Next Article in Journal
Cationic Chitooligosaccharide Derivatives Bearing Pyridinium and Trialkyl Ammonium: Preparation, Characterization and Antimicrobial Activities
Next Article in Special Issue
Analysis of the Influencing Factors of the Efficient Degradation of Waste Polyurethane and Its Scheme Optimization
Previous Article in Journal
Pyrolysis Kinetic Study of Polylactic Acid
Previous Article in Special Issue
Analysis of Factors Influencing the Efficiency of Catalysts Used in Waste PU Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, and Gas Adsorption Performance of Amine-Functionalized Styrene-Based Porous Polymers

1
Institute of Chemical Process Fundamentals, Czech Academy of Science, Rozvojova 135, 165 02 Prague, Czech Republic
2
Department of Occupational Safety and Health, Chung Shan Medical University, Taichung 402, Taiwan
3
Department of Environmental Engineering, National Chung Hsing University, Taichung 402, Taiwan
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(1), 13; https://doi.org/10.3390/polym15010013
Submission received: 5 December 2022 / Revised: 16 December 2022 / Accepted: 16 December 2022 / Published: 20 December 2022

Abstract

:
In recent years, porous materials have been extensively studied by the scientific community owing to their excellent properties and potential use in many different areas, such as gas separation and adsorption. Hyper-crosslinked porous polymers (HCLPs) have gained attention because of their high surface area and porosity, low density, high chemical and thermal stability, and excellent adsorption capabilities in comparison to other porous materials. Herein, we report the synthesis, characterization, and gas (particularly CO2) adsorption performance of a series of novel styrene-based HCLPs. The materials were prepared in two steps. The first step involved radical copolymerization of divinylbenzene (DVB) and 4-vinylbenzyl chloride (VBC), a non-porous gel-type polymer, which was then modified by hyper-crosslinking, generating micropores with a high surface area of more than 700 m2 g−1. In the following step, the polymer was impregnated with various polyamines that reacted with residual alkyl chloride groups on the pore walls. This impregnation substantially improved the CO2/N2 and CO2/CH4 adsorption selectivity.

Graphical Abstract

1. Introduction

Anthropogenic carbon dioxide is considered among the main cause of the greenhouse effect responsible for global warming. More than 65% of anthropogenic CO2 is generated from burning fossil fuels during the energy production. To limit the increase in the Earth’s temperature, reducing greenhouse gas emissions is necessary. Therefore, development of new technologies for producing clean energy with minimal CO2 emissions is necessary. CO2 can be reduced using carbon capture and storage (CCS) technologies based on the capture, transport, and storage of CO2 [1,2].
CCS technologies are considered potentially promising in reducing CO2 emissions, and thereby relieving climate change [3,4]. Adsorption is considered an effective alternative for CO2 capture, offering possible energy savings compared to other established separation technologies [5,6].
In recent years, the research interest in solid adsorbent materials has increased. Compared to traditional separation procedures, they are easy to operate and offer relatively low energy requirements [7].
Different types of solid adsorbents have been investigated for CO2 capture, such as carbon-based materials, zeolites, metal organic frameworks, and porous polymers [8,9,10,11].
Considering the direction of society towards a circular economy of materials, the use of plastic waste as a source of functional materials for various applications has become an important challenge for the future [12,13,14].
Examples of recovery of waste plastic include carbon-based materials created by carbonization [15], functional polymers prepared by post-synthetic chemical functionalization [16,17], and numerous other low-molecular products created after the decomposition of polymers [18,19].
In recent years, considerable attention has been paid to the preparation of hyper-crosslinked polymer adsorbents (HCLPs), a family of robust microporous organic materials with numerous benefits for gas storage, such as high surface area, permanent porosity, good thermal stability, and easy large-scale production [20,21,22,23,24].
Polystyrene-based polymers form a significant part of waste polymers because polystyrene is a cheap polymer precursor for the preparation of HCLPs [12,25]. Owing to the minimal biological degradation of polystyrene in nature [22], it causes many ecological problems; therefore, methods for its re-use should be investigated.
Styrenic materials, either monomers or polymers, can be easily functionalized via electrophilic aromatic substitution. Furthermore, an increased affinity of PS for CO2 compared to N2 or CH4 can be expected owing to the dipole interactions of phenyl groups containing π-electrons with quadrupolar CO2. Therefore, HCLPs based on PS and their amine group-modified forms have recently been investigated as efficient and low-cost CO2 adsorbents [12,26,27,28,29,30,31,32,33,34,35,36,37,38]. For example, Fu et al. [27,28] synthesized several types of HCLPs from waste expanded PS using different cross-linking agents; the best achieved CO2 absorption was 2.5 mmol g−1 and CO2/N2 separation efficiency of 38 at 273 K. The cross-linking of aromatic polymers was found to improve the textural properties of the prepared materials and enhance their CO2 adsorption capacity [28,33,34,35]. Wu et al. [36] prepared HCPs of different porosities from commercial PS; the porous structure of the HCLPs was tuned by varying the ratio of PS and cross-linker. The maximal CO2 adsorption capacity was 1.12 mmol g−1 at 298 K and a pressure of 1 atm.
The amine functionalization of PS adsorbents was studied by Lee et al. [32]. The influence of the length of the polyamines used in modifying the porous polymer adsorbents on their structure and selective CO2 adsorption was studied. After amination, the prepared materials showed increased CO2 adsorption capacities (the values were similar for all used polyamines, up to 1.86 mmol g−1 at 273 K and 1 bar) and higher CO2/N2 selectivity; however, the surface area of the modified polymers was considerably reduced. Similarly, Gaikar et al. [37] investigated the CO2 uptake and CO2/N2 selectivity of PS-based adsorbents by modifying chloromethylated PS with N-heterocyclic compounds.
Conversely, Merchán-Arenas et al. [38] found only a moderate CO2 uptake of 1.05 mmol g−1 at 273 K for amine-tethered PS adsorbents synthesized from waste expanded PS.
This study focuses on the synthesis of hyper-crosslinked vinylbenzyl chloride (VBC)-divinylbenzene (DVB) microporous material and its functionalization using the amino group-containing substances, simple amines, and larger dendrimer molecules to investigate their potential for application in CO2 capture and gas separation. The main objective of the work is the preparation of a series of hyper-crosslinked polymers, the characterization of their physicochemical properties and the investigation of the CO2 adsorption performance and selectivity of materials over other gases and to analyze the effect of modification by amino groups containing substances of different sizes.

