Next Article in Journal
Fibre-Reinforced Polymers and Steel for the Reinforcement of Wooden Elements—Experimental and Numerical Analysis
Previous Article in Journal
Durability Enhancement Effect of Silica Fume on the Bond Behavior of Concrete–PCM Composites under Environmental Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultramicroporous Polyphenylenes via Diels–Alder Polycondensation Approach

by
Svetlana A. Sorokina
*,
Nina V. Kuchkina
,
Alexander V. Mikhalchenko
,
Irina Yu. Krasnova
,
Dmitry A. Khanin
,
Kirill M. Skupov
and
Zinaida B. Shifrina
*
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov St., 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(9), 2060; https://doi.org/10.3390/polym15092060
Submission received: 11 April 2023 / Revised: 24 April 2023 / Accepted: 24 April 2023 / Published: 26 April 2023
(This article belongs to the Special Issue Recent Progress in Polymer Coatings)

Abstract

:
Development of new microporous organic polymers attracts significant attention due to a wide scope of promising applications. In addition, the synthesis of soluble, non-crosslinking polymers of high surface area and uniform microporosity is very challenging, and the methods for soluble microporous polymers formation are rather limited. In this work, we report a new approach to construct porous polyphenylenes which employs the Diels–Alder polycondensation of multifunctional ethynyl-containing monomers of different spatial architecture with bis(cyclopentadienone)s. The resulting polymers were soluble in common organic solvents, and their structure and properties were assessed by NMR, TGA, DSC, and SEC studies. The polymers demonstrated a specific surface area up to 751 m2·g−1 and ultramicroporous (pore size ≤ 0.6 nm) structure. N2 and CO2 adsorption–desorption data revealed that porosity parameters, e.g., specific surface area and pore sizes, can be tuned selectively by varying the type of monomers and reaction conditions.

1. Introduction

Design of new organic microporous polymers has gained considerable interest over the last two decades due to a vast diversity of potential applications, including gas capture and separation [1,2,3,4], photovoltaics [5,6], chemical sensing [7,8], and catalysis [9,10,11,12,13,14]. These materials are characterized by high specific surface area (SSA), which arises from the voids presented within the polymer architecture and formed during the synthesis due to specific structure of macromolecules and their arrangement. There are two tentatively distinguished approaches to the formation of microporous materials. The general synthetic strategy involves the network formation due to crosslinking of rigid monomers by robust covalent or strong coordination bonds. Crystalline metallic organic frameworks (MOFs) and covalent organic frameworks (COFs), as well as amorphous porous aromatic frameworks (PAFs), conjugated microporous polymers (CMPs), hyper crosslinked polymers (HCPs), etc., are the porous organic polymers (POPs) whose high SSA (up to 6000 m2/g) have been obtained due to network formation [15,16,17]. However, these compounds are insoluble in organic solvents, which significantly hampers their processability and limits the areas of application.
Another approach implies the formation of microporosity due to inefficient packing of highly rigid and contorted polymer chains. The intermolecular interactions in such polymers are precluded due to distorted structure of the monomers, which leads to the formation of interconnected voids and emergence of free volume. The process occurs without network formation. Therefore, the resulting polymers are soluble in common organic solvents [18]. The weak cohesion inter-chain interactions and free volume accessible for solvent molecules also contribute to the solubility of these polymers. This type of POP is characterized by moderate SSA (100–1000 m2/g), with the fraction of micropores usually reaching 50%.
The key advantage of the non-networking POPs is their solution processability. This enables the preparation of different materials, e.g., the feasibility in film casting, commercial roll-to-roll fabrication, membrane formation, etc., [3,8,18,19,20,21,22]. Soluble POPs also may be employed as templating molecules in the synthesis of metal nanoparticles (NPs) [9,12]. Here, the solubility of non-networking POPs offers the advantages of NPs synthesis via thermal decomposition of metal precursors in the polymer solution over the wet impregnation method since more parameters could be finely tuned during the synthesis to afford the well-defined, uniform NPs with predicted size and morphology. The resulting nanocomposites are now of particular interest for catalysis [11,12,23,24,25]. The presence of micro- and mesopores restricts the growth of NPs within the polymer matrix while allowing the easy access and diffusion of substrate molecules in the resulting catalytic composites. In contrast to MOF-based catalysts, whose hydrolytically unstable bonds limit the reaction scope where they can be applied, the exceptional thermo-, chemo-, and hydrostability of POPs allows these materials to be used under harsh reaction conditions, including high water vapor pressure, acidic or alkaline environments, etc. [11,12].
The ease of functionalization, unique topologies, extraordinary stability, and solution processability of non-networking POPs boost the research interest in expanding the chemistry underlying the non-networking POPs synthesis. Along with spirobisindane and polybenzodioxin [20,26], triptecene [27,28,29], and porphirine-based [25] monomers that are widely established structural units in the synthesis of soluble microporous polymers so-called polymers of intrinsic microporosity (PIMs), many other building blocks, such as phenantroline [30], benzylamine [31], bisphosphine [32], and phtalocyanine [33], have been utilized to create microporosity in polymers. Phenylenes of different structures are another class of compounds capable of forming micropores. The general approaches to the synthesis of porous polyphenylenes (PPPhs) are based on the crosslinking of aromatic building blocks, such as Suzuki cross-coupling of 1,2,4,5-tetrabromobenzene and 1,4-benzene diboronic acid, Friedel–Crafts alkylation, or trimerization of ethynyl groups [34,35,36,37,38]. The resulting compounds are characterized by the moderate-to-high surface areas in the range of 470–980 m2/g. The average micropore surface area constitutes half of the total SSA, while certain chemical modifications may increase this value up to 80% [36,37]. PPPhs are the attractive macromolecules due to ease of post-modification, the possibility of introducing of different functional groups, and exceptional thermal properties. They can be used as semiconductors, light harvesting molecules, and gas separation membranes, exhibiting high gas selectivity for H2 relative to CO2, CO, and CH4 [34,35,39,40].
However, the aforementioned approaches allow the synthesis of insoluble crosslinking microporous materials. In the present study we applied the Diels–Alder polycondensation reaction to synthesize soluble porous polyphenylenes. It is known that the reactions used for POPs synthesis should provide more than one covalent bond for each structural unit [18]. In this regard, the Diels–Alder reaction seems to be a promising tool for the synthesis of soluble POPs via step-growth polycondensation of multitopic aromatic monomers, yielding a new arene ring at each condensation step and leading to hyperbranched macromolecular architecture. However, despite the applicability for the synthesis of some structural units which are suitable for production of polymers with high SSA, the Diels–Alder reaction has not yet been explored for the fabrication of POPs.
In this work, we used two multitopic monomers possessing different geometries, namely tetrakis(4-ethynylphenyl)methane and 1,3,5-triethynylbenzene, as the structural units for PPPhs synthesis by the Diels–Alder polycondensation reaction with bulky phenylene-substituted bis(cyclopentadienone)s. Previously, the aforementioned ethynyl-containing compounds were used in the POP synthesis by Sonogashira–Hagihara cross-coupling and Yamamoto polymerization reactions, yielding crosslinked insoluble products [15,17,41,42]. Our approach results in soluble polyphenylenes with SSA up to 751 m2/g, which could be effectively controlled during the synthesis.

