Functionalization-Induced Breathing Control in Metal–Organic

Mar 4, 2019 - Alvarez, R. A.; Zavala-Araiza, D.; Lyon, D. R.; Allen, D. T.; Barkley, A. R.; Brandt, A. R.; Davis, K. J.; Herndon, S. C.; Jacob, D. J.;...
1 downloads 0 Views 603KB Size
Subscriber access provided by LUNDS UNIV

Article

Functionalization Induced Breathing Control in Metal-Organic Frameworks for Methane Storage with High Deliverable Capacity Tanay Kundu, Bhuvan B. Shah, Linius Bolinois, and Dan Zhao Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b05332 • Publication Date (Web): 04 Mar 2019 Downloaded from http://pubs.acs.org on March 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 6 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

Chemistry of Materials

Functionalization Induced Breathing Control in Metal-Organic Frameworks for Methane Storage with High Deliverable Capacity Tanay Kundu, Bhuvan B. Shah, Linius Bolinois, and Dan Zhao* Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, 117585 Singapore ABSTRACT: Metal-organic frameworks (MOFs) that can undergo structural flexibility upon gas sorption cycle bear enormous potential particularly as adsorbents for methane storage. A molecular level control over such flexibility can be advantageous for the development of smart adsorbents with increased deliverable capacity. Herein, we report the flexible and stable MIL-53(Al) type MOFs, whose breathing behavior can be controlled by systematic installation of hydrogen bonding sites into the frameworks. Such a control has been fine-tuned to induce high deliverable capacity in MOFs by changing gas pressure, temperature and density. The establishment of structure-property relationship for methane storage and successful pelletization of MOFs pave the way for MOF-based on-board natural gas storage.

Introduction Natural gas (NG), mostly consisting of methane, has garnered wide interests as a comparatively greener alternative for fossil-fuels.1-2 The current compressed natural gas (CNG, stored at pressures larger than 200 bar) technology suffers from low volumetric energy density and safety concerns.3 A competent adsorbed natural gas (ANG) technology relies on adsorbents to store NG at around 65 bars.4 For benchmarking purpose, the US Department of Energy (DOE) has set the volumetric NG storage target of 263 cm3(STP)cm-3 (or v/v) total uptake at 65 bar and 298 K.5 Since NG delivery is a cyclic process, considering deliverable capacity (amount of NG adsorbed at 65 bar during adsorption minus the retaining value at 5 bar during desorption) is more relevant for identifying suitable adsorbents. The Advanced Research Projects Agency-Energy (ARPA-E) has set a rather ambitious NG deliverable capacity target of 315 v/v or 500 mg g-1.6 Current stateof-the-art hollow-carbon monoliths7 could reach only 40% of this target as they trap significant amounts of NG at 5 bar due to their Type I sorption isotherms.8 Undeniably, there is an urgent need of an effective NG storage system that can achieve both high storage and deliverable capacities. Metal-organic frameworks (MOFs), crystalline porous materials formed by the coordination linkage between metal nodes and organic ligands, have emerged as the next generation adsorbent materials.9 MOFs have advantages over conventional adsorbents, such as record-high surface area,10 easily-tunable pore size and shape, and desired functionalities that suit NG storage.11-13 Till date, various MOFs have been evaluated experimentally and computationally for their methane storage capacities.14-18 One of the best performing MOFs is HKUST-1, having an experimental methane storage capacity of 270 v/v at 65 bar and 298 K (considering crystallographic density),5 while the deliverable capacity is only 190 v/v due to the sharp rise of isotherm at low pressures. The attempt to increase density and thus volumetric and deliverable capacities by pelletization failed due to partial structure collapse, affording only 180 and 135 v/v storage and deliverable

capacities, respectively.5 Recently, monolith formation of HKUST1 retained the storage capacity due to increased density of the monolith, while its general applicability remains to be further explored.19 In addition, HKUST-1 exhibits loss in gas uptake capacity over multiple sorption cycles.20 To sum up, current NG storage systems suffer from (1) low deliverable capacity, and (2) chemical and mechanical instability.

Figure 1. (a) Different phases of flexible MOFs; (b) Scheme for the synthesis of flexible MIL-53(Al) structures with different hydroxyl linkers.