2. Materials and Methods

2.1. Materials

All the chemicals required for synthesis of the hyper-crosslinked VBC-DVB microporous material and its functionalization were purchased from Sigma Aldrich, Prague, Czech Republic. Divinylbenzene (DVB; 80 vol% technical grade) and 4-vinylbenzyl chloride (VBC) were purified before use by passing them through a column of basic alumina. Other commercially available reagents were used as received.
Pure gases used for the adsorption experiments were purchased from Linde Gas with a stated purity of at least 99.995%.

2.2. Synthesis of Porous Polymer

A hyper-crosslinked VBC-DVB microporous polymer was prepared and functionalized in two steps, as shown in the schema in Figure 1a. The first step involved the suspension polymerization of DVB and 4-vinylbenzyl chloride (VBC). The resulting polymer was then hyper-crosslinked under the conditions shown in Figure 1a. During this operation, certain chloromethyl groups within the polymer network were transformed into crosslinked methylene bridges. This additional crosslinking (hyper-crosslinking) induces phase separation within the polymer gel, resulting in the creation of pores between the hyper-crosslinked nodules. However, the conversion of a chloromethyl group into a methylene bridge can proceed only if a neighboring polymer chain is available in the vicinity of the chloromethyl group. The probability of finding such a reaction partner in the surface layer of the nodules, which is on the walls of the newly created pores, is low; hence, unreacted chloromethyl groups are available for further modification on the pore walls [24]. In the second step, various polyamine species were applied to react with the alkyl chloride groups and functionalize the pore surface to improve the selectivity of the material for CO2 (see Figure 1b).

2.2.1. Preparation of Hyper-Crosslinked Porous Polymer Sample (HCLPP)

The polymerization was performed as follows: 20.2 g (130 mmol) of VBC was mixed with 1 g of technical grade (80%) DVB (corresponding to 5% crosslinking of the polymer matrix) and 0.6 g of AIBN (initiator) was added. An aqueous solution (1000 mL) containing 10 g of polyvinyl alcohol (Mowiol 4-88, MW = 31000, Sigma Aldrich, Prague, Czech Republic) and 50 g of sodium chloride was introduced into the reactor. The polymerization mixture was poured, the reactor was closed, and the mixture was stirred at 400 rpm. After one hour of stirring at room temperature, the temperature was raised to 80 °C, and the mixture was allowed to react for 24 h. Then, the reactor was opened, and the polymer was separated in a sieve, washed with ethanol, and dried. The resulting partially bonded balls were broken in a grinding mill.
Hyper-crosslinking: 15 g of polymer was suspended in 100 mL of dichloroethane. After 30 min, 20 mL of stannic chloride was added and the mixture was stirred for 5 h and then left for another 24 h. The polymer was then filtered and washed thoroughly with ethanol.

