Rusong
Sang
ab,
Yezi
Hu
a,
Zewen
Shen
a,
Guixia
Zhao
*a,
Junrong
Yue
*b and
Xiubing
Huang
*c
aMOE Key Laboratory of Resources and Environmental System Optimization, College of Environmental Science and Engineering, North China Electric Power University, Beijing 102206, PR China. E-mail: guixiazhao@ncepu.edu.cn
bState Key Laboratory of Multi Phase Complex System, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, PR China. E-mail: jryue@ipe.ac.cn
cBeijing Advanced Innovation Center for Materials Genome Engineering, Beijing Key Laboratory of Function Materials for Molecule & Structure Construction, School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, PR China. E-mail: xiubinghuang@ustb.edu.cn
First published on 15th April 2024
In light of the widespread use of fossil fuels and the resulting environmental pollution, it is crucial to develop efficient photocatalysts for renewable energy applications that utilize visible light. Organic photocatalysts based on β-ketoenamine offer several advantages, including facile preparation, high stability, structural controllability, and excellent photovoltaic properties. However, in previous studies, the synthesis of porous organic polymers (POPs) often involved long, high-temperature processes. In this study, POPs with donor (D)–acceptor (A) structure were constructed by utilizing various branched bridging groups and 2,4,6-triformylphloroglucinol, across multiple temperature gradients. Through adjustments in hydrothermal temperature, we successfully synthesized a series of POPs with varying enol–keto structure ratios. Among these POPs, the dimethoxybenzidine-POPs (DMDPOPs) with methoxy electron-rich branched chains exhibited superior photovoltaic performance, electron transfer rate, and photocatalytic activity compared to the dihydroxybenzidine-POPs (DHDPOPs) with electron-deficient hydroxyl branched chains. Notably, DMDPOP-30 demonstrated outstanding performance, achieving a conversion rate of 98% within 3 h. Additionally, other POPs exhibited favorable conversions (90%), further confirming the feasibility of this synthetic approach. Moreover, the synthesis of DMDPOP-30 was achieved under mild conditions at room temperature, highlighting its significant potential for practical applications.
Organic photocatalysts featuring β-ketoenamine as their core have gained significant attention from researchers due to their ease of synthesis, high stability, tunable structure, and impressive photovoltaic properties.17–19 This structural motif has found widespread application in covalent organic frameworks (COFs), which are metal-free crystalline organic porous materials. Research on the photocatalytic capabilities of these COFs is growing, with interest in various applications such as photocatalytic organic degradation, CO2 reduction, hydrogen production, and organic conversion.20–24 However, the conventional synthesis of COFs typically involves solvothermal reactions at elevated temperatures (120–150 °C) often under vacuum or inert atmospheres.25,26 In light of the increasing emphasis on sustainable synthesis and cost-efficiency, there is a pressing need to identify catalysts that can be prepared under low temperatures and straightforward conditions.
Enhancing photocatalytic performance involves constructing a Donor (D)–Acceptor (A) structure within the photocatalyst, facilitating the spontaneous transfer of charges from D to A.27,28 Previous research has demonstrated that D–A structures effectively reduce the band gap, enhance charge transfer rates, and prolong carrier lifetimes.29–31 Additionally, incorporating π units can impact the electron acceptor domain, expediting charge transfer and the migration of photoexcited electrons.32,33 The aim of this study was to develop β-ketoenamine-based POPs with distinct electron-rich and electron-deficient branched π-units to enhance benzylamine oxidative coupling. Our approach involved modulating thermodynamic and kinetic factors, which represents a significant stride towards efficient and sustainable photocatalytic systems.
![]() | ||
Scheme 1 Designing POP structures with electron-rich π-electron (i.e., DMDPOP-X) and electron-deficient π-electron (DHDPOP-X) systems. |
XRD analysis was performed to investigate the crystallinity of different POP catalysts. In the case of DMDPOP-30, a prominent diffraction peak is observed at 2θ = 2.98°, corresponding to the (100) crystallographic plane (Fig. 1a). This indicates a high degree of crystallinity and a well-ordered structural sequence for efficient electron transport and transfer. The Materials Studio software was used to create a simulated material with an AA stacking structure. The lattice parameters were set to a = b = 30.8 Å, c = 3.48 Å, and α = β = 90°, γ = 120° to enable a detailed analysis of the crystal structure. Table S1† provides comprehensive crystal structure information and XRD pattern comparisons between experimental and simulated data, demonstrating their high consistency. The experimental peaks at 3.12° and 14.2° were matched with the simulated crystal faces, corresponding to the (100) and (210) planes, respectively. Generally, all POPs exhibit characteristic peaks around 5°, confirming their crystalline nature as polymers (Fig. 1a). Additionally, the presence of stable diffraction peak spectra suggests that all POPs possess a stable and ordered structure. The formation of packet peaks around 26°, attributed to the π–π stacking interaction, further indicates an organic polymeric structure formation. However, the presence of packet peaks also suggests that the POPs are not highly crystalline.
![]() | ||
Fig. 1 XRD patterns of DMDPOP-X and DHDPOP-X (a); N2 adsorption–desorption isotherms of DMDPOP-X and DHDPOP-X (b); pore size distribution of DMDPOP-X and DHDPOP-X (c). |
The Brunauer–Emmett–Teller (BET) specific surface areas of DMDPOP-X were 52.4, 57.1, 86.9, 214.7, and 331.8 m2 g−1, respectively (Fig. 1b). Synthesizing DMDPOP-X under different temperature gradients revealed a correlation between higher temperatures and larger specific surface areas. This suggests that thermodynamically driven synthesis of electron-rich branched POPs yields more microporous structures with increased active sites. The lower specific surface area of DMDPOP-30 could be attributed to the collapse of the pore structure caused by constant stirring during synthesis. The specific surface areas of DHDPOP-X with the electron-deficient hydroxyl group as the branched chain were 129, 110, 85, 115.3, and 101 m2 g−1, respectively (Fig. 1b). These values exhibited similar curves and a relatively balanced distribution of specific surface areas compared to those of POPs with electron-rich branched chains. This suggests that different branched chains influence the microstructure of POPs in terms of relative specific surface area and microporous structure.34 The synthesized series of POPs in this study exhibited lower specific surface areas. This was primarily due to constraints on crystal growth, such as synthesis temperature and reaction time. Lower synthesis temperatures and shorter reaction periods limited crystal growth, resulting in reduced surface areas. Furthermore, the specific surface areas of the materials were significantly impacted by their chemical composition and structural properties. The comparison between DMDPOPs and DHDPOPs illustrated this. To increase the surface area, it is recommended to use strategies such as raising the reaction temperature, extending the reaction time, selecting specific bridging groups, and performing post-synthesis steps such as washing, drying, and activation. These measures offer a viable pathway for enhancing the surface area of POPs. Additionally, the pore size distribution diagram indicates that the polymer's pore channels primarily consist of micropores and mesopores (Fig. 1c).
FTIR analysis provides further insights into the structure and chemical composition of DMDPOP-X and DHDPOP-X. Characteristic peaks at 1268, 1580 and 1612 cm−1 indicate the presence of C–N, CC, and C
O stretching bonds, confirming the existence of the keto form (Fig. 2a).35,36 Additionally, the bending vibration peak at 1297 cm−1 corresponding to C–OH suggests the presence of the enol form.37 The weakening of the N–H stretching band at 3500–3300 cm−1 indicates the condensation of amino groups in aniline as a bridging group. These signals confirm the successful synthesis of β-ketoenamine linkages and the formation of the fundamental bridging skeleton of POPs. The peak for stretching of aromatic C
C and C–H are at 1450 cm−1 and 798 cm−1, respectively.38 Furthermore, a slight vibration at 2885 cm−1 is observed, indicating the asymmetric stretching of the methoxy C–H bond in DMDPOP. DMDPOP-30 was synthesized at room temperature and exhibits remarkable catalytic properties, holding great potential for various applications. The solid-state 13C NMR spectrum of DMDPOP-30 is presented in Fig. 2b. Notably, the keto form displays a characteristic peak at 180.8 ppm, representing the typical C
O bond,38 while the enol form exhibits a characteristic peak at 175 ppm, indicating the presence of the structural C–OH bond.39,40 These findings further support the notion that the fundamental framework of POPs comprises the keto–enol form, which aligns with the results obtained from the FTIR analysis.
