Next Article in Journal
Vertically Ti3CN@NiFe LDH Nanoflakes as Self-Standing Catalysts for Enhanced Oxygen Evolution Reaction
Previous Article in Journal
Engineering the C-Terminal Region to Enhance the Thermal Stability of Streptomyces hygroscopicus Transglutaminase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tailoring Mesoporosity of Multi-Hydroxyls Hyper-Crosslinked Organic Polymers for Reinforced Ambient Chemical Fixation of CO2

1
School of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252000, China
2
Shandong Lusoft Digital Technology Co., Ltd., Jinan 250001, China
3
College of Chemistry & Chemical Engineering, Northeast Petroleum University, Daqing 163318, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(10), 707; https://doi.org/10.3390/catal14100707
Submission received: 24 June 2024 / Revised: 7 September 2024 / Accepted: 9 September 2024 / Published: 10 October 2024
(This article belongs to the Section Catalytic Materials)

Abstract

:
Ambient condition-determined chemical CO2 fixation affords great promise for remitting the pressure of CO2 release. The construction of a microporous environment easily captures CO2 molecules around the reactive sites of the catalyst to reinforce the reaction process. Herein, multi-hydroxyl-containing hyper-crosslinked organic polymers (HCPs-OH-n) are synthesized by the polymerization of 1,4-dichlorobenzyl (DCX) and m-trihydroxybenzene in the monosaccharide form in a Friedel–Crafts alkylation hypercrosslinking process (FCAHP). By tuning the DCX ratio in the FCAHP, the structural properties can be regulated to create a more microporous surface in the HCPs-OH-n; meanwhile, the formed multi-hydroxyl species in the microporous environment could induce the easy interaction between hydroxyls and epoxides by forming a hydrogen bond, which improves the activation of epoxides during the cycloaddition reaction to synthesize the cyclic carbonates at ambient conditions. The structural properties suggest that HCPs-OH-n possess a large surface area with appreciable microporous and mesoporous distribution. As expected, the HCPs-OH-3 bearing the most abundant mesoporosity affords the highest reactivity in the chemical CO2 fixation to cyclic carbonates and is endowed with rational recoverability.

1. Introduction

The appeal of chemically fixing CO2 to produce economically competitive cyclic carbonates stems from the wide applications for these compounds [1,2,3,4,5,6,7]. The cycloaddition reaction of CO2 and epoxides can typically be triggered using a diverse array of heterogeneous catalysts concluding zeolite, metal oxides, ion-exchanged resins, mesoporous materials, nanoporous polymers, and metalorganic frameworks (MOFs) [5,6,7,8,9,10,11,12,13,14]. However, harsh condition including high pressure (>10 atm) and temperature (>70 °C) are often required to ensure the smooth occurrence of the reaction [5,6,7,8,9]. Such a reaction condition should require additive energy consumption to supply heat, which possibly leads to increased CO2 discharge. Therefore, the development of high-performance catalysts that can be desirable for his reaction at ambient conditions, thereby avoiding the generation of additional CO2, is exceedingly desirable.
In a recent development, our research has provided a molecular-level comprehensive understanding of the CO2 cycloaddition with epoxides catalyzed by ionic liquids (ILs). This work also emphasizes the rationalization of IL-based organocatalysts for the chemical fixation of CO2 [15]. The catalytic activity of [p-ArOH-IM] I arises from the incorporation of phenolic -OH groups and nucleophilic-leaving capable I anion. The former effectively serves as an efficient HBD to accelerate the activation of epoxide via a H-bond, while the latter acts as a hydrogen bond donor (HBD), while the latter serves as both a proficient nucleophile and an excellent leaving group, facilitating the ring-opening process to produce cyclic carbonates [15]. A similar activating influence can be observed within a dual-catalyst system involving pentaerythritol and a tetraalkylammonium halide, utilized for the CO2 coupling with epoxides [16]. Porous organic polymers (POPs) present themselves as promising materials to meet the demands of CO2 capture and conversion, owing to their appealing characteristics, including a substantial surface area, abundant porosity, a stable framework, and adaptability through easy customization [13,17]. A variety of versatile POPs have been synthesized by employing diverse functional building blocks [18,19,20,21,22]. Our recent work has demonstrated an efficient effect of dual hydroxyls for the activation of epoxides, which provides the inspiration for multi-hydroxyl-containing hyper-crosslinked organic polymers to explore the function of multiple hydroxyls species [17]. Apart from the affinity and activation of epoxides on the surface of the catalyst, the local CO2 concentration around the catalytic active sites possibly also determines the reaction progress at the dynamic level. Therefore, it should be emphasized on the local CO2 enrichment around the catalytic active sites for reinforcing the reaction kinetic process. However, a key role of micropores in CO2 conversion catalysts tends to be ignored because a micropore environment could ensure the enrichment of low-concentration CO2 in the reactive sites.
Recently, the utilization of Friedel–Crafts alkylation has been showcased for aromatic CO2-philic moiety monomers, demonstrating its application in synthesizing porous organic polymers (POPs) through a process akin to “knitting” [8]. These networks result from the extensive cross-linking of polymer monomers, where the inefficient packing of rigid and intricately shaped constituents gives rise to a porous microstructure. Therefore, the tactical synthesis of functional organic polymers with multiple hydroxyls confined into abundant micropores for reinforced CO2 capture and successive chemical conversion becomes essential and attractive.
Herein, we report the polymerization of 1,4-dichlorobenzyl (DCX) and a controlled amount of m-trihydroxybenzene as monomers in the monosaccharide form in a Friedel–Crafts alkylation hypercrosslinking process. These polymers feature intricately crosslinked organic frameworks, multiple hydroxyl groups, and finely distributed micropores, collectively contributing to their substantial surface area and pore volume. The adjusted microporous properties and abundant multi-hydroxyl groups were integrated into the functional polymers to afford the reaction microenvironment. The HCPs-OH-n were applied as multi-hydroxyl-containing hyper-crosslinked organic polymers for the chemical CO2 fixation to construct the cyclic carbonates. Under ambient conditions, the catalyst HCPs-OH-3, which contains hydrogen bond donor (HBD) functionality, has demonstrated remarkable activity, selectivity, and recyclability in the chemical CO2 fixation to form cyclic carbonates.

