Next Article in Journal
Formulation of Bismuth (Bi2O3) and Cerium Oxides (CeO2) Nanosheets for Boosted Visible Light Degradation of Methyl Orange and Methylene Blue Dyes in Water
Next Article in Special Issue
Core-Shell Hierarchical Fe/Cu Bimetallic Fenton Catalyst with Improved Adsorption and Catalytic Performance for Congo Red Degradation
Previous Article in Journal
Identification and Characterization of the Haloperoxidase VPO-RR from Rhodoplanes roseus by Genome Mining and Structure-Based Catalytic Site Mapping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boron and Phosphorus Co-Doped Graphitic Carbon Nitride Cooperate with Bu4NBr as Binary Heterogeneous Catalysts for the Cycloaddition of CO2 to Epoxides

Discipline Construction Office, Civil Aviation Flight University of China, Guanghan 618307, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1196; https://doi.org/10.3390/catal12101196
Submission received: 14 September 2022 / Revised: 6 October 2022 / Accepted: 7 October 2022 / Published: 8 October 2022

Abstract

:
The development of a cost-effective heterogeneous catalytic system for the cycloaddition reaction of CO2 and epoxides is of great importance. In this manuscript, three kinds of boron and phosphorus co-doping graphitic carbon nitride (BP-CN) were prepared and characterized. Among them, BP-CN-1 displayed the optimal catalytic performance in the presence of Bu4NBr (tetrabutylammonium bromide) for the CO2 cycloaddition with propylene oxide, and 95% propylene carbonate yield was obtained under a 120 °C, 2 MPa, 6 h condition. Moreover, the BP-CN-1/Bu4NBr catalytic system is compatible with various epoxides and also exhibits excellent recycling performance under metal- and solvent-free conditions. Hence, BP-CN-1 exhibited an attractive application for the efficient fixation of CO2 due to the simple, eco-friendly synthesis route and effective catalytic activity.

Graphical Abstract

1. Introduction

In recent decades, CO2 has been considered the primary culprit of the greenhouse effect, which causes global warming, rising sea levels and a series of ecological and environmental problems. CO2 is also deemed an abundant, safe, and renewable C1 building block. Hence, the increasing focus of CO2 research is its environmental friendliness and overall sustainability. In terms of current development, CO2 is a candidate for transformation into high-value-added chemicals by CO2 capture and utilization strategies [1]. Commercial chemicals mainly include urea, methanol, formic acid and organic carbonate, with CO2 as a raw material [2,3]. One approach to CO2 fixation is to synthesize cycle carbonate, which has received considerable attention due to the perspectives of “green chemistry” and “atom economy” [1]. This technology not only eliminates atmospheric CO2 concentrations to a certain extent but also provides a green route to producing cyclic carbonates, which are applied as electrolytes in lithium-ion batteries, pharmaceutical intermediates, and aprotic solvents [4,5]. However, industrialization remains a challenge due to the kinetic inertness and thermodynamic stability of CO2 [6]. To date, a plethora of catalysts that facilitate the process of CO2 cycloaddition have been developed and researched. Homogeneous catalysts and heterogeneous catalysts such as metal complex [7,8,9], porphyrin [10], metal-organic frameworks (MOFs) [11,12], covalent organic frameworks (COFs) [1], silicon-based material catalysts [13] and carbon nitride (CN) [14] were exploited for the CO2 coupling reaction.
CN, a two-dimensional (2D) layered material, is widely studied due to the metal-free, nontoxic, high chemical, thermal stability and adjustable edge functional groups in the structure, which can be synthesized by the calcination of inexpensive precursors, for instance, urea, melamine, and dicyandiamide [15]. CN possesses excellent structural stability, but CN is mostly applied in the photocatalysis field. Moreover, CN provides rich basic sites, which can be used in specific functional groups to catalyze the reaction to broaden its application field [16,17]. Nitrogen-rich structures in CN are CO2-phillic structures, guaranteeing high catalytic activity in the cycloaddition of CO2 to epoxides [18]. Many CN catalysts were reported, such as Zn2+-doped g-CN/SBA-15 [19], γ-Al2O3-supported ZnBr2 and carbon nitride heterogeneous catalysts (Zn–CN/γ-Al2O3) [18], ZnX2/CN [20]. Nevertheless, the above catalysts cannot avoid the leaching of metal elements, which is detrimental to the environment. To overcome this hitch, metal-free CN catalysts were explored for CO2 cycloaddition reactions.
Boron-doped carbon nitrides with abundantly exposed edge defects were developed using 1-butyl-3-methylimidazolium tetrafluoroborate (BmimBF4) to enhance the catalytic performance at 130 °C within 6 h [21]. In addition, boron-doped carbon nitride (B-CN), supported on mesoporous silica SBA-15, was used for CO2 cycloadditions under 130 °C, 3 MPa, 24 h conditions. The catalytic activity improvements benefited from the enhancement of surface acid and basic sites in CN [22]. Moreover, a series of phosphorus (P)-doped graphitic carbon nitrides (P-CN) showed excellent performance in CO2 cycloaddition with the presence of Bu4NBr under mild conditions (100 °C, 2 MPa, 4 h), which proved that the introduction of P could change the electronic structure of CN [23].
Hence, three kinds of boron and phosphorus co-doping graphitic carbon nitride (BP-CN-n, n = 1, 2, 3) materials were prepared and thoroughly characterized for the cycloaddition of CO2 under metal- and solvent-free conditions. Furthermore, the influence of B doping on the cycloaddition reaction was investigated. The effects of reaction time, reaction temperature and CO2 pressure on the yield of cyclic carbonates were discussed. Additionally, the reusability and catalytic reactivity to other substituted epoxides over BP-CN-1 were also investigated.