Recently, Long’s group has adopted flexible MOFs in NG storage (viz. Co-BDP and Fe-BDP), whose porosity changes as a function of methane pressure.21 The very narrow pore (vnp) phase is non-porous below 15 bar, while the large pore (lp) phase shows

ACS Paragon Plus Environment

Chemistry of Materials 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

high porosity. This approach almost equals the storage and deliverable capacity (190 v/v) due to the negligible NG uptake at 5 bar.21 In addition, functional groups can further control the gate opening pressure.22-25 Unfortunately, these MOFs suffer from longterm chemical and mechanical instability, limiting their practical applications. We envisage that designing flexible and stable MOFs with high porosity and mechanical resistance can be a feasible approach toward practical NG storage solutions. Previously, we have explored a well-known flexible and stable MOF viz. MIL-53(Al)-NH2 for methane storage.26-27 The flexibility originates from the weak hydrogen bonding between the linker amino groups and the bridging hydroxyl groups of the AlO4(OH)2 chain. Although the methane uptake at 5 bar is negligible, the NG deliverable capacity (5-120 bar, 298 K) is relatively low (133 v/v) as the np-lp phase transition is incomplete. We hypothesize that modulating the hydrogen bonding site would be a viable approach to tune the breathing behavior of the resultant MOFs. Herein, we report the molecular level control of the hydrogen bonding site and thus the flexibility in MIL-53-type MOFs by introducing different hydroxyl groups in the linker. The resultant MOFs, viz. MIL-53(Al)-OH and MIL-53(Al)-(OH)2 are constructed from 2-hydroxyterephthalate and 2,5dihydroxyterephthalate linkers, respectively (Figure 1). The MIL53(Al)-OH exhibits excellent stability, porosity and high NG storage and deliverable capacities of 164 v/v at 65 bar and 298 K, competing with most proficient MOFs reported nowadays. Introducing another hydroxyl group shows larger gate-opening pressures in MIL-53(Al)-(OH)2, while blocking the site using methoxy group for MIL-53-OCH3 results in non-flexibility of the MOF, establishing the role of hydrogen bonding sites in controlling the breathing behavior. Finally, the pristine low-density MOF powders were compressed into high-density pellet shapes with improved mechanical resilience and density without sacrificing gravimetric capacity, a significant progress towards MOF-based onboard NG storage. Result and Discussion

Figure 2. DFTB optimized crystal structures of the different breathing phases of the MOFs with corresponding channel dimensions.

MIL-53(Al)-OH and MIL-53(Al)-(OH)2 were synthesized using literature procedure with slight modifications.28-29 The synthetic procedures have been scaled up to 5 g in a single batch, paving the way for pilot-scale production (Figure S11). The density functional tight binding scheme (DFTB) optimized crystal structures of the MOFs show efficient π-π stacking of the hydroxyterephthalate linkers that stabilize the very narrow pore (vnp) phase, resulting in a pore dimension of 6 × 19 Å for both MIL-53(Al)-OH and MIL53(Al)-(OH)2 (Figure 2). Interestingly, the wine-rack type lp phase exhibits a pore dimension of 12 × 17 Å for both cases. The vnp phases show a weak hydrogen bonding interaction (ca. 3.79 Å) between the linker hydroxyl groups and the AlO4(OH)2 chains, which weakens (ca. 3.81 Å) upon transformation to lp phases for both MOFs (Figure 2). The as-synthesized MOFs remain in a mixed phase of np form and lp form, as evidenced from powder Xray diffraction (PXRD) pattern comparison with the simulated structures (Figure 3a).

Figure 3. (a) Experimental PXRD patterns of MIL-53(Al)-OH (blue) and MIL-53(Al)-(OH)2 (green) compared with the simulated patterns of MIL53(Al)-lp (red) and MIL-53(Al)-np (black); (b) N2 sorption isotherms at 77 K (solid and empty circles denote adsorption and desorption points, respectively) and pore size distributions (inset).