2. Materials and Methods

2.1. Materials

Tetraphenylmethane (95%, ABCR, Karlsruhe, Germany), 1,3,5-tribromobenzene (98%, Acros Organics, Geel, Belgium), bromine (99%, Acros Organics, Geel, Belgium) trimethylsilylacetylene (98%, ABCR, Karlsruhe, Germany), CuBr (98%, Sigma-Aldrich, Darmstadt, Germany), CuI (98%, Fluka, Dresden, Germany), tetrakis(triphenylphosphine)palladium(0) (99%, Sigma-Aldrich, Darmstadt, Germany), bis(triphenylphosphine)palladium(II) dichloride (98%, Sigma-Aldrich, Darmstadt, Germany), Bu4NF (1M solution in THF, ABCR, Karlsruhe, Germany), bis(4-bromophenyl)ether (99%, Sigma-Aldrich, Darmstadt, Germany), 1,4-diiodobenzene (98%, ABCR, Karlsruhe, Germany), phenylacetylene (98%, Acros Organics, Geel, Belgium), 1,3-diphenyl-2-propanone (99%, Sigma-Aldrich, Darmstadt, Germany), potassium hydroxide (99%, Acros Organics, Geel, Belgium), glacial acetic acid (99%, Sigma-Aldrich, Darmstadt, Germany), potassium permanganate, triethylamine (99%, Acros Organics, Geel, Belgium), toluene (99.5%, Component Reactive, Moscow, Russia), ethanol (96%, Component Reactive, Moscow, Russia), diphenylether (99%, ABCR, Karlsruhe, Germany), THF ( HPLC grade, Carlo Erba reagents, Milan, Italy), and gases: N2 (99.999%) and CO2 (99.999%) were used as received.

2.2. Measurements

1H and 13C NMR spectra were recorded on the Avance IIIHD-500 MHz NMR spectrometer (Bruker, Bremen, Germany) operating at 500.13 MHz for 1H and at 125.76 MHz for 13C. Chemical shifts are given in parts per million (ppm), using the solvent signal as a reference. CD2Cl2 was used as solvent for all NMR measurements (δ(1H) = 5.35 ppm; δ(13C) = 53.4 ppm).
Thermogravimetric analysis (TGA) was carried out on a Shimadzu DTG-60H (Shimadzu GmbH, Kyoto, Japan) at a heating rate of 10°/min in argon.
Differential scanning calorimetry (DSC) measurements were performed on a DSC-3 (Mettler Toledo, Zurich, Switzerland) at a heating rate of 10°/min in argon.
Size-exclusion chromatography (SEC) analyses were carried out using a Shimadzu LC-20AT chromatograph (Shimadzu GmbH, Kyoto, Japan ) equipped with PSS SDV analytical columns 100 Å and 100,000 Å (particle size 10 μm, 8 × 300 mm). Detection was achieved with refractive index detector. SEC was performed in THF at a flow rate of 1.0 mL min−1, with polystyrene as a standard.
N2 and CO2 adsorption–desorption isotherms were obtained at 77 and 273 K, respectively, with the Surface Area and Pore Size Analyzer System 3P Micro 200 (3P Instruments GmbH & Co. KG, Odelzhausen, Germany) in the range of 10−3–1 bar. Before the measurements, the samples were degassed at 353 K for 24 h until the residual pressure reached 5 Pa. The Brunauer–Emmett–Teller (BET) equation was applied to the nitrogen adsorption isotherm data according to the Rouquerol criteria [43]. N2 cross-sectional area A and adsorbed N2 density ρads were taken as 0.162 nm2 and 0.808 g·mL−1, respectively. Non-linear density functional theory (NLDFT) method was applied to the carbon dioxide adsorption isotherm data using NovaWin, version 11.04, Quantachrome Instruments, considering CO2 cross-sectional area A of 0.210 nm2, adsorbed CO2 density ρads of 1.044 g mL−1, saturated vapor pressure of the adsorbate p0 of 3.485 MPa, and affinity coefficient β of 0.35 [44]. Pore size distributions and pore volumes were obtained according to the NLDFT method.

