Introduction
Perchlorate (ClO4−) is a persistent toxic inorganic pollutant, while, unlike most other anions which can be precipitated out from aqueous solution by a suitable precipitator, ClO4− is freely water-soluble and there is no suitable precipitator for ClO4− to precipitate it from water. Therefore, separating ClO4− from water is a huge challenge1, 2, 3, 4, 5–6. However, ClO4− is widely used as an oxidizing agents or combustible sources in rocket propellants, explosives, fireworks, and airbag inflation systems with fast diffusion, high stability, and difficult degradation7, 8, 9, 10, 11–12. Once ClO4− enters the water systems and food chain, it will serious threat to human health and environmental security13. For instance, low concentrations of ClO4− inhibit the uptake of iodide by the thyroid gland, thus interfering with the normal functioning of the human thyroid gland and affecting the development of embryos, pregnant women, breastfeeding women, children and adolescents, resulting in serious harm to human health14,15. Therefore, the treatment methods for ClO4− pollution have become an research hotspot. By far, the reported methods for abatement ClO4− pollution mainly include biotechnology methods16, electrochemical reduction methods17, chemical catalysis methods18, ion exchange methods19 and activated carbon adsorption20 or other materials adsorption methods21, etc. Although some progress has been made, how to treat ClO4− pollution quickly, efficiently, and at low-cost without causing secondary pollution still remains a huge challenge. Due to the non-volatility, high solubility, and kinetic stability of ClO4−, it is difficult for conventional treatment techniques to remove ClO4− in the water effectively. However, removing pollutants from water through adsorption and separation is still a simple, low-cost, and easy application method22, 23–24. Therefore, the development of materials that can efficiently bind and separate ClO4− from water has important needs.
With the rapid development of supramolecular chemistry, adsorption and separation materials based on macrocycle-assembled supramolecular polymer29,34 have attracted widespread attentions because of their excellent host-guest complexation properties25, 26, 27, 28, 29, 30, 31, 32, 33, 34–35. Among various macrocyclic hosts, pillar[n]arenes, first reported by Ogoshi36, 37, 38, 39, 40, 41, 42, 43, 44–45, possesses the merits of easy synthesis and modification, adjustable cavity size, capable of providing a variety of supramolecular interactions, and strong complexing ability to the guest38,39,46, 47, 48, 49, 50–51,53. For instance, perethylated pillar[5]arene can encapsulate linear alkanes52,53, and per-hydroxylated pillar[6]arene can well complex highly toxic paraquat54, etc. These examples were achieved by taking advantage of the electron-rich cavities of pillar[n]arenes and based on the C-H···π interactions. It is worth mentioning that taking the common per-alkoxy-pillar[5]arene as an example, there are ten alkoxyl groups on the two sides of its cavity, which can provide abundant C-H···X (X = O, N, F, etc.) hydrogen bond donors and probably can effectively complex the target guest. However, the report on the utilizing the multi-alkoxy groups at both sides of the pillar[5]arene cavity to bind target guest by forming multiple hydrogen bonds is still very rare. Therefore, how to rationally and efficiently utilize these multi-alkoxy groups to enhance the binding ability of pillar[5]arene to target guests become an interesting opportunity for developing efficient macrocycle-assembled supramolecular polymer materials for binding and separation of specific guests such as ClO4−.
Given all this, and as our interests in host-guest recognition and separation27,55, 56, 57, 58, 59, 60, 61, 62–63, herein, to achieve efficient complexing and separation ClO4- from water, as shown in Fig. 1 and 2, we developed a pillar[5]arene-based hydrogen-bonded supramolecular polymer networks crystal (HBPC), which can efficiently bind ClO4− through clustered hydrogen-bonding based on two-armed-macrocycle cooperation effect and improve the complexation and separation ability of the ClO4−. Firstly, to enhance the binding property of pillar[5]arene, we rationally conjugate-functionalized pillar[5]arene on its A, A’ sites by bis-pyridyl-hydrazone groups. The bis-pyridyl hydrazone groups can act as two arms of pillar[5]arene host (PYP5) and serve as hydrogen bond donors and acceptors, which can collaborate with the eight surrounding ethoxyl groups on the pillar[5]arene cavity to form clustered hydrogen-bonding with ClO4−. Secondly, the conjugated functionalization model can enhance the rigidity of the two-arm of the PYP5 and help PYP5 molecules assemble to hydrogen-bonded supramolecular polymer networks crystal (HBPC) through multiple hydrogen bonds by the two arms of adjacent PYP5 molecules. As we expected, the HBPC-based crystal materials can efficiently bind and remove ClO4− from water.
[See PDF for image]
Fig. 1
Assembly and Single crystals structure of hydrogen-bonded supramolecular polymer networks crystal (HBPC).
a PYP5 dimer and one-dimensional catenulate supramolecular polymers. b Pore structures of HBPC. c Hydrogen-bonds in HBPC. d Single crystals structure of HBPC⊃ClO4−.
Results
Synthesis and characterization of PYP5
Synthetic details of PYP5 are shown in Supplementary Fig. 1, where the key intermediates 1 and p-Q were synthesized by previously reported methods47,48. Firstly, p-Q was obtained by the classic Suzuki coupling reaction of 1 with 4-formylphenylboronic acid in 55% yield. Subsequently, A,A’-bis-pyridyl-hydrazone-phenyl conjugate-functionalized pillar[5]arene macrocyclic molecule PYP5 was obtained by imine condensation reaction between p-Q and 2-hydrazinylpyridine under reflux conditions in 80% yield. The obtained PYP5 was fully characterized by 1H NMR, 13C NMR, high-resolution mass spectrometry (Supplementary Figs. 3–5), and single-crystal X-ray analysis (Supplementary Fig. 8).