2.2.2. Preparation of Dendrimer

Dendrimer-(COOMe)4: In a solution of 0.7 g tetraallyl silane (3.64 mmol, 1 equiv.) in dry DMF (6 mL), a catalytic amount of 2,2-Dimethoxy-2-phenylacetophenone (DMPA) was added. The reaction mixture was degassed by bubbling argon for 5 min. After addition of 2.32 g of methyl thioglycolate (0.0218 mol, 4.4 eq.), the solution was degassed again (1 min). The reaction mixture was then exposed to UV light for 30 min. The solvent and excess methyl thioglycolate were then evaporated under reduced pressure. The crude product was dissolved in diethyl ether and extracted with 1% of NaOH solution (3×), water (2×), and finally brine (1×). The organic layers were collected, dried over MgSO4, filtered, and evaporated. Crude product was purified using column chromatography (PE:EtOAc 1:4 → 1:2) and isolated as viscous yellow liquid (1.57 g, 87%).
Data for 1: NMR (CDCl3): 1H NMR (400 MHz) δ 0.56–0.61 (m, 8H, SiCH2), 1.46–1.56 (m, 8H, SiCH2CH2), 2.58 (t, J = 7.2 Hz, 8H, SiCH2CH2CH2), 3.31 (s, 8H, SCH2COOMe), 3.63 (s, 12H, Me). 13C NMR (101 MHz) δ 11.2 (SiCH2), 23.3 (SiCH2CH2), 32.6 (CH2COOMe), 35.6 (SiCH2CH2CH2), 52.0 (Me), 170.7 (CO), and 29Si NMR (79 MHz) δ 3.66.
HRMS (ESI): m/z [M + H] + theoretical for [C24H45O8S4Si]+:617.1761; found:617.1759.
Dendrimer: In a 0.5 g of dendrimer-(COOMe)4 (0.81 mmol), ethylenediamine was added under argon in large excess (10 mL). Reaction mixture was then stirred for 12 h (70 °C). Excess ethylenediamine was subsequently evaporated to obtain the product as viscous liquid (0.681 g, 98%).
Data pro 4: NMR (DMSO-d6): 1H NMR (400 MHz, H-H COSY): 0.55–0.59 (m, 8H, SiCH2), 1.45–1.53 (m, 8H, SiCH2CH2), 2.55 (t, J = 7.5 Hz, 8H, SiCH2CH2CH2), 2.57 (t, J = 6.2 Hz, 8H, CH2NH2), 3.04 (td, J = 6.2, 5.7 Hz, 8H, NHCH2), 3.08 (s, 8H, CH2CO), 8.0 (t, J = 5.7 Hz, 4H, NH). 13C {1H} NMR (101 MHz, HSQC, HMBC):11.3 (SiCH2), 23.5 (SiCH2CH2), 34.3 (COCH2), 35.6 (SiCH2CH2CH2), 41.2 (CH2NH), and 42.3 (CH2NH2). 29Si {1H} NMR (79 MHz): δ 3.56.
HRMS (ESI): m/z [M + H] + theoretical for C28H60N8O4S4Si:729.3462, found: 729.3481.

2.2.3. Functionalization of HCLPP

The prepared hyper crosslinked polymer 0.4 g was suspended in 20 mL of acetonitrile. Amine group-containing reagents (ethylenediamine EDA, diethylenetriamine DETA) or dendrimers (see Figure 1b, left) 0.2 g were added dropwise (0.1 mg in 2 mL of acetonitrile). Reaction mixture was then stirred at 50 °C overnight (12 h). After cooling to room temperature, the reaction mixture was centrifuged, washed twice with acetonitrile, and then dried under reduced pressure. Subsequently, the solid was suspended in 5 mL of 10% K2CO3. After centrifugation, the samples were washed twice with water and dried under reduced pressure.

2.3. Characterization

2.3.1. Properties of the HCLPPs

All forms of the prepared hyper-crosslinked materials were characterized using the following methods: Textural characteristics, including the specific surface area, pore volume, and pore diameter distribution, were determined based on nitrogen adsorption-desorption isotherms measured at 77 K using a commercial automated volumetric gas adsorption analyzer ASAP 2050 (Micromeritics, Norcross, GA, USA). Before the measurements, the samples were degassed under deep vacuum at 373 K for 4 h. The morphology and elemental composition were evaluated using a scanning electron microscope (SEM) equipped with an energy-dispersive X-ray (EDX) spectrometer (Quantax, Bruker, Bremen, Germany) in the range of 5–30 kV. Using this technique, the films were analyzed at an accelerating voltage of 15 kV. Thermogravimetric analysis (TGA) of samples was performed using a TG 209 F1 Libra equipment (NETZSCH, Selb, Germany) in the temperature range from 25 °C to 1000 °C in an inert argon atmosphere with a flow rate of 50 mL⋅min−1. The samples were heated from ambient temperature to 1000 °C at a heating rate of 4 °C min−1, the weight of each sample was approximately 7 mg.

2.3.2. Characterization of Dendrimers

NMR spectra were measured using a Bruker Avance 400, Bruker, Bremen, Germany (1H at 400.1 MHz; 13C at 100.6 MHz; 29Si {1H} (INEPT technique) at 79.5 MHz) at 25 °C. 1H and 13C NMR signals of the prepared compounds were assigned to the corresponding atoms using gHSQC, gCOSY, and gHMBC 2D NMR correlation spectra. 1H and 13C chemical shifts (δ/ppm) are given relative to residual solvent signals (δHC: DMSO-d6 2.50/39.52); 29Si spectra were referenced to an external standard hexamethyldisilane (−19.87 ppm). HRMS spectra were measured using a MicroTOF-QIII instrument (Bruker Daltonics, Bruker, Bremen, Germany) with an ESI or APCI ionization source in the positive mode.