2.3. Synthetic Procedures

3,3′-(1,4-Phenylene)bis(2,4,5-triphenylcyclopenta-2,4-dien-1-one) (B21) and 3,3′-(oxydi-1,4-phenylene)-bis(2,4,5-triphenylcyclopenta-2,4-dien-1-one) (B22) were synthesized as described in [45].
Tetrakis(4-ethynylphenyl)methane (A4) was synthesized according to the reported procedure [46]. 1,3,5-Triethynylbenzene (A3) was synthesized as described in the literature [47].
General Procedure for the Synthesis of PPPhs: All procedures were performed using a Schlenk flask under argon atmosphere in diphenyl ether at 165 °C using total monomer concentrations of 0.01 or 0.05 mol L−1, at equifunctional molar ratios of monomers.
In the typical procedure, a 50 cm3 Schlenk flask was charged with the bis-cyclopentadienone (B2) (0.168 mmol), tetrakis(4-ethynylphenyl)methane (A4) (0.084 mmol), or 1,3,5-triethynelbenzene (A3) (0.112 mmol) and diphenyl ether (amount for chosen concentration). The reaction was carried out for the desired time (see Table 1) at 165 °C, with stirring, in argon atmosphere, and then the mixture was allowed to cool, followed by precipitation into ethanol. The crude polymer was purified by multiple reprecipitation from chloroform into ethanol and dried in vacuo at 80 °C for 24 h. The reaction conditions and molecular weights of polymers prepared are shown in Table 1.

3. Results and Discussion

3.1. Synthesis and Characterization of PPPhs

Polyphenylenes were synthesized via Diels–Alder polycondensation reaction using two pairs of monomers (A4 + B2) and (A3 + B2) in diphenyl ether as a solvent (Figure 1). Phenyl substituted bis(cyclopentadienone)s were used as B2 monomers, and tetra- or tri-ethynyl containing compounds, namely tetrakis(4-ethynylphenyl)methane and 1,3,5-triethynylbenzene, as A4 and A3 monomers, respectively, were used as multibranching units. B2 monomers differed in the central structural group to estimate the influence of chemical composition on the surface area in molecular level, while A3 and A4 differed in spatial geometry of the molecule, namely trigonal pyramidal (A3) and tetrahedral (A4), which further implemented a macromolecular growth in different directions.
It is well-known that the polycondensation reaction of the multifunctional monomers is accompanied with the fast gelation due to the formation of polymer network. Since the aim of this work is the synthesis of non-crosslinking polymers, the reaction conditions, such as temperature, concentration of the monomers, and reaction time, were carefully adjusted to afford the formation of soluble macromolecules and to avoid the undesired crosslinking. The synthesis conditions and some characterization data are presented in Table 1. All the polymers were synthesized at equifunctional ratios of monomers, i.e., A4:B2 = 1:2 for tetrahedral A4 monomer and A3:B2 = 2:3 for trigonal A3. The molar concentration of the monomers dramatically impacts the formation of soluble polymer, and the polymer growth without fast gelation occurred in relatively dilute solution only. The effect was more pronounced for A4, where the soluble polymers were obtained at total a monomer concentration of 0.01 mol·L−1 and below. For polymers based on A3, no gelation was observed at 0.05 mol·L−1 for B22 monomer, while the reaction with more rigid B21 was more sensitive to an increase in molar concentration, yielding the insoluble product. The optimum temperature of synthesis was found to be 165 °C. The further increase in temperature led to formation of high fraction of crosslinking polymer along with soluble product. The reaction time influenced mainly the molecular weight of the resulted polymers, and no noticeable effect on the crosslinking process was observed at appropriate concentrations of the monomers.
The molecular weight of the polymers was determined by SEC. Some examples of SEC chromatograms are presented in Figure S1. Depending on the reaction conditions, the apparent molecular masses were in the range of 16–35 kDa for polymers based on tetrahedral monomer A4. For A3 monomer, the use of B22 provided the product with molecular weight up to 789 kDa. It should be noted that the use of a pair of monomers, A3 and B22, allowed the synthesis at higher concentration in comparison with the other pairs of monomers. Obviously, the increase in molar concentration dramatically affected the molecular weight of the polymers. Moreover, as can be seen from Table 1, the use of B22 allows the formation of polymers of higher molecular weight for both A3 and A4 compared with B21, apparently due to the greater reactivity of B22.
The obtained polymers were powders of a beige color. The PPPhs were achieved with high yields. The PPPhs showed good solubility in common organic solvents, such as tetrahydrofuran, chloroform, dichloromethane, toluene, o-xylol, etc.
The structure of the polymers was assessed by 1H and 13C NMR spectroscopy. Figure S2a shows the 1H NMR spectrum of Polymer 2 based on A4 and B21 bearing phenylene bridging group. The spectrum contains a set of signals related to aromatic protons at 6–8.0 ppm, which are broad and overlapping. This is consistent with the structure of the polymer composed of phenylene units. However, a signal at 7.63 ppm is clearly distinguished and ascribed to the protons of pentaphenyl-substituted benzene rings, which are newly formed during the Diels–Alder polycondensation. A weak signal at 3.17 ppm corresponds to the protons of the ethynyl groups. Obviously, the observed ethynyl groups originate from the incomplete Diels–Alder reaction of multitopic ethynyl-containing monomer. According to the integral ratio of the signals of ethynyl proton and protons of pentasubstituted benzene, their ratio is 1:14. This means that, on average, one ethynyl group is preserved in two repetitive units of polymer, and the structure can be presented as depicted in Figure S2b. The macromolecule growth occurred mainly in four directions; however, one ethynyl group per two tetraphenylmethane units remained unreacted. The results are in accordance with previous findings obtained for hyperbranched polypyridylphenylenes [48], which were synthesized by A6 and B2 Diels–Alder polycondensation. In that case, it was established that the growth of the most branched polymer has proceeded in five out of six possible directions. In the 13C NMR spectrum (Figure S2c), more signals can be distinguished. In particular, the peak at 63.95 ppm is ascribed to the tetrasubstituted carbon atom of the tetraphenylmethane unit. The low intensity peak at 200 ppm corresponds to the carbonyl groups of bis(cyclopentadienone), which are presented as terminal groups of the polymer. The carbon atoms of the acetylene groups are observed at 77.65, 78.16, and 83.17 ppm.
The 1H NMR spectrum of PPPhs based on the trifunctional unit (Figure S3a) is similar to that described above. However, the signal attributed to pentasubstituted phenylene moieties is shifted to a high field region and overlapped with the other signals, making reliable quantification of acetylene groups impossible. The 13C NMR spectrum (Figure S3b) confirms a very low content of ethynyl groups.
The thermal behavior of porous polyphenylenes were examined by TGA under nitrogen atmosphere and by DSC. As shown in Figure S4, the synthesized PPPhs exhibited exceptional thermal stability. The 5% weight loss was detected when the temperature was above 510 °C for all the samples. The results suggest that PPPhs are promising materials for high-temperature applications.
DSC results (Figure S5) determined that PPPhs are the glassy polymers, with glass transition temperature above the decomposition temperature. The results confirmed the exceptional rigidity of the synthesized structures that is a necessary feature of the high free-volume polymers.