Crystal structure of hydrogen-bonded supramolecular polymer networks HBPC
High-quality colorless transparent single crystals of PYP5 suitable for single-crystal X-ray diffraction were obtained by slowly evaporating the mixed solution of dichloromethane and acetone of PYP5 within 3–4 days at room temperature. X-ray crystallographic analysis revealed that PYP5 single crystals belong to the triclinic crystal system. In the crystal, as shown in Fig. 2a and Supplementary Fig. 8b, the two pyridyl-hydrazone arms are open toward the outside. Interestingly, a pair of N-H···N hydrogen bonds were formed between the -NH group on the hydrazone and the N atom on adjacent molecule of pyridyl group, with bond lengths of 2.235 Å and 2.125 Å, respectively (Fig. 1a). Similarly, a pair of N-H···N hydrogen bonds were also formed between another pyridyl-hydrazone arm with other adjacent molecules, thereby forming one-dimensional catenulate supramolecular polymers (Fig. 1a). More interestingly, adjacent one-dimensional catenulate supramolecular polymers were further self-assembled to form hydrogen-bonded supramolecular polymer networks crystal (HBPC) via C-H···π interactions (Fig. 1a, c and Supplementary Fig. 8c). As shown in Fig. 1b, the networks show pore structures, so, the N2 gas adsorption-desorption isotherm (Supplementary Fig. 9) was measured at 77.3 K, and the apparent surface area of HBPC was obtained to be 4.756 m2 g−1 using the Brunauer-Emmett-Teller (BET) model. Meanwhile, the free volume (FV) based on the calculation from single crystal data is 1840.128 Å3, and the free volume fraction (FVF)71 is up to 46.09% (Supplementary Fig. 10), which supplied suitable pore structures for adsorbing ClO4−.
[See PDF for image]
Fig. 2
Clustered Hydrogen-Bonds binding ClO4−: effect and mechanism.
a Single crystal structures of PYP5 and PYP5⊃2ClO4− and its side view, top view show the Clustered Hydrogen-Bonding based on two-armed-macrocycle cooperation. b ESP maps of PYP5 and PYP5⊃2ClO4−. c Visualization of Clustered Hydrogen-Bonding via the IGMH method based on single crystal data from a different view.
Separation property of HBPC crystal
To study ClO4− removal properties62 of the HBPC in aqueous solution, the sodium perchlorate standard solution was adsorbed using HBPC crystal materials. Uptake efficiency of HBPC crystal materials for ClO4− (5.0 mg/L) in aqueous solution was measured by Ion Chromatograph (IC) analysis up to about 99.24% (Fig. 3a and Supplementary Table 3). The residual ClO4− concentration was 37.8 μg/L, which is lower than the World Health Organization standard (WHO, 70 μg/L)64 for drinking water quality, indicating that the HBPC crystal materials demonstrate effective removal capabilities on ClO4− in aqueous solution. Then, we examined the uptake efficiency of HBPC crystal materials for common anions such as F−, Cl−, Br−, NO3− and SO42− in water as shown in Fig. 3b and Supplementary Table 3, and the results showed that HBPC crystal materials have good selectivity to ClO4−. Meanwhile, compared with the reported ClO4− treatment methods and materials (Supplementary Tables 8, 9), this approach with the advantage of simple operation (only needs filter), low cost, low energy consumption, convenient application, and higher uptake efficiency.
[See PDF for image]
Fig. 3
Uptake efficiency of HBPC-based crystal materials.
a The uptake effect of ClO4− in water by HBPC-based crystal materials. b The uptake efficiency of HBPC crystal materials for common anions in water.
For deeply understand the uptake behavior of HBPC crystal materials for ClO4−, we investigated the feed ratio effect between HBPC crystal materials and ClO4−. At feed molar ratios of HBPC crystals to ClO4− ranging from 10: 1 to 2: 1, the uptake efficiencies ranged from 99.24% to 91.81% (Supplementary Table 4). Subsequently, we also investigated the concentration effect of ClO4−. At concentrations of ClO4− ranging from 1 to 5 mg/L, the uptake efficiencies ranged from 94.83% to 99.24% (Supplementary Table 5), and the residual ClO4− concentration after uptake by HBPC crystals was lower than the WHO for drinking water quality standard. In addition, to further verify the role of pillar[5]arene in the clustered hydrogen-bonding based on two-armed-macrocycle cooperation strategy on binding of ClO4−, the unit model (PY) without pillar[5]arene group was designed and synthesized (Supplementary Fig. 2), and its uptake efficiency to ClO4− was investigated. The results are shown in Supplementary Tables 6, 7, the uptake efficiencies of PY to ClO4− in aqueous solution were significantly lower than that of HBPC. For example, the uptake efficiency of PY to ClO4− is 75.40%, with the feed molar ratio of 2: 1, while the uptake efficiency of HBPC to ClO4− is 91.81%. This fully proved that pillar[5]arene plays a key role in the two-armed-macrocycle cooperation based clustered hydrogen-bonding enhancing host-guest binding strategy, which remarkably improves the complexation ability of ClO4−.