2. Experimental Section

2.1. General Information

Boric acid (AR) was provided by Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Phosphoric acid (85.0%) was purchased from Tianjin Guangfu Technology Development Co., Ltd. (Tianjin, China). CO2 (99.99%) was obtained from Deyang Shengyuan Gas Co., Ltd. (Deyang, China). Ethyl acetate was supplied from Shanghai Titan Scientific Co., Ltd. (Shanghai, China). Tetrabutylammonium bromide (Bu4NBr), melamine (MA) and all epoxides (99%) were purchased from Aldrich Chemical Co., Ltd. (Shanghai, China). and directly used without further purification.

2.2. Synthesis of BP-CN Catalyst

Three kinds of B and P co-doping graphitic carbon nitride (BP-CN) catalysts were synthesized with some improvements based on previous reports [24]. First, MA (5 g, 40 mmol) and phosphorous acid (10 g, 102 mmol) with different quantities of boric acid were dissolved in 100 mL deionized water with vigorous stirring for 30 min. Subsequently, the solution was transferred for hydrothermal processes into a Teflon-lined autoclave and heated to 180 °C for 10 h. Next, the mixture was washed several times with distilled water to remove as much of the remaining H+ as possible meanwhile, the hydrogen phosphite, MA salt, and unreacted boric acid were washed and then filtered and dried at 60 °C overnight. Finally, the white solid was obtained, transferred to a nitrogen oven, and heated at 500 °C for 2 h with a heating rate of 5 °C min−1. The resulting samples were designated as BP-CN-1, BP-CN-2 and BP-CN-3 (1.0 g, 16 mmol; 2.0 g, 32 mmol and 3.0 g, 48 mmol boric acid added, respectively). Additionally, pure CN was prepared with MA in a nitrogen oven and labeled as CN. For comparison with BP-CN, P-CN was prepared without boric acid and B-CN was synthesized by replacing phosphorous acid with HCl (4 g, 36.5%, 40 mmol).

2.3. Characterizations

Fourier-transformed infrared (FT-IR) spectra were recorded on PerkinElmer Spectrum 100 FT-IR spectrometer with KBr pellet, using the range of 400–4000 cm−1. X-ray diffraction (XRD) patterns of the samples were obtained on a Bruker D8 Advance X-ray diffraction with Cu Kα radiation (λ = 1.5406 Å, 40 kV, 30 mA) irradiation. X-ray photoelectron spectroscopy (XPS) measurements were performed using a Thermo Fisher Scientic Escalab 250Xi. N2 adsorption–desorption isotherms were measured at liquid nitrogen temperature on Micromeritics ASAP2020 after the catalyst was degassed under vacuum at 180 °C for 10 h. Inductively coupled plasma–atomic emission spectroscopy (ICP-AES) analysis was carried out on ICPOES Optima 8300 from PerkinElmer. High-resolution transmission electron microscopy (HRTEM) images were obtained using a Jem F30 (FETEM, Tecnai G2 F30) with an operating voltage of 200 kV. Scanning electron microscope (SEM) images were recorded on a Hitachi S-8010 scanning electron microscope. Gas chromatography (GC) was analyzed on Agilent GC-7890A instrument with a capillary column (Agilent 19091J-413).

2.4. Catalytic Activity Test

The cycloaddition reaction was conducted in a 25 mL stainless-steel autoclave with an inner Teflon lining. Typically, 0.60 mmol Bu4NBr, 0.100 g BP-CN catalyst and 34.5 mmol propylene oxide were successively added to the autoclave with strong stirring. Then, the reactor was heated to the desired temperature and 2 MPa CO2 was injected into an oil bath for a designated period of time. After the reaction, the autoclave was allowed to cool in an ice-water bath and excessive CO2 was slowly vented. Ultimately, the products were collected by dilution into 20 mL ethyl acetate and analyzed by GC. Meanwhile, the catalyst was recovered by centrifugation and washed with ethyl acetate (5 × 4 mL) to remove the residues. The solid catalyst was collected and dried in the 60 °C oven for the next run. Moreover, Bu4NBr was partially dissolved in ethyl acetate and the influence of catalyst was mainly explored, so Bu4NBr was washed in each cycle and new Bu4NBr was added in the next cycle test.
GC test: In this experiment, the external standard method is used to determine the product yield and selectivity of the reaction. PC is considered the standard sample to test the relationship between its concentration and peak area because of the low boiling point of PO (Figure 1). For other epoxides, epoxides are selected as test samples due to their high boiling points to obtain the relationship between their concentration and peak area. Therefore, the gas phase data provided the PC yield and epoxides conversion. In addition, the detail of catalyst recycling was provided in the “Catalytic activity test”. Moreover, the selectivity was calculated on the ratio of the peak area of PC and the total peak areas of all products. GC–MS was adopted to confirm the structure of products of CO2 cycloaddition reaction (Figure S1).

