Guoxin
Cui
a,
Yan
Li
a,
Guan
Yun
a,
Yongzheng
Zhao
a,
Peng
Gu
b,
Jing
Li
a,
Jinghan
Zhang
a,
Shuxin
Cui
a,
Minghao
Liu
a,
Weiqi
Zeng
a,
Zhenlu
Wang
c and
Jian
Jiang
*a
aHeilongjiang Key Laboratory of Photoelectric Functional Materials, College of Chemistry and Chemical Engineering, Mudanjiang Normal University, Mudanjiang, 157011, P. R. China. E-mail: jianjiang@mdjnu.edu.cn
bHeilongjiang Provincial Key Laboratory of Environmental Nanotechnology and Key Laboratory of Functional Inorganic Material Chemistry, Ministry of Education of the China, School of Chemistry and Materials Science, Heilongjiang University, Harbin, 150080, P. R. China
cInstitute of Physical Chemistry, College of Chemistry, Jilin University, 2519 Jiefang Road, Changchun, 130021, P. R. China
First published on 19th December 2024
A novel porous organic polymer (BL-TP-POP) containing β-ketoenamine and ketimine moieties was prepared through Schiff-base condensation, resulting in a framework with high stability and a large surface area. BL-TP-POP emitted fluorescence in the solid state, with a quantum efficiency of 3.56%. The material displayed a high carbon dioxide uptake capacity (up to 121 mg g−1) at 273 K and 1 bar, as well as a high CO2/N2 selectivity (53.5). BL-TP-POP also proved effective as a heterogeneous catalyst for Knoevenagel condensation, achieving a reaction yield of 98.6%.
Climate change has created an urgent need to separate and capture CO2 from emission sources, such as fuel gases. Several methods, including cryogenic distillation, membranes and adsorbents, have been developed to remove CO2.9 Solid adsorbents such as zeolites,10 metal oxides,11 activated carbons,12 metal–organic frameworks (MOFs),13 and POPs14 have been studied for CO2 capture applications.15 Currently, the use of POPs with high surface area and high stability plays an important role in the efficient capture and storage of CO2. EI-Kaderi and coworkers16 reported a triptycene-derived benzimidazole-linked POP (BILP-3) with a specific surface area of 1306 cm2 g−1, and a CO2 uptake capacity of 225 mg g−1 at 273 K and 1 bar. Yavuz and coworkers17 successfully developed an ethylenediamine-grafted porous organic polymer (COF-190L-en) with a CO2 uptake capacity of 99.44 mg g−1 at 298 K. COF-190L-en exhibited a high CO2/N2 selectivity of 729 (298 K), and its isosteric heat (Qst) was as high as 97 kJ mol−1 at zero coverage. It has been known that the presence of polar groups can enhance CO2 capture efficiency through interactions between polar groups and CO2via dipole–quadrupole interactions.18 For example, it has been reported that CO2-philicity can be enhanced by incorporating N-rich functional groups (such as amine,19 imine20 and azo21) into the backbone of POPs. Another strategy to increase CO2 uptake capacity involves introducing phenolic moieties into the polymer skeleton. Bhaumik and coworkers22 reported a porous organic polymer (TrzPOP-3) functionalized with phenolic –OH groups in which the CO2 uptake at 273 K and 1 bar pressure reached up to 375.76 mg g−1. Esteves and coworkers23 synthesized a series of COFs with different numbers of phenolic groups to study their CO2 uptake behaviours, which showed, experimentally and theoretically, that the relative number of phenolic groups acted as a threshold for CO2 uptake capacities of these COFs due to the interaction between CO2 molecules and the lone electron pair of oxygen atoms from hydroxyl groups. Therefore, the introduction of N-rich functional groups and phenolic groups can be an effective method for the rational design and development of potential CO2 adsorbents.
Porous organic polymers are also attractive candidates for catalysis due to their excellent properties.24 POPs function as solid catalysts or catalyst supports for several organic transformations, such as Knoevenagel reaction,25 aldol condensation,26 oxidation,27 hydrogenation,28 transesterification reaction,29 Friedel–Crafts alkylation,30 and epoxide ring-opening reaction.31 Knoevenagel reaction, the condensation between aldehydes or ketones and active methylene compounds, is an important reaction in organic chemistry that synthesizes heterocycles, intermediates and fine chemicals by forming new C–C bonds. Homogeneous catalysts for Knoevenagel condensation are highly active but difficult to separate. POPs have shown promising applications as heterogeneous catalysts for the Knoevenagel reaction because of their high stability, high surface area and structural tunability.