2.4. Gas Adsorption Measurement

The adsorption isotherms of the tested gases were measured using the volumetric method. The adsorption of pure gas in the prepared materials was measured using a simple apparatus shown in Figure 2 constructed in our laboratory.
The thermostated part of the apparatus consisted of two chambers with calibrated volumes, V1 and V2, separated by a valve. The pressure in the apparatus was measured using a pressure transducer. A heated box maintains a constant temperature of the entire system during the absorption experiment.
Before the experiment, a weighed amount of the sample was placed in the adsorption chamber, V2. First, to remove dissolved gases and impurities, the apparatus chambers were evacuated at least 5 h at temperature 80 °C. Subsequently, the system temperature was set to the thermostat at the required experimental value. Chamber V2 was closed, while volume V1 was filled with the studied gas at the desired pressure p1 (from zero to approximately 400 kPa). At the start of the sorption measurement, volumes V1 and V2 were connected, and the pressure inside the apparatus was monitored by a computer. The pressure initially dropped immediately to a certain value because of the interconnection of the chambers. Subsequently, the pressure decreased further, albeit more slowly, owing to the adsorption of gas in the sample material. The time required to attain equilibrium was approximately 3 h.
The amount of gas adsorbed in the material was calculated from the balance as the difference between the initial and equilibrium amounts of gas in the system:
N a d s o r b e d = N g a s   i n i t i a l l y N r e s t   g a s   i n   e q u i l i b r i u m
N a d s o r b e d = p 1 V 1 p e q V 1 + V 2 V x R T
The ideal gas state Equation (3) is used for the moles in the gas-phase calculation (N gas mole number):
N = p V R T
The adsorbed amount is related to the material weight m to determine the adsorbed phase concentration q:
q = N a d s o r b e d m

3. Results and Discussion

3.1. Properties of HCLPPs

The textural properties of these materials are listed in Table 1. The results revealed that the pristine hyper-crosslinked porous polymer had the highest Brunauer-Emmett-Teller surface area (SBET = 757 m2 g−1) and its structure was partially microporous and mesoporous.
The reaction of the prepared porous material with polyamine species (amines and amine groups containing dendrimers) decreased the specific surface area and total volume of pores, while in the case of modification by dendrimers, the decrease was smaller than that for amines. The reduction in specific surface area due to amine functionalization was also reported by Lee et al. [32], see Table 6.
Similarly, the total and micropore volumes decreased significantly after amine modification. However, the modification by dendrimers reduced the total and micropore volumes only slightly, which indicated difficulties in filling the micropores with voluminous dendrimer molecules affecting only the region of mesopores and not the micropores.
To investigate the morphology and elemental composition of the prepared HCLPP, a SEM equipped with an EDX spectroscopy detector was used. From the photographs in Figure 3, the morphologies of all prepared polymers were similar, and the resulting materials had different particle sizes of irregular shape with no observable changes after functionalization by agents containing amino groups.
Elemental analyses of the prepared polymers, performed using EDX, are given in Table 2. After functionalization, the nitrogen content in the materials considerably increased. The modification by amines increased the nitrogen content to more than 9%, while with the dendrimer, the value of approximately 5% was achieved. Simultaneously, the modification lowered the chlorine content, indicating a reaction with the chloride units. The results show a more effective reaction and exchange of chloride groups for amine groups and a significantly higher density of amine groups on the surface of the porous material for amine functionalization compared to that of modification with a dendrimer. These findings were in agreement with the morphological data shown in Table 1.
Elemental analysis of the dendrimer-modified sample also revealed the presence of sulfur and silicon in the modifying agent. A small amount of silicon was also detected in other porous polymers, which was probably caused by contamination during the grinding of the porous samples in the ceramic bowl before the analyses.
Thermal stability of polymers was evaluated by TGA in the temperature range of 25 to 1000 °C under argon atmosphere with heating rate of 4 °C min−1. The TGA results are shown in Figure 4. The profiles of all the samples were very similar. A weight loss of approximately 6% at a lower temperature up to 250 °C was observed, probably caused by the evaporation of water, residual solvent, and trapped gases in the porous structure of the material. Significant weight loss followed as the temperature increased to 445 °C, corresponding to the decomposition of polystyrene-based polymers [28,40].
Compared to waste expanded PS undergoing complete weight loss in the range of 400–440 °C [27,40], the prepared porous polymers were more thermally stable, having a total weight loss at 1000 °C of only approximately 58% to 65%. This high thermal stability is because of the rigid cross-linked porous structure formed during the Friedel-Crafts reaction [27,41].

3.2. Gas Adsorption and Selectivity

The adsorption isotherms of gases (CO2, CH4, N2, H2, and O2) were measured using a simple volumetric sorption apparatus designed in our laboratory (the scheme is given in Figure 2). All sorption measurements were performed at a constant, actively controlled temperature of 25 °C.
Table 3 presents a comparison of the adsorption capacities of the unmodified and modified polymers for pure gases, determined from the measured isotherms at a pressure of 100 kPa.
The results demonstrated that polymer modification by amines led to a decrease in the amount of adsorbed gas, with the exception of oxygen. This could be attributed to the reduced surface area of the amine-functionalized forms of the polymer. However, the decrease in the adsorption capacity was not as significant as the decrease in the specific surface area or pore volume of the modified materials. Lee et al. [32] reported similar trends in CO2 and N2 capacities at 25 °C for the porous polymer network synthesized by solvothermal polymerization of DVB and VBC, see Table 6.
In contrast, in the case of the material modified with a dendrimer containing amino groups, an increase in the adsorption performance of CO2 and oxygen was observed, despite the fact that, similar to other modified forms of the polymer, the specific surface area also decreased, but to a lesser extent.
However, if the adsorption capacities of the materials were normalized to the specific surface area, the efficiency of CO2 capture was significantly improved, thus confirming the positive effect of functionalization (Table 4).
While the gas-uptake ability of the porous material decreased after functionalization, the selectivity increased in all cases. In particular, the CO2/N2 and O2/N2 separation efficiencies of the HCLPPs after amination were improved. Table 5 shows the selectivity values for different gas pairs obtained from the adsorption isotherms of pure gases at 100 kPa.
To verify the precision of the measurements using the simple apparatus constructed in our laboratory, we also performed control measurements of CO2 adsorption using the automated commercial apparatus ASAP 2050 (Micromeritics, Norcross, GA, USA). As shown in Figure 5, the agreement of the results was excellent; therefore, the apparatus provided the correct results.
The obtained CO2 adsorption data of the prepared materials were compared with other hyper-crosslinked porous polystyrene-based polymers studied for CO2 capture application reported in literature (Table 6), for examples HCLPs prepared via the Friedel-Craft reaction between waste expanded polystyrene and 1,2-dichloroethane as both the solvent and the crosslinking agent [27], similar polymers produced using four cross-linkers in order to enhance the porosity of the polymeric adsorbent [28], or amine-functionalized porous polymer network prepared by conventional radical polymerization. In all cases, the material prepared in our study showed a comparable or slightly lower CO2 adsorption capacity, but in most cases the higher selectivity.
Table 6. CO2 capture and selectivity adsorption performance of some styrene-based polymers.
Table 6. CO2 capture and selectivity adsorption performance of some styrene-based polymers.
Sample SBET, (m2 g−1) CO2, (μmol g−1)
1 bar, 273 K, 298 K
CO2/N2
Selectivity
Ref.
HCP-3A179932 9
HCP-6A2991604 22
HCP-12A4061987273 K23[27]
HCP-24A5481715 20
HCP-12B4881771 20
HCP-12C5731846 15
HCP-04251201 17
HCP-A7771473 38
HCP-B7681178298 K25[28]
HCP-C6111258 15
HCP-D6731359 15
P(DVB-VBC)-31744390 20
P(DVB-VBC)-31-EDA3891150 152
P(DVB-VBC)-31-DETA3761440298 K223[32]
P(DVB-VBC)-11134180 -
P(DVB-VBC)-11-EDA911150 -
P(DVB-VBC)-11-DETA48750 -
HCLPP7571010 28
HCLPP-EDA283910298 K81This work
HCLPP-DETA277820 71
HCLPP-DENDRIMER6161100 29

4. Conclusions

In summary, we presented a method of preparation of a new porous polymeric adsorbent by the suspension polymerization of divinylbenzene and 4-vinylbenzyl chloride, followed by modification with amines and branched dendrimers containing amino groups in order to enhance its CO2 affinity. Synthesized hyper-crosslinked porous material showed a high apparent surface area and selective CO2 adsorption over CH4, N2 and H2, which significantly increased after amines functionalization. Therefore, produced porous polymers represent promising organic porous materials for CO2 uptake from gas mixtures. The results indicated that the reaction of the prepared polymer with the dendrimer was not as effective as its reaction with amines due to the partially microporous structure being too tight for the possible interaction with the branched dendrimer molecules. Based on this finding, we will focus in the future on the development of polymers with mesoporous structure that may be more suitable for dendrimer loading. The prepared porous polymer structure and its amine-functionalized forms are considered for testing as fillers into mixed-matrix membranes with improved CO2/CH4 and CO2/N2 separation performance as well.

Author Contributions

Conceptualization: K.S. (Katerina Setnickova) and K.J.; methodology: K.J., T.S., P.U. and M.M.; investigation: K.S. (Katerina Setnickova), R.P., K.S. (Karel Soukup) and V.J.; writing—original draft preparation: K.S. (Katerina Setnickova); writing—review and editing: K.J. and H.-H.T.; supervision: K.S. (Katerina Setnickova) and K.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Czech Science Foundation (Project No. 19-23760J).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mikkelsen, M.; Jorgensen, M.; Krebs, F.C. The teraton challenge. A review of fixation and transformation of carbon dioxide. Energy Environ. Sci. 2010, 3, 43–81. [Google Scholar] [CrossRef]
  2. IPCC. Climate Change 2014: Mitigation of Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2015. [Google Scholar]
  3. Leung, D.Y.C.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443. [Google Scholar] [CrossRef] [Green Version]
  4. Zou, L.; Sun, Y.; Che, S.; Yang, X.; Wang, X.; Bosch, M.; Wang, Q.; Li, H.; Smith, M.; Yuan, S.; et al. Porous organic polymers for post-combustion carbon capture. Adv. Mater. 2017, 29, 1700229. [Google Scholar] [CrossRef] [PubMed]
  5. Figueroa, J.D.; Fout, T.; Plasynski, S.; McIlvried, H.; Srivastava, R.D. Advances in CO2 capture technology—The U.S. Department of Energy’s Carbon Sequestration Program. Int. J. Greenh. Gas Control 2008, 2, 9–20. [Google Scholar] [CrossRef]
  6. Ebner, A.D.; Ritter, J.A. State-of-the-art Adsorption and Membrane Separation Processes for Carbon Dioxide Production from Carbon Dioxide Emitting Industries. Sep. Sci. Technol. 2009, 44, 1273–1421. [Google Scholar] [CrossRef]
  7. Gray, M.L.; Champagne, K.J.; Fauth, D.; Baltrus, J.P.; Pennline, H. Performance of immobilized tertiary amine solid sorbents for the capture of carbon dioxide. Int. J. Greenh. Gas Control 2008, 2, 3–8. [Google Scholar] [CrossRef]
  8. Darunte, L.A.; Walton, K.S.; Sholl, D.S.; Jones, C.W. CO2 Capture via Adsorption in Amine-Functionalized Sorbents. Curr. Opin. Chem. Eng. 2016, 12, 82–90. [Google Scholar] [CrossRef] [Green Version]
  9. Samanta, A.; Zhao, A.; Shimizu, G.K.; Sarkar, P.; Gupta, R. Post-combustion CO2 capture using solid sorbents: A review. Ind Eng. Chem. Res. 2012, 51, 1438–1463. [Google Scholar] [CrossRef]
  10. Favvas, E.P.; Katsaros, F.K.; Papageorgiou, S.K.; Sapalidis, A.A.; Mitropoulos, A.C. Review of the latest developments in polyimide-based membranes for CO2 separation. React. Funct. Polym. 2017, 120, 104–130. [Google Scholar] [CrossRef]
  11. Tiwari, D.; Goel, C.; Bhunia, H.; Bajpai, P.K. Dynamic CO2 capture by carbon adsorbents: Kinetics, isotherm, and thermodynamic studies. Sep. Purif. Technol. 2017, 181, 107–122. [Google Scholar] [CrossRef]
  12. Teo, J.Y.Q.; Ong, A.; Tan, T.T.Y.; Li, X.; Loh, X.J.; Lim, J.Y.C. Materials from waste plastics for CO2 capture and utilization. Green Chem. 2022, 24, 6086. [Google Scholar] [CrossRef]
  13. Zhao, X.; Boruah, B.; Chin, K.F.; Đokić, M.; Modak, J.M.; Soo, H.S. Upcycling to sustainably reuse plastics. Adv. Mater. 2021, 34, 2100843. [Google Scholar] [CrossRef] [PubMed]
  14. Yeung, C.W.S.; Teo, J.Y.Q.; Loh, X.J.; Lim, J.Y.C. Polyolefins and Polystyrene as Chemical Resources for a Sustainable Future: Challenges, Advances, and Prospects. ACS Mater. Lett. 2021, 3, 1660–1676. [Google Scholar] [CrossRef]
  15. Zhuo, C.; Levendis, Y. Upcycling waste plastics into carbon nanomaterials: A review. J. Appl. Polym. Sci. 2014, 131, 39931. [Google Scholar] [CrossRef]
  16. Yeung, C.W.S.; Loh, W.W.; Lau, H.H.; Loh, X.J.; Lim, J.Y.C. Catalysts developed from waste plastics: A versatile system for biomass conversion. Mater. Today Chem. 2021, 21, 100524. [Google Scholar] [CrossRef]
  17. Williamson, J.B.; Lewis, S.E.; Johnson, R.R., III; Manning, I.M.; Leibfarth, F.A. C–H Functionalization of Commodity Polymers. Angew. Chem. Int. Ed. 2019, 58, 8654–8668. [Google Scholar] [CrossRef]
  18. Bäckström, E.; Odelius, K.; Hakkarainen, M. Designed from Recycled: Turning Polyethylene Waste to Covalently Attached Polylactide Plasticizers. ACS Sustain. Chem. Eng. 2019, 7, 11004–11013. [Google Scholar] [CrossRef] [Green Version]
  19. Gazi, S.; Đokić, M.; Chin, K.; Ng, P.R.; Soo, H.S. Visible light-driven cascade carbon–carbon bond formation for organic transformations and plastic recycling. Adv. Sci. 2019, 6, 1902020. [Google Scholar] [CrossRef] [Green Version]
  20. Tan, L.; Tan, B. Hypercrosslinked porous polymer materials: Design, synthesis, and applications. Chem. Soc. Rev. 2017, 46, 3322–3356. [Google Scholar] [CrossRef]
  21. Zhang, C.; Liu, Y.; Li, B.; Tan, B.; Chen, C.F.; Xu, H.B.; Yang, X.L. Triptycene-based microporous polymers: Synthesis and their gas storage properties. ACS Macro. Lett. 2012, 1, 190–193. [Google Scholar] [CrossRef]
  22. Ma, H.; Chen, J.J.; Tan, L.; Bu, J.H.; Zhu, Y.; Tan, B.; Zhang, C. Nitrogen-rich triptycene-based porous polymer for gas storage and iodine enrichment. ACS Macro. Lett. 2016, 5, 1039–1043. [Google Scholar] [CrossRef] [PubMed]
  23. Hankova, L.; Holub, L.; Jerabek, K. Formation of porous polymer morphology by microsyneresis during divinylbenzene polymerization. J. Polym. Sci. B Polym. Phys. 2015, 53, 774–781. [Google Scholar] [CrossRef]
  24. Veverka, P.; Jeřábek, K. Mechanism of hypercrosslinking of chloromethylated styrene–divinylbenzene copolymers. React. Funct. Polym. 1999, 41, 21. [Google Scholar] [CrossRef]
  25. Plastics: The Facts 2015: An Analysis of European Plastics Production, Demand and Waste Data; Plastics Europe: Brussels, Belgium, 2015.
  26. Kaliva, M.; Armatas, G.S.; Vamvakaki, M. Microporous Polystyrene Particles for Selective Carbon Dioxide Capture. Langmuir 2012, 28, 2690–2695. [Google Scholar] [CrossRef]
  27. Fu, Z.; Jia, J.; Li, J.; Liu, C. Transforming waste expanded polystyrene foam into hypercrosslinked polymers for carbon dioxide capture and separation. Chem. Eng. J. 2017, 323, 557–564. [Google Scholar] [CrossRef]
  28. Fu, Z.; Mohamed, I.M.A.; Li, J.; Liu, C. Novel adsorbents derived from recycled waste polystyrene via cross-linking reaction for enhanced adsorption capacity and separation selectivity of CO2. J. Taiwan Inst. Chem. Eng. 2019, 97, 381–388. [Google Scholar] [CrossRef]
  29. Flouraki, C.; Kaliva, M.; Papadas, I.T.; Armatas, G.S.; Vamvakaki, M. Nanoporous polystyrene–porphyrin nanoparticles for selective gas separation. Polym. Chem. 2016, 7, 3026–3033. [Google Scholar] [CrossRef]
  30. Sneddon, G.; McGlynn, J.C.; Neumann, M.S.; Aydin, H.M.; Yiu, H.H.P.; Ganin, A.Y. Aminated poly(vinyl chloride) solid-state adsorbents with hydrophobic function for post-combustion CO2 capture. J. Mater. Chem. A 2017, 5, 11864–11872. [Google Scholar] [CrossRef] [Green Version]
  31. Jo, D.H.; Lee, C.H.; Jung, H.; Jeon, S.; Kim, S.H. Effect of amine surface density on CO2 adsorption behaviors of amine-functionalized polystyrene. Bull. Chem. Soc. Jpn. 2015, 88, 1317–1322. [Google Scholar] [CrossRef]
  32. Lee, D.Y.; Zhang, C.Y.; Gao, H.F. Amine-functionalized porous polymer network for highly selective absorption of CO2 over N2. Macromol. Chem. Phys. 2015, 216, 489–494. [Google Scholar] [CrossRef]
  33. Yaumi, A.L.; Abu Bakar, M.Z.; Hameed, B.H. Recent advances in functionalized composite solid materials for carbon dioxide capture. Energy 2017, 124, 461–480. [Google Scholar] [CrossRef]
  34. Liu, H.; Li, Q.; Li, Q.; Jin, W.; Li, X.; Hameed, A.; Qiao, S. Rational skeletal rigidity of conjugated microporous polythiophenes for gas uptake. Chem. Polym. Chem. 2017, 8, 6733–6740. [Google Scholar] [CrossRef]
  35. Davankov, V.; Tsyurupa, M.P. Hypercrosslinked Polymeric Networks and Adsorbing Materials: Synthesis, Properties, Structure, and Applications, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2010. [Google Scholar]
  36. Wu, Y.; Li, L.; Yang, W.; Feng, S.; Liu, H. Hybrid nanoporous polystyrene derived from cubic octavinylsilsesquioxane and commercial polystyrene via the Friedel–Crafts reaction. RSC Adv. 2015, 5, 12987–12993. [Google Scholar] [CrossRef]
  37. Heer, P.K.S.; Khot, K.M.; Gaikar, V.G. Development of polystyrene adsorbents functionalized with heterocyclic ligands for selective adsorption of CO2 from CH4 and N2. Sep. Purif. Technol. 2016, 158, 212–222. [Google Scholar] [CrossRef]
  38. Merchán-Arenas, D.R.; Murcia-Patiño, A.F. Synthesis of polyamino styrene from post-consumption expanded polystyrene and analysis of its CO2 scavenging capacity. Int. J. Environ. Sci. Technol. 2021, 18, 2519–2532. [Google Scholar] [CrossRef]
  39. Harkins, W.D.; Jura, G. Surfaces of solids. XIII. A vapor adsorption method is used to determine the area of a solid without the assumption of a molecular area and the areas occupied by nitrogen and other molecules on the surface of a solid. J. Am. Chem. Soc. 1944, 66, 1366–1373. [Google Scholar] [CrossRef]
  40. Zhuang, G.L.; Tseng, H.H.; Wey, M.Y. Feasibility of using waste polystyrene as a membrane material for gas separation. Chem. Eng. Res. Des. 2016, 111, 204–217. [Google Scholar] [CrossRef]
  41. Liu, H.; Li, S.; Yang, H.; Liu, S.; Chen, L.; Tang, Z.; Fu, R.; Wu, D. Stepwise crosslinking: A facile yet versatile conceptual strategy to nanomorphology-persistent porous organic polymers. Adv. Mater. 2017, 29, 1700723. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scheme of (a) the synthesis of the hyper-crosslinked porous polymer (HCLPP) and (b) post-polyamine functionalization by ethylenediamine, diethylenetriamine, and dendrimer (c).
Figure 1. Scheme of (a) the synthesis of the hyper-crosslinked porous polymer (HCLPP) and (b) post-polyamine functionalization by ethylenediamine, diethylenetriamine, and dendrimer (c).
Polymers 15 00013 g001
Figure 2. Design and schema of the experimental apparatus for gas adsorption capacity measurement.
Figure 2. Design and schema of the experimental apparatus for gas adsorption capacity measurement.
Polymers 15 00013 g002
Figure 3. SEM images of synthesized polystyrene based HCLPPs (a) pristine HCLPP, (b) HCLPP-EDA, (c) HCLPP-DETA, and (d) HCLPP-DENDRIMER, respectively.
Figure 3. SEM images of synthesized polystyrene based HCLPPs (a) pristine HCLPP, (b) HCLPP-EDA, (c) HCLPP-DETA, and (d) HCLPP-DENDRIMER, respectively.
Polymers 15 00013 g003
Figure 4. TGA curves of synthesized polystyrene based HCLPPs.
Figure 4. TGA curves of synthesized polystyrene based HCLPPs.
Polymers 15 00013 g004
Figure 5. CO2 adsorption on the unmodified HCLPP measured at 25 °C using the laboratory constructed device and commercial apparatus ASAP 2050.
Figure 5. CO2 adsorption on the unmodified HCLPP measured at 25 °C using the laboratory constructed device and commercial apparatus ASAP 2050.
Polymers 15 00013 g005
Table 1. Texture properties of pure and functionalized porous polymers HCLPPs.
Table 1. Texture properties of pure and functionalized porous polymers HCLPPs.
SampleSBET a
(m2 g−1)
Smeso b
(m2 g−1)
Vtot c
(mm3liq g−1)
Vmicro d
(mm3liq g−1)
HCLPP757310408226
HCLPP-EDA2839315797
HCLPP-DETA27713215976
HCLPP-DENDRIMER616194341218
a SBET the specific surface area calculated by the BET method from the N2 adsorption isotherm at T = 77 K; b Smeso is the specific surface area of mesopores (t-plot method using the standard isotherm according to Harkins and Jura [39]); c Vtot the specific total volume of pores determined from the N2 adsorption isotherm at p/p0 = 0.99; d Vmicro the specific volume of micropores (t-plot method determined by the t-plot method using the standard isotherm according to Harkins and Jura).
Table 2. EDX analysis of pure and functionalized porous polymers HCLPPs.
Table 2. EDX analysis of pure and functionalized porous polymers HCLPPs.
SampleC
(%)
Cl
(%)
N
(%)
S
(%)
Si
(%)
HCLPP97.790.421.74-0.05
HCLPP-EDA90.620.089.21-0.10
HCLPP-DETA90.550.059.32-0.07
HCLPP-DENDRIMER99.930.305.370.200.21
Table 3. Gas adsorption capacity of pure and functionalized porous HCLPP at 100 kPa and 25 °C.
Table 3. Gas adsorption capacity of pure and functionalized porous HCLPP at 100 kPa and 25 °C.
SampleCO2
(μmol g−1)
CH4
(μmol g−1)
N2
(μmol g−1)
H2
(μmol g−1)
O2
(μmol g−1)
HCLPP1010310902112
HCLPP-EDA910130331547
HCLPP-DETA820130291242
HCLPP-DENDRIMER1100240631789
Table 4. Gas adsorption capacity of pure and functionalized porous HCLPP based on VBC *.
Table 4. Gas adsorption capacity of pure and functionalized porous HCLPP based on VBC *.
SampleCO2
(μmol m2)
CH4
(μmol m2)
N2
(μmol m2)
H2
(μmol m2)
O2
(μmol m2)
HCLPP1.330.410.120.030.02
HCLPP-EDA3.220.460.120.050.17
HCLPP-DETA2.960.470.100.040.15
HCLPP-DENDRIMER1.790.390.100.030.14
* at temperature 25 °C and pressure 100 kPa.
Table 5. Selectivity of pure and functionalized porous polymers HCLPP based on VBC *.
Table 5. Selectivity of pure and functionalized porous polymers HCLPP based on VBC *.
SampleCO2/CH4
(-)
CO2/N2
(-)
CO2/H2
(-)
O2/N2
(-)
CH4/N2
(-)
HCLPP5.827.674.20.14.7
HCLPP-EDA16.081.0188.91.45.1
HCLPP-DETA12.871.4172.61.55.6
HCLPP-DENDRIMER6.228.9109.31.44.7
* at temperature 25 °C and pressure 100 kPa, calculated by IAST.
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

Setnickova, K.; Jerabek, K.; Strasak, T.; Mullerova, M.; Jandova, V.; Soukup, K.; Petrickovic, R.; Tseng, H.-H.; Uchytil, P. Synthesis, Characterization, and Gas Adsorption Performance of Amine-Functionalized Styrene-Based Porous Polymers. Polymers 2023, 15, 13. https://doi.org/10.3390/polym15010013

AMA Style

Setnickova K, Jerabek K, Strasak T, Mullerova M, Jandova V, Soukup K, Petrickovic R, Tseng H-H, Uchytil P. Synthesis, Characterization, and Gas Adsorption Performance of Amine-Functionalized Styrene-Based Porous Polymers. Polymers. 2023; 15(1):13. https://doi.org/10.3390/polym15010013

Chicago/Turabian Style

Setnickova, Katerina, Karel Jerabek, Tomas Strasak, Monika Mullerova, Vera Jandova, Karel Soukup, Roman Petrickovic, Hui-Hsin Tseng, and Petr Uchytil. 2023. "Synthesis, Characterization, and Gas Adsorption Performance of Amine-Functionalized Styrene-Based Porous Polymers" Polymers 15, no. 1: 13. https://doi.org/10.3390/polym15010013

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