Next Article in Journal
Physicochemical Properties and Surface Characteristics of Ground Human Teeth
Next Article in Special Issue
Immobilization of Ionic Liquid on a Covalent Organic Framework for Effectively Catalyzing Cycloaddition of CO2 to Epoxides
Previous Article in Journal
Progress of Microencapsulated Phycocyanin in Food and Pharma Industries: A Review
Previous Article in Special Issue
Covalent Organic Frameworks Composites Containing Bipyridine Metal Complex for Oxygen Evolution and Methane Conversion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controllable Construction of Amino-Functionalized Dynamic Covalent Porous Polymers for High-Efficiency CO2 Capture from Flue Gas

1
School of Chemical Engineering and Pharmacy, Wuhan Institute of Technology, Wuhan 430205, China
2
School of Energy, Power and Mechanical Engineering, North China Electric Power University, Beijing 102206, China
3
Hubei Yihua Chemical Technology R&D Co., Ltd., Yichang 443208, China
4
Shanxi-Zheda Institute of Advanced Materials and Chemical Engineering, Taiyuan 030024, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(18), 5853; https://doi.org/10.3390/molecules27185853
Submission received: 11 August 2022 / Revised: 29 August 2022 / Accepted: 5 September 2022 / Published: 9 September 2022
(This article belongs to the Special Issue Porous Organic Materials: Design and Applications)

Abstract

:
The design of high-efficiency CO2 adsorbents with low cost, high capacity, and easy desorption is of high significance for reducing carbon emissions, which yet remains a great challenge. This work proposes a facile construction strategy of amino-functional dynamic covalent materials for effective CO2 capture from flue gas. Upon the dynamic imine assembly of N-site rich motif and aldehyde-based spacers, nanospheres and hollow nanotubes with spongy pores were constructed spontaneously at room temperature. A commercial amino-functional molecule tetraethylenepentamine could be facilely introduced into the dynamic covalent materials by virtue of the dynamic nature of imine assembly, thus inducing a high CO2 capacity (1.27 mmol·g−1) from simulated flue gas at 75 °C. This dynamic imine assembly strategy endowed the dynamic covalent materials with facile preparation, low cost, excellent CO2 capacity, and outstanding cyclic stability, providing a mild and controllable approach for the development of competitive CO2 adsorbents.

Graphical Abstract

1. Introduction

Excessive CO2 emissions from the burning of fossil fuels have caused a series of ecological and environmental problems [1,2]. Post-combustion capture of CO2 from flue gases has been considered as one of the potential strategies for reducing carbon emissions [3,4]. Various CO2 capture techniques, mainly including liquid ammonia absorption [5], membrane separation [6], and adsorption [7], have been widely developed. In particular, CO2 adsorption by solids has the advantages of high recycling rate, simple operation, low equipment corrosion, and low energy consumption; thus, it is of high promise for low-energy CO2 capture [8,9,10]. It is, therefore, of great significance for developing high-efficiency adsorbents with low cost, high capacity, and easy desorption, which yet remains a great challenge.
Great efforts have been made toward developing diverse kinds of CO2 adsorbents, e.g., inorganic porous materials, covalent organic materials, and supramolecular organic materials [11]. Inorganic materials (e.g., zeolite [12], activated carbon [13], mesoporous silica [14], and mesoporous fibers [15]), despite having the advantages of high mechanical stability and regular porosity, mainly suffer from difficult structural functionalization and intrinsic acidic sites unfavorable for CO2 adsorption [16]. Organic porous materials (e.g., polymers) have much more superior structural tunability compared to inorganic materials [17,18], but still suffer from complex preparation processes and high cost [19,20]. Alternatively, supramolecular organic materials, such as metal organic framework materials (MOFs) [21], hydrogen-bonded organic framework materials (HOFs) [22], covalent organic framework materials (COFs) [23], and coordination polymers [24], are generally constructed through dynamic covalent and/or noncovalent bonding interactions. The dynamic noncovalent nature endows supramolecular organic materials with remarkable features such as simple preparation process, adjustable structure, and easy functionalization. In particular, dynamic covalent bonding (DCC) combines the stability and dynamic reversibility of covalent bonding, thus allowing the resultant dynamic covalent functional materials assembled to possess facile preparation, high stability, and promising functionalization [25,26,27,28].
Amino functionalization of porous material has been considered as an efficient tool for improving CO2 adsorption performance by taking advantage of the interaction between the amino group and CO2 [29,30,31,32,33]. Traditionally, amino functionalization is mainly achieved through covalent grafting [34] and impregnation [35]. The former method often requires synthesis and activation of the as-prepared support (e.g., silica or zeolite) prior to grafting, complex chemical synthesis during the grafting process, and removal of solvent and/or unreacted reactants to activate pores [36]. The wet impregnation method tends to induce pore blockage and leaching of active component (e.g., organic amine or ionic liquid), and the liquid nature of active component increases both the viscosity of the composite material and the transfer resistance of CO2, thus generally inducing a relatively low CO2 adsorption and/or low stability [37]. Moreover, the control over the content of functional sites and the nature of textural structures, which highly affects the CO2 adsorption performance of the amine-functionalized materials, suffers from high challenges in both methods. Therefore, it is critical to explore a green and efficient approach for preparing amino-functionalized materials to promote CO2 adsorption.
Herein, this work proposes a facile construction approach of functional dynamic covalent polymers through the dynamic imine assembly of N-site rich motif and aldehyde-based spacers for high-efficiency CO2 capture. By virtue of the dynamic noncovalent nature, the structure of the dynamic covalent polymers can be facilely regulated through the assembly process control, thus offering a feasible approach for promoting the CO2 adsorption ability. Most importantly, amino-functional units can be facilely introduced through the imine exchange method, thus offering a mild and controllable approach for the development of competitive CO2 adsorbents.