3.2. Gas Adsorption Properties of PPPhs

The porous properties were investigated by N2 and CO2 adsorption–desorption measurements. Figure 2a,c,d show the nitrogen adsorption isotherms for tetra-branched and tri-branched polymers, respectively, and Figure 2b,e depict CO2 isotherms. Similar to other POPs, PPPhs demonstrated the hysteresis loops, which are often ascribed to the swelling of the polymer molecule in the presence of sorbate. Another possible explanation is the elastic deformation of the material which could arise at the desorption step. As a result, the penetrant molecule is captured in the matrix [18,49]. Nevertheless, similar behavior is often observed for microporous materials, especially for those having a slit-pore geometry [18,50].
Remarkably, a distinct sorption capacity was observed for the polymers on the basis of tetrahedral and trigonal monomers. The results of adsorption–desorption experiments are summarized in Table 2. In contrast to trigonal-based polymers, tetrahedral-based samples exhibited significant differences in N2 and CO2 porosity measurements. Specifically, a very low porosity measured with N2 was determined for the Polymers 3, 5, and 6. As a consequence, the calculated specific surface areas did not exceed 38 m2 g−1. However, Sample 2 demonstrated the highest specific surface area, SBET = 667 m2 g−1. The opposite situation was observed when turning to CO2 isotherms. In this case, the high values of SSA were determined for all polymers. A closer look at the porosity parameters revealed that PPPhs 3, 5, and 6 are ultramicroporous in nature. Their pore volumes are in the range of 0.042–0.110 cm2 g−1, and the pore diameters are 0.6 nm. Polymer 2 is characterized by a larger pore volume, up to 0.174 cm2 g−1, and the average pore width equals 1.0 nm, indicating that this polymer is rather microporous. The pore size distributions for the samples are given in Figure 3a. This difference explains such distinct porosity values obtained with different sorbate molecules: N2 and CO2. CO2 is a linear molecule that resolves more easily the ultramicropores. Moreover, CO2 adsorption measurements are performed at 273 K, in contrast to 77 K of N2 adsorption. The higher temperature aids in kinetic flexibility of the polymer chains and favors the diffusion of sorbate molecules, which positively influences the adsorption process. This enables easier access of CO2 to ultramicropores in comparison with N2. The results are consistent with previous reports suggesting that CO2 and H2 are more preferable for evaluation of porosity in ultramicroporous materials [18,49,50].
Nevertheless, Polymer 2 demonstrated a steep increase in the amount of adsorbed N2 at low P/P0, which is indicative of the presence of micropores. The micropore surface area, calculated by the T-plot method, was estimated to be 350 m2 g−1, indicating that almost 50% of the total surface area is composed of micropores. Considering the similar chemical structure of Polymers 2 and 3 (they are constituted of the similar building blocks (A4 and B21)), the ultramicroporous character of the latter can be attributed to a higher molecular weight. The high molecular weight may promote the compactization of the macromolecule and contribute to a denser polymer molecule arrangement [51].
When switching to monomer B22 containing the oxygen-bridging group (Polymers 5 and 6), the surface areas derived from CO2 isotherms were lower than those of the polymers composed of the more rigid B21 monomer. Obviously, this related to a larger conformational freedom, reducing the overall porosity. Nevertheless, both samples demonstrated moderate specific surface areas, S = 238 and S = 443 m2 g−1 for Polymers 5 and 6, respectively. The mean pore sizes were determined to be 0.6 nm, confirming that the polymers were still ultramicroporous. The polymer with the higher molecular weight had the larger SSA, while the pore size was affected insignificantly in comparison with the B21 monomer.
It should be noted that design of the porous organic polymers possessing uniformly sized ultramicropores is still very challenging [52,53,54]. The lack of the N2 adsorption observed for Polymers 3, 5, and 6 confirms the absence of the large pores, which could be accessible for N2 molecule. Narrow size distribution, along with ultramicroporosity that was achieved for tetrahedral-based polyphenylenes, is of particular interest for application in gas separation because of the excellent selectivity that is usually observed for uniformly sized ultramicroporous materials.
The use of 1,3,5-triethynylbenzene as a tri-branched unit also yielded the PPPhs with high surface areas up to 515 m2 g−1. Again, the values calculated from CO2 adsorption isotherms exceeded those of N2-derived ones; however, the difference was not crucial. The bis(cyclopentadienone) B21 containing the phenylene-bridging group was more potent for microporosity formation, resulting in the macromolecules of higher surface areas in comparison with the B22 monomer, as was observed earlier for tetraphenylmethane-based polymers. Trigonal geometry of the monomer A3 apparently provided the appearance of larger pores in the range of 0.8–1.2 nm, although ultramicropores were also preserved (Figure 3b). Increase in the overall pore volumes and widths is most likely related to a less dense packing of the polymer chains since a larger free volume is accessible when the macromolecule growth is in three directions rather than four.

