Abstract

•COFs are shown to be frequently unstable to in-plane mechanical shear•A method of COF monolith formation that enables hierarchical porosity is presented•Hierarchical mesopores enable improved gas uptake and selectivity•Hierarchical porosity can be accurately modeled via lattice-gas methods Porous materials are frequently used industrially as macroscopic pellets exhibiting interstitial space porosity. Although conventional methods of pelletization are suitable for mechanically robust materials, similar processing methods result in sharp losses in capacity in state-of-the-art adsorbent materials as a result of their inherently lower degrees of mechanical stability. Here, we show that COFs synthesized to date are frequently unstable to in-plane mechanical shear and further present a simple, rapid, and general process for the preparation of centimeter-scale, hierarchically porous, monolithic COFs. COF monoliths thus prepared exhibit a system of intercrystallite mesopores, which push final gas-uptake capacities to levels above those for powder or single-crystal analogs of the same material. These characteristics, without any chemical modification, enable both improved uptake capacity and mixed-gas selectivity for target gases within industrially relevant gaseous compositions. Covalent organic frameworks (COFs) have emerged as a versatile material platform for such applications as chemical separations, chemical reaction engineering, and energy storage. Their inherently low mechanical stability, however, frequently renders existing methods of pelletization ineffective, contributing to pore collapse, pore blockage, or insufficient densification of crystallites. Here, we present a process for the shaping and densifying of COFs into robust centimeter-scale porous monoliths without the need for templates, additives, or binders. This process minimizes mechanical damage from shear-induced plastic deformation and further provides a network of interparticle mesopores that we exploit in accessing analyte capacities above those achievable from the intrinsic COF structure. Using a lattice-gas model, we accurately capture the monolithic structure across the mesoporous range and tie pore architecture to performance in both gas-storage and -separation applications. Collectively, these results represent a substantial step in the practical applicability of COFs and other mechanically weak porous materials. Covalent organic frameworks (COFs) have emerged as a versatile material platform for such applications as chemical separations, chemical reaction engineering, and energy storage. Their inherently low mechanical stability, however, frequently renders existing methods of pelletization ineffective, contributing to pore collapse, pore blockage, or insufficient densification of crystallites. Here, we present a process for the shaping and densifying of COFs into robust centimeter-scale porous monoliths without the need for templates, additives, or binders. This process minimizes mechanical damage from shear-induced plastic deformation and further provides a network of interparticle mesopores that we exploit in accessing analyte capacities above those achievable from the intrinsic COF structure. Using a lattice-gas model, we accurately capture the monolithic structure across the mesoporous range and tie pore architecture to performance in both gas-storage and -separation applications. Collectively, these results represent a substantial step in the practical applicability of COFs and other mechanically weak porous materials. Porous materials capable of reducing both the cost and energy intensity of industrial chemical processes are critically needed in transitioning to a carbon-neutral energy cycle.1Sholl D.S. Lively R.P. Seven chemical separations to change the world.Nature. 2016; 532: 435-437Crossref PubMed Scopus (1789) Google Scholar,2Schoedel A. Ji Z. Yaghi O.M. The role of metal–organic frameworks in a carbon-neutral energy cycle.Nat. Energy. 2016; 1: 16034Crossref Google Scholar Constructed from Earth-abundant elements and affording a combination of chemo-structural diversity, ease of synthetic modification, and relative chemical stability, covalent organic frameworks (COFs) have emerged as attractive alternatives to existing porous materials, including activated carbons, zeolites, and metal-organic frameworks (MOFs).3Diercks C.S. Yaghi O.M. The atom, the molecule, and the covalent organic framework.Science. 2017; 355: eaal1585Crossref PubMed Scopus (1130) Google Scholar,4Slater A.G. Cooper A.I. Porous materials. Function-led design of new porous materials.Science. 2015; 348: aaa8075Crossref PubMed Scopus (861) Google Scholar,5Huang N. Wang P. Jiang D. Covalent organic frameworks: A materials platform for structural and functional designs.Nat. Rev. Mater. 2016; 1: 16068Crossref Scopus (962) Google Scholar However, a technological limit has been reached where traditional methods of adsorbent post-processing are poorly suited to COFs as a result of the low mechanical stability frequently exhibited by these materials.6Sun J. Iakunkov A. Baburin I.A. Joseph B. Palermo V. Talyzin A.V. Covalent organic framework (COF-1) under high pressure.Angew. Chem. Int. Ed. 2020; 59: 1087-1092Crossref PubMed Scopus (14) Google Scholar,7Du Y. Calabro D. Wooler B. Li Q. Cundy S. Kamakoti P. Colmyer D. Mao K. Ravikovitch P. Kinetic and mechanistic study of COF-1 phase change from a staggered to eclipsed model upon partial removal of mesitylene.J. Phys. Chem. C. 2014; 118: 399-407Crossref Scopus (24) Google Scholar,8Sick T. Rotter J.M. Reuter S. Kandambeth S. Bach N.N. Döblinger M. Merz J. Clark T. Marder T.B. Bein T. Medina D.D. Switching on and off interlayer correlations and porosity in 2D covalent organic frameworks.J. Am. Chem. Soc. 2019; 141: 12570-12581Crossref PubMed Scopus (83) Google Scholar,9Feriante C.H. Jhulki S. Evans A.M. Dasari R.R. Slicker K. Dichtel W.R. Marder S.R. Rapid synthesis of high surface area imine-linked 2D covalent organic frameworks by avoiding pore collapse during isolation.Adv. Mater. 2020; 32: e1905776Crossref PubMed Scopus (76) Google Scholar,10Zhu D. Verduzco R. Ultralow surface tension solvents enable facile COF activation with reduced pore collapse.ACS Appl. Mater. Interfaces. 2020; 12: 33121-33127Crossref PubMed Scopus (33) Google Scholar,11Zhou W. Wu H. Yildirim T. Structural stability and elastic properties of prototypical covalent organic frameworks.Chem. Phys. Lett. 2010; 499: 103-107Crossref Scopus (47) Google Scholar To date, these mechanical characteristics have been shown to limit the pressures that can be used during pelletization6Sun J. Iakunkov A. Baburin I.A. Joseph B. Palermo V. Talyzin A.V. Covalent organic framework (COF-1) under high pressure.Angew. Chem. Int. Ed. 2020; 59: 1087-1092Crossref PubMed Scopus (14) Google Scholar,12Uribe-Romo, F.J., Vazquez-Molina, D., and Harper, J.K. (2019). Mechanically shaped 2-dimensional covalent organic frameworks. US patent no. WO2018013682A1. published September 9, 2019.Google Scholar and the selection of fluids available for activation,8Sick T. Rotter J.M. Reuter S. Kandambeth S. Bach N.N. Döblinger M. Merz J. Clark T. Marder T.B. Bein T. Medina D.D. Switching on and off interlayer correlations and porosity in 2D covalent organic frameworks.J. Am. Chem. Soc. 2019; 141: 12570-12581Crossref PubMed Scopus (83) Google Scholar,9Feriante C.H. Jhulki S. Evans A.M. Dasari R.R. Slicker K. Dichtel W.R. Marder S.R. Rapid synthesis of high surface area imine-linked 2D covalent organic frameworks by avoiding pore collapse during isolation.Adv. Mater. 2020; 32: e1905776Crossref PubMed Scopus (76) Google Scholar,10Zhu D. Verduzco R. Ultralow surface tension solvents enable facile COF activation with reduced pore collapse.ACS Appl. Mater. Interfaces. 2020; 12: 33121-33127Crossref PubMed Scopus (33) Google Scholar deviations from which can result in sharp losses in capacity. Although factors such as framework topology and linker length can be synthetically tuned to target more robust architectures,13Moghadam P.Z. Rogge S.M.J. Li A. Chow C.-M. Wieme J. Moharrami N. Aragones-Anglada M. Conduit G. Gomez-Gualdron D.A. Van Speybroeck V. Fairen-Jimenez D. Structure-mechanical stability relations of metal-organic frameworks via machine learning.Matter. 2019; 1: 219-234Abstract Full Text Full Text PDF Scopus (106) Google Scholar the inverse approach—i.e., whereby a desired COF can be shaped into an industrially relevant form factor without compromising key performance metrics—has not been attempted. Here, we report a simple and rapid process for the shaping of COFs into macroscopic pellets without the use of binders, templates, or additives and without any further processing steps needed for a final application. For an archetypical two-dimensional (2D) COF, TPB-DMTP-COF,14Xu H. Gao J. Jiang D. Stable, crystalline, porous, covalent organic frameworks as a platform for chiral organocatalysts.Nat. Chem. 2015; 7: 905-912Crossref PubMed Scopus (890) Google Scholar we demonstrate control over the degree of aggregation of crystallites within pellets and systematically identify the presence of a lower limit in intercrystallite pore size for a given activation solvent. We tie this limit to the onset of capillary-action-induced, turbostratic disordering of crystallites and further confirm that mechanical damage can be avoided through the use of an activation fluid with an ultra-low surface tension. COF monoliths thus prepared are mechanically robust and exhibit low-pressure adsorption characteristics identical to those of the best-reported powder analogs. They additionally benefit from a system of interparticle mesopores that push final adsorption capacities above levels expected for single crystals. We capture these structural characteristics in a lattice-gas model, which accurately reproduces experimentally derived isotherms for COF monoliths in silico. The combination of intact crystallites, mechanical robustness, high bulk densities, and regular hierarchical mesopores is unique among COF monoliths demonstrated to date. These properties result in industrially suitable monoliths that afford better gas-adsorption performance characteristics for both pure-component gas storage (CO2 and CH4) and mixed-gas chemical separation (CO2/N2 and CO2/CH4) applications than do unprocessed powder controls. On the basis of these findings, our work provides not only a path forward for the industrial applicability of COFs but also a systematic framework through which COF microstructure and final adsorption properties can be tuned without altering the underlying COF chemistry. 2D COFs are thought to be unstable to in-plane mechanical shear.11Zhou W. Wu H. Yildirim T. Structural stability and elastic properties of prototypical covalent organic frameworks.Chem. Phys. Lett. 2010; 499: 103-107Crossref Scopus (47) Google Scholar To evaluate the mechanical properties of COFs across topologies and linkage chemistries, we first performed a high-throughput screen of all reported COFs as inventoried in the CURATED (Clean, Uniform, Refined with Automatic Tracking from Experimental Database) COF database15Ongari D. Yakutovich A.V. Talirz L. Smit B. Building a consistent and reproducible database for adsorption evaluation in covalent–organic frameworks.ACS Cent. Sci. 2019; 5: 1663-1675Crossref PubMed Scopus (52) Google Scholar and compared their bulk moduli, shear moduli, and elastic constants with those of MOFs13Moghadam P.Z. Rogge S.M.J. Li A. Chow C.-M. Wieme J. Moharrami N. Aragones-Anglada M. Conduit G. Gomez-Gualdron D.A. Van Speybroeck V. Fairen-Jimenez D. Structure-mechanical stability relations of metal-organic frameworks via machine learning.Matter. 2019; 1: 219-234Abstract Full Text Full Text PDF Scopus (106) Google Scholar (Figure 1). Within a largest cavity diameter (LCD) range of 15–40 Å, the bulk and shear moduli of COFs were found to be similar to those of MOFs, although COFs exhibited marginally higher bulk moduli and shear moduli on average. However, at lower LCD ranges characteristic of ultramicroporous (<7 Å) and microporous (<20 Å) materials, the bulk and shear moduli of COFs were found to be substantially lower than those of MOFs, suggesting an inherently greater tendency of COFs to mechanically deform even in the absence of larger (>15 Å) pores. To gain insights into the mechanical stability of these materials, we then analyzed the elastic constants (cij) of a representative COF subset, hexagonal 2D COFs, which currently account for 54% of 2D COFs and 45% of all COFs synthesized to date. Applying the stability criteria, c11 > |c12|, c33(c11 + 2c12) > 2(c13)2, c11c33 > (c13)2 and c44 > 0, we found a majority (64%) of hexagonal 2D COFs to be unstable, confirming weakness to mechanical shear as a predominating feature of these materials and possibly shedding light on the low degrees of crystallinity frequently exhibited by these materials. Given that conventional methods of powder pelletization routinely employ pressures in the range of 1–3 GPa, which are known to trigger losses in capacity within MOFs,16Howarth A.J. Liu Y. Li P. Li Z. Wang T.C. Hupp J.T. Farha O.K. Chemical, thermal and mechanical stabilities of metal–organic frameworks.Nat. Rev. Mater. 2016; 1: 15018Crossref Scopus (1083) Google Scholar,17Tian T. Zeng Z.X. Vulpe D. Casco M.E. Divitini G. Midgley P.A. Silvestre-Albero J. Tan J.C. Moghadam P.Z. Fairen-Jimenez D. A sol-gel monolithic metal-organic framework with enhanced methane uptake.Nat. Mater. 2018; 17: 174-179Crossref PubMed Google Scholar,18Connolly B.M. Aragones-Anglada M. Gandara-Loe J. Danaf N.A. Lamb D.C. Mehta J.P. Vulpe D. Wuttke S. Silvestre-Albero J. Moghadam P.Z. et al.Tuning porosity in macroscopic monolithic metal-organic frameworks for exceptional natural gas storage.Nat. Commun. 