2. Results and Discussion

2.1. Characterization of Dynamic Covalent Materials

A series of functional dynamic covalent nanomaterials were constructed via the self-assembly of 3,3′-dithiobis (propionyl hydrazine) (DTPH) and polyaldehyde at the ethyl acetate/water interface at 30 °C. Upon the self-assembly of DTPH and p-phthalaldehyde (PA), white powers were typically afforded (Scheme 1). To verify the successful imine assembly, the resultant samples were characterized by FT-IR technology. As shown in Figure 1, the N–H vibration peaks of DTPH at 3285 and 3326 cm−1 disappeared in the dynamic covalent polymers [38], and a new peak assigned to C=N bond appeared at 1670 cm−1 [39,40], indicating that DTPH was successfully assembled with PA.
Solid-state 13C-NMR characterization was used to further verify imine assembly. As shown in Figure 2, the characteristic signal at 163 ppm corresponded to the chemical shift of the C=N bond in PAP−1 [41,42]. In addition, the peak of the benzene ring derived from PA at 135 ppm and the peaks of C–C derived from DTPH at 35 ppm can observed. The phenomena above proved the successful assembly of PA and DTPH through imine bonding.
SEM observation revealed that nanoflowers with spongy pores were formed in the PAP samples with various molar ratios of PA and DTPH (Figure 3a,b), and the porous structures were considered of high promise for gas adsorption.
Similarly, the self-assembly of DTPH and 4,4′-biphenyldicarboxaldehyde (BPDA) could also be achieved (Scheme 2 and Figure 4). When the molar ratio of BPDA/DTPH was 1:0.5, porous nanospheres were formed (Figure 5a,b). With increasing content of DTPH, of great interest, nanospheres evolved into hollow nanotubes (Figure 5e,f). It can be noted that spongy porous structures were spread on the surface of the nanotubes, which was also beneficial for gas capture.
The thermodynamic and kinetic analysis of supramolecular polymerization is of vital importance for understanding the pathway of structural regulation [43]. The effect of reaction time on the formation of the polymers was studied by taking PAP-3 as an example. It was found that powders could be formed within a quite short time. SEM characterization verified that a regular nanoflower morphology was formed within 0.5 h (see Figure 6). It can be seen that the dynamic imine assembly of PA and DTPH was under thermodynamic control rather than kinetic control [44,45].
The crystal structures of the dynamic covalent polymers PAP and PBP were analyzed by X-ray diffraction (XRD). As can be seen from Figure 7, all the polymers had two broad diffraction peaks around 17° and 25°, indicating that the dynamic covalent polymer had an amorphous structure [46].
The maximum weight loss temperatures of PAP-1 and PBP-1 were both 292 °C, as shown in Figure 8, indicating the excellent thermal stability of the materials. The generation of imine from the reaction of amine and aldehyde is a reversible reaction that operates under thermodynamic control, through which the kinetically competing intermediate products are replaced by the thermodynamically most stable product after sufficient assembly time [47]. That is, imine assembly can afford a dynamic covalent material with excellent thermal stability under mild conditions.
Porosity is a crucial factor for gas adsorption. The porous structures of PAP and PBP samples were characterized using N2 adsorption-desorption. As shown in Figure 9a, the N2 adsorption-desorption curves of the PAP samples presented a typical H4-type hysteresis loop, reflecting that the samples were composited of mesopores and/or macropores, which could be further confirmed by the analysis of pore volume (Table 1). This result is consistent with the SEM observation. H4-type N2 adsorption-desorption curves with highly enlarged hysteresis loops could be observed for the PBP samples, especially for PBP-3. This phenomenon implies a relatively increased pore size in PBP-3, which was probably caused by the formation of hollow structures. It can be seen that the specific surface area of PAP-2 was higher than that of PAP-1 and PAP-3. Compared to PAP-1 and PAP-3, the molar ratio of DTPH:PA in PAP-2 was 1:1, enabling DTPH and PA to react completely to form PAP, which was helpful for a high porosity. Moreover, there was no excess branch chain from DTPH and PA blocking the pore channels, which was also helpful for a higher specific surface area of PAP-2 than that of PAP-1 and PAP-3. For PBP-3, the relatively higher specific surface area was possibly contributed by the specific hollow nanotube structures covered with spongy pores, which had much higher pore volume than PBP-1 and PBP-2.