4. Conclusions

In the present work, we showed that Diels–Alder polycondensation is a promising approach for construction of non-crosslinking microporous polymers. By varying the synthetic conditions, the soluble hyperbranched polyphenylenes possessing ultramicroporosity were successfully synthesized. The degree of branching and spatial arrangement of starting monomer had a strong impact on the porosity parameters, e.g., surface areas and pore volumes. The monomer with tetrahedral geometry was beneficial over the trigonal planar one to construct the polymers of ultramicroporous nature and, in some cases, possessed the larger surface area. Importantly, employment of tetra-branched monomer provided ultramicroporous polymers with narrow size distribution, while utilization of the tri-branched unit led to an increase in the pore sizes and formation of micropores along with ultramicropores, as confirmed by DFT analysis of pore size distributions. The results of the study open a new way for rational design of ultramicroporous non-networking polymers. Taking advantages of exceptional chemical and thermal stability and ultramicroporous structure, the developed PPPhs are promising materials for further investigation in a plethora of applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15092060/s1, Figure S1: SEC chromatograms of PPPhs 9 (a), 6 (b), and 11 (c).; Figure S2: NMR 1H (a) and 13C (b) spectra and the proposed structure of porous polyphenylene (c); Figure S3: NMR 1H (a) and 13C (b) spectra of PPPh-8; Figure S4: TGA results of the polymers; Figure S5: DSC curves for PPPhs.

Author Contributions

Conceptualization, S.A.S. and Z.B.S.; methodology and visualization, N.V.K.; investigation, I.Y.K. and A.V.M.; resources and validation, K.M.S. and D.A.K.; writing—original draft preparation, S.A.S.; writing—review and editing and supervision, Z.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 22-73-00087.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data are available upon request.