2. Results and Discussion

2.1. Synthesis and Characterization

The synthesis procedure for hydroxyl-functionalized supercrosslinked porous organic polymers is illustrated in Figure 1. Adjusting the ratio of the raw monomers enables the customization of the physicochemical characteristics of the resulting hydroxyl-functionalized supercrosslinked porous organic polymers. Through modulation of the proportion of m-trihydroxybenzene concerning DCX in the initial synthesis composition, a series of HCPs-OH-n samples (as detailed in Table 1) were synthesized. All HCPs-OH-n samples exhibit N2 adsorption and desorption isotherms of type II (as depicted in Figure 2a), indicative of the characteristic presence of both microporous and mesoporous structures. Nitrogen adsorption (P/P0 < 0.001) displays a significant increase, signifying the abundant presence of micropores within the material. Conversely, in the P/P0 = 0.5–0.9, nitrogen adsorption exhibits a steady rise, highlighting the formation of a mesoporous structure. In addition, the polymer HCPs-OH-n has a clear hysteresis loop in the high-pressure region (P/P0 > 0.9), indicating the presence of large cavities or cage-like pores also caused by inter-particle accumulation. It should be noted that HCPs-OH-3 afford the most obvious microporous adsorption behavior in the low-pressure region compared to other counterparts (Figure S1c), which is indicative of the more micropores in the resulting materials. Furthermore, the largest microporous specific surface area of 886 m2 g−1 can be observed for HCPs-OH-3 (Table 1), which outperform the other counterparts. Figure 2b shows the dominant mesopore size distribution curves of the supercrosslinked polymers, where the pore size of the polymer HD is between 3.4 and 4.2 nm. The elemental analysis (Table 1) reveals that the C, H content of HCPs-OH-n varies with changing monomers inside the polymer. The number of hydroxyl groups (Table 1, k) contained on the surface of the material HCPs-OH-n was determined by acid–base titration, and it was found that HCPs-OH-1 have the highest hydroxyl content.
As evident from the XRD findings (depicted in Figure 3a), the polymer networks exhibit an amorphous character, consistent with the characteristic disorderliness of polymers. To further assess the surface characteristics, we conducted XPS analyses on representative samples of HCPs-OH-3 and HCPs. The comprehensive survey spectra reveal the presence of similar elemental species, primarily composed of C and O (Figure 3b,c). In particular, the O element content HCPs-OH-3 affords a high fraction compared to the HCPs, revealing the successful incorporation of multiple hydroxyl groups. The O 1s XPS profile in HCPs-OH-3 determines a much higher intensity compared to that of HCPs [23], which further confirms the presence of the phenolic hydroxyl group of m-trihydroxybenzene in the skeleton according to the reported work. The C 1s XPS spectra (Figure 3d) are deconvoluted into three distinct peaks around 288.5, 286.5, and 285.1 eV of HCPs-OH-3 [24], indexed into Cl-C, C-C, and C-OH. It can be found that typical C-OH species can be demonstrated relative to HCPs. Taken together, the HCPs-OH-3 afford the typical multiple hydroxyl characteristics in the polymer networks, which can act as the active functional groups for triggering the activation of epoxides.
SEM images (Figure 4a–c) reveal that HCPs-OH-3 consist of irregular micrometer-sized particles. These primary particles within the HCPs-OH-3 samples exhibit interactions that lead to the formation of mixed spherical shapes and nanowire-like structures. This morphological characterization is further confirmed by TEM images (Figure 4d–f and Figure S2). Furthermore, a distinct randomly oriented microporous structure is evident, underscoring the presence of microporous features in the synthesized HCPs-OH-3. Elemental mapping images (Figure 4g–i) demonstrate a uniform distribution of elements C and O throughout the whole skeleton, indicating the successful incorporation of m-trihydroxybenzene within the polymer matrix. The thermal stability of HCPs-OH-3 polymer is evidenced by the TG profile, where the initial decomposition begins at temperatures exceeding 474 °C (Figure 5a).
The CO2 adsorption performance of HCPs-OH-n was evaluated via gas static adsorption, and the corresponding adsorption isotherms were obtained at 273 K, extending up to 1 bar (Figure 5b). The CO2 adsorption results reveal that HCPs-OH-n afford a CO2 uptake of 2.1–3.1 mmol g−1 at 273 K. Typically, the CO2 uptake behavior is closely linked to the adsorbent’s porosity, where a higher microporous surface area is conducive to greater CO2 adsorption. Specifically, among the samples, HCPs-OH-3 feature the highest microporous surface area and pore volume, facilitating enhanced CO2 adsorption. However, HCPs-OH-5 possess a higher BET and microporous surface area than HCPs-OH-2. Surprisingly, despite these attributes, HCPs-OH-5 exhibit a lower CO2 adsorption capacity than HCPs-OH-2. This discrepancy can be attributed to the stronger hydrogen bonding interactions between the -OH groups and CO2 in HCPs-OH-2, which enhances its carbon capture performance.

