Download PDF
Research Article  |  Open Access  |  1 Sep 2024

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO2 capture and conversion via chemical and photocatalytic methods under ambient conditions

Views: 299 |  Downloads: 35 |  Cited:  0
Chem Synth 2024;4:50.
10.20517/cs.2024.05 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

Catalytic conversion of CO2 into valuable chemicals is a promising approach to mitigate the greenhouse effect and alleviate energy shortages. Hypercrosslinked polymers (HCPs) offer a scalable and stable platform for this conversion, but they often suffer from low CO2 adsorption and activation capabilities, necessitating high temperatures and pressures for effectiveness. To overcome these limitations, nitrogen-based CO2-philic active sites have been integrated into the structure of HCPs, enhancing CO2 attraction and leading to superior adsorption performance. The incorporation of cobalt ions further bolsters CO2 affinity, with HCP-PNTL-Co-B achieving the highest observed adsorption heat of 33.0 kJ·mol-1 alongside a substantial 2.0 mmol·g-1 CO2 uptake. These modified HCPs exhibit higher yields and reaction rates in cycloaddition reactions with cocatalyst tetrabutylammonium bromide at room temperature and atmospheric pressure, while HCP-1,10-phenanthroline (PNTL)-Co-B demonstrates a higher CO production rate (2,173 μmol·g-1·h-1) and selectivity (84%) in photocatalytic reduction reaction. This research has successfully achieved outstanding carbon dioxide capture and conversion performance at room temperature and atmospheric pressure by introducing CO2-philic active sites and cobalt ions into HCPs via facile one-step polymerization. This study provides a new method to design highly efficient organic catalysts for CO2 conversion.

Keywords

Hypercrosslinked polymers, N heteroatoms, CO2 cycloaddition, CO2 photoreduction

INTRODUCTION

The transformation of CO2 into value-added chemicals offers a dual benefit: it reduces greenhouse gas emissions and recycles carbon resources[1-6]. The domain of CO2 catalytic conversion has seen the emergence of numerous homogeneous and heterogeneous catalysts[7-11]. Heterogeneous catalysts, including carbon-based materials[12,13], silica[14], metal-organic frameworks (MOFs)[15-17], covalent organic frameworks (COFs)[18,19], and porous organic frameworks (POPs)[20-24], are preferred over their homogeneous counterparts due to their ease of separation, purification, and recyclability. Hypercrosslinked polymers (HCPs), as a subcategory of porous organic polymers, are particularly notable for their high surface area, diverse porosity, affordability, facile functionalization, and robust stability[25-31]. Moreover, HCPs are adept at incorporating catalytic sites, facilitating the creation of multiphase catalytic systems, which positions them as strong candidates for CO2 conversion applications[32-39].

At mild conditions, the effectiveness of CO2 conversion is closely tied to the material’s CO2 capture aptitude, with materials exhibiting low adsorption capacities facing limitations due to reduced CO2 concentrations at catalytic sites[40-42]. Unfortunately, most metal-based HCPs reported to date possess suboptimal CO2 adsorption capabilities, leading to inefficient utilization of CO2 under ambient conditions[30,32,37,43]. Beyond merely enhancing the CO2 adsorption capacity to improve conversion[32,44,45], scant research addresses the interaction between CO2 and the catalytic material. Integrating CO2-philic sites, such as nitrogen atoms or nitrogen-functional groups, can bolster the adsorption force for CO2, amplify adsorption capacity, and concentrate CO2 around catalytic sites[46-49]. This process not only increases the concentration of CO2 but also activates the inert CO2 molecules adsorbed on the surface, thereby improving the conversion efficiency. Nonetheless, the post-functionalization approach for introducing CO2-philic sites is often labor-intensive and may compromise the integrity of the porous structures of HCPs. It also suffers from the uneven distribution of CO2-philic sites[50-52]. Alternatively, synthesizing HCPs through hypercrosslinking polymerization with monomers already containing CO2-philic sites presents a solution, potentially enabling effective CO2 conversion in ambient conditions.

Motivated by the need to improve interactions with CO2 and stabilize metal ion catalytic sites, we utilized 1,10-phenanthroline (PNTL), a nitrogen-rich ligand known for its ability to enhance the binding with acidic CO2 molecules and coordinate with metal ions[53]. We synthesized PNTL-based HCPs through the Friedel-Crafts alkylation, employing derivatives of PNTL as functional monomers [Scheme 1]. This approach strategically increased the nitrogen content within the HCP framework, circumventing the issues of inhomogeneous CO2-philic site distribution often encountered with post-functionalization. Subsequently, we created a composite catalytic system, HCP-PNTL-Co, by introducing cobalt ions (Co2+) through coordination. This system was applied to the cycloaddition of CO2 with epoxides and the photoreduction of CO2 under ambient conditions. Among the different types, HCP-PNTL-Co-B, with the highest nitrogen content of 5.6 wt%, demonstrated superior CO2 adsorption, reaching 2.0 mmol/g at 273 K and 1 bar, and an adsorption heat of 33.0 kJ/mol. This material outperformed HCP-PNTL-Co-A, which had a lower nitrogen content of 2.9 wt%, showing increased catalytic activity and efficiency in both the cycloaddition and photoreduction reactions at ambient conditions. This study illustrates that the direct introduction of CO2-philic active sites into HCPs significantly augments both CO2 adsorption capacity and activation potential, avoiding the need for post-functionalization and ensuring a more uniform distribution of CO2-philic active sites, thus facilitating the catalytic conversion of CO2. These findings highlight the potential for efficient CO2 catalytic conversion, offering a promising approach to addressing the challenges of greenhouse gas emissions and energy scarcity and paving the way for developing and designing new strategies for efficient CO2 catalytic conversion systems.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Scheme 1. The synthetic pathway of HCP-PNTL and HCP-PNTL-Co porous network structures. HCP: Hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

EXPERIMENTAL

Materials

Acetonitrile (MeCN), [Gas Chromatography (GC)], 1,3,5-trimethylbenzene (GC), formaldehyde dimethyl acetal, (3,5-diphenylphenyl)boronic acid (98%), 3,8-dibromo-1,10-phenanthroline (98%), (4-(9H-carbazol-9-yl)phenyl)boronic acid (98%), tetrakis (triphenylphosphine) palladium (99%), cobalt chloride hexahydrate [Analytical Reagent (AR)], 4-dimethylaminopyridine (99%), tetraphenylphosphonium bromide (TPPB) (99%), epoxides (AR), triethanolamine (TEOA) (AR), and tris(2,2’-bipyridyl)ruthenium(II) chloride hexahydrate (98%) were purchased from Aladdin Chemical Reagent Corp. Tetrabutylammonium bromide (TBAB) (AR), dichloromethane (AR), 1,2-dichloroethane (AR), HCl (37%), 1,4-dioxane (AR), N,N-dimethylformamide (AR), ethylacetate (AR), tetrahydrofuran (AR), anhydrous potassium carbonate (AR), magnesium sulfate anhydrous (AR), iron(III) chloride anhydrous [Chemical Reagent (CR)], ethanol (AR), and methanol (AR) were provided from Sinopharm Chemical Reagent Co., Ltd.

Synthesis procedure

Synthesis of porous polymers

The HCP-PNTL-A[54] was synthesized by adding PNTL-A, formaldehyde dimethyl acetal, and 1,2-dichloroethane to a glass flask. Anhydrous FeCl3 was then introduced under a nitrogen atmosphere, followed by stirring at 45 °C for 5 h and 80 °C for 19 h. The resulting solid polymer was collected, washed with methanol and deionized (DI) water, and extracted with methanol using a Soxhlet apparatus. The dried polymer was obtained as a brown solid. The procedure for HCP-PNTL-B[55] was similar, replacing PNTL-A with PNTL-B.

Synthesis of metal-containing polymers

For HCP-PNTL-Co-A synthesis, a mixture of HCP-PNTL-A, cobalt chloride, and MeCN was stirred and refluxed at 65 °C for 12 h under nitrogen. After purification, the product was obtained as a solid. HCP-PNTL-Co-B was synthesized using the same method with HCP-PNTL-B instead of HCP-PNTL-A.

The yields of HCP-PNTL-A, HCP-PNTL-B, HCP-PNTL-Co-A, and HCP-PNTL-Co-B were 132%, 126%, 90%, and 85%, respectively.

RESULTS AND DISCUSSION

Structural and morphological characterizations

Confirmation of successful material synthesis was achieved through Fourier-transform infrared spectroscopy (FT-IR), solid-state 13C cross-polarization/magic angle spinning nuclear magnetic resonance (13C CP/MAS NMR), and elemental analysis (EA). The FT-IR spectra [Figure 1A and B] displayed characteristic absorption peaks at 1,597 and 1,250 cm-1, indicative of the stretching vibrations of C=N and C–N in the pyridine units, respectively. This confirms the stable incorporation of the monomer structure within the polymer framework. Additionally, the presence of characteristic aliphatic C–H stretching vibrations between 2,920-2,860 cm-1 signifies the successful execution of the Friedel-Crafts alkylation, indicating the establishment of the hypercrosslinked network. The 13C CP/MAS NMR spectra [Figure 1C] exhibit aromatic carbon peaks at 138 and 131 ppm, verifying the inclusion of aromatic monomers in forming the HCPs. The signal at 39 ppm is attributed to methylene carbons, further evidencing the completion of the Friedel-Crafts alkylation reaction[25], which agrees with the FT-IR results. EA data provided in Supplementary Table 1 reveals that the nitrogen content in HCP-PNTL-B is substantially higher than in HCP-PNTL-A. This is attributed to the higher nitrogen content in the monomers used to synthesize HCP-PNTL-B and aligns with the theoretical expectations. This finding also suggests that the nitrogen content in HCPs can be controlled effectively by varying the nitrogen heteroatom content in the functional monomers used.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 1. FT-IR spectra (A) and (B), solid-state 13C NMR spectra (C), TGA (D) (measured under N2 atmosphere), and solid-state UV-visible absorption spectra (E) and (F) of HCP-PNTL and HCP-PNTL-Co. FT-IR: Fourier-transform infrared spectroscopy; NMR: nuclear magnetic resonance; TGA: thermogravimetric analysis; UV: ultraviolet; HCP: hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

The HCPs are demonstrated to possess irregular layered structures by scanning (SEM) and transmission electron microscopy (TEM) images [Supplementary Figure 1]. A uniform distribution of carbon (C), nitrogen (N), and cobalt (Co) within the HCP-PNTL-Co samples is indicated by elemental mapping [Supplementary Figures 2 and 3]. The cobalt content in HCP-PNTL-Co-A and HCP-PNTL-Co-B was determined to be approximately 1.5 wt% and 1.3 wt%, respectively, by inductively coupled plasma optical emission spectrometry (ICP-OES) analysis. Thermogravimetric analysis (TGA) suggests that both HCP-PNTL and HCP-PNTL-Co display thermal stability up to 300 °C in a nitrogen atmosphere [Figure 1D]. The light absorption ability of the polymer was analyzed using ultraviolet (UV)-visible absorption, and Figure 1E and F displays the solid-state UV absorption spectra of HCP-PNTL and HCP-PNTL-Co. All polymers demonstrate strong light absorption in the visible light region, with absorption edges beyond 500 nm. The UV absorption spectrum of HCP-PNTL-Co shows a slight redshift compared to HCP-PNTL, indicating that the introduction of Co2+ metal enhances the light absorption ability of the material. X-ray photoelectron spectroscopy (XPS) results included in Figure 2, Supplementary Figures 4 and 5 provide detailed insights into the surface chemical composition and the valence states of the elements present in the HCPs. The high-resolution N1s spectrum for HCP-PNTL-A identifies pyridinic nitrogen (binding energy of 399.3 eV) as the dominant nitrogen species [Figure 2A]. In addition, the appearance of the Co-N characteristic peak [Figure 2B], Co 2p1/2 (798.2 eV), and Co 2p3/2 (782.2 eV) [Figure 2C] confirms the successful integration of Co2+ into HCP-PNTL-A. For HCP-PNTL-B, both pyridinic and carbazolic nitrogen forms are predominant [Figure 2D]; the appearance of the Co-N characteristic peak [Figure 2E] and Co 2p [Figure 2F] also confirms the successful integration of Co2+.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 2. XPS spectra for N1s of (A) HCP-PNTL-A, (B) HCP-PNTL-Co-A, (D) HCP-PNTL-B and (E) HCP-PNTL-Co-B; Co2p of (C) HCP-PNTL-Co-A and (F) HCP-PNTL-Co-B. XPS: X-ray photoelectron spectroscopy; HCP: hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

Porosity characteristics and gas adsorption capacities

The internal porosity of HCP-PNTL and HCP-PNTL-Co was assessed using nitrogen adsorption and desorption isotherms at 77.3 K. Figure 3A shows that the N2 uptake of HCP-PNTL-A and HCP-PNTL-B sharply increased at low relative pressure (P/P0 < 0.001), indicating their microporous nature. The significant desorption hysteresis at intermediate pressure suggests the presence of mesopores. Upon loading with cobalt metal ions, the nitrogen uptake of HCP-PNTL-Co was notably lower than that of HCP-PNTL [Figure 3A], indicating pore blockage. The Brunauer-Emmett-Teller (BET) surface areas were 951 m2·g-1 for HCP-PNTL-A and 475 m2·g-1 for HCP-PNTL-B, while they decreased to 609 m2·g-1 for HCP-PNTL-Co-A and 147 m2·g-1 for HCP-PNTL-Co-B [Supplementary Table 2]. Figure 3B depicts the pore size distribution calculated by nonlocal density functional theory (NLDFT), confirming the microporous and mesoporous structures of HCP-PNTL and verifying the blockage of some pores in HCP-PNTL-Co.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 3. (A) Nitrogen adsorption and desorption isotherms of HCP-PNTL and HCP-PNTL-Co; (B) Pore size distribution calculated using DFT methods of HCP-PNTL and HCP-PNTL-Co. HCP: Hypercrosslinked polymer; PNTL: 1,10-phenanthroline; DFT: density functional theory.

The CO2 gas adsorption capacity of phenanthroline-based HCPs was assessed based on their high specific surface area, microporosity, and abundant nitrogen atoms [Supplementary Table 2]. Figure 4A and B shows that HCP-PNTL-A has the highest CO2 adsorption capacity of 3.2 mmol·g-1 (1 bar, 273 K) and 1.9 mmol·g-1(1 bar, 298 K), attributed to its highest specific surface area and micropore volume. HCP-PNTL-B with lower specific surface area and micropore volume shows a lower CO2 uptake (2.2 mmol·g-1 at 1 bar/273 K and 1.5 mmol·g-1 at 1 bar/298 K) than HCP-PNTL-A. Upon the introduction of metal Co2+, the CO2 adsorption capacity of HCP-PNTL-Co decreases due to the reduction in specific surface area and micropore volume. The CO2 uptake of HCP-PNTL-Co-A decreased to 2.2 mmol·g-1 at 1 bar/273 K and 1.4 mmol·g-1 at 1 bar/298 K [Figure 4A and B]. HCP-PNTL-Co-B still shows high CO2 adsorption capacity (2.0 mmol·g-1 at 1 bar/273 K and 1.3 mmol·g-1 at 1 bar/298 K) even though the BET specific surface area is only 147 m2·g-1, which is due to the interaction between the material and CO2 enhanced by abundant N atoms and Co2+ in the structure. The above results indicate that the prepared HCPs rich in N atoms possess good CO2 adsorption performance, which is conducive to increasing the CO2 concentration near the catalytic sites and further promoting the CO2 catalytic conversion under the ambient conditions. In addition, there is little difference in the CO2 adsorption capacity between HCP-PNTL-Co-A and HCP-PNTL-Co-B, which is beneficial for comparing the subsequent CO2 catalytic activity and identifying the factors affecting the activity. The isosteric heats of adsorption (Qst) was calculated from CO2 adsorption isotherms at different temperatures (273 and 298 K) using the Clausius-Clapeyron equation to understand the adsorption properties between phenanthroline-based HCPs and CO2 molecules [Figure 4C]. The onset Qst value of HCP-PNTL-B (31.4 kJ·mol-1) was notably higher than that of HCP-PNTL-A (27.5 kJ·mol-1), suggesting a stronger interaction between CO2 and HCP-PNTL-B, which has a higher nitrogen content. This stronger interaction is beneficial for activating CO2 to promote subsequent conversion reactions. Additionally, the Qst value increased with the introduction of metal Co2+. The onset Qst values for HCP-PNTL-Co-A and HCP-PNTL-Co-B were 30.3 and 33.0 kJ·mol-1, respectively. This result also indicates that HCP-PNTL-Co-B has stronger adsorption force on CO2 and can activate inert CO2 molecules effectively, thus improving the catalytic efficiency of CO2.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 4. CO2 adsorption-desorption isotherms at (A) 273.15 K and (B) 298.15 K of HCP-PNTL and HCP-PNTL-Co; isosteric heat of adsorption (C) of HCP-PNTL and HCP-PNTL-Co. HCP: Hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

Catalytic performance of cycloaddition

The application of HCP-PNTL-Co in the CO2 cycloaddition reaction with epoxides was initially explored using propylene oxide (PO) as the substrate. As shown in Table 1, the reaction failed to occur in the absence of catalyst and cocatalyst (entry 1). Nevertheless, the conversion yields of PO were limited without the inclusion of catalyst (21%, Table 1, entry 2) or cocatalyst (7% or 9%, Table 1, entry 3 or entry 4), indicating that both catalyst and cocatalyst are indispensable in the cycloaddition reaction. The conversion yield of PO with HCP-PNTL-Co-A reached 75% in the presence of TBAB at 48 h, realizing a relatively efficient CO2 conversion under ambient conditions (Table 1, entry 5). In contrast, HCP-PNTL-Co-B displayed higher activity than HCP-PNTL-Co-A; a high yield of up to 95% was obtained in 48 h with the turnover number (TON) [turnover frequency (TOF)] of 5,380 (112 h-1). By comparing the activity of other catalysts in the table [Supplementary Table 3], it can be seen that HCP-PNTL-Co-B is also superior to most reported heterogeneous catalysts under mild conditions. It has significant advantages in TON (5,380) compared to some catalysts with high TON values, such as 2,2’-bipyridine zinc(II)-based hierarchical meso/microporous polymers (Bp-Zn@MA) (3,378)[56]; COF-salen-Co (3,744)[57]; P-POF-Zn (33,323)[58], etc. Although the specific surface area of HCP-PNTL-Co-B is smaller than that of HCP-PNTL-Co-A, HCP-PNTL-Co-B has a higher nitrogen content. Considering the small difference in the adsorption capacity of CO2 and Co2+ content between HCP-PNTL-Co-A and HCP-PNTL-Co-B, the higher catalytic performance of HCP-PNTL-Co-B may be due to the stronger adsorption of CO2 by HCP-PNTL-Co-B with higher nitrogen content, which concentrates CO2 around the catalytic sites[46-49]. This process not only increases the concentration of CO2 but also activates the inert CO2 molecules adsorbed on the surface, thereby improving the conversion efficiency[59]. It can more effectively activate the adsorbed CO2 molecules to excited states, making them easier to convert. At the same time, the above results also reveal that the effective activation of CO2 molecules is crucial for improving the efficiency of CO2 catalytic conversion. The possible mechanism of the HCP-PNTL-Co catalyzed CO2 cycloaddition is shown in Supplementary Figure 6.

Table 1

Catalytic coupling of CO2 and PO using various catalytic systemsa

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions
EntryCatalystTBABSubstrateTime (h)Yieldb (%)TON
(TOF·h-1)
1NoneNonePO480-
2None1.2 mmolPO4821-
3HCP-PNTL-Co-ANonePO487344 (7)
4HCP-PNTL-Co-BNonePO489663 (14)
5HCP-PNTL-Co-A1.2 mmolPO48753,681 (77)
6HCP-PNTL-Co-B1.2 mmolPO48955,380 (112)

Further investigations were conducted on the effects of reaction conditions, including reaction time, various cocatalysts, and the amount of cocatalyst TBAB, using HCP-PNTL-Co-B under mild conditions (0.1 MPa, 25 °C). The yield of propylene carbonate (PC) was only 39% within the first 12 h of reaction, despite the rapid conversion of PO during this period [Figure 5A]. Moreover, the yield of product gradually increased by further prolonging the reaction time, and the PC yield could reach 95% within 48 h. Subsequently, the effects of different cocatalysts on the PO conversion were investigated under the same conditions (0.1 MPa, 25 °C). Figure 5B showed that the conversion of PO was only 46% with N,N-Dimethylpyridin-4-amine (DMAP) as the cocatalyst, while the PC yield increased dramatically to 95% or 89% when TBAB or TPPB was employed as the cocatalyst. Since the highest yield was achieved when TBAB was used as a cocatalyst, it was selected as the optimal cocatalyst for the catalytic reaction. The effects of cocatalyst (TBAB) amount on the cycloaddition reactions were also evaluated using HCP-PNTL-Co-B [Figure 5C]. The yield of PC could still reach 75% when the amount of TBAB was only 0.6 mmol, illustrating the high catalytic activity of HCP-PNTL-Co-B for cycloaddition reactions. As the amount of TBAB increased, the PC yield also increased from 75% to 95%. The preceding results indicated the optimal reaction conditions for the catalytic system as follows: a reaction duration of 48 h, employing TBAB as a cocatalyst at a quantity of 1.2 mmol.

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 5. Influence of (A) reaction time, (B) different cocatalysts and (C) amount of TBAB on the catalytic performance of HCP-PNTL-Co-B; (D) Recyclability of HCP-PNTL-Co-B for the chemical conversion of CO2 under ambient conditions of room temperature and atmospheric pressure. TBAB: tetrabutylammonium bromide; HCP: hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

Reusability is a key factor in heterogeneous catalysis; the reusability and structural stability of HCP-PNTL-Co-B were evaluated. As shown in Figure 5D, HCP-PNTL-Co-B maintained the high efficiency after six cycles, which showed good cycling stability. The structural characterization of spent HCP-PNTL-Co-B revealed identical FT-IR spectra [Supplementary Figure 7] to that of fresh HCP-PNTL-Co-B, which confirms their excellent recycling catalytic behavior. Diversifying the range of substrates with HCP-PNTL-Co-B as the catalyst, the catalytic system demonstrated effectiveness with various terminal epoxides as well [Table 2]. Various epoxides were effectively transformed into their respective cyclic carbonates, yielding high percentages of 89%-95% and high TONs of 5,040-5,380. The more inert 2, 3-epoxide propyl butyl ether (entry 6) and 1, 2-epoxide tetradecane (entry 7) can also be efficiently converted with yields of 94% and 92%, respectively. These results show that HCP-PNTL-Co-B has good universality and can catalyze different types of epoxy substrates efficiently. The products of the CO2 conversion reaction were identified by 1H NMR spectra [Supplementary Figures 8-14].

Table 2

Different substituted epoxides were catalytically coupled with CO2 using HCP-PNTL-Co-B under ambient conditions of room temperature and atmospheric pressurea

EntryEpoxidesProductsTime (h)Yieldsb (%)TON
1Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48955,380
2Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48945,323
3Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48955,380
4Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48955,380
5Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48895,040
6Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48945,323
7Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditionsSynthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions48925,213

Photocatalytic performances for CO2 reduction

To investigate the photoelectric characteristics and applicability of HCP-PNTL and HCP-PNTL-Co for photocatalytic reduction of CO2, solid-state UV-vis spectroscopy measurement was performed. As illustrated in Figure 1E and F, all polymers exhibited a wide and strong visible-light absorption region up to approximately 500 nm, while the absorption edge of HCP-PNTL-Co had a slight redshift, indicating a slight enhancement of light absorption capability with the introduction of metal Co2+. The results of the UV-vis spectroscopy indicate that HCP-PNTL-Co has strong light absorption capacity, providing a guarantee for the photocatalytic CO2 reduction under visible light.

The photocatalytic performance of HCP-PNTL-Co was studied using [Ru(bpy)3]Cl2 as the photosensitizer and TEOA as the sacrificial agent in a mixed solvent of MeCN and H2O under 1 atm CO2 and visible light irradiation. Figure 6A illustrates the time-dependent yields of CO and H2 for HCP-PNTL-Co. HCP-PNTL-Co-B demonstrated the highest photocatalytic activity, with production rates of CO and H2 reaching 2,173 and 414 μmol·g-1·h-1 after three hours, respectively, resulting in a CO selectivity of 84%.The photocatalytic activity of HCP-PNTL-Co-A was slightly inferior to that of HCP-PNTL-Co-B, exhibiting lower production rates of CO and H2 of 1,726 and 491 μmol·g-1·h-1, respectively, and a CO selectivity of 78%. Given the slight variance in CO2 adsorption capability between HCP-PNTL-Co-A and HCP-PNTL-Co-B, the higher catalytic efficiency may be attributed to the stronger adsorption force between HCP-PNTL-Co-B with higher N contents and CO2, which can activate the adsorbed CO2 molecules to excited states more effectively, thereby facilitating their conversion. The results of CO2 photocatalytic reduction were consistent with the above cycloaddition reactions, indicating that HCP-PNTL-Co-A and HCP-PNTL-Co-B with abundant N heteroatoms have high CO2 photocatalytic reduction ability. Furthermore, this clearly demonstrates the potential of metal ion doping in HCPs for photocatalytic CO2 reduction and the effectiveness of introducing CO2-philic active sites to enhance CO2 catalytic conversion efficiency. The photocatalytic performance of HCP-PNTL-B modified with other metal ions such as Fe2+, Mn2+, or Ni2+ was evaluated under identical reaction conditions, and the chemical structure, thermal stability and porosity characteristics and gas adsorption capacities were performed [Supplementary Figures 15-20]. As shown in Figure 6B, HCP-PNTL-Ni-B exhibits the highest catalytic activity, producing a large amount of CO (2,761 μmol·g-1·h-1) and a small amount of H2 (307 μmol·g-1·h-1), with a high selectivity for CO production (90%). The highest catalytic activity and selectivity of HCP-PNTL-Ni-B can be attributed to the minimal energy barrier of nickel metal in the reaction process[60]. The catalytic performance of HCP-PNTL-Ni-B is comparable to that of the majority of previously reported metal-modified COFs/covalent triazine frameworks (CTFs)-based photocatalytic systems operating under similar reaction conditions [Supplementary Table 4]. The CO production rate of HCP-PNTL-Ni-B is 2,761 μmol·g-1·h-1, significantly higher than most other systems, such as Co@COF-TVBT-Bpy (1,133 μmol·g-1·h-1)[61], 3,3’,5,5’-Tetraformyl-4,4’-biphenyldiol (TFBD)-COF-Co-salicylideneaniline (SA) (1,480 μmol·g-1·h-1)[62], etc. The CO selectivity of HCP-PNTL-Ni-B is 90%, which is comparable to other systems, such as TFBD-COF-Co-SA (90%)[62], Ni-TpBpy[constructed from triformylphloroglucinol (Tp) and 2,2’-bipyridine (Bpy)]-COF (96%)[63], and Ni-COFs (95%)[64].

Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO<sub>2</sub> capture and conversion via chemical and photocatalytic methods under ambient conditions

Figure 6. The production rate of CO and H2 evolution over (A) HCP-PNTL-Co and (B) HCP-PNTL-M-B under visible-light irradiation; (C) The production rate of CO and H2 evolution under various conditions of HCP-PNTL-Ni-B and HCP-PNTL-Co-B; (D) Cyclic stability test of HCP-PNTL-Ni-B and HCP-PNTL-Co-B. HCP: Hypercrosslinked polymer; PNTL: 1,10-phenanthroline.

Control experiments were conducted to ascertain the influence of various reaction factors in the photocatalytic system [Figure 6C]. The production rate of CO was only 136 μmol·g-1·h-1 with HCP-PNTL-B as the catalyst for photocatalytic CO2 reduction, and the selectivity for CO was only 36.8%, indicating that metal Ni2+ or Co2+ is the catalytic active site for photocatalytic CO2 reduction. Under the condition of using metal Ni2+ or Co2+ alone as the catalyst, the production rate and selectivity of CO was only 683 μmol·g-1·h-1 and 78% or 533 μmol·g-1·h-1 and 73%, much lower than that of HCP-PNTL-Ni-B or HCP-PNTL-Co-B, indicating that the existence of HCP-PNTL-B in the catalytic system improved the CO2 adsorption and activation ability and the stable dispersion of Ni2+ or Co2+ catalytic sites, which can significantly enhance the photocatalytic CO2 reduction performance. When the photosensitizer was absent from the reaction system, only trace amounts of CO and H2 were detectable, while negligible amounts were obtained without a sacrificial agent or light irradiation during the photocatalytic reduction reaction. Substituting N2 for CO2 in the photocatalytic reaction resulted in a significant amount of H2 (492 μmol·g-1·h-1) but no detectable CO. The control experiments demonstrated the essential nature of all the previously mentioned reaction factors for the photocatalytic conversion of CO2. Lastly, the impact of solvents on the photocatalytic CO2 reduction performance of HCP-PNTL-Ni-B was examined. As shown in Supplementary Figure 21, HCP-PNTL-Ni-B exhibited the optimal photocatalytic activity in MeCN. The higher photocatalytic performance may be that, compared to other solvents, MeCN has a higher solubility for CO2, which is conducive to the increase of CO2 concentration near the catalytic sites. Similar reaction mechanisms for the catalytic reduction of CO2 by metal-based POPs photocatalytic systems under catalytic conditions have been reported in related studies. Based on literature reports[64,65] and control experimental results, the possible mechanism of HCP-PNTL-Ni-B catalyzing CO2 reduction was speculated based on the action of non-conjugated metal adjacent phenanthroline-based HCPs in providing metal catalytic sites and enhancing the CO2 concentration near the catalytic sites. First, the photosensitizer [Ru(bpy)3]Cl2 generates electron-hole pairs under visible light excitation. Subsequently, the excited state Ru(bpy)32+ is quenched back to the ground state by the sacrificial agent TEOA, and some electrons are transferred to HCP-PNTL-Ni-B, where the adsorbed CO2 accepts electrons and is reduced to CO.

To confirm the source of CO, isotope-labeled 13CO2 was utilized as the substrate for the photocatalytic reaction under identical conditions, and the resulting product was identified using gas chromatography-mass spectrometry (GC-MS). As shown in Supplementary Figure 22, the presence of a peak at m/z = 29 indicated that the generated 13CO originated from gaseous 13CO2 species rather than the decomposition of other organic compounds, such as the photocatalyst or solvent, in the photocatalytic system. It can be seen from Figure 6D that there was no obvious loss of the photocatalytic activity of HCP-PNTL-Ni-B or HCP-PNTL-Co-B after four-run cycling experiments, verifying its superior reusability in the photocatalytic reaction. Furthermore, negligible changes were observed in the FT-IR spectra [Supplementary Figure 23] of HCP-PNTL-Ni-B before and after the photocatalytic cycling reactions, confirming the structural stability and photocatalytic durability of HCP-PNTL-Ni-B in the photocatalytic CO2 reduction process.

CONCLUSIONS

In summary, PNTL derivatives were selected as functional monomers for the HCP synthesis via the Friedel-Crafts reaction. Subsequently, a composite catalytic system was devised by incorporating Co2+ for use in the cycloaddition reaction of CO2 with epoxides and the photoreduction of CO2 under ambient conditions. The nitrogen content in the structure of HCPs was enhanced by directly regulating the structure of the functional monomer to increase the CO2 adsorption capacity and activation ability, which can avoid the uneven distribution of CO2-philic active sites introduced by post-functionalization. The composite catalytic system with the highest nitrogen content (5.6 wt%), HCP-PNTL-Co-B, achieved CO2 adsorption capacity and heat of adsorption of 2.0 mmol/g (273 K, 1 bar) and 33.0 kJ·mol-1, respectively. In the presence of cocatalyst TBAB, HCP-PNTL-Co-B also showed higher catalytic activity, achieving efficient catalysis of the cycloaddition reaction of CO2 with epoxides under ambient pressure and temperature with a high yield of 95% after 48 h and a TON (TOF) value of up to 5,380 (112 h-1). Meanwhile, HCP-PNTL-Co-B also achieved a high CO production rate of 2,173 μmol·g-1·h-1 and a high selectivity of 84% in the photoreduction reaction. Moreover, the composite catalytic system exhibited good structural stability in the cycling experiments. The study has shown that introducing CO2-philic active sites improves the CO2 adsorption and activation capacity of materials, offering a novel research strategy for developing and designing efficient CO2 catalytic systems. Future research efforts can be conducted in-depth and systematically around (1) enhancing the conversion efficiency of materials for CO2 under mild conditions by adjusting the microstructure of catalysts or optimizing reaction conditions; (2) improving the structural stability of materials to ensure their efficient performance during cycling; (3) elucidating the catalytic mechanism and establishing the structure-performance relationship between polymer structure and catalytic properties.

DECLARATIONS

Acknowledgments

The authors thank the Analysis and Testing Center at Huazhong University of Science and Technology for their assistance in the characterization of materials.

Authors’ contributions

Made substantial contributions to conception and design of the study and performed data analysis and interpretation: Ouyang H, Peng M

Performed data acquisition and provided administrative, technical, and material support: Song K, Wang S, Gao H, Tan B

Availability of data and materials

The authors confirm that the data supporting the findings of this study are available within its Supplementary Materials.

Financial support and sponsorship

This work was financially supported by the National Natural Science Foundation of China (Grant Nos. 22161142005, 21975086), the International S&T Cooperation Program of China (Grant No. 2018YFE0117300), Science and Technology Department of Hubei Province (No. 2019CFA008), and the Open Research Fund (No. 2023JYBKF04) of Key Laboratory of Material Chemistry for Energy Conversion and Storage (HUST), Ministry of Education.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2024.

Supplementary Materials

REFERENCES

1. Ji G, Zhao Y, Liu Z. Design of porous organic polymer catalysts for transformation of carbon dioxide. Green Chem Eng 2022;3:96-110.

2. Truong CC, Mishra DK. Catalyst-free fixation of carbon dioxide into value-added chemicals: a review. Environ Chem Lett 2021;19:911-40.

3. Cheng Y, Ding X, Han B. Porous organic polymers for photocatalytic carbon dioxide reduction. ChemPhotoChem 2021;5:406-17.

4. Zhang XY, Wang P, Zhang Y, Cheng XM, Sun WY. Facet-dependent photocatalytic behavior of Fe-soc-MOF for carbon dioxide reduction. ACS Appl Mater Interfaces 2023;15:3348-56.

5. Steinlechner C, Junge H. Renewable methane generation from carbon dioxide and sunlight. Angew Chem Int Ed Engl 2018;57:44-5.

6. Liao X, Wang Z, Kong L, et al. Synergistic catalysis of hypercrosslinked ionic polymers with multi-ionic sites for conversion of CO2 to cyclic carbonates. Mol Catal 2023;535:112834.

7. Li PZ, Wang XJ, Liu J, Lim JS, Zou R, Zhao Y. A triazole-containing metal-organic framework as a highly effective and substrate size-dependent catalyst for CO2 conversion. J Am Chem Soc 2016;138:2142-5.

8. Truong CC, Mishra DK. Recent advances in the catalytic fixation of carbon dioxide to value-added chemicals over alkali metal salts. J CO2 Util 2020;41:101252.

9. Zhao Y, Han B, Liu Z. Ionic-liquid-catalyzed approaches under metal-free conditions. Acc Chem Res 2021;54:3172-90.

10. Chen Y, Mu T. Conversion of CO2 to value-added products mediated by ionic liquids. Green Chem 2019;21:2544-74.

11. Aomchad V, Del Gobbo S, Yingcharoen P, Poater A, D’elia V. Exploring the potential of group III salen complexes for the conversion of CO2 under ambient conditions. Catal Today 2021;375:324-34.

12. Lan D, Gong Y, Tan N, et al. Multi-functionalization of GO with multi-cationic ILs as high efficient metal-free catalyst for CO2 cycloaddition under mild conditions. Carbon 2018;127:245-54.

13. Cui X, Dai X, Surkus A, et al. Zinc single atoms on N-doped carbon: an efficient and stable catalyst for CO2 fixation and conversion. Chinese J Catal 2019;40:1679-85.

14. Jayakumar S, Li H, Tao L, et al. Cationic Zn-porphyrin immobilized in mesoporous silicas as bifunctional catalyst for CO2 cycloaddition reaction under cocatalyst free conditions. ACS Sustain Chem Eng 2018;6:9237-45.

15. Miralda CM, Macias EE, Zhu M, Ratnasamy P, Carreon MA. Zeolitic imidazole framework-8 catalysts in the conversion of CO2 to chloropropene carbonate. ACS Catal 2012;2:180-3.

16. Zhu M, Srinivas D, Bhogeswararao S, Ratnasamy P, Carreon MA. Catalytic activity of ZIF-8 in the synthesis of styrene carbonate from CO2 and styrene oxide. Catal Commun 2013;32:36-40.

17. Ding M, Jiang H. Incorporation of imidazolium-based poly(ionic liquid)s into a metal-organic framework for CO2 capture and conversion. ACS Catal 2018;8:3194-201.

18. Sun Q, Aguila B, Perman J, Nguyen N, Ma S. Flexibility matters: cooperative active sites in covalent organic framework and threaded ionic polymer. J Am Chem Soc 2016;138:15790-6.

19. Zhi Y, Shao P, Feng X, et al. Covalent organic frameworks: efficient, metal-free, heterogeneous organocatalysts for chemical fixation of CO2 under mild conditions. J Mater Chem A 2018;6:374-82.

20. Sun Q, Dai Z, Meng X, Xiao FS. Porous polymer catalysts with hierarchical structures. Chem Soc Rev 2015;44:6018-34.

21. Zhang Y, Riduan SN. Functional porous organic polymers for heterogeneous catalysis. Chem Soc Rev 2012;41:2083-94.

22. Ma D, Li J, Liu K, Li B, Li C, Shi Z. Di-ionic multifunctional porous organic frameworks for efficient CO2 fixation under mild and co-catalyst free conditions. Green Chem 2018;20:5285-91.

23. Liu J, Zhao G, Cheung O, Jia L, Sun Z, Zhang S. Highly porous metalloporphyrin covalent ionic frameworks with well-defined cooperative functional groups as excellent catalysts for CO2 cycloaddition. Chemistry 2019;25:9052-9.

24. Bhanja P, Modak A, Bhaumik A. Porous organic polymers for CO2 storage and conversion reactions. ChemCatChem 2019;11:244-57.

25. Li B, Gong R, Wang W, et al. A new strategy to microporous polymers: knitting rigid aromatic building blocks by external cross-linker. Macromolecules 2011;44:2410-4.

26. Gu Y, Son SU, Li T, Tan B. Low-cost hypercrosslinked polymers by direct knitting strategy for catalytic applications. Adv Funct Mater 2021;31:2008265.

27. Tan L, Tan B. Functionalized hierarchical porous polymeric monoliths as versatile platforms to support uniform and ultrafine metal nanoparticles for heterogeneous catalysis. Chem Eng J 2020;390:124485.

28. Gao TN, Wang T, Wu W, et al. Solvent-induced self-assembly strategy to synthesize well-defined hierarchically porous polymers. Adv Mater 2019;31:1806254.

29. Wang S, Zhang C, Shu Y, et al. Layered microporous polymers by solvent knitting method. Sci Adv 2017;3:e1602610.

30. Li J, Han Y, Ji T, et al. Porous metallosalen hypercrosslinked ionic polymers for cooperative CO2 cycloaddition conversion. Ind Eng Chem Res 2020;59:676-84.

31. Liao X, Wang Z, Li Z, et al. Tailoring hypercrosslinked ionic polymers with high ionic density for rapid conversion of CO2 into cyclic carbonates at low pressure. Chem Eng J 2023;471:144455.

32. Wang S, Song K, Zhang C, Shu Y, Li T, Tan B. A novel metalporphyrin-based microporous organic polymer with high CO2 uptake and efficient chemical conversion of CO2 under ambient conditions. J Mater Chem A 2017;5:1509-15.

33. Xu W, Chen M, Yang Y, et al. Construction of aluminum-porphyrin-based hypercrosslinked ionic polymers (HIPs) by direct knitting approach for CO2 capture and in-situ conversion to cyclic carbonates. ChemCatChem 2023;15:e202201441.

34. Kunitski M, Eicke N, Huber P, et al. Double-slit photoelectron interference in strong-field ionization of the neon dimer. Nat Commun 2019;10:1.

35. Zhan Z, Wang H, Huang Q, et al. Grafting hypercrosslinked polymers on TiO2 surface for anchoring ultrafine Pd nanoparticles: dramatically enhanced efficiency and selectivity toward photocatalytic reduction of CO2 to CH4. Small 2022;18:e2105083.

36. Manigrasso J, Chillón I, Genna V, et al. Author correction: visualizing group II intron dynamics between the first and second steps of splicing. Nat Commun 2022;13:1.

37. Chen Y, Luo R, Xu Q, Zhang W, Zhou X, Ji H. State-of-the-art aluminum porphyrin-based heterogeneous catalysts for the chemical fixation of CO2 into cyclic carbonates at ambient conditions. ChemCatChem 2017;9:767-73.

38. Yuan YC, Mellah M, Schulz E, David ORP. Making chiral salen complexes work with organocatalysts. Chem Rev 2022;122:8841-83.

39. Liao X, Xiang X, Wang Z, et al. A novel crosslinker for synthesizing hypercrosslinked ionic polymers containing activating groups as efficient catalysts for the CO2 cycloaddition reaction. Sustain Energy Fuels 2022;6:2846-57.

40. Li HB, Yu MH, Wang FX, et al. Amorphous nickel hydroxide nanospheres with ultrahigh capacitance and energy density as electrochemical pseudocapacitor materials. Nat Commun 2013;4:1894.

41. Li P, Wang X, Liu J, Phang HS, Li Y, Zhao Y. Highly effective carbon fixation via catalytic conversion of CO2 by an acylamide-containing metal-organic framework. Chem Mater 2017;29:9256-61.

42. Dai Z, Sun Q, Liu X, et al. Metalated porous porphyrin polymers as efficient heterogeneous catalysts for cycloaddition of epoxides with CO2 under ambient conditions. J Catal 2016;338:202-9.

43. Li J, Han Y, Lin H, et al. Cobalt-salen-based porous ionic polymer: the role of valence on cooperative conversion of CO2 to cyclic carbonate. ACS Appl Mater Interfaces 2020;12:609-18.

44. Xie Y, Liang J, Fu Y, et al. Hypercrosslinked mesoporous poly(ionic liquid)s with high ionic density for efficient CO2 capture and conversion into cyclic carbonates. J Mater Chem A 2018;6:6660-6.

45. Ma Y, Tang Q, Sun W, et al. Assembling ultrafine TiO2 nanoparticles on UiO-66 octahedrons to promote selective photocatalytic conversion of CO2 to CH4 at a low concentration. Appl Catal B Environ 2020;270:118856.

46. Liu M, Shao L, Huang J, Liu Y. O-containing hyper-cross-linked polymers and porous carbons for CO2 capture. Micropor Mesopor Mater 2018;264:104-11.

47. Liao C, Liang Z, Liu B, Chen H, Wang X, Li H. Phenylamino-, phenoxy-, and benzenesulfenyl-linked covalent triazine frameworks for CO2 capture. ACS Appl Nano Mater 2020;3:2889-98.

48. Shao L, Sang Y, Liu N, et al. Selectable microporous carbons derived from poplar wood by three preparation routes for CO2 capture. ACS Omega 2020;5:17450-62.

49. Shao L, Liu M, Sang Y, Zhan P, Chen J, Huang J. Nitrogen-doped ultrahigh microporous carbons derived from two nitrogen-containing post-cross-linked polymers for efficient CO2 capture. J Chem Eng Data 2020;65:2238-50.

50. Yang K, Yang Z, Zhang C, et al. Recent advances in CdS-based photocatalysts for CO2 photocatalytic conversion. Chem Eng J 2021;418:129344.

51. Shen Y, Zheng Q, Zhu H, Tu T. Hierarchical porous organometallic polymers fabricated by direct knitting: recyclable single-site catalysts with enhanced activity. Adv Mater 2020;32:e1905950.

52. Wen D, Zheng Q, Yang S, Zhu H, Tu T. Direct knitting boosts the stability and catalytic activity of NHC-Au complexes towards valorization of SO2 and CO2. J Catal 2023;418:64-9.

53. Nunes P, Correia I, Marques F, et al. Copper complexes with 1,10-phenanthroline derivatives: underlying factors affecting their cytotoxicity. Inorg Chem 2020;59:9116-34.

54. Pu Y, Harding RE, Stevenson SG, et al. Solution processable phosphorescent rhenium(i) dendrimers. J Mater Chem 2007;17:4255-64.

55. Ge Z, Hayakawa T, Ando S, et al. Synthesis and properties of 3,8-Bis[4-(9H-carbazol-9-yl)phenyl]-1,10-phenanthroline for phosphorescent OLEDs. Chem Lett 2008;37:262-3.

56. Chen J, Li H, Zhong M, Yang Q. Hierarchical mesoporous organic polymer with an intercalated metal complex for the efficient synthesis of cyclic carbonates from flue gas. Green Chem 2016;18:6493-500.

57. Li H, Feng X, Shao P, et al. Synthesis of covalent organic frameworks via in situ salen skeleton formation for catalytic applications. J Mater Chem A 2019;7:5482-92.

58. Chen J, Zhong M, Tao L, et al. The cooperation of porphyrin-based porous polymer and thermal-responsive ionic liquid for efficient CO2 cycloaddition reaction. Green Chem 2018;20:903-11.

59. Ouyang H, Song K, Du J, Zhan Z, Tan B. Creating chemisorption sites for enhanced CO2 chemical conversion activity through amine modification of metalloporphyrin-based hypercrosslinked polymers. Chem Eng J 2022;431:134326.

60. Dong Y, Liu H, Wang S, Guan G, Yang Q. Immobilizing isatin-schiff base complexes in NH2 -UiO-66 for highly photocatalytic CO2 reduction. ACS Catal 2023;13:2547-54.

61. Cui J, Fu Y, Meng B, et al. A novel cobalt-anchored covalent organic framework for photocatalytic conversion of CO2 into widely adjustable syngas. J Mater Chem A 2022;10:13418-27.

62. Yang Y, Lu Y, Zhang H, et al. Decoration of active sites in covalent-organic framework: an effective strategy of building efficient photocatalysis for CO2 reduction. ACS Sustain Chem Eng 2021;9:13376-84.

63. Zhong W, Sa R, Li L, et al. A covalent organic framework bearing single Ni sites as a synergistic photocatalyst for selective photoreduction of CO2 to CO. J Am Chem Soc 2019;141:7615-21.

64. Han B, Ou X, Zhong Z, Liang S, Deng H, Lin Z. Rational design of FeNi bimetal modified covalent organic frameworks for photoconversion of anthropogenic CO2 into widely tunable syngas. Small 2020;16:e2002985.

65. Hu X, Zheng L, Wang S, Wang X, Tan B. Integrating single Co sites into crystalline covalent triazine frameworks for photoreduction of CO2. Chem Commun 2022;58:8121-4.

Cite This Article

Research Article
Open Access
Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO2 capture and conversion via chemical and photocatalytic methods under ambient conditions
Huang Ouyang, ... Bien Tan

How to Cite

Ouyang, H.; Peng M.; Song K.; Wang S.; Gao H.; Tan B. Synthesis of metal-phenanthroline-modified hypercrosslinked polymer for enhanced CO2 capture and conversion via chemical and photocatalytic methods under ambient conditions. Chem. Synth. 2024, 4, 50. http://dx.doi.org/10.20517/cs.2024.05

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

Special Issue

This article belongs to the Special Issue Synthesis and Preparation of Novel Porous Organic Materials
© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Author Biographies

Huang Ouyang
Huang Ouyang received his bachelor's degree from Hubei University in 2016. In 2019, he joined Professor Bien Tan's group and is pursuing his Ph.D. degree in polymer chemistry and physics from Huazhong University of Science and Technology. His current scientific interests include the design and synthesis of hypercrosslinked polymers for CO2 conversion.
Mengjie Peng
Mengjie Peng obtained her Bachelor from South-Central Minzu University in 2018. In the same year, she entered Huazhong University of Science and Technology to pursue a Master's degree in Polymer Chemistry and Physics, under the supervision of Professor Bien Tan. Her primary research focuses on the design and synthesis of hypercrosslinked polymers and their applications in heterogeneous catalysis.
Kunpeng Song
Kunpeng Song obtained his bachelor from Henan Normal University in 2011. In 2016, he obtained his Ph.D. from Huazhong University of Science and Technology under the supervision of Professor Tao Li. He is currently employed at China West Normal University. His main research interests are microporous organic polymers and their derived carbon-based metal catalysts for CO2 catalytic conversion and hydrogen production through water electrolysis.
Shengyao Wang
Shengyao Wang obtained his bachelor's degree from Hunan Agricultural University in 2012. In 2017, he obtained his Ph.D. in engineering from Huazhong Agricultural University. He is currently a professor in the College of Science at Huazhong Agricultural University. His main research interests are the design of photocatalytic materials for environmental and energy applications and the study of catalytic mechanisms, including N2 fixation, CO2 reduction, CH4 activation, and NO removal.
Hui Gao
Hui Gao obtained her Ph.D. from the University of Liverpool in 2021. Subsequently, she conducted postdoctoral research at the University of Oxford. In 2023, she joined Professor Bien Tan's team at the School of Chemistry and Chemical Engineering at Huazhong University of Science and Technology . Her primary research focuses on the structural design of porous organic polymers and their applications in energy storage and gas adsorption.
Bien Tan
Bien Tan obtained his Bachelor from Hubei University in 1993 and his Ph.D. from South China University of Technology (SCUT) in 1997. From 1999 to 2001, he worked as a postdoctoral fellow at the National Defense Key Laboratory of Advanced Composite Materials at the Beijing Institute of Aeronautical Materials. In 2007, he began working in the Department of Chemistry at Huazhong University of Science and Technology. He currently holds the positions of Dean of the School of Chemistry and Chemical Engineering, Level 2 Professor, and doctoral supervisor at HUST. His research interests include the synthesis of organic porous polymer materials and their applications in energy and the environment.

Data & Comments

Data

Views
299
Downloads
35
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Chemical Synthesis
ISSN 2769-5247 (Online)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/