Acknowledgments

The contribution of the Ministry of Science and Higher Education of the Russian Federation (Contract/agreement No. 075-00697-22-00) in recording NMR spectra is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Y.; Ghanem, B.S.; Ali, Z.; Hazazi, K.; Han, Y.; Pinnau, I. Recent Progress on Polymers of Intrinsic Microporosity and Thermally Modified Analogue Materials for Membrane-Based Fluid Separations. Small Struct. 2021, 2, e2100049. [Google Scholar] [CrossRef]
  2. Foster, A.B.; Tamaddondar, M.; Luque-Alled, J.M.; Harrison, W.J.; Li, Z.; Gorgojo, P.; Budd, P.M. Understanding the Topology of the Polymer of Intrinsic Microporosity PIM-1: Cyclics, Tadpoles, and Network Structures and Their Impact on Membrane Performance. Macromolecules 2020, 53, 569–583. [Google Scholar] [CrossRef]
  3. Wang, Y.; Ghanem, B.S.; Han, Y.; Pinnau, I. State-of-the-art polymers of intrinsic microporosity for high-performance gas separation membranes. Curr. Opin. Chem. Eng. 2022, 35, 100755. [Google Scholar] [CrossRef]
  4. Mohamed, M.G.; El-Mahdy, A.F.M.; Kotp, M.G.; Kuo, S.-W. Advances in porous organic polymers: Syntheses, structures, and diverse applications. Mater. Adv. 2022, 3, 707–733. [Google Scholar] [CrossRef]
  5. Bildirir, H.; Gregoriou, V.G.; Avgeropoulos, A.; Scherf, U.; Chochos, C.L. Porous organic polymers as emerging new materials for organic photovoltaic applications: Current status and future challenges. Mater. Horiz. 2017, 4, 546–556. [Google Scholar] [CrossRef]
  6. Wong, Y.-L.; Tobin, J.M.; Xu, Z.; Vilela, F. Conjugated porous polymers for photocatalytic applications. J. Mater. Chem. A 2016, 4, 18677–18686. [Google Scholar] [CrossRef]
  7. Xiong, J.-B.; Ban, D.-D.; Zhou, Y.-J.; Du, H.-J.; Zhao, A.-W.; Xie, L.-G.; Liu, G.-Q.; Chen, S.-R.; Mi, L.-W. Fluorescent porous organic polymers for detection and adsorption of nitroaromatic compounds. Sci. Rep. 2022, 12, 15876. [Google Scholar] [CrossRef]
  8. Marken, F.; Carta, M.; McKeown, N.B. Polymers of Intrinsic Microporosity in the Design of Electrochemical Multicom-ponent and Multiphase Interfaces. Anal. Chem. 2021, 93, 1213–1220. [Google Scholar] [CrossRef]
  9. Kaur, P.; Hupp, J.T.; Nguyen, S.T. Porous Organic Polymers in Catalysis: Opportunities and Challenges. ACS Catal. 2011, 1, 819–835. [Google Scholar] [CrossRef]
  10. Zhang, T.; Xing, G.; Chen, W.; Chen, L. Porous organic polymers: A promising platform for efficient photocatalysis. Mater. Chem. Front. 2020, 4, 332–353. [Google Scholar] [CrossRef]
  11. Ji, G.; Zhao, Y.; Liu, Z. Design of porous organic polymer catalysts for transformation of carbon dioxide. Green Chem. Eng. 2022, 3, 96–110. [Google Scholar] [CrossRef]
  12. Modak, A.; Ghosh, A.; Bhaumik, A.; Chowdhury, B. CO2 hydrogenation over functional nanoporous polymers and metal-organic frameworks. Adv. Colloid Interface Sci. 2021, 290, e102349. [Google Scholar] [CrossRef] [PubMed]
  13. Deng, Z.; Zhao, H.; Cao, X.; Xiong, S.; Li, G.; Deng, J.; Yang, H.; Zhang, W.; Liu, Q. Enhancing Built-in Electric Field via Molecular Dipole Control in Conjugated Microporous Polymers for Boosting Charge Separation. ACS Appl. Mater. Interfaces 2022, 14, 35745–35754. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, W.; Deng, Z.; Deng, J.; Au, C.-T.; Liao, Y.; Yang, H.; Liu, Q. Regulating the exciton binding energy of covalent triazine frameworks for enhancing photocatalysis. J. Mater. Chem. A 2022, 10, 22419–22427. [Google Scholar] [CrossRef]
  15. Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J.M.; et al. Targeted Synthesis of a Porous Aromatic Framework with High Stability and Exceptionally High Surface Area. Angew. Chem. Int. Ed. 2009, 48, 9457–9460. [Google Scholar] [CrossRef]
  16. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, S.B.; Choi, E.; Yazaydin, A.Ö.; Snurr, R.Q.; O’Keeffe, M.; Kim, J.; et al. Ultrahigh Porosity in Metal-Organic Frameworks. Science 2010, 329, 424–428. [Google Scholar] [CrossRef]
  17. Zhang, W.; Zuo, H.; Cheng, Z.; Shi, Y.; Guo, Z.; Meng, N.; Thomas, A.; Liao, Y. Macroscale Conjugated Microporous Polymers: Controlling Versatile Functionalities Over Several Di-mensions. Adv. Mater. 2022, 34, e2104952. [Google Scholar] [CrossRef]
  18. McKeown, N.B. Polymers of Intrinsic Microporosity (PIMs). Polymer 2020, 202, e122736. [Google Scholar] [CrossRef]
  19. Golubev, G.S.; Volkov, V.V.; Borisov, I.L.; Volkov, A.V. High free volume polymers for pervaporation. Curr. Opin. Chem. Eng. 2022, 36, e100788. [Google Scholar] [CrossRef]
  20. McKeown, N.B. The synthesis of polymers of intrinsic microporosity (PIMs). Sci. China Chem. 2017, 60, 1023–1032. [Google Scholar] [CrossRef]
  21. Kirk, R.A.; Putintseva, M.; Volkov, A.; Budd, P.M. The potential of polymers of intrinsic microporosity (PIMs) and PIM/graphene composites for pervaporation membranes. BMC Chem. Eng. 2019, 1, e18. [Google Scholar] [CrossRef]
  22. Ponomarev, I.I.; Volkova, Y.A.; Ponomarev, I.I.; Razorenov, D.Y.; Skupov, K.M.; Nikiforov, R.Y.; Chirkov, S.V.; Ryzhikh, V.E.; Belov, N.A.; Alentiev, A.Y. Polynaphthoylenebenzimidazoles for gas separation–Unexpected PIM relatives. Polymer 2022, 238, e124396. [Google Scholar] [CrossRef]
  23. Kathiresan, M. Metal-Ion/Metal Nanoparticle-Anchored Porous Organic Polymers as Efficient Catalysts for Organic Trans-formations–A Recent Overview. Chem. Asian J. 2023, 18, e202201299. [Google Scholar] [CrossRef]