3. Results and Discussion

3.1. Catalyst Characterization

Fourier transform infrared spectrometer (FT-IR) spectra and X-ray diffraction (XRD) patterns were employed to analyze the chemical structures of the synthesized catalysts and the results were presented in Figure 2. As shown in Figure 2A, the broad bands changed from 3000 to 3300 cm−1, indicative of primary and secondary amines ascribed to the edges of graphitic CN sheets. The band peaks at 1200–1600 cm−1 were attributed to C=N and C–N stretching vibrations of the tri-s-triazine ring. Moreover, the sharp peaks with strong intensity at 810 cm−1 were assigned to the presence of triazine units [25,26]. In Figure 2B, the two distinct diffraction peaks at 13.1° and 27.5° corresponded to the (100) plane the in-plane structural packing motif and the (002) plane characteristic of interplanar stacking structures, respectively [27,28,29]. With the B content increasing, there was no obvious change that suggested that the original structure of CN was well-preserved compared with pure CN.
X-ray photoelectron spectroscopy (XPS) was adopted to investigate the surface photo-electron state of the elements in BP-CN-1 and the results were exhibited in Figure 3. In Figure 3A, the C 1s spectra were deconvoluted into two peaks at 284.6 eV and 288.0 eV, which were ascribed to C–C and sp2 N–C=N bonds in tri-s-triazine structural units [25,30]. In Figure 3B, N 1s spectra could fit into three peaks: the peak at 398.5 eV was assigned to C–N=C in triazine or heptazine rings, the peak at 399.6 eV resulted from sp2 hybridized N, and 401.1 eV was the binding energy of sp3 terminal N [31,32]. In Figure 3C, the B 1s spectra could fit into two peaks: the 191.6 eV binding energy could be ascribed to the B atoms in the BCN network and the 192.6 eV binding energy could be ascribed to the B atoms surrounded by one N atom and two O atoms [22,33]. Figure 3D illustrated the P 2p spectra of BP-CN-1. The 133.6 eV binding energy could be ascribed to the P–N bond, indicating that P likely replaces C in triazine rings to form P–N bonds [24]. Additionally, the 134.4 eV binding energy belonged to the P–OH of phosphoric acid [23].
N2 adsorption–desorption was adopted to analyze the specific surface area (SBET) of the synthesized catalysts and the N2 adsorption–desorption curves and pore size distributions of catalysts were exhibited in Figure 4. The SBET value, BJH average pore diameter and ICP-AES analysis results of those catalysts are summarized in Table 1. The SBET of doped samples was larger than those of pure CN, indicating that the successful doping of B and P elements and larger SBET could provide more surface catalytic active sites [24]. Meanwhile, the SBET decrease observed in the doped samples might be caused by the increase in B and P dopings [34].
High-resolution transmission electron microscopy (HRTEM) and a scanning electron microscope (SEM) were adopted to reveal the morphology features of BP-CN-1 and the results are exhibited in Figure 5. TEM images (Figure 5A) clearly showed that BP-CN-1 was composed of large-size irregular sheets with some clear wrinkles, which was consistent with CN (Figure 5C). The SEM image (Figure 5B) revealed that BP-CN-1 possessed a rod structure; however, the SEM image of CN (Figure 5D) indicated irregular sheets in CN. This difference was caused by the addition of phosphorous acid [24].

3.2. Catalyst SCREENING

The catalytic activities of the synthesized catalysts were evaluated by the cycloaddition reaction of propylene oxide (PO) with CO2 to form propylene carbonate (PC). All experimental results are listed in Table 2. No PC was obtained without a catalyst (Entry 1). Only a 2% PC yield was obtained with CN (Entry 2), and a 25% PC yield could be achieved with Bu4NBr (Entry 3), which was generally applied as an effective co-catalyst in the literature [35,36]. When CN cooperated with Bu4NBr, a 26% PC yield was obtained (Entry 4). The PC yield could be elevated when B- or P-doped CN acted as a catalyst with Bu4NBr (Entries 5, 6). When B-CN/Bu4NBr acted as a catalyst, a 50% PC yield was obtained (Entry 5), which might be caused by the enhancement of active sites of the CN surface after B dopings [21]. When P-CN/Bu4NBr acted as a catalyst, a 44% PC yield was obtained (Entry 5), which might be ascribed to the electronic structure change in CN after P dopings [23]. PC yield was obviously enhanced when BP-CN-n (n = 1, 2, 3) cooperated with Bu4NBr (Entries 7–9), especially BP-CN-1, where a 64% PC yield was obtained. However, the larger the doping content, the lower the catalytic activity of BP-CN-n. This phenomenon indicated that even B or P elements could provide a more catalytic active site but could not guarantee the reaction of the catalytic active site with PO. Herein, SBET played a key role in CO2 cycloaddition reactions with the BP-CN-n/Bu4NBr catalyst [23]. Moreover, when the cocatalysts Bu4NBr/Bu4NI were used in the cycloaddition reaction at 120 °C, 55% and 59% PC yield were obtained, respectively (Entries 10–11). Meanwhile, PC yield was 99% with BP-CN-1/Bu4NI at 120 °C, which was slightly higher than BP-CN-1/Bu4NBr. However, Bu4NBr was more suitable than Bu4NI due to the economic point. Hence, BP-CN-1/Bu4NBr was employed as a catalyst to discuss the effects of reaction conditions.

3.3. Effects of Reaction Conditions

The effects of the reaction conditions were discussed and the results are exhibited in Figure 6. PC yield and reaction temperature showed a positive correlation (Figure 6A), and PC yield increased from 44% to 99% with the temperature increase from 90 to 130 °C due to the promotion of effective collisions among the catalyst and substrate [37]. When the reaction temperature was 120 °C, PC yield reached 95%. PC yield improved slightly when the temperature was further increased to 130 °C. Hence, 120 °C was regarded as a suitable reaction temperature. As depicted in Figure 6B, the PC yield increased when the CO2 pressure increased from 0.5 MPa to 2.0 MPa; however, the PC yield showed a declining trend with higher CO2 pressure. This might be ascribed to the fact that the high CO2 pressure would dilute PO, resulting in a decrease in PC yield [38,39]. The effect of reaction time was also discussed (Figure 6C). PC yield could reach 95% within 6 h. Therefore, the reaction conditions were set as 120 °C, 2 MPa, and 6 h to investigate the recyclability and universality of BP-CN-1.

3.4. Catalyst Recyclability

Recyclability was a crucial evaluation standard. As exhibited in Figure 7A, BP-CN-1 possessed excellent recyclability and there was no PC yield loss during five recycling processes. In Figure 7B, there was no obvious difference between BP-CN-1 and reused BP-CN-1 in terms of FT-IR spectra. Additionally, the supernatant was separated and characterized by ICP-AES, and no B or P was detected. All this evidence shows the structural stability of BP-CN-1.