![]() | ||
Fig. 2 FT-IR spectra of DMDPOP-X and DHDPOP-X (a); 13C CP-MAS solid-state NMR spectra of DMDPOP-30 (b). |
XPS was employed to investigate the conformational ratio of keto to enol in the prepared DMDPOP-X samples. The β-ketoenamine POPs tested are composed of C, N and O elements as shown by the total XPS spectra (Fig. 3a). Fig. 3b–f illustrate the peaks observed at approximately 400.3 eV and 399.8 eV, which correspond to the keto form (–C–NH–) and the enol form (–CN–),33 respectively. By analyzing the peak intensities, it is possible to calculate the relative contents of the keto form and enol form in a semi-quantitative manner. Kinetic control during the room temperature stirring synthesis of DMDPOP-30 results in nearly equal proportions of the keto form and enol form (Fig. 3b). As the hydrothermal synthesis temperature increases from 60 °C to 120 °C, the proportion of the keto form in DMDPOPs gradually increases. However, at a hydrothermal temperature of 150 °C, the proportion of the keto form decreases and approaches equilibrium with the enol form (Fig. 3c–f). The C 1s XPS spectrum of DMDPOP-X exhibits six distinct peaks at specific binding energies: 284.6 eV (C
C), 285.2 eV (C–C), 285.8 eV (C
N), 286.2 eV (C–N), 287.0 eV (C
O), and 288.4 eV (C–O),26 as detailed in Fig. S1.† Furthermore, the O 1s peaks predominantly arise from methoxy branched chain C–O–C bonds at around 533.6 eV, enol–imine C–OH bonds at around 533.2 eV, and ketoenamine C
O bonds at around 530.7 eV (Fig. S2†).41 These observations suggest that the relative contents of the keto form and enol form are influenced by both thermodynamics and kinetics, eventually reaching a relative equilibrium state at high temperatures.
![]() | ||
Fig. 3 (a) Full XPS and high-resolution XPS N 1s spectra of DMDPOP-30 (b), DMDPOP-60 (c), DMDPOP-90 (d), DMDPOP-120 (e) and DMDPOP-150 (f). |
The surface morphologies of the β-ketoenamine POPs were examined using SEM. The results indicate distinct morphologies for different POPs. DMDPOP-30 exhibits a coral-like shape; DMDPOP-60 shows a stacked lamellar structure; DMDPOP-90 and DMDPOP-120 display a fine-fibrous state; and DMDPOP-150 exhibits a mixed growth of lamellar and fibrous structures (Fig. 4a–e). These observations suggest a correlation between the microstructure of electron-rich branched POPs during hydrothermal synthesis and the enol–keto structures. Specifically, the enol-form tends to adopt a lamellar structure, while the keto-form tends to exhibit a fibrous morphology (Fig. 4f–j). On the other hand, the morphologies of electron-deficient branched POPs are more similar, both of which have a fiber-interlaced microstructure. This similarity in microstructural features also corresponds to the close specific surface area results. The variance in microscopic morphology between the two types of POPs may be ascribed to the effect of the nature of the branched chains on π–π superpositions.
