading led to a decrease in CPC yield.
![]() | ||
Fig. 9 Effect of catalyst loading amount on the cycloaddition reaction of CO2 to ECH. Reaction conditions: ECH (3.70 g, 40.0 mmol), CoPc/g-C3N4 (CoPc 0.66 mmol g−1), CO2 (3.0 MPa), 130 °C, 24 h. |
Scheme 1 shows a proposed mechanism of bifunctional MPc/g-C3N4 (M = Co, Cu) catalyst-promoted cycloaddition reaction of CO2 to ECH under co-catalyst and solvent free conditions. ECH is activated by a Lewis acidic center of MPc/g-C3N4 (M = Co, Cu) through coordination of oxygen atom of ECH to MPc species and subsequent polarization of C–O bond of ECH to form metal alkoxide intermediate. Owing to insufficient polycondensation, g-C3N4 moiety of MPc/g-C3N4 possesses abundant uncondensed species in the forms of primary amine (NH2) groups and secondary amine (C–NH–C) groups at the edges of graphitic sheets as edge defects.16,18,19 CO2 molecules could presumably be activated by these edge defects of g-C3N4 via O–H bonding or nucleophilic attack. Under such condition, the g-C3N4 moiety of MPc/g-C3N4 can function as co-catalyst such as organic base DMAP to activate CO2.18,19 The activated CO2 complex attacks the metal alkoxide intermediate to yield the metal carbonate that leads to CPC formation and regeneration of catalyst MPc/g-C3N4.
The process of CO2 activation and fixation was investigated by using homogeneous system of CoPc and melamine in CH3CN and monitored by UV-vis and FT-IR. CoPc–melamine–CH3CN had a weak d–d band at around 600 nm (Fig. 10a). When the solution was treated with CO2, a new band was observed at 426 nm and can be assigned to metal-to-ligand charge transfer transitions [Co (d orbitals) → CO2 (π* orbitals)], indicating the formation of an activated CO2 species.5 The observed d–d band further shifted to 494 nm when CoPc–melamine–CO2–CH3CN was reacted with ECH, suggesting the structure change of the Co complex due to the formation of CPC (Fig. 10a). The FT-IR analysis revealed that the treatment of CO2 with the CoPc–melamine system showed two new IR peaks at 1606 and 1226 cm−1, suggesting the formation of an activated CO2 complex (Fig. 10b).5 When ECH reacted with CoPc–melamine–CO2 in CH3CN, additional peaks were observed at 1803, 1168 and 1074 cm−1 with the disappearance of CO2-complex (at 1606 and 1226 cm−1), indicating the formation of CPC.
Under the optimized reaction parameters, the CoPc/g-C3N4 catalyst was further investigated for the cycloaddition of CO2 to various epoxides under solvent and co-catalyst free conditions (Table 1). In the cases of mono-substituted terminal epoxides with aliphatic, aromatic, electron-withdrawing and electron-donating substituents, the CoPc/g-C3N4 catalyst provided the corresponding cyclic carbonate products in excellent to moderate yields (Table 1, Entries 1–3). For cyclohexene oxide, the corresponding cyclic carbonate was obtained in very low yield of 11.5% (Table 1, Entry 4), presumably due to the steric hindrance of the cyclohexene oxide.
A six-cycle experiment was performed to probe the reusability of CoPc/g-C3N4 in the cycloaddition reactions. Fig. 11 shows the obtained CPC yields reduced from 97.6 to 89.5%. The decreased catalytic performance of CoPc/g-C3N4 in the recycling can presumably be related to the CoPc species leaching from 0.66 mmol g−1 for the fresh catalyst to 0.60 mmol g−1 after the cycle experiment based on ICP-AES analysis.
![]() | ||
Fig. 11 Catalyst recycling. Reaction conditions: ECH (3.70 g, 40.0 mmol), CoPc/g-C3N4 (200 mg, CoPc 0.66 mmol g−1), CO2 (3.0 MPa), 130 °C, 24 h, in a 25 mL autoclave. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra21058e |
This journal is © The Royal Society of Chemistry 2016 |