3.5. Catalyst Universality

The universality of BP-CN-1 was discussed and the results are shown in Table 3. The good yield and selectivity indicate that BP-CN-1 is adaptable to epoxides. 1,2-epoxybutane exhibited lower levels of conversion because of steric hindrance, preventing the nucleophilic attack of Br on the C atoms of epoxides (Entries 1 and 2) [40,41]. A similar result was obtained for the butyl glycidyl ether (Entry 4). The conversion of epichlorohydrin was the highest (Entry 3), as this benefited from the electron-withdrawing effect of Cl. The electron-withdrawing effect promoted the cleavage of C−O [42]. The conversion levels of styrene oxide and cyclohexene were lower than PO, even with a longer reaction time (Entries 5 and 6). For styrene oxide, the benzene ring group prevented the bulky catalyst from approaching. Additionally, the effects of the electron donor from an aromatic group hindered the polarization of C−O in styrene oxide [43,44]. For cyclohexene, the low levels of conversion might be caused by the special two-ring structure, which led to the highest hindrance of the reactant [45,46].

3.6. Comparisons of CN-Based Catalyst

The comparisons between the catalytic performance of B, P-CN-1 with various CN-based materials for the cycloaddition of CO2 to epoxides are listed in Table 4. Overall, a high temperature (>100 °C) was an essential condition to obtain an excellent yield. N, N-dimethylformamide (DMF) was added to the catalytic systems to promote the cyclic carbonate yield (Entries 1–3). Concerning single B-doped CN (Entries 4–5), not only did the temperature need to reach 130 °C, but it also took a long time to reach an outstanding yield. For the dual-doped K, B-CN-4 catalyst, a lower temperature (110 °C) was needed due to the metal K, which accelerated the epoxides’ ring-opening (Entry 6). The HCN-CN catalyst obtained more –COOH and –OH groups, leading to a lower temperature (Entry 7). The P-CN-2 catalyst sacrificed a lower temperature and needed less time to achieve an almost 100% conversion due to the change in its electronic structure (Entry 8). As a result, the B, P-CN-1 catalyst showed competitive activity in the CO2 conversion field with epoxides; hence, the B, P-CN-1 catalyst could potentially be utilized in CO2 cycloaddition reactions. To compare the activity of different modified CN catalysts under various conditions, the TOF and TON values were supplemented, which were calculated on the mass of produced cyclic carbonate per gram of catalyst per hour according to the literature [20] because the other modified CN catalysts could not provide the amounts of active sites. As a result, BP-CN-1 showed competitive activity in the CO2 conversion field with epoxides.

4. Conclusions

In conclusion, three kinds of B and P co-doped carbon nitrides (BP-CN-n, n = 1, 2, 3) were synthesized, characterized and applied in CO2 cycloaddition reactions with the co-catalyst Bu4NBr. BP-CN-1/Bu4NBr exhibited the best catalytic activity because BP-CN-1 possessed the largest specific surface area and reasonable B and P doping contents. BP-CN-1/Bu4NBr could promote the cycloaddition reaction of CO2 with propylene oxide under 120 °C, 2 MPa, 6 h, without the employment of metal or solvents. BP-CN-1/Bu4NBr also possessed remarkable recyclability and universality. Hence, BP-CN-1/Bu4NBr could be an environmentally friendly catalyst for CO2 cycloaddition reactions in the future. Moreover, in subsequent work, the nucleophilic group could be introduced into the catalyst to avoid the application of a cocatalyst and the CO2 cycloaddition reaction could be finished under mild conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12101196/s1, Figure S1: The GC-MS of (A) 1,2-epoxybutane (BO), (B) epichlorohydrin (ECH), (C) butyl glycidyl ether (BGE), (D) styrene oxide (SO) and (E) cyclohexene (CHO) before and after reaction.