![]() | ||
Fig. 4 SEM images of DMDPOP-30 (a), DMDPOP-60 (b), DMDPOP-90 (c), DMDPOP-120 (d), DMDPOP-150 (e), DHDPOP-30 (f), DHDPOP-60 (g), DHDPOP-90 (h), DHDPOP-120 (i) and DHDPOP-150 (j). |
UV-vis diffuse reflectance spectroscopy (UV-vis DRS) is a technique used to assess the light absorption capacity of POPs. The absorption of visible light in the POPs is attributed to intramolecular charge transfer and the conjugated structure of π–π stacking.42,43 Both DMDPOP-X and DHDPOP-X exhibit significant absorption capabilities in the visible light range. Notably, DMDPOP-30 demonstrates exceptional absorption in the 500–800 nm wavelength range compared to other electron-rich branched POPs, enabling it to adapt to more complex lighting conditions (Fig. 5a and b). By employing the Kubelka–Munk function, the energy gaps (Eg) of DMDPOP-X are determined as follows: 1.83, 2.03, 2.12, 2.00, and 2.07 eV, individually. These results indicate that the keto-form in electron-rich branched POPs can widen the band gap. Conversely, the energy gaps of DHDPOP-X are 1.74, 1.87, 2.03, 1.83 and 2.06 eV, respectively (Fig. 5c and d). Comparatively, POPs with hydroxyl-deficient electron-branched chains exhibit narrower band gaps and superior absorbance, which can be attributed to keto–enol tautomerism and the electron push–pull effect. Upon analysis of the Tauc plots, a linear relationship between (αhν)2 and hν was observed, confirming that the POPs in question possess the characteristics of direct bandgap semiconductors.44 This property enhances the absorption of visible light, making them useful for applications in the fields of photovoltaics and photocatalysis. Fig. S3† shows the valence band top energies of the materials under vacuum conditions, which were determined through UPS spectra analysis. The valence band top energies for the DMDPOP-X and DHDPOP-X series are 4.41, 5.18, 4.39, 4.25, 4.86, 4.15, 4.43, 4.08, 4.96, and 4.66 eV, respectively. However, it is worth noting that UPS measurements are conducted in a high-vacuum environment to prevent interactions between electrons and air molecules. In contrast, electrochemical tests are performed in electrolyte solutions, which more closely simulate actual application conditions. Therefore, future studies will continue to use electrochemical methods to determine the band structure of these materials.
![]() | ||
Fig. 5 UV-vis spectra (a) and Tauc plots (b) of DMDPOP-X; UV-vis spectra (c) and Tauc plots(d) of DHDPOP-X; the transient photocurrent densities (e) and EIS spectra (f) of POPs. |
To investigate the impact of electron-rich units and electron-deficient units on the charge separation and migration, we examined the optoelectronic properties of DMDPOP-X and DHDPOP-X, among which, DMDPOP-X demonstrated superior photocurrent response and cyclic stability, indicating its excellent ability to generate and separate photoexcited electron–hole (e−–h+) pairs (Fig. 5e).29,30 Additionally, in electrochemical impedance spectroscopy (EIS), DMDPOP-X exhibited smaller semicircle diameters, suggesting a higher efficiency of charge migration (Fig. 5f). The aforementioned observations may be ascribed to the electron-donating impact of the methoxyl groups, which promotes intensive electron delocalization and facilitates charge migration.
To estimate the conduction band (CB) positions of POPs, electrochemical tests were conducted to obtain Mott–Schottky curves (Fig. S4†). All the curves exhibited positive slopes, indicating that the POPs are n-type semiconductors with flat band potentials (Vfb) relative to Ag/AgCl. The normalized hydrogen electrode (NHE) was calculated using the equation (ENHE = EAg/AgCl + 0.22). The CB levels of DMDPOP-X and DHDPOP-X were determined to be −0.15, −0.04, −0.02, −0.08, −0.12, −0.02, −0.04, −0.40, −0.05 and −0.22 V (vs. NHE), individually. The corresponding VBs were measured as 1.68, 1.99, 2.10, 1.92, 1.95, 1.72, 1.83, 1.63, 1.78, and 1.84 V (vs. NHE), as summarized in Fig. 6. The holes generated by DMDPOP-X and DHDPOP-X are sufficient for the oxidation of benzylamine.
Benzylamines find extensive applications as valuable intermediates in the pharmaceutical and biological fields. Photocatalytic selective oxidative coupling of Benzylamines represents a green and mild production method. In this study, we conducted photocatalytic experiments using DMDPOP-X and DHDPOP-X catalysts further to investigate their differences (Fig. 7a). The electron-rich branched DMDPOP-X catalyst exhibited a relatively higher average conversion compared to DHDPOP-X. Additionally, all POPs demonstrated a selectivity of over 97% in generating oxidative coupling products. This can be attributed to the superior photocurrent response and charge migration efficiency of the electron-rich branched POPs. Notably, DMDPOP-30, DMDPOP-60, and DMDPOP-90 displayed exceptional conversion rates of 98%, 90%, and 90%. The high conversion rates observed are independent of the particular surface area and pore configuration. From a chemical structure perspective, the conversion rates are related to keto–enol tautomerism and the electron push–pull effect. When the ratio of keto–enol tautomerism is more balanced or the keto-form dominates, it facilitates electron transfer, thereby promoting the reaction.