  24. Giri, A.; Patra, A. Porous Organic Polymers: Promising Testbed for Heterogeneous Reactive Oxygen Species Mediated Photocatalysis and Nonredox CO 2 Fixation. Chem. Rec. 2022, 22, e202200071. [Google Scholar] [CrossRef]
  25. Yi, J.; Jeong, H.Y.; Shin, D.Y.; Kim, C.; Lee, S.J. Mn(III)-Porphyrin Containing Heterogeneous Catalyst based on Microporous Polymeric Constituents as a New Class of Catalyst Support. Chemcatchem 2018, 10, 3974–3977. [Google Scholar] [CrossRef]
  26. Budd, P.M.; Ghanem, B.S.; Makhseed, S.; McKeown, N.B.; Msayib, K.J.; Tattershall, C.E. Polymers of intrinsic microporosity (PIMs): Robust, solution-processable, organic nanoporous materials. Chem. Commun. 2004, 2, 230–231. [Google Scholar] [CrossRef] [PubMed]
  27. Carta, M.; Croad, M.; Malpass-Evans, R.; Jansen, J.C.; Bernardo, P.; Clarizia, G.; Friess, K.; Lanč, M.; McKeown, N.B. Triptycene Induced Enhancement of Membrane Gas Selectivity for Microporous Tröger’s Base Polymers. Adv. Mater. 2014, 26, 3526–3531. [Google Scholar] [CrossRef]
  28. Ghanem, B.S.; Swaidan, R.; Litwiller, E.; Pinnau, I. Ultra-Microporous Triptycene-based Polyimide Membranes for High-Performance Gas Separation. Adv. Mater. 2014, 26, 3688–3692. [Google Scholar] [CrossRef]
  29. Comesaña-Gándara, B.; Chen, J.; Bezzu, C.G.; Carta, M.; Rose, I.; Ferrari, M.C.; Esposito, E.; Fuoco, A.; Jansen, J.C.; McKeown, N.B. Redefining the Robeson upper bounds for CO2/CH4 and CO2/N2 separations using a series of ultrapermeable benzotriptycene-based polymers of intrinsic microporosity. Energy Env. Sci. 2019, 12, 2733–2740. [Google Scholar] [CrossRef]
  30. Gunasekar, G.H.; Yoon, S. A phenanthroline-based porous organic polymer for the iridium-catalyzed hydrogenation of carbon dioxide to formate. J. Mater. Chem. A 2019, 7, 14019–14026. [Google Scholar] [CrossRef]
  31. Masuda, S.; Mori, K.; Kuwahara, Y.; Yamashita, H. PdAg nanoparticles supported on resorcinol-formaldehyde polymers containing amine groups: The pro-motional effect of phenylamine moieties on CO2 transformation to formic acid. J. Mater. Chem. A 2019, 7, 16356–16363. [Google Scholar] [CrossRef]
  32. Gunasekar, G.H.; Padmanaban, S.; Park, K.; Jung, K.-D.; Yoon, S. An Efficient and Practical System for the Synthesis of N,N-Dimethylformamide by CO2 Hydrogenation using a Heterogeneous Ru Catalyst: From Batch to Continuous Flow. ChemSusChem 2020, 13, 1735–1739. [Google Scholar] [CrossRef] [PubMed]
  33. Mackintosh, H.J.; Budd, P.M.; McKeown, N.B. Catalysis by microporous phthalocyanine and porphyrin network polymers. J. Mater. Chem. 2008, 18, 573–578. [Google Scholar] [CrossRef]
  34. Yuan, S.; Dorney, B.; White, D.; Kirklin, S.; Zapol, P.; Yu, L.; Liu, D.-J. Microporous polyphenylenes with tunable pore size for hydrogen storage. Chem. Commun. 2010, 46, 4547–4549. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, L.; Honsho, Y.; Seki, S.; Jiang, D. Light-Harvesting Conjugated Microporous Polymers: Rapid and Highly Efficient Flow of Light Energy with a Porous Polyphenylene Framework as Antenna. J. Am. Chem. Soc. 2010, 132, 6742–6748. [Google Scholar] [CrossRef]
  36. Kovalev, A.I.; Pastukhov, A.V.; Tkachenko, E.S.; Klemenkova, Z.S.; Kuvshinov, I.R.; Khotina, I.A. Polyphenylenes with Phen-1,3,5-triyl Branching Moieties Based on p-Diacetylbenzene: Synthesis and Study of Porosity and Heat Resistance. Polym. Sci. Ser. C 2020, 62, 205–213. [Google Scholar] [CrossRef]
  37. Lee, W.-E.; Han, D.-C.; Choi, H.-J.; Sakaguchi, T.; Lee, C.-L.; Kwak, G. Remarkable Change in Fluorescence Emission of Poly(diphenylacetylene) Film via in situ Desilylation Reaction: Correlation with Variations in Microporous Structure, Chain Conformation, and Lamellar Layer Distance. Macromol. Rapid Commun. 2011, 32, 1047–1051. [Google Scholar] [CrossRef]
  38. Sedláček, J.; Sokol, J.; Zedník, J.; Faukner, T.; Kubů, M.; Brus, J.; Trhlíková, O. Homo- and Copolycyclotrimerization of Ar-omatic Internal Diynes Catalyzed with Co2(CO)8: A Facile Route to Microporous Photoluminescent Polyphenylenes with Hyperbranched or Crosslinked Architecture. Macromol. Rapid Commun. 2018, 39, 1700518. [Google Scholar] [CrossRef]
  39. Miyake, J.; Watanabe, T.; Shintani, H.; Sugawara, Y.; Uchida, M.; Miyatake, K. Reinforced Polyphenylene Ionomer Membranes Exhibiting High Fuel Cell Performance and Mechanical Durability. ACS Mater. Au 2021, 1, 81–88. [Google Scholar] [CrossRef]
  40. Li, Y.; Zhou, Z.; Shen, P.; Chen, Z. Two-dimensional polyphenylene: Experimentally available porous graphene as a hydrogen purification mem-brane. Chem. Commun. 2010, 46, 3672–3674. [Google Scholar] [CrossRef]
  41. Zang, J.; Zhu, Z.; Mu, P.; Sun, H.; Liang, W.; Yu, F.; Chen, L.; Li, A. Synthesis and Properties of Conjugated Microporous Polymers Bearing Pyrazine Moieties with Macroscopically Porous 3D Networks Structures. Macromol. Mater. Eng. 2016, 301, 1104–1110. [Google Scholar] [CrossRef]
  42. Xiang, Z.; Wang, D.; Xue, Y.; Dai, L.; Chen, J.-F.; Cao, D. PAF-derived nitrogen-doped 3D Carbon Materials for Efficient Energy Conversion and Storage. Sci. Rep. 2015, 5, 8307. [Google Scholar] [CrossRef] [PubMed]
  43. Rouquerol, J.; Rouquerol, F.; Llewellyn, P.; Maurin, G.; Sing, K.S. Adsorption by Powders and Porous Solids: Principles, Methodology and Applications; Academic Press: Amsterdam, The Netherlands, 2012. [Google Scholar]