Author Contributions

Data curation, C.Y.; formal analysis, T.Y.; investigation, C.Y.; resources, X.Z.; writing—review and editing, C.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Civil Aviation Development Fund Education Talents Project (0252101), the Key Project of Civil Aviation Flight University of China (ZJ2020-06) and the general fund of Civil Aviation Flight University of China (BSJ2021-2, BSJ2021-3, BSJ2022-6).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhi, Y.; Shao, P.; Feng, X.; Xia, H.; Zhang, Y.; Shi, Z.; Mu, Y.; Liu, X. Covalent organic frameworks: Efficient, metal-free, heterogeneous organocatalysts for chemical fixation of CO2 under mild conditions. J. Mater. Chem. A 2018, 6, 374–382. [Google Scholar] [CrossRef]
  2. Chakraborty, J.; Nath, I.; Song, S.; Kao, C.; Verpoort, F. Synergistic performance of a sub-nanoscopic cobalt and imidazole grafted porous organic polymer for CO2 fixation to cyclic carbonates under ambient pressure without a co-catalyst. J. Mater. Chem. A 2020, 8, 13916–13920. [Google Scholar] [CrossRef]
  3. Li, W.; Qi, K.; Lu, X.; Qi, Y.; Zhang, J.; Zhang, B.; Qi, W. Electrochemically assisted cycloaddition of carbon dioxide to styrene oxide on copper/carbon hybrid electrodes: Active species and reaction mechanism. Chem. Eur. J. 2022, 28, e202200622. [Google Scholar] [CrossRef] [PubMed]
  4. Yin, J.; Zhang, T.; Schulman, E.; Liu, D.; Meng, J. Hierarchical porous metallized poly-melamine-formaldehyde (PMF) as a low-cost and high-efficiency catalyst for cyclic carbonate synthesis from CO2 and epoxides. J. Mater. Chem. A 2018, 6, 8441–8448. [Google Scholar] [CrossRef]
  5. Lyubimov, S.; Zvinchuk, A.; Chowdhury, B. Iodine as an efficient and available activator of sodium and potassium halides in carbon dioxide addition to epoxides. Russian Chem. Bull. 2021, 70, 1324–1327. [Google Scholar] [CrossRef]
  6. Mujmule, R.; Chung, W.; Kim, H. Chemical fixation of carbon dioxide catalyzed via hydroxyl and carboxyl-rich glucose carbonaceous material as a heterogeneous catalyst. Chem. Eng. J. 2020, 395, 125164. [Google Scholar] [CrossRef]
  7. Qing, Y.; Liu, T.; Zhao, B.; Bao, X.; Yuan, D.; Yao, Y. Cycloaddition of di-substituted epoxides and CO2 under ambient conditions catalysed by rare-earth poly(phenolate) complexes. Inorg. Chem. Front. 2022, 9, 2969–2979. [Google Scholar] [CrossRef]
  8. Emelyanov, M.; Lisov, A.; Medvedev, M.; Maleev, V.; Larionov, V. Cobalt(III) complexes as bifunctional hydrogen-bond donor catalysts featuring halide anions for cyclic carbonate synthesis at ambient temperature and pressure: A mechanistic insight. Asian J. Org. Chem. 2022, 11, 202100811. [Google Scholar] [CrossRef]
  9. Emelyanov, M.; Stoletova, N.; Lisov, A.; Medvedev, M.; Smol’yakov, A.; Maleev, V.; Larionov, V. An octahedral cobalt(iii) complex based on cheap 1,2-phenylenediamine as a bifunctional metal-templated hydrogen bond donor catalyst for fixation of CO2 with epoxides under ambient conditions. Inorg. Chem. Front. 2021, 8, 3871–3884. [Google Scholar] [CrossRef]
  10. Campisciano, V.; Valentino, L.; Morena, A.; Santiago-Portillo, A.; Saladino, N.; Gruttadauria, M.; Aprile, C.; Giacalone, F. Carbon nanotube supported aluminum porphyrin-imidazolium bromide crosslinked copolymer: A synergistic bifunctional catalyst for CO2 conversion. J. CO2 Util. 2022, 57, 101884. [Google Scholar] [CrossRef]
  11. Li, Y.; Zhang, X.; Lan, J.; Xu, P.; Sun, J. Porous Zn(Bmic)(AT) MOF with Abundant Amino Groups and Open Metal Sites for Efficient Capture and Transformation of CO2. Inorg. Chem. 2019, 58, 13917–13926. [Google Scholar] [CrossRef]
  12. Ahmad, A.; Khan, S.; Tariq, S.; Luque, R.; Verpoort, F. Self-sacrifice MOFs for heterogeneous catalysis: Synthesis mechanisms and future perspectives. Mater. Today 2022, 55, 137–169. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Liu, K.; Wu, L.; Zhong, H.; Luo, N.; Zhu, Y.; Tong, M.; Long, Z.; Chen, G. Silanol-enriched viologen-based ionic porous hybrid polymers for efficient eatalytic CO2 fixation into cyclic carbonates under mild conditions. Chem. Eng. 2019, 7, 16907–16916. [Google Scholar]
  14. Biswas, T.; Mahalingam, V. g-C3N4 and tetrabutylammonium bromide catalyzed efficient conversion of epoxide to cyclic carbonate under ambient conditions. New J. Chem. 2017, 41, 14839–14842. [Google Scholar] [CrossRef]
  15. Ong, W.; Tan, L.; Ng, Y.; Yong, S.; Chai, S. Graphitic carbon nitride (g-C3N4)-based photocatalysts for artificial photosynthesis and environmental Remediation: Are we a step closer to achieving sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  16. Li, Y.; Zhu, S.; Kong, X.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. In situ synthesis of a novel Mn3O4/g-C3N4 p-n heterostructure photocatalyst for water splitting. J. Colloid Interf. Sci. 2021, 586, 778–784. [Google Scholar] [CrossRef]