2.2. Catalytic Activity Analysis

The economically efficient reaction between CO2 and epoxides represents an appealing pathway to obtain cyclic carbonates, which hold significant importance in various applications, including organic synthesis, utilization as electrolyte solvents in batteries, degreasers, and fuel additives, among others [25,26]. Given the abundant presence of H-bond donor sites in HCPs-OH-n, which can facilitate the activation of epoxides, we performed the catalytic activity of the cycloaddition of epoxides with CO2 [27]. We initially selected the CO2 cycloaddition with epichlorohydrin as our test reaction to assess the optimized conditions and the corresponding catalytic characteristics of HCPs-OH-n (Table 1). To our delight, HCPs-OH-n demonstrated efficient catalytic activity in facilitating the cyclization of epichlorohydrin with CO2 when combined with tetrabutylammonium iodide (TBAI) under ambient conditions, resulting in epichlorohydrin conversions ranging from 70% to 82% (Table 1, entries 1–5). The most favorable catalytic performance, achieving an 82% conversion rate, was observed with HCPs-OH-3 when employing TBAI as the co-catalyst and maintaining a CO2 pressure of 1 bar at room temperature for a duration of 10 h (entry 3). Notably, when tetrabutyl ammonium bromide (TBAB) was used as the co-catalyst, the conversion rate remained at a respectable 74% (entry 4). To provide context, control experiments were conducted (Table 1, entries 6–10). These experiments revealed that HCPs-OH-3 exhibited minimal activity at such an ambient condition (Table 1, entry 10). Furthermore, in the absence of HCPs-OH-3, a significant reduction in substrate conversion was observed (entries 7). The above results can be deduced that the resulting excellent activity of HCPs-OH-3/TBAI might be attributed to the synergistic effect between HCPs-OH-3 and TBAI. Normally, there was trace amount of residual catalyst (FeCl3) in the HCPs-OH. For example, the Fe content was below 0.1 wt% for the sample HCPs-OH-3. The activity of FeCl3 and HCPs-OH-3 in the presence of FeCl3 was investigated to clarify the effects of the residual FeCl3 (Table S1). The yield was 81% by using HCPs-OH-3 in the presence of FeCl3 (entry 2), nearly the same as the one in the absence of FeCl3. These results indicated the rare influence of the residual FeCl3 on the reaction. Especially, the presence of functional groups -OH in HCPs-OH-3 mainly determines the enhanced catalytic activity of the reaction. The combination of homogeneous 1,3,5 phloroglucinol with TBAI showed similar activities to that of HCPs-OH-3/TBAI (Table 1, entries 3, 8). These outcomes suggest that the majority of -OH groups within HCPs-OH-3 are readily accessible and efficiently engaged. This accessibility can be attributed to the porous structure of HCPs-OH-3, which does not hinder the interaction between -OH groups and epichlorohydrin (PO). These -OH groups are capable of forming hydrogen bonds with epichlorohydrin, thereby facilitating its activation to a certain degree and promoting its subsequent conversion [28,29]. The phenolic resin rich in OH functional groups was used as a comparison sample with a significantly lower conversion rate of 62% than that of HCPs-OH-3, which confirms that the high specific surface enhances mass transfer and thus activity. Additionally, HCPs-OH-3/TBAI afford an obvious superiority in the catalytic activity compared to other counterparts with poor microporous structure properties. This factor, mainly ascribed to the microporous environment, could reinforce the CO2 enrichment around reactive active sites, thereby improving the reaction process and performance of the catalyst.
The stability and durability of a catalyst play pivotal roles in assessing its practical applicability [30,31,32,33,34,35]. To illustrate the stability and recyclability of HCPs-OH-3 in the CO2 cycloaddition of epichlorohydrin, we conducted a series of recycling experiments. HCPs-OH-3 were readily recovered through filtration and maintained its catalytic activity effectively during a five-run recycling test under ambient conditions (Figure 6a). Remarkably, the catalyst recovered in the fifth run displayed a kinetic profile nearly identical to that of the fresh catalyst throughout the reaction duration (Figure 6b), indicating minimal deactivation, even in the early stages of the reaction. This observation further underscores the excellent reusability of HCPs-OH-3 as a catalyst. Additionally, N2 adsorption isotherm data and SEM images of the recovered HCPs-OH-3 catalyst suggest that it retained its textural properties to a great extent compared to the fresh catalyst, further substantiating its impressive recyclability (Figure S3).
The epoxide substrates scope experiments were then screened using multi-hydroxyl-containing hyper-crosslinked organic polymers HCPs-OH-3 as the catalyst. Table 2 shows that the resulting reaction system was effective when in the presence of a variety of epoxides under ambient conditions (Table 2, entries 1–5). Furthermore, at 70 °C and 1 bar CO2 (entries 2, 3, 5), HCPs-OH-3 can efficiently and rapidly convert epoxides. We suggest that the constructed catalytic system successfully insert CO2 into a variety of epoxides with high yield and selectivity.