ACS Paragon Plus Environment

Page 2 of 6

Page 3 of 6 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

Chemistry of Materials The Fourier-transform infrared spectroscopy (FT-IR) spectra of MIL-53(Al)-OH and MIL-53(Al)-(OH)2 clearly indicate the formation of MOFs without free linker in the pore (Figure S3). The Brunauer–Emmett–Teller (BET) surface areas of the MOFs were calculated to be 1284 m2 g-1 and 1193 m2 g-1 for MIL-53(Al)OH and MIL-53(Al)-(OH)2, respectively, based on nitrogen sorption isotherms at 77 K (Figure 3b). The pore size distributions exhibit micropores of ca. 1.18 nm for MIL-53(Al)-OH and ca. 1.21 nm for MIL-53(Al)-(OH)2 (Figure 3b inset). Interestingly, the MIL-53(Al)-(OH)2 exhibits enhanced N2 uptake and pore volume than literature values, probably arising from different activation conditions (Section S3 in SI).29 The thermogravimetric analyses (TGA) of MIL-53(Al)-OH and MIL-53(Al)-(OH)2 show their thermal stability up to 523 K (Figure S4).

Figure 4. Comparison of gravimetric CH4 sorption isotherms at different temperatures for (a) MIL-53(Al)-OH and (b) MIL-53(Al)-(OH)2.

The methane storage and deliverable capacities of the above MOFs were evaluated by methane sorption isotherms under high pressures. In the case of MIL-53(Al)-OH, the absolute volumetric methane isotherms show large hysteresis, indicating gate openingclosure triggered by methane pressure. At a lower temperature of 273 K, three distinct phases are well-resolved in isotherms with characteristic plateau. The vnp phase is stable up to 8 bar of methane pressure, exhibiting a negligible gas uptake. The np phase becomes stable in 14-17 bar of methane, while increasing pressure finally leads to the fully expanded lp phase. Upon desorption, the lp to np transition is not observed, while the vnp phase is reached below 3.3 bar. On the contrary, the methane isotherms of MIL53(Al)-(OH)2 do not show such a distinguished plateau for np phase. The vnp phase gradually transforms to lp phase above 30 bar during adsorption, and the vnp phase was observed below 2.7 bar during desorption. At ambient temperature (298 K), MIL-53(Al)OH shows a vnp-to-lp transition at above 15 bar, while reverse transition happens below 4.1 bar. In the case of MIL-53(Al)(OH)2, the higher stability of the vnp phase results in gate opening at above 46 bar, and gate closure below 5.8 bar. Such a huge difference in gate opening pressure between MIL-53(Al)-OH and MIL-53(Al)-(OH)2 can be attributed to the efficient π-π stacking and reinforced dual site for hydrogen bonding between the dihydroxyterephthalate linker in MIL-53(Al)-(OH)2 and the acidic bridging hydroxyl groups, while only one site is available for hydroxyterephthalate linker in MIL-53(Al)-OH. Further increase in temperature (313 K) minutely increases the gate opening pressure, while the absolute methane uptake is reduced compared to 298 K. However, both MIL-53(Al)-OH (156 cm3 g-1 for 5-65 bar) and MIL-53(Al)-(OH)2 (138 cm3 g-1 for 5-120 bar) exhibit the highest gravimetric deliverable capacities at 298 K (Figure S22 & S23). Either increase or decrease in temperature (e.g., 313 K and 273 K) decreases the deliverable capacity value. This observation is in contrary with conventional NG adsorbents, where increasing temperature generally decreases the deliverable capacity. The

prime reason behind this peculiarity is the position of the gate closing pressure, which lies near 5 bar at room temperature. Thus, the increase in gravimetric uptake upon decreasing temperature is offset by a shift in desorption branch (Figure 4a-b).

Figure 5. (a) Scheme for removing the hydrogen bonding site; (b) PXRD pattern comparison; (c) N2 sorption isotherm at 77 K and pore size distribution (inset) of MIL-53(Al)-OCH3; (d) Comparison of gravimetric CH4 uptake of the MIL-53(Al)-OH and MIL-53(Al)-OCH3 at 298 K.

Regarding the gravimetric deliverable capacity, the MIL-53(Al)OH clearly exhibits a more useful uptake profile than MIL-53(Al)(OH)2. For example, the np-lp phase transition is incomplete for MIL-53(Al)-(OH)2 at 65 bar, while the same phase transition is achieved in MIL-53(Al)-OH at 35 bar considering the 298 K isotherms. For the calculation and comparison of volumetric uptakes, a total uptake calculation is necessary including density and pore volume values. Although the density for calculation should be the tap density, the crystallographic density can be used for preliminary evaluation purpose. Considering the crystallographic density of MIL-53(Al)-OH (1.49 g cm-3 and 1.01 g cm-3 for vnp and lp phase, respectively), its absolute CH4 uptake at 298 K is 167 v/v or 119 mg g-1 at 35 bar, 217 v/v or 155 mg g-1 at 65 bar, and 53 v/v or 37 mg g-1 retained at 5.8 bar (Figure S14). The deliverable capacity of MIL-53(Al)-OH is 114 v/v or 82 mg g-1 at 35 bar and 164 v/v or 118 mg g-1 at 65 bar. These values are higher than that of HKUST-1 (143 v/v), comparable to the monolithic HKUST-1 (172 v/v),19 but lower than Co-BDP (197 v/v).21 For MIL-53(Al)-(OH)2, the 5-65 bar and 5-120 bar deliverable capacities are 71 v/v or 51 mg g-1 and 203 v/v or 145 mg g-1, respectively.

Figure 6. Optical images of (a) MIL-53(Al)-OH and (b) MIL-53(Al)(OH)2 under various conditions.

Further, we tried to comprehend the role of hydrogen bonding in determining the breathing behavior of the flexible MOFs under methane pressure. A control experiment was designed by replacing

ACS Paragon Plus Environment

Chemistry of Materials 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

the hydrogen bond forming hydroxyl group with methoxy one. The modified MOF viz. MIL-53(Al)-OCH3 shows similar crystallinity, stability, and porosity (Figure 5a-c). However, the high-pressure methane excess isotherm measurements reveal that unlike flexible MIL-53(Al)-OH, the MIL-53(Al)-OCH3 only exhibits Type I isotherm without hysteresis (Figure 5d). Thus, it could be established that the existence of hydrogen bonding site at the ortho-position of the ligand is crucial for designing MIL-53(Al) based flexible adsorbents for NG storage. Adsorbents for practical NG storage should meet the following criteria: 1) high NG storage and deliverable capacity, 2) decent chemical and mechanical stability, 3) low cost, and 4) good pelletizability to facilitate downstream processes. Regarding NG storage and deliverable capacity, the predicted optimized pore size and geometry for methane uptake is 1.12 nm slit-shaped pore.30 Interestingly, MIL-53(Al) structures have similar porous characteristics, making them ideal for further developments.29 In addition, MIL-53(Al) structures are well known for their high chemical and mechanical stability (Figure S5, S6, and Section S9).31 Furthermore, MIL-53(Al) based MOFs have been prepared by inexpensive Al-salts and functionalized benzenedicarboxylate linkers. The synthetic process can be easily scaled up to pilot scale,32 an important aspect for the test of on-board NG storage.

Figure 7. Gravimetric methane uptake comparison of powder and pelleted samples of (a) MIL-53(Al)-OH and (b) MIL-53(Al)-(OH)2 at 298 K.

Finally, pelletization has been commonly used to increase material density, and should be able to increase NG storage and deliverable capacities accordingly. However, high performing MOFs such as HKUST-1 show partial pore collapse upon pelletization with up to 35% decreased performance,5 making it the deciding criterion for selecting suitable MOF-based adsorbents. In addition, a proper density measurement is crucial for the accurate calculation of storage and deliverable capacity. Notably, most of the literature data of volumetric capacities are calculated using ideal single-crystal density, which is typically higher than the tap density or pellet density of the corresponding MOFs, leading to overestimated results. For example, the crystallographic and tap densities of HKUST-1 are 0.88 and 0.43 g cm-3, respectively. The tap density of MIL-53(Al)-OH and MIL-53(Al)-(OH)2 was measured to be 0.20 g cm-3, contrary to the ideal crystallographic densities of 1.01 g cm-3 and 1.04 g cm-3, respectively. In order to improve the material density and accordingly the volumetric NG uptake and deliverable capacity, MIL-53(Al)-OH and MIL-53(Al)(OH)2 were compacted into 8.5 mm pellets using a lab-scale pelletizer under an applied pressure of 6.9 kPa without the use of any external binder (Figure 6a-b). The pellets can still maintain the crystallinity, stability, and porosity of the MOF powders, as evidenced by PXRD, FT-IR, TGA, SEM and gas sorption analyses (Section S9 in SI). As expected, the pellet densities were increased up to 0.77 and 0.83 g cm-3 for MIL-53(Al)-OH and MIL-53(Al)(OH)2, respectively (Figure 6a-b). When the MIL-53(Al)-OH