Complexation mechanism
For investigating the complexation mechanism of PYP5 on ClO4−, firstly, we performed HR-MS characterization and found that the complexation ratio between PYP5 and ClO4− was 1: 2 (Supplementary Fig. 11). Fortunately, we obtained the single crystals of HBPC-ClO4− complex HBPC⊃ClO4−, and the binding mode of PYP5 and ClO4− in HBPC can be understood more directly through the single crystal structure. In the crystal structure of PYP5 (Fig. 2a, left), the bis-pyridyl-hydrazone stretches towards the outside of the pillar[5]arene, where one of the pyridine N atoms is away from the pillar[5]arene cavity. When combined with ClO4−, in the single crystals of HBPC⊃ClO4−, the pyridyl-hydrazones on the PYP5 underwent adaptive rotation due to the inducing of ClO4− (Fig. 2a right, and Supplementary Fig. 12), and the two-pyridyl-hydrazone arms and the pillar[5]arene cavity cooperatively bind the ClO4− with multiple clustered hydrogen bonds. Among them, the pyridine groups on the bis-pyridyl-hydrazone have been protonated by obtaining the proton from solvent and formed N-H···O hydrogen bonds with ClO4−; meanwhile, the methylene and methyl groups on the multiple ethoxyl groups of pillar[5]arene formed multiple C-H···O hydrogen bonds with ClO4−; moreover, the -NH groups on the hydrazone group of adjacent molecule of PYP5 also form N-H···O hydrogen bonds with ClO4− (Supplementary Fig. 13), all these hydrogen bonds surrounded the ClO4− and formed Clustered Hydrogen-Bonding to achieve very effective binding for ClO4− (Fig. 2a, right). In addition, for deeply theoretical study and visualization of the Clustered Hydrogen-Bonding, the hydrogen bonds and weak interaction forces between PYP5 and ClO4− were calculated based on crystal data through the IGMH method70. As shown in Fig. 2c and Supplementary Fig. 14, ClO4− forms multiple clustered N-H···O and C-H···O hydrogen bonds with -NH groups on pyridyl-hydrazone, as well as the methyl and methylene groups on ethoxyl groups. It was further demonstrated that PYP5 could produce effective complexation of ClO4− via Clustered Hydrogen-Bonding. Interestingly, the HBPC assembled from PYP5 after complexing ClO4− produced a single-crystal to single-crystal transformation. As shown in Fig. 5a, the PXRD pattern of HBPC showed observable changes after uptake of ClO4−, which confirmed that the uptake of ClO4− caused the transformation of crystal structure30,46. At this time, the HBPC has assembled through π···π stacking interactions (Supplementary Figs. 15–17), C-H···π interactions, and N-H···O hydrogen bonds (Fig. 1d and Supplementary Figs. 18, 19), and the networks structure of the HBPC still maintained, which endow the beneficial property for HBPC on efficient separation of ClO4−.
To further explore the complexation mechanism based on Clustered Hydrogen-Bonding, 1H NMR titration experiments were carried out based on a constant concentration of PYP5 and varying concentrations of ClO4− at room temperature (Fig. 4). As the concentration of ClO4− increased, the signal peaks of protons Hd, Hf, Hb, and Ha on the conjugated bis-pyridyl-hydrazone arms shifted to the low fields (blue dashed arrows), and the signal peaks of protons He shown slight upfield shift (red dashed arrows), indicating that N-H···O hydrogen bonds formed between the conjugated bis-pyridyl-hydrazone arms and ClO4−. In the meantime, the signal peaks of protons Hc and Hg on the pyridyl-hydrazone functionalized two-arms shown downfield shifts, indicating that C-H···O hydrogen bonds formed between the conjugated bis-pyridyl-hydrazone arms and ClO4−. And the signal peaks of protons Hj and Hk on the ethoxyl groups at the ends of the pillar[5]arene cavity also displayed obvious chemical shifts, which were attributed to the formation of multiple C-H···O hydrogen bonds between the ethoxyl groups at the ends of the pillar[5]arene cavity and ClO4−. 1H NMR titration experiments also support the Clustered Hydrogen-Bonding between PYP5 and ClO4− based on two-armed-macrocycle cooperation effect. Moreover, since PYP5 is a two-armed macrocyclic host, PYP5 and ClO4− could gradually form 1: 1 and 1: 2 complexes with corresponding complexation constants Ka1 and Ka263. The complexation constants Ka1, Ka2, and Ka between PYP5 and ClO4− were obtained by 1H NMR titration experiments based on the method of non-linear curve-fitting according to the SupraFit program65 as (2.7 ± 0.0010) × 104 M−1, (2.1 ± 0.01) × 102 M−1, and (5.7 ± 0.03) × 106 M−1, respectively, and the ΔG1, ΔG2, ΔG as −25.32 kJ/mol, −13.24 kJ/mol, and −38.56 kJ/mol, respectively, which indicated that the stable complex has been formed between PYP5 and ClO4− and the binding of PYP5 to ClO4− was spontaneous. Meanwhile, as shown in Supplementary Table 10, compared with the reported results, the PYP5 shows considerable complexation constants and can firmly bind ClO4−.
[See PDF for image]
Fig. 4
1H NMR titration evidence of Clustered Hydrogen-Bonding for ClO4−.
Partial 1H NMR spectra (DMSO‑d6, 400 MHz, 298 K) of PYP5 in the presence of various molar equivalents of ClO4−.
To gain a deeper understanding of the Clustered Hydrogen-Bonding based on two-armed-macrocycle cooperation effect, the complexation mechanism of PYP5 with ClO4− was further studied using density functional theory (DFT)66, 67–68. Firstly, in order to clearly present the surface charge distribution before and after the host-guest complexation, the electrostatic potential surface (ESP) of PYP5, ClO4− and PYP5⊃2ClO4− were calculated (Fig. 2b)69. The ESP showed that the eight ethoxyl groups on the pillar[5]arene and part of bis-pyridyl-hydrazone arms supply an electron-deficient environment, whereas ClO4− is electron-rich, which is favorable for the binding of the PYP5 to ClO4−. Meanwhile, the electrons were redistributed during the formation of the host-guest complex PYP5⊃2ClO4−. Subsequently, the complex mechanism also has been investigated by frontier molecular orbitals data for host and host-guest complexes (highest occupied molecular orbital and lowest unoccupied molecular orbitals, HOMO and LUMO). As shown in Fig. 5b, the HOMO orbitals of PYP5 are mainly distributed on the pillar[5]arene cavity, while in the host-guest complexes PYP5⊃2ClO4−, the HOMO orbitals are distributed on the guest ClO4−. Meanwhile, the LUMO orbitals were transferred from the pillar[5]arene functionalized two-arms of PYP5 to the single-arm of PYP5⊃2ClO4−. And the HOMO-LUMO gap after forming the host-guest complexes PYP5⊃2ClO4− is 0.81 eV, which is less than PYP5 (3.41 eV). These changes are caused by the charge transfer between molecules during the formation of host-guest complex and the formation of multiple hydrogen bonds. Moreover, the binding energy between PYP5 and ClO4− was calculated to be −183.34 kcal/mol, indicating that PYP5 has strong binding ability to ClO4− 72. The above DFT calculation results provide a clear explanation and theoretical basis for the formation of Clustered Hydrogen-Bonding between PYP5 and ClO4−. Therefore, crystal structure analysis, 1H NMR titration experiments, and DFT calculations fully demonstrated that Clustered Hydrogen-Bonding based on two-armed-macrocycle cooperation effect was formed between PYP5 and ClO4−, which enhanced significantly the adsorption and separation ability of PYP5 to ClO4−.