In this paper, we successfully synthesized a novel POP (BL-TP-POP) using 3,6-bis((E)-((4-aminophenyl)imino)(phenyl)methyl)-benzene-1,2-diol (BL) and 2,4,6-triformylphloroglucinol (TP) as building blocks (Scheme 1). BL-TP-POP was synthesized through Schiff-base condensation, followed by irreversible enol-to-keto tautomerization to form a β-ketoenamine structure. The building block BL was synthesized through ketimine condensation between 3,6-dibenzoylcatechol and p-phenylenediamine, which introduced phenolic and ketimine groups into the polymer. Therefore, both ketimine and ketoenamine moieties embedded in the backbone of BL-TP-POP. BL-TP-POP was characterized by 13C CP-MAS NMR, FT-IR, DRS, BET, SEM and HR-TEM. The material exhibited a high BET surface area of 611 m2 g−1, and the CO2 adsorption reached 121 mg g−1.
![]() | ||
Fig. 1 (a) FTIR spectra of BL, TP and BL-TP-POP. (b) 13C solid-state nuclear magnetic resonance spectrum of BL-TP-POP. |
The material was amorphous as indicated by the powder X-ray diffraction profile of BL-TP-POP (Fig. S6, ESI†) under this synthesis condition (Table S1, ESI†). The low crystallinity of the layer structure was believed to be caused by the nonplanar twisted geometry of building block BL (Fig. S3, ESI†) and sp3 nitrogen node in the keto-enamine moiety, which resulted in the flexibility and twisting of the layers. Field-emission scanning electron microscopy (FE-SEM) was performed to investigate the morphology of BL-TP-POP. As shown in Fig. 2a, SEM revealed that BL-TP-POP was mainly composed of spherical particles of different sizes from about 0.2 μm to 2 μm in diameter. Thermogravimetric analysis (TGA) indicated that BL-TP-POP possessed high thermal stability with minimal weight loss of up to 400 °C (Fig. 2b). Chemical stability was checked by immersing BL-TP-POP samples in THF, CH2Cl2, hexane, water, aqueous HCl (3 M) and NaOH (3 M) at room temperature. After 7 days, the FT-IR spectra of these immersed samples (except for the sample in NaOH) showed little change, indicating the unchanged composition. The stability test in NaOH showed retention of FT-IR peak positions after treatment for 1 day and became unstable afterwards (Fig. S20, ESI†). The exceptional thermal and chemical stability was attributed to the irreversible β-ketoenamine structure and stable ketoimine linkage.32 In addition, introducing –OH functionalities adjacent to the Schiff-base –CN– centres further enhanced the stability of the structure due to the formation of intramolecular O–H⋯N
C hydrogen bonding, which helped to protect the basic imine nitrogen from hydrolysis in the presence of water and acid.33
![]() | ||
Fig. 2 (a) SEM image of the BL-TP-POP bulk material. (b) TGA curve of BL-TP-POP. (c) Nitrogen sorption isotherm curve measured at 77 K. The insert shows the pore-size distribution for BL-TP-POP. |
The permanent porous structure of BL-TP-POP was investigated using its nitrogen sorption isotherm at 77 K. Nitrogen adsorption/desorption measurement revealed that the isotherm of the material fit the typical type I sorption model (Fig. 2c) and gave steep nitrogen uptake in the low-pressure range (P/P0 = 0–0.01), which suggested that the POP showed microporosity. The Brunauer–Emmett–Teller (BET) surface area and pore volume were calculated to be 611 m2 g−1 and 0.53 cm3 g−1, respectively. The pore size calculated using the quenched solid density functional theory (QSDFT) method was 0.93 nm.
UV-vis diffuse reflectance spectroscopy (DRS) was conducted to determine the electronic properties of BL-TP-POP. BL-TP-POP showed broad absorptions in the range of 300–550 nm (Fig. 3a). The corresponding band gap (Eg) was determined to be 2.09 eV by the Tauc plot. The semiconductor nature of the material was further investigated by cyclic voltammetry (CV) in a three-electrode system. As shown in Fig. 3b, the onset of reduction and oxidation potentials were −0.81 and 1.32 V (vs. AgCl/Ag), respectively. Accordingly, the HOMO and LUMO energies of BL-TP-POP were calculated to be −5.68 eV and −3.55 eV, respectively, and the band gap was inferred to be 2.13 eV, which is consistent with the value from DRS.
The optical property of the material was analysed using a fluorescence spectrometer. Under 380 nm light excitation, the BL-TP-POP powder exhibited yellow fluorescence at 580 nm (Fig. S17, ESI†) with a quantum efficiency of 3.56%, and the average lifetime was measured to be 2.24 ns (Fig. S18, ESI†). The solid powder of BL-TP-POP exhibited visible fluorescence upon 365 nm irradiation with a UV lamp. The high emission of the material in a solid state can be attributed to its nonplanar and twisted molecular structure, which avoids the stacking interaction of π–π during aggregation.