2.2. CO2 Adsorption Performance

To mimic the post-combustion CO2 capture, the CO2 adsorption performances of the materials were detected from a simulated flue gas (a mixture of CO2 and N2 with 15 vol.% CO2) at 75 °C in a fixed-bed flow system equipped with a gas analyzer (see Figure 10a). On the whole, all samples presented a rapid adsorption within a short time and reached saturated adsorption within 300 s (Figure 10b). For PAP samples, the CO2 capacity was increased following the order of PBP-1 < PBP-2 < PAP-1< PAP-3 < PAP-2 < PBP-3, which was overall positively correlated with their specific surface areas (see Table 1). Amongst these samples, PBP-3 presented the maximum CO2 capacity (0.84 mmol·g−1) although its surface area was somewhat lower than those of PAP-1 and PAP-2 (see Table 2). This result was probably due to the competitive effect of physisorption and chemisorption [48]. On one hand, the low specific surface area was not in favor of physisorption. On the other hand, however, chemisorption was gradually enhanced caused by the relatively high content of N-site rich motif DPTH, which could counteract the decrease in physisorption.
To further improve the CO2 performance of the dynamic covalent samples, amino-functional molecule tetraethylenepentamine (TEPA) was further introduced on PBP. It can be noted that the direct assembly of TEPA and DTPH failed to afford any solid powder. By virtue of the dynamic nature of imine assembly, TEPA can be facilely grafted on PBP or even replace DPTH through imine exchange [49,50]. Benefiting from the many more active sites in TEPA than that in DPTH, as expected, the CO2 capacity of the TEPA-modified PBP-3 sample (named TEPA-PBP) could reach up to 1.27 mmol·g−1 (Figure 10b), 1.5 times that of bulk PBP-3. The high adsorption capacity for TEPA-PBP was mainly caused by the porous property and good affinity of amino/acylamide groups. The amino groups could act as chemisorbed sites to react with CO2 to form ammonium carbamate [51], and acylamide and imine groups could adsorb CO2 under the effect of hydrogen bonding and electrostatic attraction, respectively (see Scheme 3) [51].
The cyclic stability was also crucial for the industrial application of an adsorbent. The cyclic stability of TEPA-PBP was investigated by performing 10 cycles of CO2 adsorption at 75 °C. It can be seen that the CO2 adsorption capacity of TEPA-PBP remained constant in each cycle (Figure 10c), reaching 1.21 mmol·g−1 of CO2 uptake after 10 cycles, verifying the excellent adsorption reversibility of TEPA-PBP.
The CO2 adsorption performance of the designed material TEPA-PBP was compared with typical reported adsorption materials. As shown in Table 3, TEPA-PBP exhibited competitive CO2 adsorption capacity from flue gas. Superior to most reported adsorbents, moreover, the material developed in this work exhibits facile preparation, low cost, and a metal-free nature, thus promoting its availability in practical application.

3. Materials and Methods

3.1. Materials

Dimethyl 3,3′-dithiobispropionate (98%) and tetraethylenepentamine (98%) were purchased from Shanghai Titan Technology Co., Ltd., Shanghai, China. p-Phthalaldehyde (PA, 98%) and 4,4′-biphenyldicarboxaldehyde (BPDA, 98%) were purchased from Innochem Chemical Co., Beijing, China. CO2 (99.99%) and N2 (99.99%) were purchased from Taiyuan Steel Co., Shanxi, China.

3.2. Synthesis of Materials

3.2.1. Synthesis of 3,3′-Dithiobis(Propionyl Hydrazine)

Dimethyl 3,3′-dithiodipropionate (12.01 g, 50.4 mmol) was added to anhydrous methanol (90 mL), followed by the addition of hydrazine hydrate (20.59 g, 403.2 mmol). After stirring at room temperature for 24 h, a white solid (3,3′-dithiobis(propionyl hydrazine), DTPH) was obtained by centrifugation, washed with methanol (30 mL × 2) and diethyl ether (30 mL × 2), and then dried in vacuum at 50 °C for 12 h. 1H-NMR (δppm, 400 MHz, DMSO-d6): 9.09 (s, 1H), 4.20 (m, 2H), 2.88 (m, 2H), 2.40 (m, 2H).

3.2.2. Synthesis of PAP-X

PA (0.78 g, 3.3 mmol) was dissolved in 15 mL of ethyl acetate at 30 °C, and the solution was slowly added to 15 mL of DTP aqueous solution (0.45 g, 3.3 mmol). The two-phase system was sealed and left standing at 30 °C for 24 h. White powder was collected by centrifugation, washed with water (30 mL × 3) and ethyl acetate (30 mL × 3), and then dried in vacuum at 70 °C for 12 h. The samples with different molar ratios of PA/DTP (i.e., 1:0.5, 1:1, and 1:2) were prepared and named PAP-1, PAP-2, and PAP-3, respectively.

3.2.3. Synthesis of PBP-X

PBP-1 was synthesized following a similar procedure with PAP-1 under the reaction of BPDA and DTP at a molar ratio of 1:0.5. The samples with different molar ratios of BPDA/DTP (i.e., 1:0.5, 1:1, and 1:2) were prepared and named PBP-1, PBP-2, and PBP-3, respectively.

3.2.4. Synthesis of TEPA-PBP

Tetraethylenepentamine (TEPA, 0.03 g, 158 mmol) was dissolved in 15 mL of methanol and stirred at room temperature for 2 h. Then, 0.50 g of PBP-3 was added to the above solution under stirring at room temperature for 8 h. The precipitate was collected by centrifugation, washed with methanol (5 mL × 2) and water (5 mL × 2) in turn, and dried at 60 °C for 12 h.

3.3. Characterizations

The microscopic structures of samples were observed by scanning electron microscopy (SEM) on a ZEISS MERLIN CoMPact (Zeiss, Jena, Germany). Fourier-transform infrared (FT-IR) spectra were recorded on a Tensor 27 spectrometer (Bruker, Billerica, MA, USA) over a KBr pellet in the region of 4000–400 cm−1. X-ray diffraction (XRD) patterns were acquired in the range of 2θ = 10°–80° with scanning rate of 5°/min on PANayltical Empyrean (Almelo, The Netherlands). Thermogravimetry analysis (TGA) was conducted using an STA 449 F3 Jupiter® instrument (Netzsch, Selb, Germany) at a heating rate of 10 °C/min in N2 atmosphere. The specific surface areas and pore structures of samples were measured on a Brunauer-Emmett-Teller (BET) apparatus (JW-BK200, Beijing, China). The micropore size distribution was calculated using the t-plot method, the mesopore size distribution was calculated using the BJH method, and the specific surface area was analyzed using the BET equation.