[See PDF for image]
Fig. 5
Host-guest study.
a PXRD patterns of single crystalline samples of HBPC: I, before ClO4− uptake; II, after uptake of ClO4−. b HOMO and LUMO of PYP5 and PYP5⊃2ClO4− were analyzed by DFT calculations based on crystal data.
Discussion
In conclusion, to achieve convenient removal ClO4− from water, we successfully developed a hydrogen-bonded supramolecular polymer networks crystal (HBPC) based on reasonable clustered hydrogen-bonding enhancing host-guest binding strategy to improve the complexation and separation ability of the hosts to ClO4−. The clustered hydrogen-bonding is realized through two-armed-macrocycle cooperation effect based on a bis-pyridyl hydrazone functionalized pillar[5]arene PYP5. Interestingly, the PYP5 can form hydrogen-bonded supramolecular polymer networks crystal (HBPC) through hydrogen-bond self-assembly, which can efficiently separate ClO4− from water with an uptake efficiency of up to 99.24%, and the residual concentration of ClO4− is below the WHO standard for drinking water quality. Compared with the reported ClO4− treatment materials, the HBPC-based crystal materials possess considerably higher removal efficiency as well as with the advantage of easy operation and convenient application. The excellent ClO4− binding and removal properties of the HBPC is originate from the cooperation of the two pyridyl hydrazone arms and the eight ethoxyl groups on pillar[5]arene of PYP5 by forming multiple clustered hydrogen-bonding, the proposed binding mechanism has been carefully investigated and verified by single crystals XRD analysis and theoretical calculation. The proposed clustered hydrogen-bonding based on two-armed-macrocycle cooperation is a reasonable and useful way to enhance the binding property of supramolecular polymer, which not only offers a rational and feasible solution for separating water-soluble pollutants or anions from water but also provides experience and design strategies for macrocycle-assembled supramolecular polymer materials.
Methods
Materials
Starting materials and reagents including ammonium ceric nitrate (AR, 99%), sodium dithionite (AR, 90%), pyridine (99.5%), trifluoromethanesulfonic anhydride (AR, ≥99.5%), tetrakis(triphenylphosphine)palladium (99%), 4-formylphenylboronic acid (98%), 2-hydrazinopyridine (98%), were purchased from commercial suppliers and used without further purification unless stated otherwise.
Synthesis of 1
Compound 1 was prepared according to a previously reported literature. 47-48 Firstly, EtP5 (1.78 g, 2.00 mmol) was dissolved in 100 mL of dichloromethane (CH2Cl2). Diammonium cerium (IV) nitrate (CAN, 2.19 g, 4.00 mmol) in 20 mL of water was added drop by drop. The mixture was stirred at room temperature for 30 min. The organic layer was separated, washed with water three times and dry with Na2SO4, then the organic layer was purified by silica gel chromatography to yield EtP[4]Q[1] as a red solid (1.00 g, yield 60%). Secondly, EtP[4]Q[1] (1.67 g, 2.00 mmol) was dissolved in 100 mL of dichloromethane (CH2Cl2), and a solution of sodium dithionite (1.74 g, 10.00 mmol) in 40 mL of water was added. The mixture was stirred under a nitrogen atmosphere at room temperature overnight. The red solution turned into colorless. The organic layer was separated and washed with water three times. After concentration, a light-yellow solid was obtained, which was used in the next step immediately. Finally, trifluoromethanesulfonic anhydride (Tf2O, 1.50 mL, 3.00 mmol) was added dropwise to a mixture of the light-yellow solid obtained in the previous step and pyridine (dry, 5 mL) in CH2Cl2 (dry, 50 mL) at 0 °C under a nitrogen atmosphere. The mixture was stirred at room temperature for 24 h before being quenched with water. After washing the organic phase with water three times, the organic layer was concentrated under vacuum. And the crude product was purified via silica gel column chromatography to yield 1 as a white powder (1.80 g, yield 82%).
Synthesis of p-Q
Compound p-Q was prepared according to a previously reported literature. 48 Under the protection of argon, 1 (1.10 g, 1.00 mmol), 4-formylphenylboronic acid (37.49 mg, 2.50 mmol), Pd(PPh3)4 (28.99 mg, 0.25 mmol) and K2CO3 (55.28 mg, 4.00 mmol) in 40 mL of a mixed solvent (THF / H2O, V / V = 3: 1). The mixture was stirred and heated at reflux for 24 h. After the reaction was deemed complete, the volatiles were removed under reduced pressure, and the crude product was purified using silica gel column chromatography to yield p-Q as a white powder (558 mg, yield 55%).
Synthesis of PYP5
In a 100 mL round bottom flask, p-Q47,48 (0.20 g, 0.20 mmol), 2-hydrazinopyridine (47.48 mg, 0.44 mmol), and anhydrous ethanol (20 mL) were added. Then add the catalytic amount of glacial acetic acid. The reaction mixture was stirred under reflux temperature until p-Q was completely reacted. After the reaction, the solvent was removed by evaporation, and the precipitate was washed with anhydrous ethanol to obtain PYP5 as a white solid (189 mg, yield 80%).
Synthesis of PY
In a 100 mL round bottom flask, 4,4-p-terphenyl dicarboxaldehyde (143.17 mg, 0.50 mmol), 2-hydrazinopyridine (120.05 mg, 1.10 mmol), and anhydrous ethanol (20 mL) were added. Then add the catalytic amount of glacial acetic acid. The reaction mixture was stirred at reflux temperature until 4,4-p-terphenyl dicarboxaldehyde was completely reacted. After the reaction, the solvent was removed by evaporation, and the precipitate was washed with anhydrous ethanol to obtain PY as a yellow solid (182 mg, yield 78%).