pellets were subjected to high pressure methane sorption measurements, they exhibited similar breathing phase transitions as the powdered sample, with only 8% decrease in gravimetric uptake capacity measured at 65 bar at 298 K (Figure 7a). On the contrary, the gravimetric CH4 uptake of pelletized MIL-53(Al)-(OH)2 maintained breathing behavior without decrease in capacity, as measured at 100 bar and 298 K (Figure 7b). Unfortunately, both MIL-53(Al)-OH and MIL-53(Al)-(OH)2 pellets were partially disintegrated after high pressure methane sorption measurements (Figure 6a & 6b), resulting decrease in density up to 0.56 and 0.54 g cm-3, respectively. This can be attributed to the adsorptiondesorption induced stress generated by phase change of these flexible MOFs.21 Further studies should be targeted at increasing the stability of MOF tablets as well as the engineering design of NG storage tanks to accommodate possible adsorbent expansion and shrinkage during practical NG storage applications. Conclusion In conclusion, we investigate the flexibility and stability of hydroxyfunctionalized MIL-53(Al) MOFs for methane storage with high deliverable capacity. The positioning of the hydroxy groups promotes weak hydrogen bonding, which along with efficient π-π stacking control the breathing behavior of these MOFs. A promising deliverable capacity of 164 v/v was achieved in MIL53(Al)-OH at 65 bar and 298 K (considering crystallographic density), a value close to the record of 197 v/v by Co-BDP. The deliverable capacity has been further tuned by pressure, temperature, and density of the MOF pellets. Thus, a unique structure-property relationship is established to envisage flexible and stable MOFs for on-board NG storage after taking the engineering issues into consideration.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Experimental methods and synthesis procedures of the MOFs, FT-IR, TGA, SEM, TEM and XRD characterization, Low pressure and high pressure gas adsorption measurements (PDF)

AUTHOR INFORMATION Corresponding Author *Correspondence: [email protected]

ACKNOWLEDGMENT This work was supported by the National Research Foundation Singapore (NRF2018-NRF-ANR007 POCEMON), the Ministry of Education - Singapore (MOE AcRF Tier 1 R-279-000-472-112, R-279000-540-114), and the Agency for Science, Technology and Research (PSF 1521200078, IRG A1783c0015, and IAF-PP A1789a0024).

REFERENCES 1. U.S. Energy Information Administration. https://www.eia.gov/todayinenergy/detail.php?id=9991 (accessed February 28, 2019). 2. Alvarez, R. A.; Zavala-Araiza, D.; Lyon, D. R.; Allen, D. T.; Barkley, A. R.; Brandt, A. R.; Davis, K. J.; Herndon, S. C.; Jacob, D. J.; Karion, A.; Kort, E. A.; Lamb, B. K.; Lauvaux, T.; Maasakkers, J. D.; Marchese, A. J.; Omara, M.; Pacala, S. W.; Peischl, J.; Robinson, A. L.; Shepson, P. B.; Sweeney, C.; Townsend-Small, A.; Wofsy, S. C.; Hamburg, S. P. Assessment of methane emissions from the U.S. oil and gas supply chain. Science 2018, 361, 186188.

ACS Paragon Plus Environment

Page 4 of 6

Page 5 of 6 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

Chemistry of Materials 3. Semin, R. A. B., A technical review of compressed natural gas as an alternative fuel for internal combustion engines. Am. J. Eng. Appl. Sci. 2008, 1 (4), 302-311. 4. Yang, X.; Zheng, Q.; Gu, A.; Lu, X., Experimental studies of the performance of adsorbed natural gas storage system during discharge. Appl. Therm. Eng. 2005, 25 (4), 591-601. 5. Peng, Y.; Krungleviciute, V.; Eryazici, I.; Hupp, J. T.; Farha, O. K.; Yildirim, T., Methane Storage in Metal–Organic Frameworks: Current Records, Surprise Findings, and Challenges. J. Am. Chem. Soc. 2013, 135 (32), 11887-11894. 6. The Advanced Research Projects Agency – Energy (ARPA-E) of the US Department of Energy. https://arpa-efoa.energy.gov/Default.aspx?Archive=1 (accessed February 28, 2019). 7. Casco, M. E.; Martínez-Escandell, M.; Gadea-Ramos, E.; Kaneko, K.; Silvestre-Albero, J.; Rodríguez-Reinoso, F., High-pressure methane storage in porous materials: are carbon materials in the pole position? Chem. Mater. 2015, 27 (3), 959-964. 8. Menon, V. C.; Komarneni, S., Porous adsorbents for vehicular natural gas storage: a review. J. Porous Mater. 1998, 5 (1), 43-58. 9. Li, H. L.; Eddaoudi, M.; O'Keeffe, M.; Yaghi, O. M., Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402 (6759), 276-279. 10. Hönicke, I. M.; Senkovska, I.; Bon, V.; Baburin, I. A.; Bönisch, N.; Raschke, S.; Evans, J. D.; Kaskel, S., Balancing Mechanical Stability and Ultrahigh Porosity in Crystalline Framework Materials. Angew. Chem., Int. Ed., 2018, 57, 13780-13783. 11. He, Y.; Zhou, W.; Qian, G.; Chen, B., Methane storage in metalorganic frameworks. Chem. Soc. Rev. 2014, 43 (16), 5657-5678. 12. Li, B.; Wen, H.-M.; Zhou, W.; Xu, J. Q.; Chen, B., Porous MetalOrganic Frameworks: Promising Materials for Methane Storage. Chem 2016, 1 (4), 557-580. 13. Yuan, D.; Zhao, D.; Sun, D.; Zhou, H.-C., An Isoreticular Series of Metal–Organic Frameworks with Dendritic Hexacarboxylate Ligands and Exceptionally High Gas-Uptake Capacity. Angew. Chem. Int. Ed. 2010, 49 (31), 5357-5361. 14. Mason, J. A.; Veenstra, M.; Long, J. R., Evaluating metal-organic frameworks for natural gas storage. Chem. Sci. 2014, 5 (1), 32-51. 15. Bloch, E. D.; Murray, L. J.; Queen, W. L.; Chavan, S.; Maximoff, S. N.; Bigi, J. P.; Krishna, R.; Peterson, V. K.; Grandjean, F.; Long, G. J.; Smit, B.; Bordiga, S.; Brown, C. M.; Long, J. R., Selective Binding of O2 over N2 in a Redox-Active Metal-Organic Framework with Open Iron(II) Coordination Sites. J. Am. Chem. Soc. 2011, 133 (37), 14814-14822. 16. Simon, C. M.; Kim, J.; Gomez-Gualdron, D. A.; Camp, J. S.; Chung, Y. G.; Martin, R. L.; Mercado, R.; Deem, M. W.; Gunter, D.; Haranczyk, M.; Sholl, D. S.; Snurr, R. Q.; Smit, B., The materials genome in action: identifying the performance limits for methane storage. Energy Environ. Sci. 2015, 8 (4), 1190-1199. 17. Li, B.; Wen, H.-M.; Wang, H.; Wu, H.; Tyagi, M.; Yildirim, T.; Zhou, W.; Chen, B. A Porous Metal-Organic Framework with Dynamic Pyrimidine Groups Exhibiting Record High Methane Storage Working Capacity. J. Am. Chem. Soc. 2014, 136, 6207-6210. 18. Wen, H.-M.; Li, B.; Li, L.; Lin, R.-B.; Zhou, W.; Qian, G.; Chen, B. A Metal-Organic Framework with Optimized Porosity and Functional Sites for High Gravimetric and Volumetric Methane Storage Working Capacities. Adv. Mater. 2018, 30, 1704792. 19. Tian, T.; Zeng, Z.; 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. 2017, 17, 174-179.

20. Hu, Z.; Sun, Y.; Zeng, K.; Zhao, D., Structural-failure resistance of metal-organic frameworks toward multiple-cycle CO2 sorption. Chem. Commun. 2017, 53 (62), 8653-8656. 21. Mason, J. A.; Oktawiec, J.; Taylor, M. K.; Hudson, M. R.; Rodriguez, J.; Bachman, J. E.; Gonzalez, M. I.; Cervellino, A.; Guagliardi, A.; Brown, C. M.; Llewellyn, P. L.; Masciocchi, N.; Long, J. R., Methane storage in flexible metal-organic frameworks with intrinsic thermal management. Nature 2015, 527 (7578), 357-361. 22. Yang, Q.-Y.; Lama, P.; Sen, S.; Lusi, M.; Chen, K.-J.; Gao, W.-Y.; Shivanna, M.; Pham, T.; Hosono, N.; Kusaka, S.; Perry, J. J.; Ma, S.; Space, B.; Barbour, L. J.; Kitagawa, S.; Zaworotko, M. J., Reversible Switching between Highly Porous and Nonporous Phases of an Interpenetrated Diamondoid Coordination Network That Exhibits Gate-Opening at Methane Storage Pressures. Angew. Chem. Int. Ed. 2018, 57 (20), 56845689. 23. Elsaidi, S. K.; Mohamed, M. H.; Banerjee, D.; Thallapally, P. K., Flexibility in Metal–Organic Frameworks: A fundamental understanding. Coord. Chem. Rev. 2018, 358, 125-152. 24. Taylor, M. K.; Runčevski, T.; Oktawiec, J.; Gonzalez, M. I.; Siegelman, R. L.; Mason, J. A.; Ye, J.; Brown, C. M.; Long, J. R., Tuning the Adsorption-Induced Phase Change in the Flexible Metal–Organic Framework Co(bdp). J. Am. Chem. Soc. 2016, 138 (45), 15019-15026. 25. Yang, H.; Guo, F.; Lama, P.; Gao, W.-Y.; Wu, H.; Barbour, L. J.; Zhou, W.; Zhang, J.; Aguila, B.; Ma, S., Visualizing Structural Transformation and Guest Binding in a Flexible Metal–Organic Framework under High Pressure and Room Temperature. ACS Cent. Sci. 2018, 4 (9), 1194-1200. 26. Bolinois, L.; Kundu, T.; Wang, X.; Wang, Y.; Hu, Z.; Koh, K.; Zhao, D., Breathing-induced new phase transition in an MIL-53(Al)-NH2 metalorganic framework under high methane pressures. Chem. Commun. 2017, 53 (58), 8118-8121. 27. Stavitski, E.; Pidko, E. A.; Couck, S.; Remy, T.; Hensen, E. J. M.; Weckhuysen, B. M.; Denayer, J.; Gascon, J.; Kapteijn, F., Complexity behind CO2 Capture on NH2-MIL-53(Al). Langmuir 2011, 27 (7), 39703976. 28. Himsl, D.; Wallacher, D.; Hartmann, M., Improving the HydrogenAdsorption Properties of a Hydroxy-Modified MIL-53(Al) Structural Analogue by Lithium Doping. Angew. Chem. Int. Ed. 2009, 48 (25), 46394642. 29. Biswas, S.; Ahnfeldt, T.; Stock, N., New Functionalized Flexible AlMIL-53-X (X = -Cl, -Br, -CH3, -NO2, -(OH)2) Solids: Syntheses, Characterization, Sorption, and Breathing Behavior. Inorg. Chem. 2011, 50 (19), 9518-9526. 30. Chen, X. S.; McEnaney, B.; Mays, T. J.; Alcaniz-Monge, J.; CazorlaAmoros, d.; Linares-Solano, A., Theoretical and experimental studies of methane adsorption on microporous carbons. Carbon 1997, 35 (9), 12511258. 31. Yot G, P.; Yang, K.; Guillerm, V.; Ragon, F.; Dmitriev, V.; Parisiades, P.; Elkaïm, E.; Devic, T.; Horcajada, P.; Serre, C.; Stock, N.; Mowat P. S., J.; Wright A, P.; Férey, G.; Maurin, G., Impact of the Metal Centre and Functionalization on the Mechanical Behaviour of MIL‐53 Metal–Organic Frameworks. Eur. J. Inorg. Chem. 2016, 2016 (27), 4424-4429. 32. Ren, J.; Dyosiba, X.; Musyoka, N. M.; Langmi, H. W.; Mathe, M.; Liao, S., Review on the current practices and efforts towards pilot-scale production of metal-organic frameworks (MOFs). Coord. Chem. Rev. 2017, 352, 187-219.

ACS Paragon Plus Environment

Chemistry of Materials 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 6

SYNOPSIS TOC

ACS Paragon Plus Environment

6