3.4. CO2 Adsorption Capacity Evaluation

The CO2 adsorption performances were determined in a fixed-bed flow system equipped with a gas analyzer (Gasboard-3100, Wuhan Cubic Optoelectronics Co., Ltd., Wuhan, China.). Typically, 0.5 g samples were placed in a column with diameter of 0.8 cm and heated to 100 °C in N2 (99.999%) at a flow rate of 100 mL·min−1 for 1 h. Then, the column was cooled to 75 °C, and a gas mixture of CO2 (15 vol.%) and N2 flowed through the column. The total pressure was 1 bar, and the CO2 pressure was 0.15 bar. The CO2 concentration was recorded every 1.0 s until returning to 15 vol.%. After adsorption, the column was heated to 120 °C for 1 h in N2 (99.999%) at a flow rate of 100 mL·min−1 to achieve CO2 desorption. The CO2 adsorption capacity (qm, mmol·g−1) of the samples at a certain time (t, s) was calculated using Equation (1) [57].
q m = Q M a d 0 t ( C 0 C t ) d t × T 0 T × 1 V m
where Q is the flow of the mixture (mL·min−1), Mad is the mass of adsorbents, 𝑡 is the adsorption time (s), C0 and Ct are the CO2 concentrations at the inlet and outlet, T is the adsorption temperature (K), and Vm is the standard molar volume (22.4 L·mol−1, T0 = 0 °C).

4. Conclusions

Through dynamic covalent assembly, N-site rich polymers were successfully constructed via the self-assembly of 3,3′-dithiobis (propionyl hydrazine) and polyaldehyde at the ethyl acetate/water interface at 30 °C. By regulating the type and ratio of polyaldehyde, nanospheres and hollow nanotubes with spongy pores were controllably formed. The dynamic covalent materials presented good thermal stability with a maximum decomposition temperature of ca. 292 °C. Tetraethylenepentamine with rich amino-functional groups was successfully grafted onto PBP at room temperature and provided rich chemisorption sites, such that the functional dynamic covalent polymer had a high CO2 capacity (1.27 mmol·g−1) in flue gas at 75 °C. This strategy has the advantages of a simple, green, and efficient preparation process, and it provides a new idea for the controllable design of highly active CO2 adsorbent.

Author Contributions

Investigation, methodology, visualization, writing—original draft, and formal analysis, M.Q.; methodology, H.W.; visualization, Y.H.; formal analysis, H.G., D.G. and F.P.; supervision, conceptualization, funding acquisition, and writing—review and editing, L.S. and Q.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Knowledge Innovation Program of Wuhan Basic Research (2022020801010354), the Joint Fund of the Yulin University and the Dalian National Laboratory for Clean Energy (Grant YLU-DNL Fund 2021021), the National Natural Science Foundation of China (51776133; U1810125), and the Shanxi-Zheda Institute of Advanced Materials and Chemical Engineering (2022SX–TD015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Doney, S.C.; Busch, D.S.; Cooley, S.R.; Kroeker, K.J. The Impacts of Ocean Acidification on Marine Ecosystems and Reliant Human Communities. Annu. Rev. Environ. Resour. 2020, 45, 83–112. [Google Scholar] [CrossRef]
  2. Zhang, T.; Zhang, W.; Yang, R.; Liu, Y.; Jafari, M. CO2 Capture and Storage Monitoring Based on Remote Sensing Techniques: A Review. J. Clean. Prod. 2021, 281, 124409. [Google Scholar] [CrossRef]
  3. Ahmed, D.S.; El-Hiti, G.A.; Yousif, E.; Hameed, A.S.; Abdalla, M. New Eco-Friendly Phosphorus Organic Polymers as Gas Storage Media. Polymers 2017, 9, 336. [Google Scholar] [CrossRef] [PubMed]
  4. Hadi, A.G.; Jawad, K.; Yousif, E.; El-Hiti, G.A.; Alotaibi, M.H.; Ahmed, D.S. Synthesis of Telmisartan Organotin(IV) Complexes and their use as Carbon Dioxide Capture Media. Molecules 2019, 24, 1631. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, R.; Liu, S.; Li, Q.; Zhang, S.; Wang, L.; An, S. CO2 Capture Performance and Mechanism of Blended Amine Solvents Regulated By N-methylcyclohexyamine. Energy 2021, 215, 119209. [Google Scholar] [CrossRef]
  6. He, X. Polyvinylamine-Based Facilitated Transport Membranes for Post-Combustion CO2 Capture: Challenges and Perspectives from Materials to Processes. Engineering 2021, 7, 124–131. [Google Scholar] [CrossRef]
  7. Li, Z.L.; Zhou, Y.L.; Yan, W.; Luo, L.; Su, Z.Z.; Fan, M.Z.; Wang, S.R.; Zhao, W.G. Cost-Effective Monolithic Hierarchical Carbon Cryogels with Nitrogen Doping and High-Performance Mechanical Properties for CO2 Capture. ACS. Appl. Mater. Inter. 2020, 12, 21748–21760. [Google Scholar] [CrossRef]
  8. Xu, L.; Xing, C.Y.; Ke, D.; Chen, L.; Qiu, Z.J.; Zeng, S.L.; Li, B.J.; Zhang, S. Amino-Functionalized beta-Cyclodextrin to Construct Green Metal-Organic Framework Materials for CO2 Capture. ACS Appl. Mater. Inter. 2020, 12, 3032–3041. [Google Scholar] [CrossRef]
  9. Faria, A.C.; Trujillano, R.; Rives, V.; Miguel, C.V.; Rodrigues, A.E.; Madeira, L.M. Alkali metal (Na, Cs and K) Promoted Hydrotalcites for High Temperature CO2 Capture from Flue Gas in Cyclic Adsorption Processes. Chem. Eng. J. 2022, 427, 131502. [Google Scholar] [CrossRef]
  10. Omer, R.M.; Al-Tikrity, E.T.B.; El-Hiti, G.A.; Alotibi, M.F.; Ahmed, D.S.; Yousif, E. Porous Aromatic Melamine Schiff Bases as Highly Efficient Media for Carbon Dioxide Storage. Processes 2019, 8, 17. [Google Scholar] [CrossRef] [Green Version]
  11. Shi, X.; Xiao, H.; Azarabadi, H.; Song, J.; Wu, X.; Chen, X.; Lackner, K.S. Sorbents for the Direct Capture of CO2 from Ambient Air. Angew. Chem. Int. Ed. 2020, 59, 6984–7006. [Google Scholar] [CrossRef]
  12. Choi, H.J.; Min, J.G.; Ahn, S.H.; Shin, J.; Hong, S.B.; Radhakrishnan, S.; Chandran, C.V.; Bell, R.G.; Breynaert, E.; Kirschhock, C.E.A. Framework Flexibility-Driven CO2 Adsorption on a Zeolite. Mater. Horiz. 2020, 7, 1528–1532. [Google Scholar] [CrossRef]
  13. Rattanaphan, S.; Rungrotmongkol, T.; Kongsune, P. Biogas Improving by Adsorption of CO2 on Modified Waste Tea Activated Carbon. Renew. Energy 2020, 145, 622–631. [Google Scholar] [CrossRef]
  14. Chen, C.; Xu, H.; Jiang, Q.; Lin, Z. Rational Design of Silicas with Meso-Macroporosity as Supports for High-Performance Solid Amine CO2 Adsorbents. Energy 2021, 214, 119093. [Google Scholar] [CrossRef]
  15. Vadillo, J.M.; Gómez-Coma, L.; Garea, A.; Irabien, A. Hollow Fiber Membrane Contactors in CO2 Desorption: A Review. Energy Fuels 2020, 35, 111–136. [Google Scholar] [CrossRef]
  16. Lee, R.J.; Jawad, Z.A.; Ahmad, A.L.; Chua, H.B. Incorporation of Functionalized Multi-Walled Carbon Nanotubes (MWCNTs) into Cellulose Acetate Butyrate (CAB) Polymeric Matrix to Improve the CO2/N2 Separation. Process Saf. Environ. 2018, 117, 159–167. [Google Scholar] [CrossRef]
  17. Chen, M.; Xu, J.; You, Z.; Dong, Y.; Yuan, H.; Yin, J.; Zhou, X. Synergetic Effect of Liquid/Plasma on Wet Wheat Straw Porous Carbon: Pore Structure, Heteroatom Doping and CO2 adsorption. Mater. Lett. 2020, 262, 127185. [Google Scholar] [CrossRef]
  18. Mutyala, S.; Jonnalagadda, M.; Ibrahim, S.M. Effect of Modification of UiO-66 for CO2 Adsorption and Separation of CO2/CH4. J. Mol. Struct. 2021, 1227, 129506. [Google Scholar] [CrossRef]
  19. Gautam, S.; Cole, D. CO2 Adsorption in Metal-Organic Framework Mg-MOF-74: Effects of Inter-Crystalline Space. Nanomaterials 2020, 10, 2274. [Google Scholar] [CrossRef]
  20. Wang, J.; Wang, F.; Duan, H.; Li, Y.; Xu, J.; Huang, Y.; Liu, B.; Zhang, T. Polyvinyl Chloride-Derived Carbon Spheres for CO2 Adsorption. ChemSusChem 2020, 13, 6426–6432. [Google Scholar]
  21. Avci, G.; Erucar, I.; Keskin, S. Do New MOFs Perform Better for CO2 Capture and H2 Purification? Computational Screening of the Updated MOF Database. ACS. Appl. Mater. Inter. 2020, 12, 41567–41579. [Google Scholar] [CrossRef] [PubMed]
  22. Hisaki, I.; Xin, C.; Takahashi, K.; Nakamura, T. Designing Hydrogen-Bonded Organic Frameworks (HOFs) with Permanent Porosity. Angew. Chem. Int. Ed. 2019, 58, 11160–11170. [Google Scholar] [CrossRef] [PubMed]
  23. Zeng, Y.; Zou, R.; Zhao, Y. Covalent Organic Frameworks for CO2 Capture. Adv. Mater. 2016, 28, 2855–2873. [Google Scholar] [CrossRef] [PubMed]
  24. Inukai, M.; Tamura, M.; Horike, S.; Higuchi, M.; Kitagawa, S.; Nakamura, K. Storage of CO2 into Porous Coordination Polymer Controlled by Molecular Rotor Dynamics. Angew. Chem. Int. Ed. 2018, 57, 8687–8690. [Google Scholar] [CrossRef]
  25. Lin, Y.; Chen, Y.; Yu, Z.; Huang, Z.; Lai, J.-C.; Tok, J.B.H.; Cui, Y.; Bao, Z. Reprocessable and Recyclable Polymer Network Electrolytes via Incorporation of Dynamic Covalent Bonds. Chem. Mater. 2022, 34, 2393–2399. [Google Scholar] [CrossRef]
  26. Afrose, S.P.; Mahato, C.; Sharma, P.; Roy, L.; Das, D. Nonequilibrium Catalytic Supramolecular Assemblies of Melamine- and Imidazole-Based Dynamic Building Blocks. J. Am. Chem. Soc. 2022, 144, 673–678. [Google Scholar] [CrossRef]
  27. Blanke, M.; Postulka, L.; Ciara, I.; D’Acierno, F.; Hildebrandt, M.; Gutmann, J.S.; Dong, R.Y.; Michal, C.A.; Giese, M. Manipulation of Liquid Crystalline Properties by Dynamic Covalent Chemistry horizontal line En Route to Adaptive Materials. ACS. Appl. Mater. Inter. 2022, 14, 16755–16763. [Google Scholar] [CrossRef]
  28. Chen, S.; Liu, M.; Zhang, J.; Zhang, Z.; Zhu, J.; Pan, X.; Zhu, X. Photoresponsive Dynamic Covalent Bond Based On Addition–Fragmentation Chain Transfer of Allyl Selenides. Polym. Chem. 2021, 12, 1622–1626. [Google Scholar] [CrossRef]
  29. Vo, T.K.; Kim, W.-S.; Kim, J. Ethylenediamine-incorporated MIL-101(Cr)-NH2 Metal-Organic Frameworks for Enhanced CO2 Adsorption. Korean J. Chem. Eng. 2020, 37, 1206–1211. [Google Scholar] [CrossRef]
  30. Sanz-Pérez, E.S.; Lobato, B.; Lopez-Anton, M.A.; Arencibia, A.; Sanz, R.; Martínez-Tarazona, M.R. Effectiveness of Amino-Functionalized Sorbents for CO2 Capture in the Presence of Hg. Fuel 2020, 267, 117250. [Google Scholar] [CrossRef]
  31. Guo, Y.; Luo, L.; Zheng, Y.; Zhu, T. Optimization of CO2 Adsorption on Solid-Supported Amines and Thermal Regeneration Mode Comparison. ACS Omega 2020, 5, 9641–9648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wang, D.; Xue, Y.; Wang, C.; Ji, J.; Zhou, Z.; Tang, C. Improved Capture of Carbon Dioxide and Methane via Adding Micropores Within Porous Boron Nitride Fibers. J. Mater. Sci. 2019, 54, 10168–10178. [Google Scholar] [CrossRef]
  33. Rehman, A.; Farrukh, S.; Hussain, A.; Fan, X.; Pervaiz, E. Adsorption of CO2 on amine-functionalized green metal-organic framework: An interaction between amine and CO2 molecules. Environ. Sci. Pollut. Res. Int. 2019, 26, 36214–36225. [Google Scholar] [CrossRef] [PubMed]
  34. Putz, A.M.; Ciopec, M.; Negrea, A.; Grad, O.; Ianasi, C.; Ivankov, O.I.; Milanovic, M.; Stijepovic, I.; Almasy, L. Comparison of Structure and Adsorption Properties of Mesoporous Silica Functionalized with Aminopropyl Groups by the Co-Condensation and the Post Grafting Methods. Materials 2021, 14, 628. [Google Scholar] [CrossRef]
  35. Lawson, S.; Griffin, C.; Rapp, K.; Rownaghi, A.A.; Rezaei, F. Amine-Functionalized MIL-101 Monoliths for CO2 Removal from Enclosed Environments. Energy Fuels 2019, 33, 2399–2407. [Google Scholar] [CrossRef]
  36. De ON Ribeiro, J.; Nunes, E.H.M.; Vasconcelos, D.C.L.; Vasconcelos, W.L.; Nascimento, J.F.; Grava, W.M.; Derks, P.W.J. Role of the Type of Grafting Solvent and Its Removal Process on APTES Functionalization onto SBA-15 Silica for CO2 Adsorption. J. Porous Mater. 2019, 26, 1581–1591. [Google Scholar] [CrossRef]
  37. Ji, C.; Huang, X.; Li, L.; Xiao, F.; Zhao, N.; Wei, W. Pentaethylenehexamine-Loaded Hierarchically Porous Silica for CO2 Adsorption. Materials 2016, 9, 835. [Google Scholar] [CrossRef]
  38. Xie, W.; Huang, S.; Tang, D.; Liu, S.; Zhao, J. Biomass-derived Schiff Base Compound Enabled Fire-Safe Epoxy Thermoset with Excellent Mechanical Properties and High Glass Transition Temperature. Chem. Eng. J. 2020, 394, 123667. [Google Scholar] [CrossRef]
  39. Liu, X.; Liang, L.; Lu, M.; Song, X.; Liu, H.; Chen, G. Water-Resistant Bio-Based Vitrimers Based on Dynamic Imine Bonds: Self-healability, Remodelability and Ecofriendly Recyclability. Polymer 2020, 210, 123030. [Google Scholar] [CrossRef]
  40. Wang, Y.; Xiao, G.; Peng, Y.; Chen, L.; Fu, S. Effects of Cellulose Nanofibrils on Dialdehyde Carboxymethyl Cellulose Based Dual Responsive Self-Healing Hydrogel. Cellulose 2019, 26, 8813–8827. [Google Scholar] [CrossRef]
  41. He, X.; Yang, Y.; Wu, H.; He, G.; Xu, Z.; Kong, Y.; Cao, L.; Shi, B.; Zhang, Z.; Tongsh, C.; et al. De Novo Design of Covalent Organic Framework Membranes toward Ultrafast Anion Transport. Adv. Mater. 2020, 32, e2001284. [Google Scholar] [CrossRef] [PubMed]
  42. Fan, C.; Wu, H.; Guan, J.; You, X.; Yang, C.; Wang, X.; Cao, L.; Shi, B.; Peng, Q.; Kong, Y.; et al. Scalable Fabrication of Crystalline COF Membranes from Amorphous Polymeric Membranes. Angew. Chem. Int. Ed. 2021, 60, 18051–18058. [Google Scholar] [CrossRef] [PubMed]
  43. Wehner, M.; Würthner, F. Supramolecular polymerization through kinetic pathway control and living chain growth. Nat. Rev. Chem. 2019, 4, 38–53. [Google Scholar] [CrossRef]
  44. Adelizzi, B.; Aloi, A.; Markvoort, A.J.; Ten Eikelder, H.M.M.; Voets, I.K.; Palmans, A.R.A.; Meijer, E.W. Supramolecular block copolymers under thermodynamic control. J. Am. Chem. Soc. 2018, 140, 7168–7175. [Google Scholar] [CrossRef] [PubMed]
  45. Vantomme, G.; Ter Huurne, G.M.; Kulkarni, C.; Ten Eikelder, H.M.M.; Markvoort, A.J.; Palmans, A.R.A.; Meijer, E.W. Tuning the length of cooperative supramolecular polymers under thermodynamic control. J. Am. Chem. Soc. 2019, 141, 18278–18285. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, S.; Jiang, C.; Fan, J.; Hong, S.; Mei, P.; Yao, R.; Liu, Y.; Zhang, S.; Li, H.; Zhang, H.; et al. Hydrophilicity gradient in covalent organic frameworks for membrane distillation. Nat. Mater. 2021, 20, 1551–1558. [Google Scholar] [CrossRef]
  47. Belowich, M.E.; Stoddart, J.F. Dynamic imine chemistry. Chem. Soc. Rev. 2012, 41, 2003–2024. [Google Scholar] [CrossRef]
  48. Alves, L.D.S.; Moreira, B.R.A.; Viana, R.D.S.; Dias, E.S.; Rinker, D.L.; Pardo-Gimenez, A.; Zied, D.C. Spent Mushroom Substrate Is Capable of Physisorption-Chemisorption of CO2. Environ. Res. 2022, 204, 111945. [Google Scholar] [CrossRef]
  49. Vitaku, E.; Dichtel, W.R. Synthesis of 2D Imine-Linked Covalent Organic Frameworks through Formal Transimination Reactions. J. Am. Chem. Soc. 2017, 139, 12911–12914. [Google Scholar] [CrossRef]
  50. Shan, Z.; Wu, X.; Xu, B.; Hong, Y.L.; Wu, M.; Wang, Y.; Nishiyama, Y.; Zhu, J.; Horike, S.; Kitagawa, S.; et al. Dynamic Transformation between Covalent Organic Frameworks and Discrete Organic Cages. J. Am. Chem. Soc. 2020, 142, 21279–21284. [Google Scholar] [CrossRef]
  51. Mao, H.Y.; Tang, J.; Day, G.S.; Peng, Y.C.; Wang, H.Z.; Xiao, X.; Yang, Y.F.; Jiang, Y.W.; Chen, S.; Halat, D.M.; et al. A scalable solid-state nanoporous network with atomic-level interaction design for carbon dioxide capture. Sci. Adv. 2022, 8, eabo6849. [Google Scholar] [CrossRef] [PubMed]
  52. Ma, X.; Yang, Y.; Wu, Q.; Liu, B.; Li, D.; Chen, R.; Wang, C.; Li, H.; Zeng, Z.; Li, L. Underlying Mechanism of CO2 Uptake onto Biomass-Based Porous Carbons: Do Adsorbents Capture CO2 Chiefly Through Narrow Micropores? Fuel 2020, 282, 118727. [Google Scholar] [CrossRef]
  53. Kong, W.; Liu, J. Ordered Mesoporous Carbon with Enhanced Porosity to Support Organic Amines: Efficient Nanocomposites for the Selective Capture of CO2. New J. Chem. 2019, 43, 6040–6047. [Google Scholar] [CrossRef]
  54. Alkadhem, A.M.; Elgzoly, M.A.A.; Onaizi, S.A. Novel Amine-Functionalized Magnesium Oxide Adsorbents for CO2 Capture at Ambient Conditions. J. Environ. Chem. Eng. 2020, 8, 103968. [Google Scholar] [CrossRef]
  55. Fang, Q.; Huang, W.; Wang, H. Role of Additives in Silica-Supported Polyethylenimine Adsorbents for CO2 Adsorption. Mater. Res. Express 2020, 7, 035026. [Google Scholar] [CrossRef]
  56. Mohamad, N.A.; Zubair, N.A.; Lotf, E.A.; Nasef, M.M.; Ahmad, A.; Ali, R.R.; Abdullah, T.A.T.; Ting, T.M. Effect of Ligand Type on CO2 Adsorption over Amine Functionalized Fibrous Adsorbents. IOP Conf. Ser. Mater. Sci. Eng. 2020, 808, 012009. [Google Scholar] [CrossRef]
  57. Rong, X.; Ettelaie, R.; Lishchuk, S.V.; Cheng, H.; Zhao, N.; Xiao, F.; Cheng, F.; Yang, H. Liquid Marble-Derived Solid-Liquid Hybrid Superparticles for CO2 Capture. Nat. Commun. 2019, 10, 1854. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Schematic diagram of the synthesis of PAP at the ethyl acetate/water interface at 30 °C.
Scheme 1. Schematic diagram of the synthesis of PAP at the ethyl acetate/water interface at 30 °C.
Molecules 27 05853 sch001
Figure 1. FT-IR spectra of PA, DTPH, and the resultant dynamic covalent materials PAP-X (X = 1, 2, 3).
Figure 1. FT-IR spectra of PA, DTPH, and the resultant dynamic covalent materials PAP-X (X = 1, 2, 3).
Molecules 27 05853 g001
Figure 2. Solid-state 13C-NMR spectrum of PAP-1.
Figure 2. Solid-state 13C-NMR spectrum of PAP-1.
Molecules 27 05853 g002
Figure 3. SEM images of (a,b) PAP-1, (c,d) PAP-2, and (e,f) PAP-3.
Figure 3. SEM images of (a,b) PAP-1, (c,d) PAP-2, and (e,f) PAP-3.
Molecules 27 05853 g003
Figure 4. FT-IR spectra of BPDA, DTPH, and the resultant dynamic covalent materials PBP-X (X = 1, 2, 3).
Figure 4. FT-IR spectra of BPDA, DTPH, and the resultant dynamic covalent materials PBP-X (X = 1, 2, 3).
Molecules 27 05853 g004
Scheme 2. Schematic diagram of the synthesis of PBP at the ethyl acetate/water interface at 30 °C.
Scheme 2. Schematic diagram of the synthesis of PBP at the ethyl acetate/water interface at 30 °C.
Molecules 27 05853 sch002
Figure 5. SEM images of (a,b) PBP-1, (c,d) PBP-2, and (e,f) PBP-3.
Figure 5. SEM images of (a,b) PBP-1, (c,d) PBP-2, and (e,f) PBP-3.
Molecules 27 05853 g005
Figure 6. SEM images of dynamic covalent material PAP-3 formed at 0.5 h (a) and 3 h (b).
Figure 6. SEM images of dynamic covalent material PAP-3 formed at 0.5 h (a) and 3 h (b).
Molecules 27 05853 g006
Figure 7. XRD spectra of (a) PAP and (b) PBP.
Figure 7. XRD spectra of (a) PAP and (b) PBP.
Molecules 27 05853 g007
Figure 8. TGA and DTG curves of (a) PAP-1 and (b) PBP-1.
Figure 8. TGA and DTG curves of (a) PAP-1 and (b) PBP-1.
Molecules 27 05853 g008
Figure 9. N2 adsorption and desorption curves of PAP (a) and PBP (b) samples.
Figure 9. N2 adsorption and desorption curves of PAP (a) and PBP (b) samples.
Molecules 27 05853 g009
Figure 10. (a) Schematic diagram of the fixed-bed flow system. (b) CO2 adsorption curves of PAP and PBP samples from a gas mixture of CO2 (15 vol.%)/N2 at 75 °C. (c) Ten cycles of CO2 capacity of TEPA-PBP from a gas mixture of CO2 (15 vol.%)/N2 at 75 °C.
Figure 10. (a) Schematic diagram of the fixed-bed flow system. (b) CO2 adsorption curves of PAP and PBP samples from a gas mixture of CO2 (15 vol.%)/N2 at 75 °C. (c) Ten cycles of CO2 capacity of TEPA-PBP from a gas mixture of CO2 (15 vol.%)/N2 at 75 °C.
Molecules 27 05853 g010
Scheme 3. Mechanistic diagram of CO2 absorption on dynamic covalent material TEPA-PBP.
Scheme 3. Mechanistic diagram of CO2 absorption on dynamic covalent material TEPA-PBP.
Molecules 27 05853 sch003
Table 1. Specific surface area and pore volume of PAP and PBP samples.
Table 1. Specific surface area and pore volume of PAP and PBP samples.
SampleSBET a (m2·g−1)Vtotal b (cm3·g−1)Vmeso c (cm3·g−1)
PAP-172.570.270.27
PAP-277.910.360.36
PAP-358.200.280.28
PBP-117.270.090.19
PBP-233.120.160.16
PBP-363.120.320.32
a BET surface area. b Single-point pore volume calculated at relative pressure p/p0 of 0.99. c Vtotal − Vmicro.
Table 2. CO2 adsorption capacity of different adsorbents.
Table 2. CO2 adsorption capacity of different adsorbents.
EntryAdsorbentCO2 Capacity (mmol·g−1)CO2 Capacity (wt.%)
1PAP-10.381.67
2PAP-20.632.77
3PAP-30.421.85
4PBP-10.160.70
5PBP-20.220.97
6PBP-30.843.70
7TEPA-PBP1.275.59
Table 3. Comparison of the CO2 adsorption performance of TEPA-PBP with advanced adsorbents.
Table 3. Comparison of the CO2 adsorption performance of TEPA-PBP with advanced adsorbents.
EntryAdsorbentAdsorption Temperature (°C)CO2 Capacity (mmol·g−1)Ref.
1TEPA-PBP751.27This work
2OC700251.36[52]
30.44PEI@AOMC751.14[53]
4PEI-MgO300.54[54]
5SA-PEI-TEA700.61[55]
6PE/PP-g-PVBC-EDA502.7[56]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qiu, M.; Wu, H.; Huang, Y.; Guo, H.; Gao, D.; Pei, F.; Shi, L.; Yi, Q. Controllable Construction of Amino-Functionalized Dynamic Covalent Porous Polymers for High-Efficiency CO2 Capture from Flue Gas. Molecules 2022, 27, 5853. https://doi.org/10.3390/molecules27185853

AMA Style

Qiu M, Wu H, Huang Y, Guo H, Gao D, Pei F, Shi L, Yi Q. Controllable Construction of Amino-Functionalized Dynamic Covalent Porous Polymers for High-Efficiency CO2 Capture from Flue Gas. Molecules. 2022; 27(18):5853. https://doi.org/10.3390/molecules27185853

Chicago/Turabian Style

Qiu, Mingyue, Haonan Wu, Yi Huang, Huijuan Guo, Dan Gao, Feng Pei, Lijuan Shi, and Qun Yi. 2022. "Controllable Construction of Amino-Functionalized Dynamic Covalent Porous Polymers for High-Efficiency CO2 Capture from Flue Gas" Molecules 27, no. 18: 5853. https://doi.org/10.3390/molecules27185853

Article Metrics

Back to TopTop