![]() | ||
Fig. 4 (a) CO2 and N2 isotherms for BL-TP-POP at 273 K and 298 K. (b) Isosteric heat of adsorption of CO2 for BL-TP-POP. |
To test the recyclability of the adsorbent, recycling experiments were performed and showed that BL-TP-POP could be reused without loss of adsorption capacity of CO2 after four cycles (Fig. S21, ESI†).
The binding sites of CO2 on BL-TP-POP (CO2@BL-TP-POP) were simulated with the Materials Studio Package. From the calculation with the sorption module, it was found that CO2 molecules preferred to adsorb near ketonic oxygens on β-ketoenamine moieties, and attraction around phenolic hydroxyl groups was less favoured (Fig. S26, ESI†). In addition, we found that CO2 molecules were reluctant to bind to nitrogen on the backbone. Therefore, only oxygen atoms were available for binding CO2 molecules. As shown in Fig. 5a, from the simulation, when the number of CO2 was more than the number of oxygen atoms in a unit cell, all oxygen atoms interacted with CO2, and excess CO2 molecules (shown by arrow) were distributed in the periphery of the frame, instead of around nitrogen atoms. To illuminate the selectivity of the binding sites, a Hirshfeld population analysis of the density functional electronic charge distribution of BL-TP-POP was performed with the DMol3 module to map the partial atomic charges. As depicted in Fig. 5b, the charges of ketonic oxygens (around −0.294) were more negative than those of oxygens on phenolic hydroxyl groups (−0.198). Higher electron density of ketonic oxygens facilitated the attraction of CO2 molecule whose carbon atoms showed a partial positive charge. The absolute values of the partial atomic charges of the nitrogen atoms were smaller than those of the oxygens; in particular, the nitrogen atoms on β-ketoenamine moieties exhibited a charge of −0.042, whose absolute value was only slightly higher than that of carbon on the phenyl rings (around −0.030). Although the nitrogen on ketoimines showed a higher negative charge (−0.106), the hydrogen bond formed with nitrogen (O–H⋯N) and steric hindrance due to the hydrogen atoms of adjacent groups prevented CO2 from accessing this binding site.
The selective uptakes of CO2 and N2 were studied based on the corresponding sorption isotherm curves (Fig. 4a). The selectivity of CO2/N2 was determined by Henry's law based on the initial slope calculation. The ideal adsorption selectivity of CO2/N2 was calculated to be 53.5 at 273 K and 41.4 at 298 K. The high selectivity could be attributed to abundant polar groups on the pore wall that were beneficial to bind CO2 with the quadrupole moment. The decrease in selectivity with an increase in temperature in the Henry regime could be attributed to the attenuation of the interaction between the host and guest molecules at a higher temperature.
Entry | Catalyst | Mass (mg) | Solvent | Time (h) | T (°C) | Yieldb (%) |
---|---|---|---|---|---|---|
a Reaction conditions: benzaldehyde (1 mmol), malononitrile (1 mmol), benzene (2 ml). b The yield of the product was determined by GC–MS using benzylidenemalononitrile as the external standard. The selectivity of the condensed product was greater than 99% in all cases. | ||||||
1 | — | — | Benzene | 8 | 80 | <1 |
2 | BL | 25 | Benzene | 8 | 80 | 41.4 |
3 | TP | 25 | Benzene | 8 | 80 | <1 |
4 | BL-TP-POP | 15 | Benzene | 8 | 80 | 7.5 |
5 | BL-TP-POP | 20 | Benzene | 8 | 80 | 82.5 |
6 | BL-TP-POP | 25 | Benzene | 8 | 80 | 98.6 |
7 | BL-TP-POP | 25 | THF | 8 | 66 | 84.4 |
8 | BL-TP-POP | 25 | EtOH | 8 | 78 | 76.7 |
9 | BL-TP-POP | 25 | Benzene | 6 | 80 | 78.9 |
10 | BL-TP-POP | 25 | Benzene | 8 | 60 | 43.6 |
With the increase in the loading amount of BL-TP-POP, the productivity of benzylidenemalononitrile was enhanced. It was found that the productive rate could reach up to 98.6% when 25 mg of BL-TP-POP was used to catalyse 1 mmol of benzaldehyde and 1 mmol of malononitrile at 80 °C for 8 h in benzene. Condensations in different solvents, such as THF and EtOH, were investigated, and experimental results showed that yields of benzylidenemalononitrile in these solvents were lower than those in benzene, which could be attributed to the solvation effect of the active sites by the polar solvent.38 The catalytic performances of the monomers were also evaluated. It was observed that TP exhibited poor catalytic activity and BL showed moderate yield, which further demonstrated that base groups played a key role in the catalysis. We noticed that the catalytic efficiency of BL was much lower than expected because there were more base groups in BL than in BL-TP-POP with the same weight. The low catalytic efficiency of BL could be attributed to its low solubility in benzene and the small surface area of the solid compound (Fig. S23, ESI†).
The investigation of the reusability of BL-TP-POP showed that the catalytic activity did not decrease considerably after repeating four times (Fig. 6a). The FT–IR spectra showed no significant difference between the prepared sample and the 4th cycle sample, indicating that BL-TP-POP was stable after 4 cycles (Fig. 6b).
![]() | ||
Fig. 6 (a) Recyclability study of the catalytic activities of BL-TP-POP. (b) IR spectra of as-synthesized BL-TP-POP after 4 cycles. |
X-ray powder diffraction (PXRD) patterns were recorded using a Rigaku Miniflex 600 diffractometer with Cu Kα radiation (λ = 1.54056 Å) over the 2θ range of 2.5–30° and using a SAXSess mc2 diffractometer with Cu Kα radiation (λ = 1.54056 Å) over the 2θ range of 1–5°. FT-IR spectra were measured using a PerkinElmer Frontier infrared spectrometer FTIR-650 ranging from 4000 to 400 cm−1. Thermogravimetric analysis was performed using a Netzsch TG 209 F1 Libra thermogravimetric analyzer with a ramp rate of 10 K min−1 under an atmosphere of nitrogen. A Shimadzu UV-2600 UV-vis spectrophotometer, equipped with integration sphere ISR-2200, was used to measure the UV-vis diffuse reflectance spectrum of solid powders at room temperature. SEM images were obtained using a JSM-6480LV at 5.0 kV. An FEI (Jeol FEG 2100F) high-resolution transmission electron microscope (HRTEM) equipped with a field emission source operating at 300 kV was used to record the TEM images. The nitrogen adsorption and desorption isotherms were measured at 77 K using an Autosorb-iQ (Quantachrome) surface area size analyzer. The Brunauer–Emmett–Teller (BET) method was utilized to calculate the specific surface area. CO2 adsorption measurements were performed at 273 and 298 K, and 1 bar. Before measurement, the samples were degassed in a vacuum at 120 °C for 6 h. 1H NMR spectra were recorded using a Bruker Avance NEO 400 MHz NMR spectrometer, where chemical shifts (δ in ppm) were determined with a residual proton of solvent as standard. Solid state cross polarization magic angle spinning (CP/MAS) NMR was performed using an AVIII 500 MHz solid-state NMR spectrometer. Electrical measurement was performed using a CHI 900C electrochemical workstation produced by Chenhua Instrument Co., Ltd. The MALDI-TOF mass spectrum was obtained using the Shimadzu MALDI-8020. Fluorescence spectra were measured using a Hitachi F-7000 Photoluminescence Spectrometer. The fluorescence lifetime and quantum yield were measured using EI FLS980 Fluorescence Spectrometers produced by Edinburgh Instruments. GC-MS measurement was performed using Agilent 7890A-5975C GC/MS equipped with an HP-5MS column with helium as the carry gas. The melting point was measured using an RY-1G melting point instrument.
Data for 1: 1H NMR (400 MHz, DMSO-d6): δ 7.66 (br, 4H), 7.55 (br, 10H), 7.43 (br, 6H), 7.33 (br, 2H).
Data for 2: 1H NMR (400 MHz, DMSO-d6): δ 7.82 (d, 4H), 7.77 (s, 2H), 7.66 (m, 6H), 7.56 (m, 6H), 7.36 (t, 4H).
Data for 3: 1H NMR (400 MHz, DMSO-d6): δ 10.35 (s, 2H), 7.79 (d, 4H), 7.68 (t, 2H), 7.57 (t, 4H), 6.95 (s, 2H).
Data for BL: 1H NMR (400 MHz, CDCl3): δ (ppm) 15.53 (s, 2H), 7.33 (m, 6H), 7.16 (m, 4H), 6.63 (d, 4H), 6.47 (d, 4H), 6.33 (s, 2H), 3.58 (s, 4H). IR (KBr): υ = 3476, 1613, 1586, 1518, 1468, 1431, 1255, 1155, 1081, 880, 834, 711, 537 cm−1. MALDI-TOF MS (dithranol matrix) m/z = 499.2 (M + H)+. Anal. calcd for C32H26N4O2: C (77.09%), H (5.26%), N (11.24%); found: C (77.28%), H (5.22%), N (11.15%). MP = dec. at 285 °C (13C NMR is not available for BL due to its low solubility).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nj04667f |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |