Polymers of Intrinsic Microporosity Chemical Sorbents Utilizing

Aug 1, 2019 - Here, we present novel chemical sorbents based on polymers with intrinsic ... which can be synthesized inexpensively and under mild reac...
0 downloads 0 Views 489KB Size
Subscriber access provided by RUTGERS UNIVERSITY

Energy, Environmental, and Catalysis Applications

Polymers with intrinsic microporosity chemical sorbents utilizing primary amine appendance through acid-base and hydrogen bonding interactions Ali Kemal Sekizkardes, Sonia Hammache, James S. Hoffman, and David P. Hopkinson ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.9b09856 • Publication Date (Web): 01 Aug 2019 Downloaded from pubs.acs.org on August 5, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Polymers of intrinsic microporosity chemical sorbents utilizing primary amine appendance through acid-base and hydrogen bonding interactions Ali K. Sekizkardes,a,b* Sonia Hammache,a,b James S. Hoffman,a David Hopkinsona aNational

Energy Technology Laboratory, 626 Cochrans Mill Road, P.O. Box 10940, Pittsburgh, PA 15236-0940, USA bLeidos Research Support Team, 626 Cochrans Mill Road, P.O. Box 10940, Pittsburgh, PA 15236-0940, USA Supporting Information Placeholder ABSTRACT: Here we present novel chemical sorbents based on polymers with intrinsic microporosity (PIM). For the first time, alkyl amines were incorporated in PIMs through an acid-base interaction to create a chemi-sorbent. The amine-appended PIM polymers not only showed a nearly four-fold enhancement in CO2 loading capacity (36.4 cc/g at 0.15 bar and 298 K) and very high CO2/N2 selectivity compared to neat PIM-1, but they also proved to have stable performance when cycled between adsorption and desorption under both dry and humid conditions that are typical for post-combustion CO2 capture. KEYWORDS: PIM-1, porous organic polymers, amine impregnation, chemical sorbent, CO2 capture INTRODUCTION

Porous sorbents are one class of material being studied for use in CO2 capture applications.1 An increasing number of sorbent materials have been reported in the literature, from silica-based sorbents to activated/porous carbon, and more recently metal-organic frameworks.2 In the last decade, a new class of porous materials, porous organic polymers (POPs) have emerged, including porous aromatic frameworks (PAFs), porous polymeric networks (PPNs), benzimidazole linked polymers (BILPs) and hyper crosslinked polymers (HPCs).3 In general, POPs have been reported as high surface area materials with a highly stable polymer structure resulting from the covalent bonding between the monomers.4 However, the CO2 uptake capacity of most POPs is not able to exceed 20 cc/g (at 0.15 bar CO2 and 298 K), as the interaction between CO2 and POPs is primarily due to physisorption. Although there have been several efforts to append primary amines to POPs through either amineimpregnation or grafting methods, drawbacks such as harsh synthesis, poor scalability, and poor processability have been a hurdle for POPs as a breakthrough for CO2 capture.5,6 Polymers with intrinsic microporosity (PIMs) are wellstudied porous organic polymers which can be synthesized inexpensively and under mild reaction conditions.7 In contrast to most POPs, PIMs can be

processed into thin films and fibers.8 Consequently the majority of the studies on PIMs have focused on gas separation membrane applications in which they feature exceptionally high permeability and moderate selectivity for several different light gas pairs.9,10 Although PIMbased membranes have been among the best performing gas separation materials, few studies have examined PIMs as solid sorbents for CO2 capture or other gas separations.11-13 While PIMs possess the high surface area and permanent microporosity desired for a sorbent, they also suffer from low CO2 adsorption capacity (1nm) nonpolar micropores as well as some mesopores.14 Here we synthesized the most studied PIM, PIM-1, and post synthetically functionalized PIM-1 with carboxylic acid (-COOH) and amide (CONH2) functional groups to create a sorbent media with a moderate surface area and strong bonding sites for primary amines (Figure 1, SI-1). We report the gas separation performance for this sorbent using both dry and humidified feed gas under conditions that are relevant to post-combustion CO2 capture. DISCUSSION

PIM-1 was synthesized using the method reported by Budd et al. and was modified using a reaction temperature of 58 oC and a shorter duration of reaction (40hr) to afford higher surface area (SI, experimental section).7,15 As-synthesized PIM-1 showed a high surface area of 840 m2/g with a pore size distribution around 1 nm (Figure SI-2). Subsequently, PIM-1 was postsynthetically functionalized with carboxylic acid groups to afford the hydrolyzed polymer, PIM-1-C.16 The degree of the polymer functionalization was controlled by adjusting the time and temperature of the reaction. Hydrolyzed polymers were denoted as PIM-1-C1, PIM1-C2, and PIM-1-C3 for the reaction times of 24 hr at 25 oC, 1.5 hr at 120 oC, and 3.5 hr at 120 oC, respectively. A distinct product color change was observed from fluorescent yellow to off-white as the degree of functionalization increased (Figure SI-1).

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 7

Figure 1. (a) Schematic synthesis of PIM-1-C3 and PIM-1-C3-TA, (b) FT-IR spectra for neat PIM-1, PIM-TA, PIM-C3 and PIM-C3-TA; and (c) pore size distribution of neat PIM-1, PIM-TA, PIM-C3 and PIM-C3-TA. more stable sorbent system with stronger interaction. Fourier transform infrared spectroscopy (FT-IR) was Polymers were treated with the same amount of TAEA performed on the polymers to characterize the hydrolysis by maintaining the same preparation conditions (solvent reaction of nitrile (-CN) functional groups of PIM-1 concentration, temperature, etc.). Final sorbents were (Figure SI-4). The corresponding FT-IR stretching band denoted PIM-1-TA, PIM-1-C1-TA, PIM-1-C2-TA, and intensity for -CN at 2250 cm-1 was found to decrease with PIM-1-C3-TA in the order of increasing hydrolysis of higher degrees of polymer hydrolysis, while the broad nitrile groups in PIM-1 (Figure SI-1). peak intensities between 3000-3500 cm-1 and at 1660 cm1, representing -COOH and -C=O bonds, were found to FT-IR analysis was performed on the amine-loaded be more prominent. Additional absorption bands at 1612 polymers as well. Amine loaded polymers showed a cm-1 showed that a significant amount of amide similar trend in their FT-IR absorption bands with respect functional groups were also present in the polymer.17 The to their post-synthetic functionalization degree (Figure functionalization degree of hydrolyzed PIMs was SI-5). FT-IR spectra of the most highly hydrolyzed PIMcalculated as 6%, 48% and 92 % for PIM-1-C1, PIM-11 sample, PIM-1-C3 and its TAEA appended sorbent, C2 and PIM-1-C3, respectively, from FT-IR absorption PIM-1-C3-TA, are depicted in Figure 1b. As a control bands of -CN at 2240 cm-1, relative to -CH bands in 2800sample, neat PIM-1 was also loaded with TAEA under 3010 cm-1 region.18 the same conditions so that the degree of interaction with Functionalized PIM-1 and PIM-1-C polymers were the amine could be compared between the hydrolyzed treated with the primary amine, tris(2-aminoethyl)amine PIM-1 (PIM-1-C) and neat PIM-1. The unfunctionalized (TAEA). In the literature, primary amines were PIM-1 sample loaded with TAEA (PIM-1-TA) showed impregnated in sorbents through a solvent evaporation nearly the same FT-IR absorption spectrum as the neat PIM-1. Weak and broad peak intensities observed at method wherein the loaded amines were trapped in a sorbent media.19 In this method, the desired amount of 3200-3500 cm-1 range suggested very minimal amineprimary amine is dissolved in solvents and porous retention in the PIM-1 polymer, implying that most of the TAEA leached out from PIM-1 during the solvent substrates are mixed with these solutions. Subsequently, washing purification step. This can be attributed to the solvent evaporation affords an amine-impregnated non-polar (hydrophobic) polymer structure of PIM-1 sorbent. We followed a less common sorbent preparation method in which an amine solution (in methanol) was which does not interact with polar/basic guest molecules such as TAEA. stirred with the polymer, excess amine solution was decanted, and the sorbents were thoroughly washed with solvent.20 The former method provides higher amineloading as nearly all of the impregnated amine becomes trapped within the sorbent. The latter method, on the other hand, retains less primary amine but provides a

ACS Paragon Plus Environment

2

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

On the other hand, amine appendence in the hydrolyzed PIM-1 sample, PIM-1-C3-TA, showed a distinct FT-IR spectrum compared to PIM-1-C3. Characteristic N-H stretching bands for primary amines (-NH2) at 3200-3500 cm-1 were observed for PIM-1-C3-TA. Also, FT-IR absorption bands of PIM-1-C3-TA showed a noticeable shift to higher wavenumbers at 1606 cm-1 and 1672 cm-1 when compared with PIM-1-C3 (Figure SI-6). The FT-IR spectrum of the sorbent PIM-1-C3-TA not only showed that TAEA was successfully appended in the sorbent, but it also suggests that primary amines will only interact strongly with PIM based sorbents that have been functionalized with a compatible appendage such as carboxylic acid. Thermogravimetric analysis (TGA) was performed on sorbents to quantify the amine loading in polymers. TGA showed that the amine amount in PIM-1C1-TA, PIM-1-C2-TA and PIM-1-C3-TA was 10.6, 12.2 and 13.8 wt%, respectively (Figure SI-7). Surface area analysis can also be used to confirm polymer functionalization and appending of amines. A polymer with higher surface area can accommodate a greater number of functional groups, which will increase the performance of the sorbent material. Previous reports indicated that hydrolysis of PIM-1 into carboxylic acid and amide functionalized PIM-1 resulted in a reduction of surface area after functionalization.21 Brunauer– Emmett–Teller (BET) surface areas of the polymers were calculated from nitrogen isotherms collected at 77K. An increasing degree of PIM-1 hydrolysis into PIM-1-C decreased the surface area, which was evidence that the

functionalization was successful. Notably, PIM-1-C polymers had high surface area (495-696 m2/g) in every case (Figure SI-2). N2 adsorption (at 77K) of sorbents was found to be lower compared to the functionalized PIMs, indicating the amine intercalation in pores of sorbents (Figure SI-3). N2 adsorption isotherms were used to calculate pore size distributions (PSDs) using non-local density functional theory (NLDFT). Hydrolyzed PIM-1-C showed a slight decrease in pore size relative to PIM-1 (Figure 1c), indicating that some pore space was occupied by functional groups. Appending amines to the hydrolyzed PIMs caused a shift in the PSD to larger pore sizes, indicating that primary amines were mostly immobilized in the smaller micro-pore region (compare PIM-1-C3-TA to PIM-1-C3 in Figure 1C). More importantly, the shift in the PSD was more pronounced with greater degrees of carboxylic acid functionalization (Figure SI-3). This supports the hypothesis that primary amines are retained in the sorbents based on the concentration of carboxylic acid functional groups. In the most extreme case, we calculated the PSD of PIM-1-TA without the presence of any carboxylic acid functional groups (Figure 1C). The high concentration of micropores in PIM-1-TA indicates that the pores remain unblocked by amine guest molecules. This comparison suggests that primary amines were only immobilized in the sorbents possessing amine-interaction sites (e.g. COOH, amide, etc.) found in functionalized PIM-1.

Figure 2. (A) CO2 adsorption (filled circles) and desorption (open circles) isotherms of neat PIM-1, PIM-1-C3 and PIM-1C3-TA; (B) CO2 adsorption cyclability test at 0.15 bar and 298 K; and (C) CO2 isosteric heats of adsorption for PIM-1, PIM1-C3, and PIM-1-C3-TA.

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table 1: Summary of CO2 uptake performance and Qst of sorbents. Sorbent CO2 uptake (cc/g) CO2 uptake CO2 uptake at 0.15 bar and 298 (cc/g) at 1 bar (cc/g) at 0.15 K and 298 K bar and 313 K PIM-1 PIM-1-TA PIM-1-C1-TA PIM-1-C2-TA PIM-1-C3-TA

9.7 12.7 30.8 32.4 36.4

38.0 41.0 49.8 48.9 50.9

The CO2 uptake of the sorbents was evaluated from CO2 adsorption isotherms collected using a volumetric gas adsorption analyzer (Table 1). Neat PIM-1 showed low CO2 adsorption capacity (9.7 cc/g, 298 K) at 0.15 bar, which is a typical partial pressure of CO2 in coal derived post-combustion flue gas (Figure 2A).22 Hydrolyzed PIM-1 (PIM-1-C3) performed slightly better compared to neat PIM-1, due to the fact that carboxylic acid and amide functional groups can interact with CO2 molecules.23 Loading primary amines into the hydrolyzed PIMs resulted in a drastic increase in CO2 uptake. PIM-1-C1TA showed higher CO2 capacity compared to PIM-1-C2TA at high CO2 pressure, despite having less hydrolysis functionalization (Table 1, Figure SI-9 and SI-10). This result can be attributed to higher surface area of the former (Figure SI-2). On the other hand, PIM-1-C2-TA showed steeper CO2 uptake at low pressure, suggesting more amine loading in the sorbent compared to PIM-1C1-TA. PIM-1-C3-TA, which has the highest degree of nitrile hydrolysis showed the highest CO2 uptake performance of 36.4 cc/g at 0.15 bar and 298K. This value is nearly four-fold higher than the CO2 uptake of neat PIM-1 and is the highest amount by any PIM-based sorbent that has been reported to date.11-13,24 Six adsorption and regeneration cycles were performed for PIM-1-C3-TA where adsorption was at 0.15 bar and 298 K and desorption was at vacuum and 358 K. Figure 2B shows that the sorbent retained its full capacity after each desorption step. A noticeable hysteresis between adsorption and desorption isotherms also indicates that the primary mode of adsorption for PIM-1-C-TA sorbents is chemisorption (Figure 2A). A high isosteric heat of adsorption (Qst) for CO2 can serve as an indicator of chemisorption. Qst values were calculated with the virial equation by fitting CO2 isotherms collected at 298 K and 313 K.25 High Qst values (75.8-86.5 kJ/mol) showed that PIM-1-C-TA sorbents retained a considerable amount of primary and secondary amine functional groups (Table 1). On the other hand, the Qst of neat PIM-1 and hydrolyzed PIM-1-C3 sorbents were calculated as 32 and 35 kJ/mol, respectively (Figure 2C), similar to the reported values.12 Nitrogen adsorption was also measured to calculate the CO2/N2 selectivity. Henry`s law initial slope calculations

5.9 24.8 26.7 31.1

Page 4 of 7

CO2 uptake (cc/g) at 1 bar and 313 K

Qst for CO2 (kJ/mol)

25.2 39.4 40.3 44.8

32.4 75.8 83.8 86.5

showed that PIM-1-C3-TA has a CO2/N2 selectivity of over 500 (Figure SI-14). Compared to the CO2/N2 selectivity of PIM-1 (18, Figure SI-12) and hydrolyzed PIM-1 (28, Figure SI-13), the very high CO2/N2 selectivity in PIM-1-C3-TA is a result of two factors: (i) amine functionality, and (ii) diminished surface area which provides much less adsorption media for inert gasses such as N2.26 We further evaluated the CO2 capture properties of the best performing sorbent, PIM-1-C3-TA, with 3% humidity using dynamic breakthrough curves at 308 K (Figure SI-15). The sorbent was tested under five adsorption/regeneration cycles, shown in Figure 3. The performance of the sample was first assessed under 10% CO2/He (dry conditions) to establish a baseline performance, and afterwards it was cycled twice under humid conditions (10% CO2, 3% H2O/He) until saturation of CO2 and H2O was achieved. The sorbent was then tested again under dry conditions to determine if exposure to humidity had caused any lasting changes to the material. Regeneration was carried out at 358 K under helium for all five cycles. The breakthrough results showed that CO2 uptake was 37.4 cc/g under dry conditions, and increased slightly to 39.4 cc/g in the presence of H2O.

Figure 3: Dry (green) and humid (blue) CO2 adsorption breakthrough results on PIM-1-C3-TA at 308 K.

ACS Paragon Plus Environment

4

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

After the humid cycles, testing the sorbent under dry conditions showed that CO2 uptake remained stable at 36.0 cc/g. For the limited number of cycles used, the sorbent appears to be stable in a humid environment. More cycles are needed to thoroughly assess the stability of the sorbent. CONCLUSION

In summary, the synthesis, characterization, and performance testing of new chemisorbent materials based on carboxylic acid functionalized PIM-1 microporous polymers has been presented. Primary amines were appended to the polymers through acid-base and hydrogen bonding interactions. The sorbents showed the highest CO2 uptake performance achieved to date in all reported functionalized and non-functionalized PIM based polymers (Table SI-2). Moreover, high CO2 uptake performance was maintained even in humid conditions and after multiple cycles. Finally, the reported sorbents are soluble in common solvents and are processable into a variety of geometries. ASSOCIATED CONTENT

Supporting Information The Supporting Information is available free of charge on the ACS Publications website. The document contains materials list, experimental, characterization description and figures. AUTHOR INFORMATION

Corresponding Author [email protected]

Phone: 412-386-4773 Funding Sources This project was funded by the Department of Energy, National Energy Technology Laboratory, an agency of the United States Government, under the Carbon Capture Field Work Proposal and in part through a support contract with Leidos Research Support Team (LRST, contract 89243318CFE000003). Neither the United States Government nor any agency thereof, nor any of their employees, nor LRST, nor any of their employees, makes any warranty, expressed or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. REFERENCES

(1) Pelley, J. Better Carbon Capture through Chemistry. ACS Cent. Sci. 2015, 1, 412-415. (2) Mason, J. A.; McDonald, T. M.; Bae, T.-H.; Bachman, J. E.; Sumida, K.; Dutton, J. J.; Kaye, S. S.; Long, J. R. Application of a High-Throughput Analyzer in Evaluating Solid Adsorbents for Post-Combustion Carbon Capture via Multicomponent Adsorption of CO2, N2, and H2O. J. Am. Chem. Soc. 2015, 137, 4787-4803. (3) Zou, L.; Sun, Y.; Che, S.; Yang, X.; Wang, X.; Bosch, M.; Wang, Q.; Li, H.; Smith, M.; Yuan, S.; Perry, Z.; Zhou, H.-C. Porous Organic Polymers for Post-Combustion Carbon Capture. Adv. Mater. 2017, 29, 1700229-1700264. (4) Dawson, R.; Stöckel, E.; Holst, J. R.; Adams, D. J.; Cooper, A. I. Microporous organic polymers for carbon dioxide capture. Energy Environ. Sci. 2011, 4, 4239-4245. (5) Dogan, N. A.; Ozdemir, E.; Yavuz, C. T. Direct Access to Primary Amines and Particle Morphology Control in Nanoporous CO2 Sorbents. ChemSusChem 2017, 10, 2130-2134. (6) Lu, W.; Sculley, J. P.; Yuan, D.; Krishna, R.; Wei, Z.; Zhou, H.-C. Polyamine-Tethered Porous Polymer Networks for Carbon Dioxide Capture from Flue Gas. Angew. Chem. Int. Ed. 2012, 51, 7480-7484. (7) Budd, P. M.; Elabas, E. S.; Ghanem, B. S.; Makhseed, S.; McKeown, N. B.; Msayib, K. J.; Tattershall, C. E.; Wang, D. Solution-Processed, Organophilic Membrane Derived from a Polymer of Intrinsic Microporosity. Adv. Mater. 2004, 16, 456-459. (8) Zou, X.; Zhu, G. Microporous Organic Materials for Membrane-Based Gas Separation. Adv. Mater. 2018, 30,1700750-1700784. (9) Low, Z.-X.; Budd, P. M.; McKeown, N. B.; Patterson, D. A. Gas Permeation Properties, Physical Aging, and Its Mitigation in High Free Volume Glassy Polymers. Chem. Rev. 2018, 118, 5871-5911. (10) Sekizkardes, A. K.; Kusuma, V. A.; Dahe, G.; Roth, E. A.; Hill, L. J.; Marti, A.; Macala, M.; Venna, S. R.; Hopkinson, D. Separation of Carbon Dioxide from Flue Gas by Mixed Matrix Membranes Using Dual Phase Microporous Polymeric Constituents. Chem. Commun. 2016, 52, 11768-11771. (11) Pang, S. H.; Jue, M. L.; Leisen, J.; Jones, C. W.; Lively, R. P. PIM-1 as a Solution-Processable “Molecular Basket” for CO2 Capture from Dilute Sources. ACS Macro Lett. 2015, 4, 1415-1419. (12) Hu, Z.; Wang, Y.; Wang, X.; Zhai, L.; Zhao, D. Solution-Reprocessable Microporous Polymeric Adsorbents for Carbon Dioxide Capture. AIChE J. 2018, 64, 3376-3389. (13) Wang, X.; Liu, Y.; Ma, X.; Das, S. K.; Ostwal, M.; Gadwal, I.; Yao, K.; Dong, X.; Han, Y.; Pinnau, I.; Huang, K.-W.; Lai, Z. Soluble Polymers with Intrinsic Porosity for Flue Gas Purification and Natural Gas Upgrading. Adv. Mater. 2017, 29, 1605826-1605832. (14) Patel, H. A.; Yavuz, C. T. Noninvasive functionalization of polymers of intrinsic microporosity for enhanced CO2 capture. Chem. Commun. 2012, 48, 9989-9991. (15) Song, J.; Du, N.; Dai, Y.; Robertson, G. P.; Guiver, M. D.; Thomas, S.; Pinnau, I. Linear High Molecular Weight Ladder Polymers by Optimized Polycondensation of Tetrahydroxytetramethylspirobisindane and 1,4Dicyanotetrafluorobenzene. Macromolecules 2008, 41, 7411-7417. (16) Du, N.; Robertson, G. P.; Song, J.; Pinnau, I.; Guiver, M. D. High-Performance Carboxylated Polymers of Intrinsic Microporosity (PIMs) with Tunable Gas Transport Properties. Macromolecules 2009, 42, 6038-6043. (17) Satilmis, B.; Budd, P. M. Base-catalysed hydrolysis of PIM-1: amide versus carboxylate formation. RSC Adv. 2014, 4, 52189-52198. (18) Yanaranop, P.; Santoso, B.; Etzion, R.; Jin, J. Facile Conversion of Nitrile to Amide on Polymers of Intrinsic Microporosity (PIM-1) Polymer 2016, 98, 244-251.

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 7

(19) Hicks, J. C.; Drese, J. H.; Fauth, D. J.; Gray, M. L.; Qi, G.; Jones, C. W. Designing Adsorbents for CO2 Capture from Flue Gas-Hyperbranched Aminosilicas Capable of Capturing CO2 Reversibly. J. Am. Chem. Soc. 2008, 130, 2902-2903. (20) Li, H.; Wang, K.; Feng, D.; Chen, Y.-P.; Verdegaal, W.; Zhou, H.-C. Incorporation of Alkylamine into Metal–Organic Frameworks through a Brønsted Acid–Base Reaction for CO2 Capture. ChemSusChem 2016, 9, 2832-2840. (21) Halder, K.; Neumann, S.; Bengtson, G.; Khan, M. M.; Filiz, V.; Abetz, V. Polymers of Intrinsic Microporosity Postmodified by Vinyl Groups for Membrane Applications. Macromolecules 2018, 51, 7309-7319. (22) Sekizkardes, A. K.; Culp, J. T.; Islamoglu, T.; Marti, A.; Hopkinson, D.; Myers, C.; El-Kaderi, H. M.; Nulwala, H. B. An ultra-microporous organic polymer for high performance carbon dioxide capture and separation. Chem. Commun. 2015, 51, 13393-13396. (23) Jeon, J. W.; Kim, D.-G.; Sohn, E.-h.; Yoo, Y.; Kim, Y. S.; Kim, B. G.; Lee, J.-C. Highly CarboxylateFunctionalized Polymers of Intrinsic Microporosity for CO2Selective Polymer Membranes. Macromolecules 2017, 50, 80198027. (24) Mason, C. R.; Maynard-Atem, L.; Heard, K. W. J.; Satilmis, B.; Budd, P. M.; Friess, K.; Lanc̆, M.; Bernardo, P.; Clarizia, G.; Jansen, J. C. Enhancement of CO2 Affinity in a Polymer of Intrinsic Microporosity by Amine Modification. Macromolecules 2014, 47, 1021-1029. (25) Sekizkardes, A. K.; İslamoğlu, T.; Kahveci, Z.; El-Kaderi, H. M. Application of pyrene-derived benzimidazolelinked polymers to CO2 separation under pressure and vacuum swing adsorption settings. J. Mater. Chem. A 2014, 2, 12492-12500. (26) Milner, P. J.; Siegelman, R. L.; Forse, A. C.; Gonzalez, M. I.; Runčevski, T.; Martell, J. D.; Reimer, J. A.; Long, J. R. A Diaminopropane-Appended Metal–Organic Framework Enabling Efficient CO2 Capture from Coal Flue Gas via a Mixed Adsorption Mechanism. J. Am. Chem. Soc. 2017, 139, 1354113553. For Table of Graphics Only

ACS Paragon Plus Environment

6

PIM-1 based chemisorbent for CO2 capture

Page 7 ofACS 7 Applied Materials & Interfaces 31 kJ/mol 35 kJ/mol 78 kJ/mol ΔHCO2

too low

1 2 3 4 5 6

too high

4x more CO2uptake

PIM-1

PIM-1-C

PIM-1-C-TA

Acidic functional groups

Primary amine groups

ACS Paragon Plus Environment Neat PIM-1