2019; 10: 2345Crossref PubMed Scopus (105) Google Scholar,19Connolly B.M. Madden D.G. Wheatley A.E.H. Fairen-Jimenez D. Shaping the future of fuel: Monolithic metal–organic frameworks for high-density gas storage.J. Am. Chem. Soc. 2020; 142: 8541-8549Crossref PubMed Scopus (108) Google Scholar,20Chapman K.W. Halder G.J. Chupas P.J. Pressure-induced amorphization and porosity modification in a metal−organic framework.J. Am. Chem. Soc. 2009; 131: 17546-17547Crossref PubMed Scopus (312) Google Scholar,21Kandambeth S. Dey K. Banerjee R. Covalent organic frameworks: Chemistry beyond the structure.J. Am. Chem. Soc. 2019; 141: 1807-1822Crossref PubMed Scopus (560) Google Scholar we sought a revised approach for COF processing and pelletization. To permit ease of experimental benchmarking and analysis, we identified TPB-DMTP-COF as a representative 2D COF with an LCD of 25 Å and excellent known crystallinity. Upon screening a variety of synthesis solvent systems, we identified acetonitrile and a 1:1 (v/v) mixture of 1,3,5-trimethylbenzene (mesitylene) and 1,4-dioxane (dioxane) as two systems capable of both solubilizing the starting materials and producing crystalline samples of TPB-DMTP-COF. However, whereas the 1:1 (v/v) mixture of mesitylene and dioxane produced powder samples consisting of aggregated particles > 500 nm in diameter (Figures 2E and S5A), the acetonitrile system produced dense pellets consistent with those previously described for MOF monoliths and composed of particles of approximately 40 nm in diameter (Figures 2E and S5H)—well within the limits previously established for monolith formation in MOFs17Tian T. Zeng Z.X. Vulpe D. Casco M.E. Divitini G. Midgley P.A. Silvestre-Albero J. Tan J.C. Moghadam P.Z. Fairen-Jimenez D. A sol-gel monolithic metal-organic framework with enhanced methane uptake.Nat. Mater. 2018; 17: 174-179Crossref PubMed Google Scholar,18Connolly B.M. Aragones-Anglada M. Gandara-Loe J. Danaf N.A. Lamb D.C. Mehta J.P. Vulpe D. Wuttke S. Silvestre-Albero J. Moghadam P.Z. et al.Tuning porosity in macroscopic monolithic metal-organic frameworks for exceptional natural gas storage.Nat. Commun. 2019; 10: 2345Crossref PubMed Scopus (105) Google Scholar (i.e., <120 nm). Taking these two systems as extremes, we used solvent compositions consisting of different fractions of each to prepare pellets as follows: (1) reaction for a fixed amount of time (typically 30 min), (2) centrifugation, (3) purification and solvent exchange to methanol, and (4) controlled drying and activation (Figure 2). Upon processing, scanning electron microscopy (SEM) of the finished pellets revealed a gradual progression in microstructure from larger, loosely aggregated particles to densely packed monoliths exhibiting conchoidal fracture and little to no interparticle free volume (Figures 2E and S5). Analysis of the nitrogen adsorption isotherms (Figures 3A and S3) collected for these pellets, however, revealed a striking trend. Whereas we observed a monotonic increase in Brunauer-Emmett-Teller (BET) area—calculated by BET surface identification (BETSI)22Osterrieth J.W.M. Rampersad J. Madden D. Rampal N. Skoric L. Connolly B. Allendorf M.D. Stavila V. Snider J.L. Ameloot R. et al.How reproducible are surface areas calculated from the BET equation?.Adv. Mater. 2022; 34: e2201502Crossref PubMed Scopus (17) Google Scholar—for pellets synthesized in solvent systems containing acetonitrile fractions ranging from 0.000 to 0.750 (v/v), we observed a sharp decrease in BET area to 4 m2 g−1 for samples prepared at higher acetonitrile fractions (Figure 3C). As a result, the highest BET area that could be obtained for TPB-DMTP-COF with methanol as the activation solvent was 1,122 m2 g−1, suggesting the presence of a lower limit in intercrystallite pore size beyond which pore disruption takes place. To test whether this pore disruption was being induced by capillary action,10Zhu D. Verduzco R. Ultralow surface tension solvents enable facile COF activation with reduced pore collapse.ACS Appl. Mater. Interfaces. 2020; 12: 33121-33127Crossref PubMed Scopus (33) Google Scholar we prepared a further sample in a pure acetonitrile solvent system and processed it as before but this time dried and activated it in supercritical carbon dioxide (scCO2) instead of in methanol and air. The finished pellet not only recovered full porosity but also lay on the monotonic trend previously described in that it exhibited a BET area of 2,125 m2 g−1—slightly above those previously described for powder analogs of TPB-DMTP-COF.14Xu H. Gao J. Jiang D. Stable, crystalline, porous, covalent organic frameworks as a platform for chiral organocatalysts.Nat. Chem. 2015; 7: 905-912Crossref PubMed Scopus (890) Google Scholar When a higher rate of scCO2 pressure release (8 versus 3 bar h−1) was used during the activation of an identically prepared 1.000 acetonitrile pellet, a reduction in BET area to 1,439 m2 g−1 was observed, further suggesting that losses in BET area occur as a result of damage induced by capillary action. Collectively, these results suggest that intercrystallite pore size is modulated during synthesis by control over particle size—where interparticle pores are the void space created by the fundamentally imperfect packing of approximate hard spheres.23Song C. Wang P. Makse H.A. A phase diagram for jammed matter.Nature. 2008; 453: 629-632Crossref PubMed Scopus (698) Google Scholar Where characteristic interparticle pore size falls below a certain threshold, capillary action during drying causes damage to crystallites. To gain deeper insights into the structural changes accompanying these bulk characteristics, we used a combination of pair distribution function (PDF) and X-ray diffraction (XRD) (supplemental information section S6). Non-negative matrix factorization of the PDF-XRD data revealed three independent underlying components that we attribute to non-crystalline-layer COF content, residual starting-material content, and multilayer (i.e., crystalline) COF content: components A, B, and C, respectively (Figure 3G). For methanol-activated pellets below an acetonitrile fraction of 0.75, a respective decrease and increase in components A and C were observed as the acetonitrile fraction was increased, indicating that TPB-DMTP-COF crystallinity gradually improves before the onset of mechanical damage. Above an acetonitrile fraction of 0.750 (v/v), crystallinity sharply declines, resulting in an increased content of non-crystalline-layer TPB-DMTP-COF, as seen from the increasing weighting of component A. When scCO2 is used during drying and activation, the multilayer content is recovered—an observation consistent with findings from nitrogen adsorption studies and providing clear evidence for a correlation between mechanical disruption of COF crystallites during post-processing and observable gas-uptake capacities as previously noted for powdered COF systems.8Sick T. Rotter J.M. Reuter S. Kandambeth S. Bach N.N. Döblinger M. Merz J. Clark T. Marder T.B. Bein T. Medina D.D. Switching on and off interlayer correlations and porosity in 2D covalent organic frameworks.J. Am. Chem. Soc. 2019; 141: 12570-12581Crossref PubMed Scopus (83) Google Scholar To better understand the mechanism of crystallite disordering into non-crystalline layers within COF monoliths, we performed a combination of high-resolution transmission electron microscopy (HR-TEM) and NanoBlitz indentation studies on a methanol-activated 1.000 acetonitrile control sample for which crystallites are sufficiently disrupted to afford a BET area of 4 m2 g−1 (Figure 4). Analysis of the microstructure both within the bulk and within a few layers of the monolith indicated a series of small multilayer crystalline domains bridged by less-ordered regions. Fourier transforms of the image (Figure 4C, inset) further revealed that these features result in a single diffuse band corresponding to a real-space length of 0.36 nm—consistent with interlayer spacing values obtained from analysis of components A (0.37 nm) and C (0.35 nm) derived from the PDF-XRD data. These results suggest that a crystalline-to-turbostratic-disordering mechanism, similar to that observed in mechanically milled graphite, might be responsible for losses in observed porosity within monoliths. At the macroscopic level, NanoBlitz indentation mapping revealed heterogeneities in both the indentation modulus and the mechanical hardness across a 200 × 200 μm region of the material. Given that turbostratic disordering is triggered by capillary action, differences in the local structure of the monolith during drying could give rise to regions of greater or lesser disruption, resulting in macroscopic domains with slightly differing mechanical properties in the finished pellet. Collectively, these findings both confirm the presence of disrupted crystallites in non-porous monoliths and suggest that a turbostratic disordering mechanism is responsible for such observable losses in porosity. After establishing post-processing conditions capable of explicitly avoiding crystallite damage, we then used a combination of mercury porosimetry and small-angle X-ray scattering (SAXS) to gain insights into the structure of monolith free-volume elements across the mesoscale. Pore-size distributions derived from mercury intrusion curves for a scCO2-activated 1.000 acetonitrile monolith revealed the presence of sharp mesoporosity at 18.7 nm attributable to narrow and regular interparticle free-volume elements (Figure 3A, inset). Broader macroporosity centered at a pore width of around 3 μm was also observed. By contrast, a non-monolithic powder control prepared via the method of Xu et al. (BET area of 1,985 m2 g−1)14Xu H. Gao J. Jiang D. Stable, crystalline, porous, covalent organic frameworks as a platform for chiral organocatalysts.Nat. Chem. 2015; 7: 905-912Crossref PubMed Scopus (890) Google Scholar exhibited no regular meso- or macroporosity. Analysis of the respective mercury areas for accessible pore widths down to 3.9 nm (above that of the intrinsic framework, i.e., 2.5 nm) further showed an area of 504 m2 g−1 for the monolith and 196 m2 g−1 for the powder. These results are comparable to those for classical mesoporous templated silicas and carbons30Lee J. Sohn K. Hyeon T. Fabrication of novel mesocellular carbon foams with uniform ultralarge mesopores.J. Am. Chem. Soc. 2001; 123: 5146-5147Crossref PubMed Scopus (278) Google Scholar,31Zhang F.-A. Lee D.-K. Pinnavaia T.J. PMMA/mesoporous silica nanocomposites: Effect of framework structure and pore size on thermomechanical properties.Polym. Chem. 2010; 1: 107-113Crossref Google Scholar and are consistent with those derived from SAXS (Figure 3D and supplemental information section S7). The scCO2-processed monolith was well fit by a spheroidal particle model with two log-normalized-distribution models with mean diameters of 25.8 nm (σ = 0.4) and 99.8 nm (σ = 0.2), indicating the presence of mesoporous interparticle free-volume elements and providing evidence for additional macroporosity. By contrast, the non-monolithic powder control was found to possess an interparticle size distribution beyond the 0.5–100 nm range and, consequently, could not be fitted. These results suggest that COF processing into monoliths can not only be used to avoid pore collapse but also provide additional mesoporosity (inaccessible from powders) that can be used for tuning final uptake performance characteristics—potentially beyond those of purely crystalline systems. To examine the impact of crystallite disordering on mesoporous free-volume elements, we also analyzed a scCO2-processed monolith activated at an accelerated depressurization rate of 8 bar h−1 (BET area of 1,439 m2 g−1) by using SAXS (Figure S4). The sample was fit by three spheroidal size-distribution models exhibiting mean diameters of 14.7 nm (σ = 0.3), 21.1 nm (σ = 0.6), and 98.5 nm (σ = 0.1). The emergence of a third, narrow free-volume element along with an overall shift in mesopore distribution to smaller values suggests that disruption of crystallites is concomitant with a reduction in interparticle free volume. Because this reduction in interparticle pore size can be controlled by the scCO2 pressure release rate, future opportunities exist for top-down control over monolith microstructure and gas-adsorption properties. To assess the extent to which COF monoliths can be used as industrial pellets, we carried out nanoindentation studies, from which we derived mechanical indentation moduli and hardness values. For a 3 bar h−1 scCO2-activated 1.000 (v/v) pellet, we obtained an indentation modulus of 3.71 ± 0.20 GPa and a hardness of 0.18 ± 0.02 GPa (Figures 3B and 3E). These values are significantly higher than those previously obtained for COF aerogel pellets26Martín-Illán J.Á. Rodríguez-San-Miguel D. Castillo O. Beobide G. Perez-Carvajal J. Imaz I. Maspoch D. Zamora F. Macroscopic ultralight aerogel monoliths of imine-based covalent organic frameworks.Angew. Chem. Int. Ed. 2021; 60: 13969-13977Crossref PubMed Scopus (29) Google Scholar and slightly above those known for high-molecular-weight polyethylene.32Iqbal T. Camargo S.S. Yasin S. Farooq U. Shakeel A. Nano-indentation response of ultrahigh molecular weight polyethylene (UHMWPE): a detailed analysis.Polymers. 2020; 12: 795Crossref Google Scholar A full comparison of the mechanical properties of the COF monolith and those of other COF bodies reported in the literature is included in Table S6. These results suggest mechanical robustness and potential industrial suitability, possibly as a result of weak, non-crystalline-layer interfaces between COF crystallites, which serve to dissipate stress.33Cook J. Gordon J. A mechanism for the control of crack propagation in all-brittle systems.Proc. R. Soc. Lond. A. 1964; 282: 508-520Crossref Google Scholar,34Clegg W.J. Kendall K. Alford N.M. Button T.W. Birchall J.D. A simple way to make tough ceramics.Nature. 1990; 347: 455-457Crossref Scopus (739) Google Scholar By contrast, powder controls crumbled readily and could not be mounted in the instrument to yield reliable results. Further densification of monoliths with the use of higher-surface-tension activation solvents yielded slight increases in both indentation modulus and hardness. For a methanol-activated 1.000 (v/v) monolith, where pore collapse is complete, the indentation modulus and hardness values were 4.21 ± 0.37 and 0.34 ± 0.03 GPa, respectively (Figures 4D–4H). However, because these values represent modest improvements over the fully porous analog, potential design trade-offs between mechanical properties and porosity are likely to favor porous monoliths. Collectively, the combination of mechanical robustness, high bulk densities, high surface areas, and regular hierarchical mesopores is unique among COF-shaped bodies demonstrated to date. This is illustrated in Figure 3F and Table S6, in which the 3 bar h−1 scCO2-activated 1.000 (v/v) monolith is compared with other COF bodies (including those featuring additives and binders) on the basis of density, BET area, and mechanical figures of merit. To accurately capture the adsorption characteristics of TPB-DMTP-COF in silico, we carried out grand canonical Monte Carlo (GCMC) simulations on TPB-DMTP-COF crystalline fragments exhibiting varying degrees of interlayer slip (supplemental information section S2). Starting from perfect AA stacking (0% slip), one of two sequential layers of the COF was gradually shifted until perfect AB stacking was achieved (100% slip). Using cells derived from 0%, 25%, 40%, 50%, 75%, and 100% slipped starting structures (Figure S1), we then used GCMC simulations to generate predicted nitrogen isotherms at 77 K. Upon comparison of the respective low-pressure regions and mesoporous steps of the experimental adsorption isotherms to those derived from theory, a 40% slipped structure was found to provide the best agreement with experiment, giving almost identical low-pressure adsorption characteristics up to the mesoporous step (Figure 5D). Above the mesoporous step, however, whereas experimental isotherms for TPB-DMTP-COF powders maintained a reasonable agreement with those calculated from the 40% slipped structure until saturation, substantial deviations from theory were observed for experimental isotherms derived from TPB-DMTP-COF monoliths as a result of interparticle mesoporosity. Given that these deviations ultimately push total nitrogen uptake within the monolith above levels expected for purely crystalline systems, the ability to accurately capture such deviations computationally is critical in evaluating and subsequently tuning final gas-uptake characteristics for a desired target application. To model contributions to total gas uptake arising from interparticle mesopores, we moved to a lattice-gas model of the TPB-DMTP-COF monolith (supplemental information section S4). Lattice-gas models have been extensively used in the past for studying the nature of sorption hysteresis for fluids in confined interconnected void spaces of porous glasses.35Evans R. Fluids adsorbed in narrow pores: Phase equilibria and structure.J. Phys. Condens. Matter. 1990; 2: 8989-9007Crossref Scopus (655) Google Scholar,36Gelb L.D. Gubbins K.E. Radhakrishnan R. Sliwinska-Bartkowiak M. Phase separation in confined systems.Rep. Prog. Phys. 1999; 62: 1573-1659Crossref Scopus (1474) Google Scholar,37Kierlik E. Rosinberg M.L. Tarjus G. Pitard E. Mean-spherical approximation for a lattice model of a fluid in a disordered matrix.Mol. Phys. 1998; 95: 341-351Crossref Scopus (44) Google Scholar We numerically reconstructed the structural model of the monolith (Figure 5B) used in the lattice-gas model from the SAXS data for the TPB-DMTP-COF monolith by means of generating a two-point correlation function S2(r) and using it in the reconstruction algorithm. A 3D reconstructed structure and its 2D slice used in the lattice-gas model for the TPB-DMTP-COF monolith activated by scCO2 are shown in Figures 5A and 5C, respectively. To model the trajectory of the system in the grand canonical ensemble, we subsequently employed kinetic Monte Carlo (kMC) simulations from which nitrogen adsorption isotherms at 77 K could be obtained. The numerically generated isotherms show an excellent agreement with experimental data for the TPB-DMTP-COF monolith within the high-pressure region of the adsorption isotherms, providing complementary data to the GCMC-calculated isotherms and demonstrating the applicability of lattice-gas models in capturing the interparticle mesoporosity of COF monoliths. Collectively, these results suggest that the hierarchical porosity of COF monoliths can be accurately described computationally across the micro- and mesoporous ranges, enabling robust future predictions of adsorption characteristics. To demonstrate the utility of monolithic processing of COFs in gas-storage applications, we performed pure-component adsorption studies on TPB-DMTP-COF powders and monoliths. Low-pressure isotherms collected at 298 K revealed good CO2 (Figures 6A and 6D ) uptake for both powders and monoliths with modest to low CH4 (Figure 6A) and N2 (Figure 6D) uptake, respectively, for both systems. However, up to pressures of 1 bar, although higher CO2 uptake was obtained for monoliths than for powders, lower uptake for both CH4 and N2 was obtained for monoliths than for powders. These results suggest that the presence of interparticle mesopores in monoliths can be used not only for improving final storage capacities for a single component but also for favorably or disfavorably influencing final uptake characteristics of various components within a mixed feed. To examine these characteristics within the context of chemical separations, we evaluated adsorption selectivities for industrially relevant compositions of CO2, CH4, and N2 mixtures. From pure-component adsorption isotherms and using the ideal absorbed solution theory (IAST), we calculated selectivities for 15% CO2/85% N2 (Figure 6E) and 50% CO2/50% CH4 (Figure 6B) (v/v) mixtures (supplemental information section S8). At low pressures, the selectivity for CO2 relative to other components was substantially improved, providing evidence that monolithic COF structuring can be used to provide separation enhancements relative to unstructured COF powders. To confirm this, we performed dynamic breakthrough studies on TPB-DMTP-COF monoliths and powders by using mixed-gas feeds. For the 15% CO2/85% N2 mixture (Figure 6F), although comparable separations were achieved for the monolith and powder (and some additional evidence for axial dispersion was observed), the total CO2 uptake was found to be 13.4% higher for the monolith. For the 50% CO2/50% CH4 mixture (Figure 6C), a markedly sharper separation for the monolith than for the powder was observed, and an additional improvement in CO2 capacity of 8.6% was achieved. Tables S6 and S7 compare these results against those for similar separations reported in the literature. Collectively, these results not only demonstrate the utility of monolithic processing for adsorbent-based chemical storage and separation but also afford additional degrees of freedom through which the properties of COFs can be systematically designed and tuned. Using a simple and general processing workflow, we introduce methods for the preparation of hierarchically porous COF monoliths without the need for additional materials or processing components. We show that such processing methods are compatible with mechanically weak materials and further afford degrees of design freedom in the control of both extrinsic and intrinsic porosities. These characteristics endow monolithic COFs with properties that are distinct from both powder and single-crystal analogs, which we accurately capture in silico by using a lattice-gas model. We envision that such computational approaches can be used in the future to predict gas-uptake properties for broad classes of monolithic mesoporous materials. The extrinsic porosity present in COF monoliths can further be leveraged for simultaneously increasing and decreasing the final uptake capacities for various gas constituents relative to powder benchmarks, which we make use of in demonstrating improved separation performance for industrially relevant gas compositions. We believe that this study not only opens up new possibilities for the practical applicability of COFs but also provides a pathway forward for tuning sorbent-analyte interactions where changes to the underlying framework chemistry might not be possible or synthetically accessible.

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call