  17. Zhou, X.; Luo, J.; Jin, B.; Wu, Z.; Yang, S.; Zhang, S.; Tian, Y.; Fang, Y.; Hou, Y.; Zhou, X. Sustainable synthesis of low-cost nitrogen-doped-carbon coated Co3W3C@g-C3N4 composite photocatalyst for efficient hydrogen evolution. Chem. Eng. J. 2021, 426, 131208. [Google Scholar] [CrossRef]
  18. Liang, H.; Wang, J.; Li, Q.; Liang, C.; Feng, Y.; Kang, M. Supported ZnBr2 and carbon nitride bifunctional complex catalysts for the efficient cycloaddition of CO2 with diglycidyl ethers. New J. Chem. 2018, 42, 16127–16137. [Google Scholar] [CrossRef]
  19. Huang, Z.; Li, F.; Chen, B.; Lu, T.; Yuan, Y.; Yuan, G. Well-dispersed g-C3N4 nanophases in mesoporous silica channels and their catalytic activity for carbon dioxide activation and conversion. Appl. Catal. B-Environ. 2013, 136, 269–277. [Google Scholar] [CrossRef]
  20. Xu, J.; Gan, Y.; Hu, P.; Zheng, H.; Xue, B. Comprehensive insight into the support effect of graphitic carbon nitride for zinc halides on the catalytic transformation of CO2 into cyclic carbonates. Catal. Sci. Technol. 2018, 8, 5582–5593. [Google Scholar] [CrossRef]
  21. Su, Q.; Yao, X.; Cheng, W.; Zhang, S. Boron-doped melamine-derived carbon nitrides tailored by ionic liquids for catalytic conversion of CO2 into cyclic carbonates. Green Chem. 2017, 19, 2957–2965. [Google Scholar] [CrossRef]
  22. Zhu, J.; Diao, T.; Wang, W.; Xu, X.; Sun, X.; Carabineiro, S.; Zhao, Z. Boron doped graphitic carbon nitride with acid-base duality for cycloaddition of carbon dioxide to epoxide under solvent-free condition. Appl. Catal. B-Environ. 2017, 219, 92–100. [Google Scholar] [CrossRef]
  23. Lan, D.; Wang, H.; Chen, L.; Au, C.; Yin, S. Phosphorous-modified bulk graphitic carbon nitride: Facile preparation and application as an acid-base bifunctional and efficient catalyst for CO2 cycloaddition with epoxides. Carbon 2016, 100, 81–89. [Google Scholar] [CrossRef]
  24. Guo, S.; Deng, Z.; Li, M.; Jiang, B.; Tian, C.; Pan, Q.; Fu, H. Phosphorus-doped carbon nitride tubes with a layered micro-nanostructure for enhanced visible-light photocatalytic hydrogen evolution. Angew. Chem. Int. Edit. 2016, 55, 1830–1834. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.; Liu, X.; Hou, L.; Guo, X.; Fu, R.; Sun, J. Construction of covalent bonding oxygen-doped carbon nitride/graphitic carbon nitride Z-scheme heterojunction for enhanced visible-light-driven H2 evolution. Chem. Eng. J. 2020, 383, 123132. [Google Scholar] [CrossRef]
  26. Raza, A.; Haidry, A.; Saddique, J. In-situ synthesis of Cu2ZnSnS4/g-C3N4 heterojunction for superior visible light-driven CO2 reduction. J. Phys. Chem. Solids 2022, 165, 110694. [Google Scholar] [CrossRef]
  27. Samanta, S.; Srivastava, R. A novel method to introduce acidic and basic bi-functional sites in graphitic carbon nitride for sustainable catalysis: Cycloaddition, esterification, and transesterification reactions. Sustain. Energ. Fuels 2017, 1, 1390–1404. [Google Scholar] [CrossRef]
  28. Liu, Y.; Chen, Y.; Liu, Y.; Chen, Z.; Yang, H.; Yue, Z.; Fang, Q.; Zhi, Y.; Shan, S. Zn and N co-doped porous carbon nanosheets for photothermally-driven CO2 cycloaddition. J. Catal. 2022, 407, 65–76. [Google Scholar] [CrossRef]
  29. Sakuna, P.; Ketwong, P.; Ohtani, B.; Trakulmututa, J.; Kobkeatthawin, T.; Luengnaruemitchai, A.; Smith, S. The influence of metal-doped graphitic carbon nitride on photocatalytic conversion of acetic acid to carbon dioxide. Front. Chem. 2022, 10, 825786. [Google Scholar] [CrossRef]
  30. Zhu, T.; Song, Y.; Ji, H.; Xu, Y.; Song, Y.; Xia, J.; Yin, S.; Li, Y.; Xu, H.; Zhang, Q.; et al. Synthesis of g-C3N4/Ag3VO4 composites with enhanced photocatalytic activity under visible light irradiation. Chem. Eng. J. 2015, 271, 96–105. [Google Scholar] [CrossRef]
  31. Wang, X.; Yang, L.; Fu, G.; Chen, Y.; Yang, C.; Sun, J. Experimental and theoretical investigation for the cycloaddition of carbon dioxide to epoxides catalyzed by potassium and boron co-doped carbon nitride. J. Colloid Interf. Sci. 2022, 609, 523–534. [Google Scholar] [CrossRef]
  32. Ma, X.; Tian, J.; Wang, M.; Shen, M.; Zhang, L. Polymeric carbon nitride supported Bi nanoparticles as highly efficient CO2 reduction electrocatalyst in a wide potential range. J. Colloid Interf. Sci. 2022, 608, 1676–1684. [Google Scholar] [CrossRef]
  33. Li, Y.; Zhai, G.; Liu, Y.; Wang, Z.; Wang, P.; Zheng, Z.; Cheng, H.; Dai, Y.; Huang, B. Synergistic effect between boron containing metal-organic frameworks and light leading to enhanced CO2 cycloaddition with epoxides. Chem. Eng. J. 2022, 437, 135363. [Google Scholar] [CrossRef]