Theoretical calculations
All calculations were performed by density functional theory (DFT) using the B3LYP hybrid function combined with 6-31 G (d, p) basis set under Gaussian 09 D.01 program. Using single crystals as input files, visualized by Multiwfn 3.8 program and Visual Molecular Dynamics software.
Acknowledgements
This work was supported by the NSFC (Nos. 22471222, 22461040, 22061039, 22069031, 22001214, 22165027), the Top Leading Talents Project of Gansu Province, the Key R & D Program of Gansu Province (No. 21YF5GA066), the Science Fund for Distinguished Young Scholars of Gansu Province (22JR5RA131), Gansu Province College Industry Support Plan Project (No. 2022CYZC-18), the Fundamental Research Funds for the Central Universities (No. 31920230145), Natural Science Foundation of Gansu Province (Nos. 2020-0405-JCC-630, 20JR10RA088, 21JR1RA220), Gansu Province Science Foundation for Youths (23JRRA690), and Northwest Normal University Young Scholars Research Capacity Improvement Program (NWNU-LKQN2023-05), Northwest Normal University Postgraduate Research Funding Project (No. 2022KYZZ-S162), Gansu Province Outstanding Postgraduate Students “The Star of Innovation” Project (No. 2023CXZX-349).
Author contributions
J.T.: Molecular synthesis, Experiments, Data analysis, Writing – original draft. X.H.: Experiments, Data analysis. H.-R. Y.: Data analysis. J.L.: Funding acquisition. J.-F.C.: Data analysis. T.-B.W.: Experiments, Data analysis. H.Y.: Data analysis. W.-J.Q.: Data analysis. B.S.: Funding acquisition. Q.L.: Conceived and designed the experiments, Data analysis, Supervision, Conceptualization, Funding acquisition, Manuscript writing and revision. All authors discussed and commented on the manuscript.
Peer review
Peer review information
Nature Communications thanks Tomoki Ogoshi, Ying-Wei Yang, and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
All data supporting the findings of this study are available from the article and its Supplementary Information or available from the corresponding author upon request. The X-ray crystallographic coordinates for structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre (CCDC), under deposition numbers 2379654 (PYP5), and 2379655 (HBPC⊃ClO4−). These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. Source data are provided with this paper. Atomic coordinates for DFT experiments are provided with the paper.
Competing interests
The authors declare no competing interests.
Supplementary information
The online version contains supplementary material available at https://doi.org/10.1038/s41467-025-61910-y.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
1. Cao, F et al. Worldwide occurrence and origin of perchlorate ion in waters: A review. Sci. Total Environ.; 2019; 661, pp. 737-749.2019ScTEn.661.737C1:CAS:528:DC%2BC1MXhvVequr4%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30684841]
2. Kumar, KS; Kavitha, S; Parameswari, K; Sakunthala, A; Sathishkumar, P. Environmental occurrence, toxicity and remediation of perchlorate – A review. Chemosphere; 2023; 311, 137017.1:CAS:528:DC%2BB38XivVOiu7rN [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36377118]
3. Hu, J et al. Perchlorate occurrence in foodstuffs and water: Analytical methods and techniques for removal from water – A review. Food Chem.; 2021; 360, 1:CAS:528:DC%2BB3MXhtFertrrP [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34034057]130146.
4. Hu, Y et al. Adsorption and recovery of phosphate using sodium carbonate as co-precipitant synthesized La&Zr dual-metal modified material: Adsorption mechanism and practical application. Chemosphere; 2024; 363, 142878.1:CAS:528:DC%2BB2cXhsF2nt7vJ [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/39032732]
5. Tzioumis, NA; Cullen, DA; Jolliffe, KA; White, NG. Selective removal of sulfate from water by precipitation with a rigid Bis-amidinium compound. Angew. Chem. Int. Ed.; 2023; 62, e202218360.1:CAS:528:DC%2BB3sXjtlKjtbg%3D
6. Houtun, F; Chuncheng, HJ; Limin, Z; Fang, C. Reduction of chloride matrix effect using silver oxide as a precipitating reagent for the determination of trace anions in chloride-rich samples via ion chromatography. J. Sep. Sci.; 2007; 30, pp. 693-698. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/17461108]
7. Jacobs, PWM; Whitehead, HM. Decomposition and combustion of ammonium perchlorate. Chem. Rev.; 1969; 69, pp. 551-590.1:CAS:528:DyaF1MXkslanur8%3D
8. Trache, D; Klapötke, TM; Maiz, L; Abd-Elghany, M; DeLuca, LT. Recent advances in new oxidizers for solid rocket propulsion. Green. Chem.; 2017; 19, pp. 4711-4736.1:CAS:528:DC%2BC2sXhsVersrrO
9. Samanta, J et al. Tripodal organic cages with unconventional C-H···O interactions for perchlorate remediation in water. J. Am. Chem. Soc.; 2023; 145, pp. 21723-21728.1:CAS:528:DC%2BB3sXitVSqur%2FE [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37769032]
10. Coates, JD; Achenbach, LA. Microbial perchlorate reduction: rocket-fueled metabolism. Nat. Rev. Microbiol.; 2004; 2, pp. 569-580.1:CAS:528:DC%2BD2cXkvVSru7c%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/15197392]
11. Steinhauser, G; Klapotke, TM. “Green” pyrotechnics: a chemists’ challenge. Angew. Chem. Int. Ed.; 2008; 47, pp. 3330-3347.1:CAS:528:DC%2BD1cXmtlWrtbk%3D
12. Su, Z; Li, Y; Li, J; Li, K; Dou, X. Ultrasensitive dual-mode visualization of perchlorate in water, soil and air boosted by close and stable Pt–Pt packing endowed low-energy absorption and emission. J. Mater. Chem. A; 2022; 10, pp. 8195-8207.1:CAS:528:DC%2BB38Xot1GjsLY%3D
13. Maffini, MV; Trasande, L; Neltner, TG. Perchlorate and Diet: Human Exposures, Risks, and Mitigation Strategies. Curr. Environ. Health Rep.; 2016; 3, pp. 107-117.1:CAS:528:DC%2BC28Xht1agsrrF [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27029550]
14. De Groef, B; Decallonne, BR; Van der Geyten, S; Darras, VM; Bouillon, R. Perchlorate versus other environmental sodium/iodide symporter inhibitors: potential thyroid-related health effects. Eur. J. Endocrinol.; 2006; 155, pp. 17-25. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/16793945]
15. Zhang, T et al. Placental transfer of and infantile exposure to perchlorate. Chemosphere; 2016; 144, pp. 948-954.2016Chmsp.144.948Z1:CAS:528:DC%2BC2MXhsFygt7zL [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26432537]
16. Levakov, I; Ronen, Z; Dahan, O. Combined in-situ bioremediation treatment for perchlorate pollution in the vadose zone and groundwater. J. Hazard. Mater.; 2019; 369, pp. 439-447.1:CAS:528:DC%2BC1MXjsVens7c%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30784974]
17. Wang, DM; Lin, HY; Shah, SI; Ni, CY; Huang, CP. Indirect electrochemical reduction of perchlorate and nitrate in dilute aqueous solutions at the Ti–water interface. Sep. Purif. Technol.; 2009; 67, pp. 127-134.1:CAS:528:DC%2BD1MXmtVKhtr4%3D
18. Choe, JK; Shapley, JR; Strathmann, TJ; Werth, CJ. Influence of Rhenium speciation on the stability and activity of Re/Pd bimetal catalysts used for perchlorate reduction. Environ. Sci. Technol.; 2010; 44, pp. 4716-4721.2010EnST..44.4716C1:CAS:528:DC%2BC3cXmtlWmu7s%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20481620]
19. Yoon, I-H et al. Perchlorate adsorption and desorption on activated carbon and anion exchange resin. J. Hazard. Mater.; 2009; 164, pp. 87-94.2009AIPC.1195..87Y1:CAS:528:DC%2BD1MXislGlsrk%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/18789577]
20. Mahmudov, R; Huang, CP. Perchlorate removal by activated carbon adsorption. Sep. Purif. Technol.; 2010; 70, pp. 329-337.1:CAS:528:DC%2BD1MXhsFynu7%2FN
21. Srinivasan, R; Sorial, GA. Treatment of perchlorate in drinking water: A critical review. Sep. Purif. Technol.; 2009; 69, pp. 7-21.1:CAS:528:DC%2BD1MXhtVKntLvO
22. Liang, L; Zhao, W; Yang, X-J; Wu, B. Anion-coordination-driven assembly. Acc. Chem. Res.; 2022; 55, pp. 3218-3229.1:CAS:528:DC%2BB38Xisl2gtrzI [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36331808]
23. He, M et al. Biomimetic charge-neutral anion receptors for reversible binding and release of highly hydrated phosphate in water. Angew. Chem. Int. Ed.; 2024; 63, 1:CAS:528:DC%2BB2cXhsFSlurfM e202406946.
24. Liang, L et al. Peripheral control of the assembly and chirality of anion-based octanuclear cubes by Cation–π networks. J. Am. Chem. Soc.; 2024; 146, pp. 10908-10916.1:CAS:528:DC%2BB2cXnsV2hurw%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38579155]
25. Jie, K; Zhou, Y; Li, E; Huang, F. Nonporous adaptive crystals of Pillararenes. Acc. Chem. Res.; 2018; 51, pp. 2064-2072.1:CAS:528:DC%2BC1cXhtlWntbvF [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30011181]
26. Zhu, H et al. Applications of macrocycle-based solid-state host–guest chemistry. Nat. Rev. Chem.; 2023; 7, pp. 768-782.1:CAS:528:DC%2BB3sXitVeqsr%2FF [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37783822]
27. Shi, B et al. Clamparene: synthesis, structure, and its application in spontaneous formation of 3D porous crystals. J. Am. Chem. Soc.; 2024; 146, pp. 2901-2906.1:CAS:528:DC%2BB2cXisF2mtrk%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38271666]
28. Wu, J-R; Wu, G; Yang, Y-W. Pillararene-inspired macrocycles: from extended Pillar[n]arenes to Geminiarenes. Acc. Chem. Res.; 2022; 55, pp. 3191-3204.1:CAS:528:DC%2BB38Xis1Kgtb7F [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36265167]
29. Wang, Z-Q; Wang, X; Yang, Y-W. Pillararene-based supramolecular polymers for adsorption and separation. Adv. Mater.; 2024; 36, 2301721.1:CAS:528:DC%2BB3sXis1Cgsr3K
30. Chen, J et al. Separation of benzene and toluene associated with vapochromic behaviors by hybrid[4]arene-based co-crystals. Nat. Commun.; 2024; 15, 2024NatCo.15.1260C1:CAS:528:DC%2BB2cXksVWjtL4%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38341431][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10858960]1260.
31. Li, B; Wang, Y; Liu, L; Dong, M; Li, C. Separation of Cyclohexanone and Cyclohexanol by adaptive Pillar[5]arene Cocrystals accompanied by vapochromic behavior. JACS Au; 2023; 3, pp. 1590-1595.1:CAS:528:DC%2BB3sXpvFCqu7o%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37388695][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10301796]
32. Li, E; Zhu, W; Fang, S; Jie, K; Huang, F. Reimplementing guest shape sorting of nonporous adaptive crystals via substituent-size-dependent solid-vapor postsynthetic modification. Angew. Chem. Int. Ed.; 2022; 61, e202211780.1:CAS:528:DC%2BB38XitlGlt7jP
33. Ren, S; Qiao, G-Y; Wu, J-R. Supramolecular-macrocycle-based functional organic cocrystals. Chem. Soc. Rev.; 2024; 53, pp. 10312-10334.1:CAS:528:DC%2BB2cXhvFWksbrP [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/39240538]
34. Li, Z; Yang, Y-W. Macrocycle-based porous organic polymers for separation, sensing, and catalysis. Adv. Mater.; 2022; 34, 2107401.1:CAS:528:DC%2BB38XitlGj
35. Shi, B et al. Pillar[5]arene-based molecular recognition induced Crystal-to-Crystal transformation and its application in adsorption of Adiponitrile in water. ACS Mater. Lett.; 2019; 1, pp. 111-115.1982MatL..1.111S1:CAS:528:DC%2BC1MXhtVWlur7K
36. Ogoshi, T; Kanai, S; Fujinami, S; Yamagishi, T-a; Nakamoto, Y. para-Bridged symmetrical Pillar[5]arenes: Their Lewis acid catalyzed synthesis and host–guest property. J. Am. Chem. Soc.; 2008; 130, pp. 5022-5023.1:CAS:528:DC%2BD1cXjslSnsb4%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/18357989]
37. Cao, D et al. A facile and efficient preparation of Pillararenes and a Pillarquinone. Angew. Chem. Int. Ed.; 2009; 48, pp. 9721-9723.1:CAS:528:DC%2BD1MXhsFGru7%2FE
38. Ogoshi, T; Yamagishi, T-a; Nakamoto, Y. Pillar-shaped macrocyclic hosts Pillar[n]arenes: New key players for supramolecular chemistry. Chem. Rev.; 2016; 116, pp. 7937-8002.1:CAS:528:DC%2BC28XhtVaksL7P [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27337002]
39. Kakuta, T; Yamagishi, T-a; Ogoshi, T. Stimuli-responsive supramolecular assemblies constructed from Pillar[n]arenes. Acc. Chem. Res.; 2018; 51, pp. 1656-1666.1:CAS:528:DC%2BC1cXhtV2ksL3K [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29889488]
40. Yang, J; Lou, X-Y; Dai, D; Shi, J; Yang, Y-W. Desymmetrized pillar[8]arenes: High-yield synthesis, functionalization, and host-guest chemistry. Chin. Chem. Lett.; 2025; 36, 109818.1:CAS:528:DC%2BB2cXit1GntbfI
41. Wu, Y et al. Pillararene incorporated metal–organic frameworks for supramolecular recognition and selective separation. Nat. Commun.; 2023; 14, 2023NatCo.14.4927W1:CAS:528:DC%2BB3sXhslaltLbF [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37582786][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10427641]4927.
42. Xu, L et al. A conjugated polymeric supramolecular network with aggregation-induced emission enhancement: an efficient light-harvesting system with an ultrahigh antenna effect. Angew. Chem. Int. Ed.; 2019; 59, pp. 9908-9913.
43. Zhou X. H. et al. Catalytic Enantioselective synthesis of planar chiral Pillar[5]arenes via Asymmetric Sonogashira Coupling. Angew. Chem. Int. Ed. e202415190 (2024).
44. Lou X. Y. et al. Self-assembled Nanohelixes driven by host-guest interactions and metal coordination. Angew. Chem. Int. Ed. e202414611 (2024).
45. Wang, Z et al. A Pillar[5]arene-containing metal-organic framework for rapid and highly capable adsorption of a mustard gas simulant. J. Am. Chem. Soc.; 2024; 146, pp. 23330-23337.1:CAS:528:DC%2BB2cXhslClt7zI [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/39110895]
46. Li, Q; Zhu, H; Huang, F. Alkyl chain length-selective vapor-induced Fluorochromism of Pillar[5]arene-based nonporous adaptive crystals. J. Am. Chem. Soc.; 2019; 141, pp. 13290-13294.1:CAS:528:DC%2BC1MXhsFKlurfP [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31405274]
47. Lou, X-Y; Yang, Y-W. Pyridine-Conjugated Pillar[5]arene: From molecular crystals of blue luminescence to red-emissive coordination nanocrystals. J. Am. Chem. Soc.; 2021; 143, pp. 11976-11981.1:CAS:528:DC%2BB3MXhs1KqsrbL [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34319726]
48. Zhu, H et al. Substrate-responsive Pillar[5]arene-based organic room-temperature phosphorescence. J. Am. Chem. Soc.; 2023; 145, pp. 11130-11139.1:CAS:528:DC%2BB3sXptlyltbc%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37155275]
49. Mi, Y et al. Guest-binding-induced interhetero hosts charge transfer crystallization: selective coloration of commonly used organic solvents. J. Am. Chem. Soc.; 2021; 143, pp. 1553-1561.1:CAS:528:DC%2BB3MXot1ShtA%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33432813]
50. Ji, J et al. The more the slower: self-inhibition in supramolecular chirality induction, memory, erasure, and reversion. J. Am. Chem. Soc.; 2022; 144, pp. 1455-1463.1:CAS:528:DC%2BB38XhtVSlsLw%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/35029384]
51. Tuo, D-H et al. Helical-sense matching facilitates supramolecular copolymerization of helical-chiral Pillar[5]arenes. J. Am. Chem. Soc.; 2024; 146, pp. 31816-31824.1:CAS:528:DC%2BB2cXisVCmt7bE [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/39527493]
52. Ogoshi, T et al. Host-guest complexation of Perethylated Pillar[5]arene with alkanes in the crystal state. Angew. Chem. Int. Ed.; 2015; 54, pp. 9849-9852.1:CAS:528:DC%2BC2MXhtFWgu73F
53. Ohtani, S; Kato, K; Fa, S; Ogoshi, T. Hos-Guest chemistry based on solid-state pillar[n]arenes. Coord. Chem. Rev.; 2022; 462, 214503.1:CAS:528:DC%2BB38XovVamurY%3D
54. Ma, Y et al. per-Hydroxylated Pillar[6]arene: Synthesis, X-ray crystal structure, and host guest complexation. Org. Lett.; 2012; 14, pp. 1532-1535.1:CAS:528:DC%2BC38XjsFegurs%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22401142]
55. Jia, Y et al. Novel conductive metallo-supramolecular polymer AIE gel for multi-channel highly sensitive detection of hydrazine hydrate. Chin. Chem. Lett.; 2023; 34, 1:CAS:528:DC%2BB3sXht1amtLnN 108082.
56. Guan, W-L et al. Macrocycles-assembled AIE supramolecular polymer networks. Coord. Chem. Rev.; 2024; 507, 1:CAS:528:DC%2BB2cXlt1Oru74%3D 215717.
57. Tang, Z-D et al. Pillar[n]arenes-based materials for detection and separation of pesticides. Chin. Chem. Lett.; 2023; 34, 1:CAS:528:DC%2BB3sXhsFynt7w%3D 107698.
58. Zhang, Y-F et al. Novel tripodal-pillar[5]arene-based chemical sensor for efficient detection and removal paraquat by synergistic effect. Sens. Actuators, B.; 2021; 327, 1:CAS:528:DC%2BB3cXitVKgurzM 128885.
59. Wang, Z-H et al. Linear tri-pillar[5]arene-based acceptor for efficiently separate paraquat from water through collaboration effect. Mater. Sci. Eng. C.; 2021; 118, 1:CAS:528:DC%2BB3cXhs1OksrzJ 111358.
60. Jia, Y et al. Rational tuning of binding properties of Pillar[5]arene-based crystalline material by synergistic effect and its application for fluorescent detection and adsorption of 1,2-Ethylenediamine. ACS Sustain. Chem. Eng.; 2021; 9, pp. 16203-16209.1:CAS:528:DC%2BB3MXisFWgtLbF
61. Lin, Q et al. Spongy materials based on supramolecular polymer networks for detection and separation of broad-spectrum pollutants. ACS Sustain. Chem. Eng.; 2019; 7, pp. 14775-14784.1:CAS:528:DC%2BC1MXhsFekur3E
62. Zhang, Q et al. Supramolecular AIE polymer-based rare earth metallogels for the selective detection and high efficiency removal of cyanide and perchlorate. Polym. Chem.; 2021; 12, pp. 2001-2008.2021espc.book...Z1:CAS:528:DC%2BB3MXkvVaqsr8%3D
63. Sun, X-W et al. Tri-pillar[5]arene-based multifunctional stimuli-responsive supramolecular polymer network with conductivity, aggregation-induced emission, thermochromism, fluorescence sensing, and separation properties. Macromolecules; 2020; 54, pp. 373-383.2021MaMol.54.373S
64. Cotruvo, JA. 2017 WHO Guidelines for Drinking Water Quality: First Addendum to the Fourth Edition. J. AWWA; 2017; 109, pp. 44-51.
65. Hübler, C. SupraFit - An Open Source Qt based fitting application to determine stability constants from titration experiments. Chem. Methods; 2022; 2, e202200006.
66. Frisch M. J. et al. Gaussian09, http://www.gaussian.com/ (2009).
67. Frisch, MJ; Pople, JA; Binkley, JS. Self-consistent molecular orbital methods. J. Chem. Phys.; 1984; 80, pp. 3265-3269.1984JChPh.80.3265F1:CAS:528:DyaL2cXhvFOqu7k%3D
68. Humphrey, W; Dalke, A; Schulten, KVMD. Visual molecular dynamics. J. Mol. Graph.; 1996; 14, pp. 33-38.1:CAS:528:DyaK28Xis12nsrg%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/8744570]
69. Lu, T; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem.; 2011; 33, pp. 580-592. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22162017]
70. Lu, T; Chen, Q. Independent gradient model based on Hirshfeld partition: A new method for visual study of interactions in chemical systems. J. Comput. Chem.; 2022; 43, pp. 539-555.1:CAS:528:DC%2BB38XislGkurc%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/35108407]
71. Yao, H et al. Ultralong room temperature phosphorescence and broad color-tunability persistent luminescence via new strategy. Chin. Chem. Lett.; 2025; 36, 1:CAS:528:DC%2BB2MXnt1Kitrc%3D 110375.
72. Yin, M et al. Selective adsorption of aldehyde impurities from propylene oxide using perethylated pillar[5]arene enabled by spatial size matching effect. Sep. Purif. Technol.; 2025; 354, 1:CAS:528:DC%2BB2cXhvV2itbnJ 129401.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© The Author(s) 2025. This work is published under http://creativecommons.org/licenses/by-nc-nd/4.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Perchlorate is a toxic, explosive, and water-soluble pollutant but while efficient and low-cost removal of ClO4− from water is an important challenge, suitable methods for precipitation of ClO4− from water are underdeveloped. Here, we demonstrate a hydrogen-bonded supramolecular polymer network crystal (HBPC) for efficient complexation of ClO4−. The HBPC network is constructed by an A,A’-bis-pyridyl-hydrazone-phenyl conjugate-functionalized pillar[5]arene (PYP5), which self-assembles via the formation of clustered hydrogen-bonds. Two-pyridyl-hydrazone moieties provide coordination sites while the eight ethoxyl moieties on the pillar[5]arene enable multiple hydrogen bonding motifs to tightly bind to the ClO4−. The HBPC network therefore adsorbs ClO4− ions and facilitates the separation of these ions from water.
Perchlorate is a toxic, explosive pollutant, and is difficult to precipitate due to its water solubility. Here, the authors report a pillar[5]arene-based supramolecular polymer network that can bind with perchlorate by clustered hydrogen-bonds for removal from water.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Northwest Normal University, Key Laboratory of Eco-functional Polymer Materials of the Ministry of Education, Key Laboratory of Eco-environmental Polymer Materials of Gansu Province, College of Chemistry and Chemical Engineering, Lanzhou, China (GRID:grid.412260.3) (ISNI:0000 0004 1760 1427)
2 Northwest Minzu University (Northwest University for Nationalities), Key Laboratory of Environment-Friendly Composite Materials of the State Ethnic Affairs Commission, Gansu Provincial Biomass Function Composites Engineering Research Center, College of Chemical Engineering, Lanzhou, China (GRID:grid.412264.7) (ISNI:0000 0001 0108 3408)