2.3. Insights into Reaction Mechanism

Numerous studies have proposed a mechanistic pathway for the cycloaddition of epoxide and CO2 triggered by hydroxyl groups [27,28,29]. In alignment with these findings, we present a similar mechanistic hypothesis (Figure 7). Firstly, the coupling reaction commences with the polarization of the C-O bond within the epoxide through hydrogen bonding interactions with the hydroxyl groups of HCPs-OH-3, which serve as Lewis acidic sites. This initial step activates the epoxy ring (1). Subsequently, the iodide anion (I) initiates an attack on the C-O bond of the coordinated epoxides, leading to ring opening (2). The resulting oxyanion species is postulated to be stabilized by hydrogen-bond donors. Following this, the electrophilic addition of CO2 to the oxygen atom forms an alkyl carbonate anion, which subsequently undergoes ring closure to yield the corresponding cyclic carbonate. Concurrently, the catalyst is regenerated. The above analysis underscores the reinforcing effect of hydrogen bonding, with the hydroxyl groups playing a crucial role in the initial activation of the epoxide and the stabilization of intermediates and transitional states throughout the reaction. This dual function reduces both the time and pressure required for the reaction.

3. Materials and Methods

3.1. Materials and Chemicals

1,4-dichlorobenzyl (C8H8Cl2, AR), m-trihydroxybenzene (C6H6O3, AR), Iron (III) chlorideanhydrous (FeCl3, AR), and 1,2-dichloroethane (C2H4Cl2, AR) were purchased from Aladdin Industrial Corporation (Shanghai, China). Epichlorohydrin (C3H5ClO, AR), epibromohydrin (C3H5BrO, AR), glycidyl phenyl Ether (C9H10O2, AR), allyl glycidyl ether (C6H10O2, AR), and styrene oxide (C8H8O, AR) were purchased from TCI (Shanghai, China) Chemical Industry Development Co., Ltd. (Shanghai, China). All reagents (AR grade) were commercially purchased and used as received without further purification.

3.2. Synthesis of Catalysts

The synthesis of catalysts was carried out using the Friedel–Crafts alkylation hypercrosslinking process (FCAHP). The supercrosslinked organic polymers derived from polymerization of 1,4-dichlorobenzyl (DCX) and m-trihydroxybenzene with Iron (III) chlorideanhydrous as a catalyst in 20 mL 1,2-dichloroethane solvent. The synthesis process is depicted in Figure 1, and the products were named as HCPs-OH-n, where n stands for the molar ratio of DCX added. In a typical procedure, 3 mmol of DCX, 5 mmol of m-trihydroxybenzene, and 10 mmol of anhydrous Iron (III) chloride were dissolved in 20 mL of anhydrous dichloroethane under N2 protection. The above mixture was stirred for 1 h and subsequently subjected to reflux conditions at 80 °C for 24 h. Following the completion of the reaction, the resulting brown-black precipitate was isolated, filtered, and subjected to three methanol washes. It was then subjected to a Soxhlet extraction process using anhydrous methanol as the solvent at 80 °C for 24 h. The resulting products, labeled as HCPs-OH-3, were subsequently dried at 100 °C for 12 h. Additionally, the HCP samples were synthesized using 3 mmol of DCX.

3.3. Catalytic Performance Assessment

The pressurized catalytic cycloaddition of CO2 was conducted within a 25 mL Schlenk flask equipped with a CO2 balloon. In a standard procedure, the reaction occurred at ambient condition by introducing a mixture comprising 5 mmol of epoxide, 30 mg of catalyst, and 90 mg of auxiliary tetrabutylammonium iodide as co-catalysts into the Schlenk flask, which was then sealed with a CO2 balloon and maintained at the preset temperature. Following the reaction, n-dodecane (0.5 g) was introduced to an internal standard, which was diluted with ethyl acetate. Subsequently, the products were analyzed using gas chromatography. Reusability was assessed through a series of five consecutive reaction cycles. The used catalyst was extracted from the reaction solution via filtration, washed with ethyl acetate, dried in vacuum conditions, and then employed in the next reaction cycle, with the addition of fresh co-catalysts for reutilization.