  34. Wang, K.; Fu, J.; Zheng, Y. Insights into photocatalytic CO2 reduction on C3N4: Strategy of simultaneous B, K co-doping and enhancement by N vacancies. Appl. Catal. B Environ. 2019, 254, 270–282. [Google Scholar] [CrossRef]
  35. Liu, J.; Yang, G.; Liu, Y.; Zhang, D.; Hu, X.; Zhang, Z. Efficient conversion of CO2 into cyclic carbonates at room temperature catalyzed by Al-salen and imidazolium hydrogen carbonate ionic liquid. Green Chem. 2020, 22, 4509–4515. [Google Scholar] [CrossRef]
  36. Ge, S.; Cheng, G.; Ke, H. Triethanolamine borate as bifunctional Lewis pair catalyst for the cycloaddition of CO2 with epoxides. J. CO2 Util. 2022, 57, 101873. [Google Scholar] [CrossRef]
  37. Yang, C.; Chen, Y.; Wang, X.; Sun, J. Polymeric ionic liquid with carboxyl anchored on mesoporous silica for efficient fixation of carbon dioxide. J. Colloid Interf. Sci. 2022, 618, 44–55. [Google Scholar] [CrossRef]
  38. Cheng, X.; Zhao, P.; Zhang, M.; Wang, S.; Liu, M.; Liu, F. Fabrication of robust and bifunctional cyclotriphosphazene-based periodic mesoporous organosilicas for efficient CO2 adsorption and catalytic conversion. Chem. Eng. J. 2021, 418, 129360. [Google Scholar] [CrossRef]
  39. Qu, Y.; Chen, Y.; Sun, J. Conversion of CO2 with epoxides to cyclic carbonates catalyzed by amino acid ionic liquids at room temperature. J. CO2 Util. 2022, 56, 101840. [Google Scholar] [CrossRef]
  40. Xie, Y.; Sun, Q.; Fu, Y.; Song, L.; Liang, J.; Xu, X.; Wang, H.; Li, J.; Tu, S.; Lu, X.; et al. Sponge-like quaternary ammonium-based poly(ionic liquid)s for high CO2 capture and efficient cycloaddition under mild conditions. J. Mater. Chem. A 2017, 5, 25594–25600. [Google Scholar] [CrossRef]
  41. Yang, C.; Chen, Y.; Xu, P.; Yang, L.; Zhang, J.; Sun, J. Facile synthesis of zinc halide-based ionic liquid for efficient conversion of carbon dioxide to cyclic carbonates. Mol. Catal. 2020, 480, 110637. [Google Scholar] [CrossRef]
  42. Shi, X.; Chen, Y.; Duan, P.; Zhang, W.; Hu, Q. Conversion of CO2 into organic carbonates over a fiber-supported ionic liquid catalyst in impellers of the agitation system. ACS Sustain. Chem. Eng. 2018, 6, 7119–7127. [Google Scholar] [CrossRef]
  43. Zhong, H.; Gao, J.; Sa, R.; Yang, S.; Wu, Z.; Wang, R. Carbon dioxide conversion upgraded by host-guest cooperation between nitrogen-rich covalent organic framework and imidazolium-based ionic polymer. ChemSusChem 2020, 13, 6323–6329. [Google Scholar]
  44. Song, H.; Wang, Y.; Xiao, M.; Liu, L.; Liu, Y.; Liu, X.; Gai, H. Design of novel poly(ionic liquids) for the conversion of CO2 to cyclic carbonates under mild conditions without solvent. ACS Sustain. Chem. Eng. 2019, 7, 9489–9497. [Google Scholar] [CrossRef]
  45. Zhang, X.; Geng, W.; Yue, C.; Wu, W.; Xiao, L. Multilayered supported ionic liquids bearing a carboxyl group: Highly efficient catalysts for chemical fixation of carbon dioxide. J. Environ. Chem. Eng. 2016, 4, 2565–2572. [Google Scholar] [CrossRef]
  46. Yang, C.; Chen, Y.; Qu, Y.; Zhang, J.; Sun, J. Phase-controllable polymerized ionic liquids for CO2 fixation into cyclic carbonates. Sustain. Energ. Fuels 2021, 5, 1026–1033. [Google Scholar] [CrossRef]
  47. Xu, J.; Gan, Y.; Pei, J.; Xue, B. Metal-free catalytic conversion of CO2 into cyclic carbonate by hydroxylfunctionalized graphitic carbon nitride materials. Mol. Catal. 2020, 491, 110979. [Google Scholar] [CrossRef]
  48. Xu, J.; Wu, F.; Jiang, Q.; Shang, J.; Li, Y. Metal halides supported on mesoporous carbon nitride as efficient heterogeneous catalysts for the cycloaddition of CO2. J. Mol. Catal. A-Chem. 2015, 403, 77–83. [Google Scholar] [CrossRef]
  49. Zhang, D.; Xu, T.; Li, C.; Xu, W.; Wang, J.; Bai, J. Synthesis of carbon fibers support graphitic carbon nitride immobilize ZnBr2 catalyst in the catalytic reaction between styrene oxide and CO2. J. CO2 Util. 2019, 34, 716–724. [Google Scholar] [CrossRef]
  50. Wang, X.; Shi, L.; Chen, Y.; Yang, C.; Yang, L.; Sun, J. Facile synthesis of carboxyl- and hydroxyl-functional carbon nitride catalyst for efficient CO2 cycloaddition. Mol. Catal. 2021, 515, 111924. [Google Scholar] [CrossRef]
Figure 1. Standard curve of peak area versus PC concentration.
Figure 1. Standard curve of peak area versus PC concentration.
Catalysts 12 01196 g001
Figure 2. (A) FT-IR spectra, (B) XRD patterns of synthesized catalysts.
Figure 2. (A) FT-IR spectra, (B) XRD patterns of synthesized catalysts.
Catalysts 12 01196 g002
Figure 3. (A) C 1s; (B) N 1s; (C) B 1s; (D) P 2p XPS spectra of BP-CN-1.
Figure 3. (A) C 1s; (B) N 1s; (C) B 1s; (D) P 2p XPS spectra of BP-CN-1.
Catalysts 12 01196 g003
Figure 4. N2 adsorption–desorption isotherms of the synthesized catalysts. (A) N2 adsorption–desorption isotherms; (B) pore size distribution.
Figure 4. N2 adsorption–desorption isotherms of the synthesized catalysts. (A) N2 adsorption–desorption isotherms; (B) pore size distribution.
Catalysts 12 01196 g004
Figure 5. (A) HRTEM; (B) SEM images of BP-CN-1; (C) TEM; and (D) SEM images of CN.
Figure 5. (A) HRTEM; (B) SEM images of BP-CN-1; (C) TEM; and (D) SEM images of CN.
Catalysts 12 01196 g005
Figure 6. Effects of reaction parameters on CO2 cycloaddition reactions over BP-CN-1. Reaction conditions: PO 34.5 mmol, 100 mg BP-CN-1, 100 mg Bu4NBr, (A) effect of reaction temperature, 2.0 MPa, 6 h; (B) effect of CO2 pressure, 120 °C, 6 h; (C) effect of reaction time, 120 °C, 2.0 MPa.
Figure 6. Effects of reaction parameters on CO2 cycloaddition reactions over BP-CN-1. Reaction conditions: PO 34.5 mmol, 100 mg BP-CN-1, 100 mg Bu4NBr, (A) effect of reaction temperature, 2.0 MPa, 6 h; (B) effect of CO2 pressure, 120 °C, 6 h; (C) effect of reaction time, 120 °C, 2.0 MPa.
Catalysts 12 01196 g006
Figure 7. (A) Recyclability of BP-CN-1, conditions: PO 34.5 mmol, 100 mg BP-CN-1, 100 mg Bu4NBr, 120 °C, CO2 2.0 MPa, 6 h; (B) FT-IR spectra of BP-CN-1 and reused BP-CN-1.
Figure 7. (A) Recyclability of BP-CN-1, conditions: PO 34.5 mmol, 100 mg BP-CN-1, 100 mg Bu4NBr, 120 °C, CO2 2.0 MPa, 6 h; (B) FT-IR spectra of BP-CN-1 and reused BP-CN-1.
Catalysts 12 01196 g007
Table 1. SBET, BJH average pore diameter and B, P mass fraction of catalysts.
Table 1. SBET, BJH average pore diameter and B, P mass fraction of catalysts.
SamplesSBET (m2 g−1)Average Pore Diameter (nm)B (wt%)P (wt%)
CN2.20.01
B-CN13.824.51.21
P-CN12.514.11.27
BP-CN-111.912.10.101.37
BP-CN-29.313.70.483.56
BP-CN-35.520.70.965.00
Table 2. Catalytic performance of catalysts a.
Table 2. Catalytic performance of catalysts a.
EntryCatalystCoCatalystYield b (%)Selectivity b (%)
10
2CN2
3Bu4NBr2598
4CNBu4NBr2699
5B-CNBu4NBr5099
6P-CNBu4NBr4498
7BP-CN-1Bu4NBr6498
8BP-CN-2Bu4NBr5898
9BP-CN-3Bu4NBr5398
10 cBu4NBr5598
11 cBu4NI5998
12 cBP-CN-1Bu4NI9998
a Reaction: PO 34.5 mmol, 100 °C, CO2 2 MPa, 6 h, 100 mg catalyst, 100 mg Bu4NBr; b Yield and selectivity were determined by GC; c 120 °C.
Table 3. Universality of BP-CN-1 for catalytic conversion of various epoxides a.
Table 3. Universality of BP-CN-1 for catalytic conversion of various epoxides a.
EntryEpoxidesProductsResults a
Conversion b (%)Yield b (%)Selectivity b (%)
1 Catalysts 12 01196 i001 Catalysts 12 01196 i002969599
2 Catalysts 12 01196 i003 Catalysts 12 01196 i004878699
3 Catalysts 12 01196 i005 Catalysts 12 01196 i006999999
4 Catalysts 12 01196 i007 Catalysts 12 01196 i008747398
5 c Catalysts 12 01196 i009 Catalysts 12 01196 i010858297
6 c Catalysts 12 01196 i011 Catalysts 12 01196 i012666396
a Reaction condition: PO 34.5 mmol, 120 °C, CO2 2 MPa, 6 h, 100 mg BP-CN-1, 100 mg Bu4NBr; b Conversion and selectivity were determined by GC; c 12 h.
Table 4. Comparisons of B, P-CN-1 catalyst with reported CN catalysts in the CO2 cycloaddition with epoxides a.
Table 4. Comparisons of B, P-CN-1 catalyst with reported CN catalysts in the CO2 cycloaddition with epoxides a.
EntryCatalystsSolventReaction ConditionYield (%)TOFTONReference
Tem. (°C)Press (MPa)T(h)
1eg-CN-OHDMF1402.06.0605.835.0[47]
2mpg-CN/ZnCl2(10)DMF1402.56.0727.042.0[48]
3Zn-CN/CDMF1402.08.0745.443.3[49]
4 bB0.1CN/SBA-151303.024951.228.5[22]
5MCNB(0.01)1302.86.0311.59.1[21]
6K,B-CN-4(Bu4NBr)1102.06.0875.130.7[31]
7HCN-FD(Bu4NI)1102.06.0925.432.4[50]
8P-CN-2 (Bu4NBr)1002.04.0994.819.3[23]
9B,P-CN-1(Bu4NBr)1202.06.0955.533.1This work
a PO; b SO; TOF: the mass of produced PC per gram catalyst per hour; TON: the mass of produced PC per gram catalyst.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, C.; Zhao, X.; Yang, T. Boron and Phosphorus Co-Doped Graphitic Carbon Nitride Cooperate with Bu4NBr as Binary Heterogeneous Catalysts for the Cycloaddition of CO2 to Epoxides. Catalysts 2022, 12, 1196. https://doi.org/10.3390/catal12101196

AMA Style

Yang C, Zhao X, Yang T. Boron and Phosphorus Co-Doped Graphitic Carbon Nitride Cooperate with Bu4NBr as Binary Heterogeneous Catalysts for the Cycloaddition of CO2 to Epoxides. Catalysts. 2022; 12(10):1196. https://doi.org/10.3390/catal12101196

Chicago/Turabian Style

Yang, Chaokun, Xin Zhao, and Tuantuan Yang. 2022. "Boron and Phosphorus Co-Doped Graphitic Carbon Nitride Cooperate with Bu4NBr as Binary Heterogeneous Catalysts for the Cycloaddition of CO2 to Epoxides" Catalysts 12, no. 10: 1196. https://doi.org/10.3390/catal12101196

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