Guishan
Hu
a,
Binyu
Zhou
a,
Yu
Zhen
a,
Junyong
Zhu
*a,
Jingwei
Hou
b,
Yatao
Zhang
*a,
Yong
Wang
*c and
Bart
Van der Bruggend
d
aSchool of Chemical Engineering, Zhengzhou University, Zhengzhou, 450001, P. R. China. E-mail: zhujunyong@zzu.edu.cn; zhangyatao@zzu.edu.cn
bSchool of Chemical Engineering, The University of Queensland, St Lucia, QLD 4072, Australia
cSchool of Energy and Environment, Southeast University, Nanjing, 210096, P. R. China. E-mail: yongwang@seu.edu.cn
dDepartment of Chemical Engineering, KU Leuven, Celestijnenlaan 200F, B-3001, Leuven, Belgium
First published on 2nd September 2025
Heterogeneous catalysis has emerged as a promising approach for sustainable and efficient water purification due to its high catalytic efficacy and low energy consumption. However, the complex nature of nontraditional water resources (e.g., trace toxic metals, organic molecules, and high salinity) leads to gradual catalyst deactivation, hindering large-scale implementation. Here we introduce the concept of sieving-coupled nanoconfined catalytic water pollutant conversion for rapid purification of complex water sources. This dual-function system was achieved by integrating in situ recycled palladium nanoclusters within covalent organic framework (COF) membranes. The strong interaction between Pd and the pyrazine nitrogen in the COF facilitates the formation of a Pd layer within a 40 nm-thick COF nanofilm. The resultant Pd0–TpPz membrane exhibited a high permeability of 85.4 L m−2 h−1 bar−1 while achieving 99.8% for Eriochrome black T. This precise-sieving effect of the membrane enables the efficient catalytic reduction of various pollutants such as rhodamine B (RhB), Cr(VI), and 4-nitroaniline in complex systems. The reported Pd0–TpPz membrane evinced favorable long-term stability, recyclability, antimicrobial activity, and acid–base resistance (pH = 2–12), demonstrating its high potential for water treatment. This work paves the way towards the development of sieving-coupled nanoconfined catalysis for rapid water purification.
Membrane-based technologies play a key role in water treatment and purification primarily due to their high selectivity, low cost, and broad applicability.10–13 Dense ultrafiltration and loose nanofiltration are among the most common membrane-based processes utilized for the fractionation of diverse solutes with varying molecular/ionic sizes.14,15 However, these approaches are incapable of retaining small toxic molecules and high-valent metals (e.g., 4-NP and CrVI), necessitating additional costly separation units to remove these contaminants.16,17 The widespread large-scale application of these technologies is further hindered by the limited physical and chemical tunability of the pore structure in conventional polymeric membranes. Of note, the empirical-dominated processing of common linear polymers frequently yields high packing-density membranes with limited free volume and dynamic porosity, leading to a trade-off between flux and selectivity.18 Rapid interfacial polymerization of two highly reactive monomers (diamine and triacyl chloride) is effective in fabricating thin selective films but lacks fine control of the membrane microstructure.19 The resultant wide pore size distribution limits membrane performance and increases the cost of water treatment. While substantial improvements in water purification membranes have been achieved using optimized film processing strategies, their surface fouling propensity, especially biofouling, is often overlooked.20–22 Once a mature biofilm forms on the membrane surface, there is a significant decline in membrane performance and lifespan, hindering its practical and large-scale use.23 In this context, new membrane materials fabricated through molecular-level design methods are crucial for advancing membrane-based water purification technologies.
Unlike traditional linear polymers, covalent organic frameworks (COFs) have emerged as crystalline porous polymers that feature structural periodicity, well-arranged nanopores, high surface area, and robust water stability.24–26 These properties, combined with their designable pore architectures and tunable apertures, make them promising candidates for selective adsorption and separation applications with a level of precision that is not achievable using conventional materials.27–29 Furthermore, the unique COF architecture provides confined spaces for the interplay of ions or guest molecules, thus evincing unique properties and functions. Their ordered nanopores offer stable anchoring sites for metal catalysts, ensuring nanoconfined catalytic conversion of small molecules and ions.30,31 In addition, by tailoring their pore size and functional groups, the dispersion and catalytic activity of metal catalysts can be optimized, thereby enhancing the selectivity and efficiency of the reaction.32 Importantly, the diversity of building blocks and covalent linkages makes COFs highly promising for selective capture of precious metal catalysts from nontraditional water sources.33 This unique function enables the usage of precious metal-containing wastewater to upcycle COF catalytic membranes with recycled metal nano-catalysts.34 More importantly, pore engineering of COF membranes allows them to precisely sieve the organics with large molecular weights, facilitating continuous catalytic conversion of toxic ions and small molecules that can enter the nanopores.35 Therefore, COFs offer a promising opportunity to leverage precise sieving and heterogeneous catalysis for potentially improved water purification efficiency.36,37
To this end, we sought to develop a separation-synergistic catalytic membrane platform for use in multifunctional water-purification applications, based on the introduction of recycled Pd nanoclusters into COF separation membranes. Unlike heterogeneous catalysis using 2D laminar membranes, which may suffer from the fouling of catalytic sites, our proposed nanoconfined membrane reaction platform fully unlocks the advantages of ordered nanopores and chemical tunability. This new mode enables simultaneous rejection of large organic dyes and catalytic conversion of small organics or toxic high-valence metals when dealing with complex wastewater (Fig. 1a). A TpPz-COF membrane was fabricated in situ on a Kevlar substrate via an IP reaction between 1,3,5-triformylphloroglucinol (Tp) and 1,4-diaminopyrazine (Pz). This pyrazine-based COF membrane exhibited the highest selectivity for the capture of PdCl42− from simulated electronic wastewater containing diverse competing ions (Fig. 1b). This is likely due to the high affinity between palladium and dual-nitrogen atoms from the pyrazine-based skeleton. The subsequent in situ reduction feasibly transformed the adsorptive membrane into a catalytic COF membrane rich in recycled Pd nanoclusters, without diminishing its original filtration performance. This Pd-appended COF membrane not only exhibited high removal efficiency of large dye molecules (e.g., Congo red, methyl blue, and Direct Red 23) but also enabled catalytic reduction of toxic molecules and high-valence metals (rhodamine B, 4-nitroaniline, methyl orange, 4-nitrophenol, and CrVI, Fig. 1c). By combining robust long-term stability, recyclability, antimicrobial activity, and acid–base resistance (pH 2–12), this membrane shows great promise for real-world water purification applications.
Fourier transform-infrared (FTIR) spectra of organic linkers and TpPz (Fig. 2c) reveal the successful condensation between Tp and Pz linkers. The disappearance of the aldehyde (1633 cm−1) and the amine (3259 cm−1) peaks, coupled with the emergence of a new peak at 1578 cm−1 CC stretching, provides strong evidence for the formation of a β-ketoenamine-linked COF film. Solid-state 13C CP-MAS NMR spectroscopy (Fig. 2d) further elucidates the atomic-level structure of the TpPz nanofilm. The signal at 184 ppm is indicative of keto C
O bond formation, likely resulting from tautomerization during the reaction. The signal at 108 ppm corroborates the presence of C
C–N bonds and further supports the keto structure. In addition, the peak that emerged in the range of 144–153 ppm for the TpPz film is assigned to the C–NH (nonheterocyclic N) carbon, while the signal at ∼152 ppm corresponds to the C–N (heterocyclic N) carbon.39
Nitrogen sorption isotherms at 77 K confirm the permanent porosity of TpPz and Pd0–TpPz films. TpPz exhibits a BET surface area of 87.4 m2 g−1 and a primary pore size of 2.05 nm, consistent with an AA-eclipsed framework geometry (Fig. 2e and f). Pd0–TpPz shows a reduced BET surface area (48.3 m2 g−1) and a smaller pore size (1.83 nm), likely due to the significant adsorption of Pd within the pores. In contrast, Pd0–TpPz displayed a primary pore size of 1.83 nm, likely attributed to the significant adsorption of Pd within the TpPz pores, resulting in a decline in both pore size and BET surface area.40 Free-standing COF nanofilms were transferred onto PAN substrates and silicon wafers to gain insight into their internal microstructure. Scanning electron microscopy (SEM) imaging reveals a uniform surface morphology for the TpPz nanofilm (Fig. 2g). Transmission electron microscopy (TEM) images demonstrate the regular structure of the TpPz nanofilm, while energy-dispersive X-ray spectroscopy (EDS) mapping confirms a uniform distribution of carbon, nitrogen, and oxygen elements (Fig. 2h). Additionally, atomic force microscopy (AFM) height profile measurements evince a thickness of 39.3 nm for the TpPz nanofilm (Fig. S4).
The surface morphology of the as-fabricated membranes was examined using SEM images. Compared with the Kevlar substrate, the TpPz membranes demonstrated a more compact and uniform surface morphology without visible defects (Fig. S7a and S8a). AFM analysis further revealed a uniform TpPz surface with a low roughness (Ra = 15.0 nm, Fig. S7b and S8b). Cross-sectional TEM images distinctly depicted a 44.4 nm-thick nanofilm (Fig. S9a and S10a–b), consistent with the thickness observed for the freestanding film by AFM analysis. The well-defined boundaries and an even elemental distribution confirm the formation of a uniform COF nanofilm at the top surface of the Kevlar support (Fig. S10c).
Given the strong affinity of N-heterocyclic groups to Pd, TpPz membranes were explored for Pd capture. A noticeable color change from dark red to bright red for the TpPz surface confirmed Pd uptake from acidic leaching liquor. Leveraging the strong binding interaction between pyrazine groups and PdClnX−, TpPz membranes with rich nanochannels were explored as templates for Pd nanocluster growth. In situ reduction with NaBH4 yielded brown Pd0–TpPz membranes with a uniform, low-roughness (21.2 nm, Fig. S8c and d) surface. Cross-sectional TEM images of the Pd0–TpPz membranes revealed a 40.7 nm-thick film with highly dispersed nanoclusters (Fig. 2i). High-resolution TEM confirms the (111) crystal plane of the Pd nanoclusters with a crystal plane spacing of 0.225 nm (Fig. S9b). EDS mapping further confirmed their presence once again. These Pd nanoclusters evinced an average size of 1.96 nm in diameter with a narrow size distribution (Fig. 2j and S10d–e).
X-ray photoelectron spectroscopy (XPS) was used to determine the elemental composition and oxidation state of Pd within the TpPz membranes (Fig. 2k–l and S11–S13). XPS data revealed a Pd loading of 4.0% in the Pd0–TpPz membrane (Table S1). Deconvolution of the C 1s, N 1s, and O 1s narrow scan spectra confirmed the successful formation of keto-enamine based TpPz (Fig. S12a–c). For Pd0–TpPz membranes, the Pd 3d region displayed new characteristic peaks at 336.1 eV and 341.4 eV (Fig. 2k), corresponding to the 3d5/2 and 3d3/2 states of zero-valent Pd nanoclusters. However, weak peaks at 337.2 eV and 342.1 eV corresponding to the 3d5/2 and 3d3/2 states of PdII and PdO appeared, suggesting incomplete reduction and surface oxidation. This interpretation is further supported by the higher oxygen content observed in Pd0–TpPz compared to TpPz. Thermogravimetric analysis (TGA) revealed high thermal stability for all membranes, with Pd0–TpPz showing a higher residual weight (∼15%) due to the presence of Pd (Fig. S14). Zeta potential measurements (Fig. S15) revealed the less electronegativity of the Pd0–TpPz membrane, primarily due to the abundant nitrogen content and the presence of reduced Pd species. As shown in Fig. S16, the hydrogel substrate had the lowest WCA and the highest surface energy due to its high hydration and water content. As the TpPz layer formed on the Kevlar surface, the WCA gradually increased. In contrast, the introduction of Pd nanoclusters increased the surface energy of the TpPz membranes, resulting in a lower WCA and a higher hydrophilicity for the Pd0–TpPz membranes.
As the predominant palladium species at pH 2.5 are PdCl3− or PdCl42−, these anionic species are effectively adsorbed onto the TpPz membrane surface through strong local electrostatic interactions. Fig. S21a illustrates the impact of feed pH on Pd adsorption. A substantial increase in Pd adsorption capacity was observed as pH rose from 1.0 to 3.0. This enhancement stems from the competitive adsorption between excessive chloride ions and protonated sites for PdCl42− at lower pH levels. When the pH exceeded 3, the precipitation of palladium as Pd(OH)+/Pd(OH)2 may compromise the accurate determination of adsorption capacity.41 Furthermore, a slight increase in the adsorption capacity, from 100.3 to 113.6 μg cm−2, was observed as the temperature increased from 15 °C to 45 °C, suggesting the endothermic nature of the palladium adsorption process (Fig. S22b). Notably, the TpPz membrane exhibited remarkable selectivity for PdCl42−, even in the presence of competitive ions including Na+, K+, Ca2+, Mg2+, and Ni2+ (Fig. 3f). Following palladium elution with thiourea, the bright red Pd(II)-saturated TpPz membrane reverted to its original dark red color (Fig. S22). Over twenty adsorption–desorption cycles, the membrane exhibited minimal loss in adsorption capacity and efficiency, indicating its excellent recyclability (Fig. 3g).
Compared to the pristine TpPz membranes, the peak at 3259 cm−1 (pyrazine N) in Pd0–TpPz is significantly diminished, likely due to coordination between pyrazine N and palladium, which weakens the original pyrazine N absorption peak (SI, Fig. S23). In N 1s spectra of TpPz (Fig. S12b), two binding energy peaks at 399.3 and 398.4 eV are attributed to pyrazine N and C–NH, respectively. After in situ reduction treatment, the N 1s spectrum of Pd0–TpPz shows that these peaks shift to higher binding energies by 1.33 eV and 1.62 eV, respectively (Fig. S13b). The O 1s peaks in the Pd0–TpPz membrane (Fig. S12c and S13c) also shift to higher binding energies by 1.61 eV and 1.48 eV. These changes indicate that the Pd adsorption onto TpPz is a chemical process, where both pyrazine-N and ketone groups in TpPz are coordinated with PdII. Fluorescence imaging reveals red fluorescence of the TpPz nanofilm, attributed to its highly conjugated structure. An obvious decline in fluorescence intensity of palladium-absorbed TpPz suggests its strong affinity for palladium (Fig. S24), resulting in fluorescence quenching.42 Electronic localization function (ELF) contour maps (Fig. 3i and j) reveal a higher localized charge density at the pyrazine-N site compared to the oxygen atoms in the TpPz skeleton. Analysis of the electronic structure suggests limited lone pair delocalization at pyrazine-N atoms, resulting in stronger electron donation from these nitrogen centers.43 This enhanced electron-donating ability of the pyrazine-N site contributes to selective palladium adsorption. Furthermore, the electrostatic surface potential of TpPz (Fig. S25) indicates the presence of abundant positive and weak negative charges around the N atom, further facilitating palladium capture. Density functional theory (DFT) calculations were performed to determine the adsorption potential energies of various metal ions on the rational N sites (Fig. 3h and S26). The binding energy (Eads) between PdCl42− and TpPz was calculated to be −0.03 eV, indicating a strong Pd affinity. In contrast, the positive ΔEads values for the other metals revealed the repulsive force against TpPz, primarily attributed to electrostatic interactions between the cations and the positively charged pyrazine and secondary amine groups under acidic conditions.44
![]() | ||
Fig. 4 Separation and catalytic performance of Pd0–TpPz membranes. (a) Separation and catalytic performance of Pd0–TpPz membranes prepared with different Pd2+ adsorption times. (b) Separation performance of Pd0–TpPz membranes for different dyes; (c) catalytic conversion performance of Pd0–TpPz membranes for small molecule pollutants; (d) comparison of dynamic and static catalytic properties of Pd0–TpPz membranes for RhB; (e) RhB conversion and feed flow rate under different pressures; (f) plot of −ln(CP/CF) versus Pd0–TpPz membrane contact time for the dynamic catalysis of RhB, showing first-order kinetics; (g) separation catalytic cycle experiments of Pd0–TpPz membranes for CR and 4-NA mixtures; (h) performance comparison of the Pd0–TpPz membranes with other reported materials. Cu/Cu2O network CMMR;47 Cu–Ag PES CMMR;48 AuNPs PVDF CMMR;49 Ag/ZIF-8 CMMR;50 Nano Cu/ZIF-8 CMMR;36 Pd-TpBpy CMMR;42 PNG/Ag CMMR;51 BNPC-1000 nanoparticles;52 Co@NG-900 nanoparticles;53 0.2 wt% Pd/TiO2 nanoparticles;54 Cu1.5-FCLL nanoparticles.55 |
Pd0–TpPz membranes demonstrated exceptional catalytic activity for both CrVI and RhB reductions under static conditions (Fig. 4d and S37). Complete reduction of CrVI was achieved within 30 min, while RhB reduction required 60 min. Both processes followed first-order kinetics, with calculated Kcat values of 6.75 × 10−2 min−1 for CrVI and 6.15 × 10−2 min−1 for RhB, respectively (Fig. S38a and S39). These values are an order of magnitude higher than those reported for other catalytic membranes under similar conditions.45 Despite their excellent catalytic activity under static conditions, the limited contact with reactants/reductants hinders their suitability for continuous industrial water treatment. However, the uniform, defect-free surface and evenly distributed Pd nanoclusters within the COF layer position the Pd0–TpPz membrane as a promising candidate for superior performance in continuous flow catalytic reactions. Notably, the Pd0–TpPz membrane exhibited exceptional performance in both water permeation and dynamic catalysis. When applied at 4.0 bar, Pd0–TpPz demonstrated an excellent water permeability of 85.4 L m−2 h−1 bar−1 and a remarkably high conversion rate of 98.6% for CrVI and 98.8% for RhB. Moreover, CrVI and RhB conversion rates remained above 95% even at an elevated pressure of 7 bar (Fig. S39b and 4e). The ultrathin TpPz nanofilm, with rich evenly dispersed and ultrafine Pd nanoclusters, facilitated rapid mass transfer and reaction kinetics. The apparent rate constants, Kcat, for CrVI and RhB were determined to be 4.99 ms−1 and 5.46 ms−1, respectively (Fig. S39c and 4f). These values are significantly higher than those observed in static catalysis (6.75 × 10−2 min−1 and 6.15 × 10−2 min−1), highlighting the advantages of dynamic catalysis. By rapidly removing products, dynamic catalysis can overcome thermodynamic limitations and enhance catalytic efficiency.46 Moreover, the Pd0–TpPz membrane exhibits exceptional pH tolerance, maintaining over 99% catalytic efficiency for 10 cycles in both strongly alkaline (0.03 M NaBH4, pH 12) and acidic (0.4 M formic acid, pH 2) environments. This demonstrated stability across extreme pH ranges (2–12) confirms its suitability for real wastewater treatment. Pd–TpPz membranes exhibited minimal Pd leaching (∼8.0 ppb, 0.015%) over 96 h (Fig. S40), confirming their outstanding structural stability.
To address the challenges posed by real-world wastewater, we assessed the water decontamination efficiency of Pd0–TpPz membranes using a mixture of CR and CrVI as feed. The membrane demonstrated impressive efficiency in sieving CR while simultaneously catalyzing the reduction of CrVI. Superior performance was also achieved for treating a mixture of 4-NA and CR (Fig. S41 and S42). The cyclic regeneration ability of the membrane was evaluated through ten consecutive cycles of operation with mixed wastewaters containing CR, CrVI or 4-NA. Remarkably, the removal rates remained consistently above 99.0% (Fig. 4g and S43). For the 4-NA and CR mixture, the water flux increased from 78.5 to 115.3 L m−2 h−1 bar−1, likely stemming from the swelling of the Kevlar substrate under alkaline conditions. In contrast, for the Cr VI and CR mixture, the water flux decreased from 79.0 to 45.8 L m−2 h−1 bar−1, potentially caused by pore blockage from metal complexes. After ten catalytic cycles, a slight increase in Pd nanoparticles was observed on the surface of the Pd0–TpPz membrane (Fig. S44), likely due to the reduction of residual PdII during the catalytic process. These results highlight the robust stability of the Pd0–TpPz membrane, with both catalytic performance and structural integrity remaining intact under prolonged testing. Fig. S45 depicts that the MWCO of the Pd0–TpPz membrane was determined to be 35 kDa. In comparison to the pore size of 1.8 nm measured through N2 adsorption, the Stokes diameter of Pd0–TpPz significantly increased to 4.2 nm. This pore expansion is closely related to the reduction process, where interactions with methanol and bubble formation caused partial swelling of the Kevlar substrate. As a result, the weakened interaction between the Pd–COF layer and the Kevlar substrate led to a substantial increase in membrane pore size. The Pd0–TpPz membrane demonstrated superior catalytic performance for 4-NP reduction, even at low NaBH4 ratios, compared to other reported materials (Fig. 4h).
The Kevlar substrate and TpPz membranes, lacking antibacterial components, exhibited minimal antibacterial activity against Gram-negative Escherichia coli (E. coli) (Fig. S46). In contrast, Pd0–TpPz membranes demonstrated remarkably high antibacterial properties, achieving a 98.1% bacteriostasis rate (BR) as measured by the plate-counting method. This high efficacy stems from the generation of intracellular reactive oxygen species (ROS) induced by the Pd nanoclusters, leading to bacterial inactivation.
DFT calculations reveal that the rate-limiting step in pathway 1 is the hydrogenation from ST-3 to ST-4, with an energy barrier of 0.87 eV (Fig. 5h). In pathway 2, the rate-limiting step for generating product II is the hydrogenation from ST-I to ST-II (0.53 eV). For product IV, the rate-limiting step is the hydrogenation from ST-III to ST-IV (0.37 eV). Lower energy barriers favor the formation of ST-II and ST-IV. Once ST-II forms, its subsequent conversion to ST-III is thermodynamically favorable. In addition, the energy barrier for the transition from ST-III to ST-IV is lower than that from ST-I to ST-II. From an overall reaction perspective, the formation of ST-IV from ST-II is energetically downhill. Thus, we anticipate a higher yield of ST-IV compared to ST-II. However, since ST-II is an intermediate in the formation of ST-IV, it is expected to be present in the final mixture. While the rate-limiting step for the formation of ST-4 has a higher energy barrier than those for ST-II and ST-IV, the overall barrier is still relatively low (less than 1.0 eV), suggesting that ST-4 may also be present in the final mixture.58 These DFT calculation results align well with the findings from the LC-MS analysis.
This journal is © The Royal Society of Chemistry 2025 |