4. Conclusions

In conclusion, we have successfully fabricated a series of HCPs-OH-n, which are multi-hydroxyl-containing hyper-crosslinked organic polymers with finely tunable microporous properties. These HCPs-OH-n materials exhibit high surface areas and are rich in hydrogen bond donor sites, attributes that are proven to benefit the activation of epoxides. Consequently, the constructed microporous catalysts have demonstrated remarkable efficiency in reaction of CO2 with epoxides under ambient conditions. The phenolic hydroxyl groups within these materials plays a pivotal role in facilitating the reaction, particularly when combined with tetrabutylammonium iodide. The microporous framework enhances the concentration of CO2 around the catalytically active sites, thus improving the overall catalytic performance in the cycloaddition of epoxides and CO2. Furthermore, our catalyst exhibits outstanding recyclability, allowing for five consecutive reuse cycles, and is compatible with solvent-free and metal-free conditions. Ongoing efforts are directed towards the development of even more applicable practical catalysts for ambient condition CO2 fixation.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal14100707/s1. Figure S1: Micropore size distribution of various samples derived from HK method. Figure S2: TEM images of HCPs-OH-3. Figure S3: Characterizations of the reused catalyst HCPs-OH-3. Table S1: Cycloaddition of CO2 with epichlorohydrin under ambient conditions.

Author Contributions

Z.G. conceived the project. S.N. and H.H. supervised the project. Z.G. conceived the idea, designed the experiments, analyzed the data, and wrote the manuscript. Z.G. carried out most of the experiments. S.X., Y.Z. and Y.D. participated in the partial experiment, data analysis, and manuscript writing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Shandong Province (ZR2019BB075), the Development Project of Youth Innovation Team in Shandong Colleges and Universities (2019KJC031), Guangyue Young Scholar Innovation Team of Liaocheng University (LCUGYTD2022-02), and the Development Program of Youth Innovation Team in Colleges and Universities of Shandong Province (Grant no. 2022KJ109).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