  44. Linares-Solano, A. Commentary on the paper “On the adsorption affinity coefficient of carbon dioxide in microporous car-bons” by E.S. Bickford et al. (Carbon 2004; 42: 1867–71). Carbon 2005, 43, 658–660. [Google Scholar] [CrossRef]
  45. Rusanov, A.L.; Keshtov, M.L.; Keshtova, S.V.; Petrovskii, P.V.; Shchegolikhin, A.N.; Kirillov, A.A.; Kireev, V.V. New bis-tetraarylcyclopentadienones. Russ. Chem. Bull. 1998, 47, 318–320. [Google Scholar] [CrossRef]
  46. Galoppini, E.; Gilardi, R. Weak hydrogen bonding between acetylenic groups: The formation of diamondoid nets in the crystal structure of tetrakis(4-ethynylphenyl)methane. Chem. Commun. 1999, 2, 173–174. [Google Scholar] [CrossRef]
  47. Luo, J.; Liu, X.; Ma, M.; Tang, J.; Huang, F. Dendritic poly(silylene arylacetylene) resins based on 1,3,5-triethynylbenzene. Eur. Polym. J. 2020, 129, 109628. [Google Scholar] [CrossRef]
  48. Kuchkina, N.V.; Zinatullina, M.S.; Serkova, E.S.; Vlasov, P.S.; Peregudov, A.S.; Shifrina, Z.B. Hyperbranched pyridylphenylene polymers based on the first-generation dendrimer as a multifunc-tional monomer. RSC Adv. 2015, 5, 99510–99516. [Google Scholar] [CrossRef]
  49. Minelli, M.; Paul, D.R.; Sarti, G.C. On the interpretation of cryogenic sorption isotherms in glassy polymers. J. Membr. Sci. 2017, 540, 229–242. [Google Scholar] [CrossRef]
  50. Breunig, M.; Dorner, M.; Senker, J. Ultramicroporous polyimides with hierarchical morphology for carbon dioxide separation. J. Mater. Chem. A 2021, 9, 12797–12806. [Google Scholar] [CrossRef]
  51. Zhang, Z.; Wang, Q.; Liu, H.; Li, T.; Ren, Y. Ultramicroporous Organophosphorus Polymers via Self-Accelerating P–C Coupling Reactions: Kinetic Effects on Crosslinking Environments and Porous Structures. J. Am. Chem. Soc. 2022, 144, 11748–11756. [Google Scholar] [CrossRef]
  52. Das, S.; Heasman, P.; Ben, T.; Qiu, S. Porous Organic Materials: Strategic Design and Structure–Function Correlation. Chem. Rev. 2017, 117, 1515–1563. [Google Scholar] [CrossRef] [PubMed]
  53. Lee, J.-S.M.; Cooper, A.I. Advances in Conjugated Microporous Polymers. Chem. Rev. 2020, 120, 2171–2214. [Google Scholar] [CrossRef] [PubMed]
  54. Tian, Y.; Zhu, G. Porous Aromatic Frameworks (PAFs). Chem. Rev. 2020, 120, 8934–8986. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Proposed structures of polyphenylenes based on multifunctional monomers by Diels–Alder polycondensation.
Figure 1. Proposed structures of polyphenylenes based on multifunctional monomers by Diels–Alder polycondensation.
Polymers 15 02060 g001
Figure 2. The nitrogen adsorption and desorption isotherms of PPPhs measured at 77 K (a,c,d). CO2 adsorption and desorption isotherms of PPPhs measured at 273 K (b,e).
Figure 2. The nitrogen adsorption and desorption isotherms of PPPhs measured at 77 K (a,c,d). CO2 adsorption and desorption isotherms of PPPhs measured at 273 K (b,e).
Polymers 15 02060 g002
Figure 3. Pore size distribution curves of PPPhs calculated by NLDFT. (a) PPPh-2,3,5 and (b) PPPh-8,12.
Figure 3. Pore size distribution curves of PPPhs calculated by NLDFT. (a) PPPh-2,3,5 and (b) PPPh-8,12.
Polymers 15 02060 g003
Table 1. Reaction conditions and molecular characteristics.
Table 1. Reaction conditions and molecular characteristics.
Polymer aMonomer AMonomer BC b, mol L−1Time, hMw c·10−3,
g·mol−1
Mw/MnYield,
%
1A4B210.0510Insoluble-99.0
2A4B210.011016.31.6976.2
3A4B210.011735.12.1188.6
4A4B220.056Insoluble-99.0
5A4B220.011023.22.2188.5
6A4B220.012047.42.5695.4
7A3B210.0510Insoluble-99.2
8A3B210.011017.22.3578.7
9A3B210.011723.42.4796.3
10A3B220.17Insoluble-99.1
11A3B220.0510317.33.5375.3
12A3B220.0522789.320.6198.0
a Reactions were conducted in diphenyl ester at 165 °C; b Total molar concentration of the monomers A3 (or A4) + B2; c Apparent molecular weights determined by SEC with polystyrene as a standard.
Table 2. Surface areas and porosity parameters of porous polyphenylenes.
Table 2. Surface areas and porosity parameters of porous polyphenylenes.
PolymerSBET a (N2),
m2 g−1
SDFT b (CO2),
m2 g−1
Vpore c (CO2),
cm3 g−1
D d, nm
26677510.1741.0
3195530.1100.6
5382380.0420.6
6234430.0970.6
82473320.0610.8–1.2
93134370.0970.8–1.2
113564680.0980.8–1.2
124045150.1070.8–1.2
a Specific surface areas are calculated on the basis of the BET equation. b SSA are calculated on the basis of NLDFT with a slit-pore mode. c Pore volumes are calculated by DFT method from CO2 isotherms. d Mean diameters are calculated by DFT method from CO2 isotherms.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sorokina, S.A.; Kuchkina, N.V.; Mikhalchenko, A.V.; Krasnova, I.Y.; Khanin, D.A.; Skupov, K.M.; Shifrina, Z.B. Ultramicroporous Polyphenylenes via Diels–Alder Polycondensation Approach. Polymers 2023, 15, 2060. https://doi.org/10.3390/polym15092060

AMA Style

Sorokina SA, Kuchkina NV, Mikhalchenko AV, Krasnova IY, Khanin DA, Skupov KM, Shifrina ZB. Ultramicroporous Polyphenylenes via Diels–Alder Polycondensation Approach. Polymers. 2023; 15(9):2060. https://doi.org/10.3390/polym15092060

Chicago/Turabian Style

Sorokina, Svetlana A., Nina V. Kuchkina, Alexander V. Mikhalchenko, Irina Yu. Krasnova, Dmitry A. Khanin, Kirill M. Skupov, and Zinaida B. Shifrina. 2023. "Ultramicroporous Polyphenylenes via Diels–Alder Polycondensation Approach" Polymers 15, no. 9: 2060. https://doi.org/10.3390/polym15092060

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop