Elsevier

Polymer

Volume 233, 26 October 2021, 124192
Polymer

Effect of pore structure on the adsorption capacities to different sizes of adsorbates by ferrocene-based conjugated microporous polymers

https://doi.org/10.1016/j.polymer.2021.124192Get rights and content

Highlights

  • Two conjugated microporous polymers based on carbazole and ferrocene units were synthesized successfully.

  • In comparison with PDCPF, PDCF has similar pore size distribution in the region below 3 nm and extra pores in the region between 10 and 100 nm.

  • The extra mesopores in PDCF exhibit better adsorption capacities to adsorbates with larger sizes.

  • Dye molecules adsorbed in the micropores are harder to desorption than that in the mesopores.

Abstract

Conjugated microporous polymers (CMPs) are potential materials for gas uptake, pollutant adsorption and photocatalysis. However, their relatively wide pore size distribution (PSD) makes it hard to establish the relationship between adsorption capacities and pore structure. Here, we synthesized two CMPs with similar chemical structure based on carbazole and ferrocene units. Poly[1,1′-di(9-carbazolyl)ferrocene] (PDCF) showed hierarchical pore structure with pore sizes distributed in three regions: 0.7–0.85 nm, 1–3 nm and 10–100 nm, while the pore sizes of poly{1,1′-di[4-(9-carbazolyl))phenyl]ferrocene} (PDCPF) were mostly less than 3 nm. The results of gas uptake and dye adsorption capacities show that the extra mesopores of PDCF exhibit better adsorption capacities to adsorbates with larger sizes. Besides, dye molecules adsorbed in the micropores are harder to desorption than that in the mesopores. The intermolecular forces between porous polymers and adsorbates were used to explain these results successfully. Our findings indicate that designing the right pore sizes for porous polymers when adsorbing gases or dyes with a specific molecules size is very important in the future.

Introduction

Porous organic polymers (POPs) have potential applications in gas uptake and separation [[1], [2], [3]], pollutant adsorption [4,5], heterogeneous catalysis [6], etc [7]. During the past few decades, a series of POPs, such as conjugated microporous polymers (CMPs) [8], polymers of intrinsic microporosity (PIMs) [9], hyper-cross-linked polymers (HCPs) [10] and covalent organic frameworks (COFs) [11], have been intensively studied. Among them, CMPs are three-dimensional amorphous materials that combine extended π-conjugation with a permanently microporous skeleton [8]. The conjugated structure offers CMPs stable pore structure and special photoelectric performance.

Carbazole is a commonly used building unit for linear conjugated polymers [12] and CMPs [13]. Han et al. reported the formation of microporous polycarbazole via oxidative coupling polymerization in 2012. The Brunauer–Emmett–Teller (BET) specific surface area of microporous polycarbazole was higher than 2000 m2/g [14]. After that, a series of carbazole-based POPs have been successfully synthesized, and they exhibited good gas uptake and dye removal capacities [[15], [16], [17], [18]].

Ferrocene has a sandwich structure consisting of two cyclopentadienyl rings bound on opposite sides of a central metal atom [19], which makes it easy to form a three-dimensional structure when the hydrogen atoms on the cyclopentadienyl rings are replaced. Therefore, ferrocene is a promising building unit for POPs. Several groups have used ferrocene to build POPs, and the products showed good gas uptake and pollutant adsorption capacities [[20], [21], [22], [23], [24], [25]]. The cyclopentadienyl ring in ferrocene is a conjugated structure, and thus ferrocene could form CMPs when connected by other conjugated units. For example, Yu et al. chose ferrocene and s-triazine as the building units, and the resulting CMP showed good gas uptake capabilities for H2, CH4 and CO2 [26]. Yu et al. also found that CMP based on ferrocene had a high affinity towards I2 [27]. Besides, like many CMPs without ferrocene, CMPs based on ferrocene also had good photocatalysis performance, as shown by Wang et al. [28].

In addition to chemical structure, the pore structure also has a big influence on the adsorption capacity of POPs [2]. Microporous materials (with pore sizes smaller than 2 nm) are generally regarded as promising gas adsorbing materials because their pore sizes are on the order of the molecular size of small gases and their adsorption capacities are high due to their large surface areas [2,29]. For porous carbons, Gogotsi's group showed that smaller pores increase both the heat of adsorption and total volume of adsorbed H2 [30]. Gogotsi's group also found that CO2 sorption is limited by pores smaller than a certain diameter, which decreased from 0.8 nm to 0.5 nm when the pressure decreased from 1 bar to 0.1 bar [31]. Teng et al. studied phenol, iodine and tannic acid adsorption capacities of carbons with different mesopore volumes, and found that the adsorption capacity is an increasing function of the mesopore volume [32]. Thus, understanding the relationship between the adsorption capacity and pore structure is very important. However, most POPs (except COFs) are amorphous structures without exact chemical structure, which makes it hard to study the relationship between the adsorption capacity and pore structure accurately. The synthesized POPs generally have a wide pore size distribution (PSD) in the range of micropore and mesopore [23,33,34].

Herein, two CMPs based on carbazole and ferrocene units were synthesized. In comparison with poly{1,1′-di[4-(9-carbazolyl))phenyl]ferrocene} (PDCPF), poly[1,1′-di(9-carbazolyl)ferrocene] (PDCF) has a similar PSD in the micropore region and a much wider PSD in the mesopore and macropore regions. The results of gas uptake and dye adsorption show that the extra mesopores of PDCF are beneficial to the adsorption of adsorbates with larger sizes. Moreover, dye molecules adsorbed in the micropores are harder to desorption than that in the mesopores.

Section snippets

Materials

1,1-Dibromoferrocene (FcBr) and 4-(9H-Carbozol-9-yl)phenylboronic acid (NBAPC) were purchased from Tokyo Chemical Industry, Japan. Carbazole, tetrakis (triphenylphosphine) palladium(0), cuprous iodide (CuI), Iron(III) chloride (FeCl3), 1,10-phenanthroline monohydrate, potassium bromide (KBr), methanol, acetone, chloroform, tetrahydrofuran, N, N-dimethylformamide (DMF), petroleum ether, dichloromethane and 1,2-Dichloroethane (DCE) were purchased from Aladdin Reagents Co., Ltd., Shanghai, China.

Chemical structure of PDCF and PDCPF

Two CMPs based on ferrocene and carbazole were synthesized here, as shown in Scheme 1. Their chemical structure is similar, PDCPF has two extra p-phenyls in the building units in comparison with PDCF.

The chemical structure of two monomers was confirmed by 1H NMR spectra, 13C NMR spectra and mass spectra, as shown in Figs. S1–S5. DCF and DCPF gave satisfactory spectroscopic data corresponding to their expected molecular structures. FT-IR spectra of PDCF and PDCPF is shown in Fig. S6. The FT-IR

Conclusion

Two CMPs based on ferrocene and carbazole were successfully synthesized. In comparison with PDCF, PDCPF has two extra p-phenyls in the building units. PDCF shows hierarchical pore structure with pore sizes distributed in three regions: 0.7–0.85 nm, 1–3 nm and 10–100 nm. However, the pore sizes of PDCPF are mostly less than 3 nm. In comparison with PDCPF, the extra mesopores in PDCF is beneficial to the adsorption of dye molecules and gas molecules with a larger size, while they have little

CRediT authorship contribution statement

Xinxiu Cao: Writing, Resources. Ruiyuan Wang: Synthetic materials, Characterization. Qi Peng: Synthetic materials, Characterization. Hongwei Zhao: Reviewing. Hui Fan: Reviewing. Huan Liu: Characterization. Qingquan Liu: Conceptualization, Resources.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This work was partly supported by the National Natural Science Foundation of China (Grant No. 51778226), the Natural Science Foundation of Hunan Province, China (Grant No. 2019JJ50164), the Research Foundation of Education Bureau of Hunan Province, China (Grant No. 19C0750).

References (43)

  • G. Kupgan et al.

    Modeling amorphous microporous polymers for CO2 capture and separations

    Chem. Rev.

    (2018)
  • D.S. Ahmed et al.

    Design and synthesis of porous polymeric materials and their applications in gas capture and storage: a review

    J. Polym. Res.

    (2018)
  • X. Wang et al.

    Imidazolium salt-incorporated postcross-linked porous polymers for efficient adsorption of rhodamine B and Cd2+ from aqueous solution

    J. Chem. Eng. Data

    (2020)
  • B. Altava et al.

    Chiral catalysts immobilized on achiral polymers: effect of the polymer support on the performance of the catalyst

    Chem. Soc. Rev.

    (2018)
  • J. Wu et al.

    Porous polymers as multifunctional material platforms toward task-specific applications

    Adv. Mater.

    (2019)
  • J.M. Lee et al.

    Advances in conjugated microporous polymers

    Chem. Rev.

    (2020)
  • L. Tan et al.

    Hypercrosslinked porous polymer materials: design, synthesis, and applications

    Chem. Soc. Rev.

    (2017)
  • M.S. Lohse et al.

    Covalent organic frameworks: structures, synthesis, and applications

    Adv. Funct. Mater.

    (2018)
  • J. Li et al.

    Carbazole-based polymers for organic photovoltaic devices

    Chem. Soc. Rev.

    (2010)
  • Q. Cao et al.

    Recent advance in organic porous polycarbazoles: preparation and properties

    Acta Chim. Sinica

    (2015)
  • Q. Chen et al.

    Microporous polycarbazole with high specific surface area for gas storage and separation

    J. Am. Chem. Soc.

    (2012)
  • Cited by (23)

    View all citing articles on Scopus
    View full text