DMDPOP-30 is highly regarded for its mild synthesis conditions and exceptional photoelectric and catalytic properties. In the current study, the reaction kinetics were investigated using DMDPOP-30 as a catalyst. As depicted in Fig. 7b, the conversion consistently increases with reaction time, aligning with a level 0 kinetic model. Furthermore, Fig. 7c demonstrates that even after three cycles, DMDPOP-30 maintains a conversion rate of over 80%, indicating its commendable cycling performance and reaction stability. The FT-IR and SEM images post-cycling reveal minimal alterations in the chemical structure and microstructure compared to pre-cycling (Fig. S5 and S6†). In essence, the observations indicate that DMDPOP-30 exhibits commendable chemical stability. Additionally, the thermal stability of DMDPOP-30 is evident in Fig. S7,† where the catalyst experiences a weight loss of only approximately 10% when heated to around 300 °C, showing its favorable thermal stability. Considering its high crystallinity, mild synthesis conditions, excellent performance, good cycling performance, and thermal stability, DMDPOP-30 exhibits significant potential for various applications.
To elucidate the mechanism of selective oxidation using DMDPOP-30, a series of comparative experiments were conducted (Fig. 7d). Under dark conditions, the conversion of the reaction significantly decreased, highlighting the importance of light in the reaction. The conversion of 60% in the absence of oxygen implies that oxygen is not a prerequisite for the reaction to occur. To investigate the role of reactive oxygen species (ROS), various agents were added to the reaction. The addition of 1,4-benzoquinone (BQ, O2˙− scavenger), TEMPO (1O2 scavenger), and tertbutyl alcohol (TBA, ˙OH scavenger) did not effectively inhibit the reaction, suggesting that reactive oxygen species do not primarily drive the reaction. However, the addition of AgNO3 and KI, which act as photogenerated electron and hole scavengers respectively, effectively inhibited the reaction. Fig. 8 illustrates the proposed mechanism of selective oxidative coupling of benzylamine by DMDPOP-30 photocatalyst. Within the valence band (VB), benzylamine undergoes oxidation, leading to the formation of nitrogen-centered radical cations. Through a proton transfer process, the hydrogen protons within these nitrogen-centered radical cations readily undergo cleavage, resulting in the generation of carbon-centered radicals. These carbon-centered radicals then react with the existing nitrogen-centered free radicals, ultimately yielding N-benzylidene benzylamines.
To expand the application potential of DMDPOP-30, the oxidation of other amines with diverse substituents was also conducted. The results demonstrate excellent conversion for all target substrates, indicating a consistent pattern across different substituents. Particularly, the para-methoxy substituent exhibits superior enhancement to the photocatalytic performance compared to the neighboring methoxy substituent, attributed to the spatial site resistance effect (Table 1, entries 1 and 2). Furthermore, the conversion rate increases with the electronegativity of halogen substituents, owing to the varying electron density electronic effect of different substituents on the benzyl group (Table 1, entries 3–5). Additionally, amines containing heteroatom N also display enhanced conversion (Table 1, entry 6), further emphasizing the universal efficacy of DMDPOP in the photocatalytic oxidative coupling process of diverse amines.
Entry | Substrate | T (h) | Conversion (%) | Selectivity (%) |
---|---|---|---|---|
Reaction conditions: DMDPOP-30 (5 mg), substituted benzylamine (0.2 mmol), blue LEDs (460 ± 15 nm), 1 atm O2, CH3CN (1 mL). | ||||
1 | 4-Methoxybenzylamine | 3.5 | 98 | 98 |
2 | 2-Methoxybenzylamine | 3.5 | 92 | 97 |
3 | 4-Methylbenzylamine | 3.5 | 99 | 97 |
4 | 4-Fluorobenzylamine | 3.5 | 99 | 99 |
5 | 4-Bromobenzylamine | 3.5 | 92 | 98 |
6 | 2-Picolylamine | 3.5 | 84 | 99 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nr00391h |
This journal is © The Royal Society of Chemistry 2024 |