Author Shuguang Ning is employed by the company Shandong Lusoft Digital Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Jiang, X.; Guo, X.; Nie, X.; Song, C.; Chen, J. Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef] [PubMed]
  2. Bragato, N.; Perosa, A.; Selva, M.; Fiorani, G. Dihydroxybenzene-derived ILs as Halide-Free, Single-component Organocatalysts for CO2 Insertion Reactions. ChemCatChem 2023, 15, e202201373. [Google Scholar] [CrossRef]
  3. Gao, D.; Arán-Ais, R.; Jeon, H.; Cuenya, B. Rational catalyst and electrolyte design for CO2 electroreduction towards multicarbon products. Nat. Catal. 2019, 2, 198–210. [Google Scholar] [CrossRef]
  4. Li, H.; Qiu, C.; Ren, S.; Dong, Q.; Zhang, S.; Zhou, F.; Liang, X.; Wang, J.; Li, S.; Yu, M. Na+-gated water-conducting nanochannels for boosting CO2 conversion to liquid fuels. Science 2020, 367, 667–671. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, X.T.; Hu, C.H.; Wu, J.J.; Cui, P.; FWei, Y. Defective NH2-UiO-66 (Zr) effectively converting CO2 into cyclic carbonate under ambient pressure, solvent-free and co-catalyst-free conditions. Chin. J. Chem. Eng. 2022, 43, 222–229. [Google Scholar] [CrossRef]
  6. Lu, Z.; He, J.; Guo, B.G.; Zhao, Y.L.; Cai, J.Y.; Xiao, L.Q.; Hou, L.X. Efficient homogenous catalysis of CO2 to generate cyclic carbonates by heterogenous and recyclable polypyrazoles. Chin. J. Chem. Eng. 2022, 43, 110–115. [Google Scholar] [CrossRef]
  7. Jiang, B.W.; Liu, J.; Yang, G.Q.; Zhang, Z.B. Efficient conversion of CO2 into cyclic carbonates under atmospheric by halogen and metal-free poly(ionic liquid)s. Chin. J. Chem. Eng. 2023, 55, 202–211. [Google Scholar] [CrossRef]
  8. Li, J.; Jia, D.G.; Guo, Z.J.; Liu, Y.Q.; Lyu, Y.N.; Zhou, Y.; Wang, J. Imidazolinium based porous hypercrosslinked ionic polymers for efficient CO2 capture and fixation with epoxides. Green Chem. 2017, 19, 2675–2686. [Google Scholar] [CrossRef]
  9. Tortajada, A.; Juliá-Hernández, F.; Borjesson, M.; Moragas, T.; Martin, R. Transition-Metal-Catalyzed Carboxylation Reactions with Carbon Dioxide. Angew. Chem. Int. Ed. 2018, 57, 15948–15982. [Google Scholar] [CrossRef]
  10. Xie, Y.; Wang, T.; Liu, X.; Zou, K.; Deng, W. Capture and conversion of CO2 at ambient conditions by a conjugated microporous polymer. Nat. Commun. 2013, 4, 1960–1966. [Google Scholar] [CrossRef]
  11. Scheffen, M.; Marchal, D.G.; Beneyton, T.; Schuller, S.K.; Klose, M.; Diehl, C.; Lehmann, J.; Pfister, P.; Carrillo, M.; He, H. A new-to-nature carboxylation module to improve natural and synthetic CO2 fixation. Nat. Catal. 2021, 4, 105–115. [Google Scholar] [CrossRef]
  12. Huang, C.; Dong, J.; Sun, W.; Xue, Z.; Ma, J.; Zheng, L.; Liu, C.; Li, X.; Zhou, K.; Qiao, X.Z. Coordination mode engineering in stacked nanosheet metal–organic frameworks to enhance catalytic reactivity and structural robustness. Nat. Commun. 2019, 10, 2779–2788. [Google Scholar] [CrossRef]
  13. Zhang, X.; Liu, H.; An, P.; Shi, Y.; Tang, Z. Delocalized electron effect on single metal sites in ultrathin conjugated microporous polymer nanosheets for boosting CO2 cycloaddition. Sci. Adv. 2020, 6, 4824–4833. [Google Scholar] [CrossRef]
  14. Takaishi, K.; Nath, B.; Yamada, Y.; Kosugi, H.; Ema, T. Unexpected Macrocyclic Multinuclear Zinc and Nickel Complexes that Function as Multitasking Catalysts for CO2 Fixations. Angew. Chem. Int. Ed. 2019, 58, 9984–9988. [Google Scholar] [CrossRef]
  15. Guo, Z.J.; Hu, Y.H.; Dong, S.; Chen, L.; Ma, L.; Zhou, Y.; Wang, L.; Wang, J. “Spring-loaded” mechanism for chemical fixation of carbon dioxide with epoxides. Chem. Catal. 2022, 2, 519–530. [Google Scholar] [CrossRef]
  16. Wilhelm, M.E.; Anthofer, M.H.; Cokoja, M.; Markovits, I.I.E.; Herrmann, W.A.; Kühn, F.E. Cycloaddition of Carbon Dioxide and Epoxides using Pentaerythritol and Halides as Dual Catalyst System. ChemSusChem 2014, 7, 1357–1360. [Google Scholar] [CrossRef] [PubMed]
  17. Guo, Z.; Jiang, Q.; Shi, Y.; Li, J.; Yang, X.; Hou, W.; Zhou, Y.; Wang, J. Tethering Dual Hydroxyls into Mesoporous Poly (ionic liquid)s for Chemical Fixation of CO2 at Ambient Conditions: A Combined Experimental and Theoretical Study. ACS Catal. 2017, 7, 6770–6780. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Chen, G.; Wu, L.; Liu, K.; Zhong, H.; Long, Z.; Tong, M.; Yang, Z.; Dai, S. Two-in-one: Construction of hydroxyl and imidazolium-bifunctionalized ionic networks in one-pot toward synergistic catalytic CO2 fixation. Chem. Commun. 2020, 56, 3309–3312. [Google Scholar] [CrossRef] [PubMed]
  19. Kuznetsova, S.A.; Rulev, Y.A.; Larionov, V.A.; Smol’yakov, A.F.; Zubavichus, Y.V.; Maleev, V.; Li, H.; North, M.; Saghyan, A.S.; Belokon, Y.N. Self-Assembled Ionic Composites of Negatively Charged Zn(salen) Complexes and Triphenylmethane Derived Polycations as Recyclable Catalysts for the Addition of Carbon Dioxide to Epoxides. ChemCatChem 2019, 11, 511–519. [Google Scholar] [CrossRef]
  20. Xiao, L.; Lai, Y.; Song, Q.; Cai, J.; Zhao, R.; Hou, L. POSS-based polyionic liquids for efficient CO2 cycloaddition reactions under solvent- and cocatalyst-free conditions at ambient pressure. Dalton Trans. 2023, 52, 9282–9293. [Google Scholar] [CrossRef] [PubMed]
  21. Chen, Y.; Li, Y.; Wang, H.; Chen, Z.; Lei, Y. Facile Construction of Carboxyl-Functionalized Ionic Polymer towards Synergistic Catalytic Cycloaddition of Carbon Dioxide into Cyclic Carbonates. Int. J. Mol. Sci. 2022, 23, 10879. [Google Scholar] [CrossRef] [PubMed]
  22. 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 Catalytic CO2 Fixation into Cyclic Carbonates under Mild Conditions. ACS Sustain. Chem. Eng. 2019, 7, 16907–16916. [Google Scholar] [CrossRef]
  23. Yamamoto, S.; Andersson, K.; Bluhm, H.; Ketteler, G.; Starr, D.E.; Schiros, T.; Ogasawara, H.; Pettersson, L.G.M.; Salmeron, M.; Nilsson, A. Hydroxyl-Induced Wetting of Metals by Water at Near-Ambient Conditions. J. Phys. Chem. C 2007, 111, 7848–7850. [Google Scholar] [CrossRef]
  24. Zhang, W.L.; Ma, F.P.; Ma, L.; Zhou, Y.; Wang, J. Imidazolium-functionalized ionic hypercrosslinked porous polymers for efficient synthesis of cyclic carbonates from simulated flue gas. ChemSusChem 2020, 19, 341–350. [Google Scholar] [CrossRef]
  25. Martín, C.; Fiorani, G.; Kleij, A. Recent Advances in the Catalytic Preparation of Cyclic Organic Carbonates. ACS Catal. 2015, 5, 1353–1370. [Google Scholar] [CrossRef]
  26. Shaikh, R.; Pornpraprom, S.; D’Elia, V. Catalytic Strategies for the Cycloaddition of Pure, Diluted, and Waste CO2 to Epoxides under Ambient Conditions. ACS Catal. 2018, 8, 419–450. [Google Scholar] [CrossRef]
  27. Ding, M.L.; Jiang, H.L. One-step assembly of a hierarchically porous phenolic resin-type polymer with high stability for CO2 capture and conversion. Chem. Commun. 2016, 52, 12294–12297. [Google Scholar] [CrossRef]
  28. Li, H.Y.; Meng, B.; Mahurin, S.M.; Chai, S.H.; Nelson, K.M.; Baker, D.C.; Liu, H.L.; Dai, S. Carbohydrate based hyper-crosslinked organic polymers with -OH functional groups for CO2 separation. J. Mater. Chem. A 2015, 3, 20913–20918. [Google Scholar] [CrossRef]
  29. Jia, S.H.; Ding, X.S.; Yu, H.T.; Han, B.H. Multi-hydroxyl-containing porous organic polymers based on phenol formaldehyde resin chemistry with high carbon dioxide capture capacity. RSC Adv. 2015, 5, 71095–71101. [Google Scholar] [CrossRef]
  30. Peng, X.; Chen, L.; Li, Y. Ordered macroporous MOF-based materials for catalysis. Mol. Catal. 2022, 529, 112568. [Google Scholar] [CrossRef]
  31. Mabaleha, S.S.; Gholizadeh, F.; Kalita, P. Recent advances in Ni-based stable catalysts for methane dry reforming: Stable catalysts’ preparation review. Mol. Catal. 2023, 547, 113398. [Google Scholar] [CrossRef]
  32. Liu, Y.; Li, S.; Yu, X.; Chen, Y.; Tang, X.; Hu, T.; Shi, L.; Pudukudy, M.; Shan, S.; Zhi, Y. Cellulose nanofibers (CNF) supported (Salen)Cr(III) composite as an efficient heterogeneous catalyst for CO2 cycloaddition. Mol. Catal. 2023, 547, 113344. [Google Scholar] [CrossRef]
  33. Dong, X.; Yuan, S.; Aizudin, M.; Wang, X.; Zhou, Y.; Song, H.; Yu, C.; Yuan, A.; Yang, F.; Ang, E.H. Heterogeneous Reductive C-N Coupling of Nitroarenes: Remarkable Activity Reconstruction of MoS2 Basal Plane by Gradient Oxygen-injection. Appl. Surf. Sci. 2023, 624, 157152. [Google Scholar] [CrossRef]
  34. Yang, F.; Lu, Y.; Liu, M.; Zhong, X.; Tu, W.; Zhang, W.; Zhu, C.; Guo, Z.; Yuan, A. Carbon-framework-encapsulated CoMn2O4 spinel derived from electrospun nanofiber coupling via the photothermal approach reinforces PMS activation to eliminate 2,4-dichlorophenol. Mater. Chem. Front. 2022, 6, 2810–2825. [Google Scholar] [CrossRef]
  35. Dong, X.; Yang, Y.; Shen, Y.; Yuan, A.; Guo, Z.; Song, H.; Yang, F. Enabling room-temperature reductive C−N coupling of Nitroarenes: Combining homogeneous with heterogeneous synergetic catalysis mediated by light. Green Chem. 2022, 24, 4012–4025. [Google Scholar] [CrossRef]
Figure 1. Synthesis of hydroxyl functionalized hyper-crosslinked bischloromethyl-based porous organic polymers.
Figure 1. Synthesis of hydroxyl functionalized hyper-crosslinked bischloromethyl-based porous organic polymers.
Catalysts 14 00707 g001
Figure 2. N2 adsorption isotherms (a) and BJH mesopore size distribution curves (b) of various HCPs-OH materials.
Figure 2. N2 adsorption isotherms (a) and BJH mesopore size distribution curves (b) of various HCPs-OH materials.
Catalysts 14 00707 g002aCatalysts 14 00707 g002b
Figure 3. (a) XRD patterns of HCPs-OH-3 and HCPs, (b) XPS surveys of HCPs-OH-3 and HCPs, (c) O1s XPS spectra of HCPs-OH-3 and HCPs, (d) C1s XPS of HCPs-OH-and HCPs.
Figure 3. (a) XRD patterns of HCPs-OH-3 and HCPs, (b) XPS surveys of HCPs-OH-3 and HCPs, (c) O1s XPS spectra of HCPs-OH-3 and HCPs, (d) C1s XPS of HCPs-OH-and HCPs.
Catalysts 14 00707 g003
Figure 4. SEM images (ac), TEM images (df) of HCPs-OH-3, (gi) Annular dark-field STEM image and corresponding elemental mapping images.
Figure 4. SEM images (ac), TEM images (df) of HCPs-OH-3, (gi) Annular dark-field STEM image and corresponding elemental mapping images.
Catalysts 14 00707 g004
Figure 5. (a) TG curve of HCPs-OH-3, (b) CO2 adsorption isotherms of HCPs-OH materials.
Figure 5. (a) TG curve of HCPs-OH-3, (b) CO2 adsorption isotherms of HCPs-OH materials.
Catalysts 14 00707 g005
Figure 6. (a) The reusability test of HCPs-OH-3 in cycloaddition of CO2 with epichlorohydrin. (b) Kinetic curves of the fresh and fifth recovered catalyst. Reaction conditions: epichlorohydrin 5 mmol, catalyst HCPs-OH-3 0.03 g, RT, 24 h, in the presence of 4.9 mol% n-Bu4NI as co-catalyst.
Figure 6. (a) The reusability test of HCPs-OH-3 in cycloaddition of CO2 with epichlorohydrin. (b) Kinetic curves of the fresh and fifth recovered catalyst. Reaction conditions: epichlorohydrin 5 mmol, catalyst HCPs-OH-3 0.03 g, RT, 24 h, in the presence of 4.9 mol% n-Bu4NI as co-catalyst.
Catalysts 14 00707 g006
Figure 7. The proposed possible reaction mechanism for cycloaddition of CO2 with epoxide by HCPs-OH-3.
Figure 7. The proposed possible reaction mechanism for cycloaddition of CO2 with epoxide by HCPs-OH-3.
Catalysts 14 00707 g007
Table 1. Textural properties of various acidic hypercrosslinked polymers a and conversions yields in the CO2 coupling with epichlorohydrinwith b.
Table 1. Textural properties of various acidic hypercrosslinked polymers a and conversions yields in the CO2 coupling with epichlorohydrinwith b.
EntryCatalystDCX (mmol)C%H%K cSBET d
(m2 g−1)
Smicro e (m2 g−1)Vp f
(cm3 g−1)
Dave g
(nm)
Con. (%)Sel. (%)
1HCPs-OH-1160.53.97.83004140.263.507099
2HCPs-OH-2262.34.42.95076030.524.087799
3HCPs-OH-3364.83.92.46798860.674.008299
4HCPs-OH-4464.54.12.05346450.564.207699
5HCPs-OH-5565.83.71.65826800.503.407599
6HCPs-OH-3364.83.92.46798860.674.0074 j99
7HCPs-OH-3364.83.92.46798860.674.000 k0
8HCPs-BA368.94.0-6257740.544.186699
9 hPR----- --6299
10 i1,3,5-THB----- --8399
11TBAI----- --2299
a Synthetic condition: the samples prepared in solvent of dichloroethane = 20 mL, 1,3,5 benzenetriols 5 mmol, DCX x mmol, 80 °C, 24 h. b Reaction conditions: epichlorohydrin 5 mmol, CO2 0.1 MPa (balloon), catalyst 0.03 g, n-Bu4NI 0.09 g, RT, reaction time (10 h). c The number of hydroxyl groups per 1 nm2 material surface, determination of hydroxyl content by acid–base titration, d BET surface area. e Total pore volume. Micropore surface area calculated from the nitrogen adsorption isotherm using the t-plot method. f Total pore volume. g Average pore size. h Phenolic Resin. i Homogeneous 1,3,5 phloroglucinol. j n-Bu4NBr 4.9 mol%. k absence of n-Bu4NI.
Table 2. Substrate scope with various epoxides catalyzed by HCPs-OH-3 a.
Table 2. Substrate scope with various epoxides catalyzed by HCPs-OH-3 a.
Catalysts 14 00707 i001
EntryEpoxide1Product 2t (h)Con. (%)Sel. (%)
1Catalysts 14 00707 i002Catalysts 14 00707 i0032499>99
2Catalysts 14 00707 i004Catalysts 14 00707 i00548/15 b87/81>99
3Catalysts 14 00707 i006Catalysts 14 00707 i00748/10 b95/86>99
4Catalysts 14 00707 i008Catalysts 14 00707 i0092495>99
5Catalysts 14 00707 i010Catalysts 14 00707 i011120/24 b93/96>99
a Reaction conditions: epoxides 5 mmol, catalyst HCPs-OH-3 0.03 g, RT, in the presence of 4.9 mol% n-Bu4NI as co-catalyst, b 70 °C.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, Z.; Ning, S.; Xu, S.; Zhang, Y.; Dong, Y.; Han, H. Tailoring Mesoporosity of Multi-Hydroxyls Hyper-Crosslinked Organic Polymers for Reinforced Ambient Chemical Fixation of CO2. Catalysts 2024, 14, 707. https://doi.org/10.3390/catal14100707

AMA Style

Guo Z, Ning S, Xu S, Zhang Y, Dong Y, Han H. Tailoring Mesoporosity of Multi-Hydroxyls Hyper-Crosslinked Organic Polymers for Reinforced Ambient Chemical Fixation of CO2. Catalysts. 2024; 14(10):707. https://doi.org/10.3390/catal14100707

Chicago/Turabian Style

Guo, Zengjing, Shuguang Ning, Shicheng Xu, Yongying Zhang, Yifan Dong, and Hongjing Han. 2024. "Tailoring Mesoporosity of Multi-Hydroxyls Hyper-Crosslinked Organic Polymers for Reinforced Ambient Chemical Fixation of CO2" Catalysts 14, no. 10: 707. https://doi.org/10.3390/catal14100707

APA Style

Guo, Z., Ning, S., Xu, S., Zhang, Y., Dong, Y., & Han, H. (2024). Tailoring Mesoporosity of Multi-Hydroxyls Hyper-Crosslinked Organic Polymers for Reinforced Ambient Chemical Fixation of CO2. Catalysts, 14(10), 707. https://doi.org/10.3390/catal14100707

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