Benchmark Study of Hydrogen Storage in Metal–Organic Frameworks

Mar 1, 2018 - NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg , Maryland 20899 , United States. ⊥ Dep...
5 downloads 8 Views 853KB Size
Benchmark Study of Hydrogen Storage in Metal−Organic Frameworks under Temperature and Pressure Swing Conditions

C

Table 1. Experimental Excess H2 Stored in Different Materials

arbon dioxide emission is a primary driver of anthropogenic climate change.1 In 2015, the transportation sector alone emitted 1733 million metric tons of CO2, which accounted for 32% of the total CO2 emissions in the United States.1 One approach to decreasing vehicular CO2 emissions is to transition away from fossil fuels and toward alternative sources of energy. The use of hydrogen as an alternative fuel, where the only byproducts are water and heat, would result in lower CO2 emissions.2 To advance the move toward a hydrogen economy, the current goal is to fabricate a fuel cell vehicle (FCV) that can have a driving range of 300 miles3 and thus requires 5.6 kg of H2 which must be stored and delivered in a safe, efficient manner.4 Toward this goal, the U.S. Department of Energy (DOE) has set 2020 (4.5 wt %; 30 g/L), 2025 (5.5 wt %; 40 g/L), and ultimate (6.5 wt %; 50 g/L) targets for onboard H2 storage systems (including tank, materials, and all other components) to be used in lightduty vehicles.4,5 Hydrogen can be stored as a pressurized gas, cryogenic liquid, or solid fuel through physisorption or chemisorption onto (or into) solid materials.6 H2 FCVs, using both compressed hydrogen gas (CHG) and cryo-compression methods, are emerging in the light-duty vehicle market: BMW is currently working on a cryo-compression prototype, i8 (350 bar operating pressure), while Toyota and Honda have commercialized the CHG Mirai and Clarity models, respectively (700 bar operating pressures).7 However, these high-pressure H2 storage methods require bulky carbon-fiberreinforced tanks.8 The use of adsorbent materials for H2 storage systems offers a less energy intensive way of storing H2 at near liquid densities compared to CHG methods, while still providing quick filling, use, and reuse due to the fast kinetics of H2 adsorption and desorption.9 Additionally, by using a solid adsorbent for H2, the pressure required for storage could be decreased, reducing the complexity and cost of FCVs10 and ultimately leading to safer and more fuel-efficient vehicles.11 These characteristics are also relevant in H2 transport because cryogenic liquid tanker trucks or gaseous tube trailers are used to deliver H2 from the production point to the end-use site.12 By using adsorption, the energetic cost of compression or liquefaction can be avoided, thus making it an appealing alternative to current H2 transport technologies. Many different adsorbents for H2 storage have been studied, including carbon materials,13 zeolites,14 covalent−organic frameworks (COFs), 15,16 porous aromatic frameworks (PAFs),16 and metal−organic frameworks (MOFs). Examples of materials in each class are shown in Table 1.18 Among the candidates studied as adsorbent materials, MOFs19−21 are particularly promising for gas storage. Their © XXXX American Chemical Society

material

excess H2 stored (wt %)

ZTC NaA COF-105 ILCOF-1 PAF-1 NU-100

6 2 10 5 7 10

conditions 77 77 77 77 77 77

K/24 K/15 K/80 K/40 K/48 K/56

bar bar bar bar bar bar

ref. 13 14 15 16 17 18

structural diversity and tunability allow for the systematic study of how pore size and volume,22 open metal sites,23 etc. affect adsorption and deliverable capacity, to ultimately design an optimal material. MOFs have been studied for storage of H2,24−26 CO2,27 and CH428,29 among other gases as well as for numerous gas or liquid separations and purifications.30−32 As a testament to their potential applicability for H2 storage, MOFs have exceeded the DOE 2020 onboard gravimetric H2 storage technical targets (4.5 wt %) at 77 K and 100 bar, although they are yet to do so at room temperature, for which MOFs have been reported to adsorb 1−2 wt % at 100 bar.33 Despite its importance,7,34 MOF volumetric storage has not been studied as much as gravimetric storage, although a handful of MOFs reach the 2020 target at 77 K and 100 bar.35 Indeed, through the use of cryo-adsorption storage, the amount of H2 delivered from MOF-based adsorption can be increased.36 The DOE center Hydrogen Storage Engineering Center of Excellence (HSECoE) has proposed designing tanks for cryoadsorption storage that operate with H2 loading and adsorption occurring at 77 K and 100 bar and discharge and desorption occurring at 160 K and 5 bar.37 When a combined temperature and pressure swing is performed to load and deliver the H2 from these systems, the amount of H2 remaining in the tank after use is minimized; therefore, the deliverable H2 is maximized.37 H2 storage in MOFs using pressure and temperature swing methods has yet to be systematically studied under these conditions experimentally. Herein, a series of MOFs with varying structural and physical properties (topology, porosity, presence of open metal sites, surface area, etc.) is studied experimentally and computationally to standardize comparisons and identify high-performing sorbents for volumetric H2 storage under the new temperature and pressure swing tank design conditions (Figure 1). All MOFs were first characterized using powder X-ray diffraction (PXRD) and compared to simulated patterns, to Received: January 30, 2018 Accepted: February 5, 2018

748

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

Cite This: ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters

Figure 1. Volumetric H2 deliverable capacity of the MOFs studied under different temperature and pressure conditions.

confirm crystallinity and phase purity (Figures S7−S17). Additionally, porosity was measured by N2 adsorption at 77 K and was consistent with literature values. Both simulated and experimental values for Brunauer−Emmett−Teller (BET) area and pore volume obtained from simulated and measured N2 isotherms (Figures S18−S28) are summarized in Table 2. All MOFs with the exception of CYCU-3-Al and NU-1000 show an almost perfect linear correlation between BET area and pore volume due to Type I N2 isotherm (Figure 2). The exceptions to this correlation are CYCU-3-Al and NU-1000, which present a large step in the N2 isotherm that corresponds to so-called “pore filling”38 of mesoporous channels. Simulations (Figure S29) reveal that for CYCU-3-Al and NU-1000 around 60% and 40% of all adsorbed N2 correspond to “monolayer formation” and “pore filling,” respectively. However, for all other MOFs, pore filling corresponds to only less than 20% of all adsorbed N2, making saturation loadings and loadings corresponding to monolayer formation rather similar. Hydrogen uptake studies were performed at 77, 160, and 296 K with pressures ranging from 0 to 120 bar (Figure 3). The details of the measured and simulated H2 isotherms are presented in Figures S34−S44. Generally, agreement between simulations and experiments was excellent.39 A summary of the major H2 uptake characteristics is presented in Tables S4 and S5 for all the MOFs studied. As expected, in every case, roomtemperature H2 uptake was substantially lower than any of the

Figure 2. Plot of BET area against pore volume. Roughly 1 cc/g pore volume gives about 2500 m2/g BET area.

Figure 3. Experimental (circles) and simulated (triangles) total hydrogen adsorption isotherms at 77 K (green), 160 K (blue), and 296 K (red) for NU-125.

Table 2. BET Area, Pore Volume, and Density for the MOFs Discussed BET area (m2/g)

pore volume, Vpore (cc/g)

MOF

measured

calculated

measured

calculated

density, ρ (g/cm3)

HKUST-1 NOTT-112 NU-125 rht-MOF-7 Cu-MOF-74 PCN-250 NU-1000b UiO-67 UiO-68-Ant CYCU-3-Al Zn2(bdc)2(dabco) NU-1101a NU-1102a NU-1103a

1980 3440 3230 1950 1270 1780 2200 2360 3030 2450 2020 4340 3720 6245

1950 4220 3140 2270 1190 1685 2350 2910 2930 2865 1990 4440 4710 6885

0.75 1.44 1.33 0.79 0.47 0.71 1.48 0.91 1.17 1.56 0.76 1.72 1.65 2.72

0.79 1.84 1.36 0.90 0.47 0.68 1.36 1.03 1.17 1.67 0.78 1.72 2.05 2.97

0.879 0.446 0.578 0.789 1.323 0.896 0.571 0.688 0.607 0.447 0.873 0.459 0.403 0.298

a

The values for NU-1101, NU-1102, and NU-1103 were obtained from Gómez-Gualdrón et al.35 bThe model with the best correlation and thus the one reported had 25% occupancy of hexagonal channels. 749

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters

MOFs. This MOF also has a large gravimetric deliverable capacity (8.5 wt %), which exceeds the 6.5 wt % ultimate system target. NU-1000 and UiO-68-Ant also exhibit high gravimetric and volumetric deliverable capacities. NU-1000, with the largest void fraction of the MOFs studied, has a gravimetric capacity of 8.3 wt % and 48 g/L, the second highest volumetric deliverable among the MOFs. NOTT-112, NU-125, and rht-MOF-7 are Cu(II)-based MOFs with rht topology, but their different linkers yield different cage sizes: 8.1, 10.3, and 16.6 Å for rht-MOF-7;46 11, 15, 16, and 24 Å for NU-125;29 and 13, 13.9, and 20 Å for NOTT-112.47 As a result, these MOFs have different areas and H2 uptake, despite containing the same inorganic node. As the BET area increases from 1950 m2/g in rht-MOF-7 to 3230 m2/g in NU-125 and 3440 m2/g in NOTT-112, the gravimetric deliverable capacity increases from 4.7 wt % to 8.5 and 9.1 wt %, respectively. In general, the gravimetric deliverable capacity of the MOFs studied is proportional to the pore volume and gravimetric surface area as expected (Figure 4).29,35 NOTT-112 demon-

DOE targets. However, the total gravimetric uptake is linearly proportional to surface area, roughly 0.5 wt % for every 1000 m2/g surface area reaching up to 3 wt % for NU-1103 (Figure S31). Cryogenic conditions showed promising uptake, with NU-1103 presenting the highest uptake. H2 uptake studies at 77 K have been reported for some of the MOFs discussed, and our results are consistent.9,35,40−45 At 77 K the well-known Chahine rule (i.e., 1 wt % excess H2 adsorption for every 500 m2/g surface area) holds for MOFs only up to 3000 m2/g. Above this surface area, we have roughly 0.5 wt % H2 excess adsorption for additional 1000 m2/g (Figure S30). Note that the validity of the Chahine rule has been discussed recently.34,35 Deliverable capacities were calculated by taking the difference between the total amount of H2 uptake at 100 bar and 77 K and uptake at 5 bar and 160 K, subscribing to the tank design criteria studied by the HSECoE.37 Volumetric H2 storage and deliverable capacities were calculated using monolithic densities. When the deliverable capacities of the MOFs are compared, considering only isothermal pressure swing at 77 K compared to the capacity with a combined temperature and pressure swing, a significant increase is observed (Table 3). Table 3. Experimental Deliverable Capacity and Isosteric Heat of Adsorption of the MOFs Discussed ΔH2@100 bar/ 77 K → 5 bar/77 K

ΔH2@100 bar/ 77 K → 5 bar/160 K

MOF

wt %

g/L

wt %

g/L

Qst kJ/mol

HKUST-1 NOTT-112 NU-125 rht-MOF-7 Cu-MOF-74 PCN-250 NU-1000 UiO-67 UiO-68-Ant CYCU-3-Al Zn2(bdc)2(dabco) NU-1101a NU-1102a NU-1103a

2.0 5.3 4.1 1.8 1.0 1.8 5.2 2.9 4.3 5.5 1.6 6.1 6.9 10.1

17 24 24 14 13 16 30 20 26 27 14 30 31 33

5.2 9.1 8.5 4.7 3.0 5.2 8.3 6.0 7.8 8.7 4.8 9.1 9.6 12.6

46 41 49 37 39 47 48 41 47 41 42 47 44 43

6.5 5.1 5.1 5.9 5.6 6.6 5.0 5.8 6.0 4.5 4.9 5.5 4.5 3.8

Figure 4. Plot of pore volume (red circles) and BET area (blue triangles) against gravimetric deliverable capacity. Approximately a linear relationship exists between increased gravimetric deliverable capacity with increasing pore volume and BET area.

a

The values for NU-1101, NU-1102, and NU-1103 were obtained from Gómez-Gualdrón et al.35

strated high uptake at 77 K and 100 bar (48 mmol/g) and lower uptake at 160 K and 5 bar (2.5 mmol/g) compared to NU-125, but the lower volumetric BET area (1534 m2/cm3) and void fraction (0.64) of NOTT-112 results in a volumetric deliverable capacity that is lower than that of NU-125 (41 vs 49 g/L). Because gravimetric BET area and pore volume are generally proportional to each other, gravimetric uptake is proportional to pore volume or surface area and varies greatly in the studied series. Volumetric uptake, on the other hand, is found to be nearly constant (45 ± 5 g/L) (Figure S32), likely because of being affected by both void fraction and volumetric surface area as previously discussed (Figure S33).35 Nevertheless, most of the MOFs studied showed volumetric uptakes that meet the DOE’s target at cryogenic temperatures under pressure and temperature swing conditions. To better understand the effect of the metal nodes as an adsorption site, the isosteric heats of adsorption (Qst) were calculated from the H2 isotherms (Figures 2 and S28−S38). The experimental Qst values for the series of MOFs studied range from 4.5 to 6.6 kJ/mol (Table 3). The values for the

Among the most notable increases are those for Cu-MOF-74, PCN-250, and Zn2(bdc)2(dabco), where the volumetric deliverable capacity triples when using the pressure and temperature swing method. Cu-MOF-74, for example, has a deliverable volumetric capacity under the new conditions of 39 g/L, a value that is much higher than the 13 g/L obtained under only pressure swing conditions. The additional capacity for Cu-MOF-74 attributed to the temperature swing results in H2 storage high enough to compete with current lightweight vehicles. All of the 14 MOFs studied exceed the DOE 2020 volumetric system target (30 g/L). NU-125 presents the highest volumetric deliverable capacity of all MOFs studied at 49 g/L, owing in part to a favorable combination of volumetric surface area and void fraction. The measured volumetric BET area (1870 m2/cm3) and void fraction (0.77) of NU-125 are the second and third largest, respectively, among the studied 750

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters copper paddlewheel MOFs (Cu-MOF-74, NOTT-112, HKUST-1, and NU-125) are independent of the volumetric density of metal centers. The similarities within the Qst values calculated indicate that for the conditions evaluated, the presence of accessible open metal sites is a determining factor neither in the interactions observed nor in the H2 deliverable capacity. The 14 MOFs tested present high deliverable capacities which meet the DOE system targets without considering any loss in capacity due to packing. Taking into account packing density, at least a 25% loss in capacity should be considered.48 As a result, with the exception of rht-MOF-7, the MOFs tested meet the volumetric DOE 2020 target. With this decrease in capacity, NU-125 has a volumetric deliverable capacity equal to that of current CHG technologies (37 g/L).49 Fourteen MOFs were evaluated for H2 deliverable capacity under conditions corresponding to a tank design studied by the HSECoE: charge at 100 bar and 77 K and discharge at 5 bar and 160 K. We have observed a reasonable linear correlation between pore volume and deliverable gravimetric capacity as well as a relatively constant trend for volumetric deliverable capacity. Some of the MOFs studied here present promising volumetric and gravimetric deliverable capacities which exceed the ultimate DOE technical targets, even when considering packing density. For example, the gravimetric and volumetric deliverable capacities for NU-125 (49 g/L, 8.5 wt %), NU-1000 (48 g/L, 8.3 wt %), and UiO-68-Ant (47 g/L, 7.8 wt %) are promising for applications in H2 storage and delivery. Additionally, the tank design requirements result in a significant 86% pressure reduction compared to current CHG technologies, making these MOFs attractive for further studies including scale-up and pelletization.

Points along the H2 isotherm were calculated using 2000 cycles for equilibration followed by 2000 cycles for data collection. For points along the nitrogen isotherm, 20 000 cycles for equilibration and 20 000 cycles for data collections were used. A cycle corresponds to N Monte Carlo moves, where N is defined as the number of adsorbates in the simulation supercell with a maximum number of 20. The equal-probability Monte Carlo moves correspond to translation, rotation, swap, and reinsertion. During simulation, MOF atoms were fixed at their crystallographic positions. MOF crystallographic structures were obtained from the Cambridge Structural Database (CSD)59 and prepared for simulation (e.g., eliminating crystallographic disorder, adding experimentally undetected hydrogen atoms, and removing solvent molecules) using Materials Studio.60 In the case of NU-1000, Materials Studio was also used to construct structures with various “occupancies” of the hexagonal channels (Figure S6). A Lennard-Jones (LJ) plus Coulomb potential was used to describe nonbonded interactions. To account for quantum effects at low temperature, Feymann−Hibbs corrections61 to the LJ potential were included for interactions involving H2 molecules. They were described using the Darkrim−Levesque model,62 whereas nitrogen molecules were described using the TraPPE model.63 MOF atoms were described by the Universal Force Field (UFF).64 Lorentz−Berthelot mixing rules were used to obtain adsorbate−adsorbent LJ parameters. Dispersion interactions were neglected after a 12.8 Å cutoff. The smallest l × n × m simulation supercell needed to satisfy the minimum image convention was used for each MOF. Electrostatic interactions were considered only for adsorbate−adsorbate interactions. Ewald summation was used to compute long-range electrostatic interactions. BET areas65 were calculated following the BET consistency criteria as discussed in previous work.38 Additional computational details can be found in the Supporting Information.



EXPERIMENTAL METHODS The MOFs, HKUST-1,29 NU-125,40 rht-MOF-7,46 NOTT112,40,47 Cu-MOF-74, PCN-250,50 NU-1000,41,42 UiO-67,51,52 UiO-68-Ant,53 CYCU-3-Al,54 and Zn2(bdc)2(dabco),55 were synthesized either using slightly modified literature methods or as detailed in the Supporting Information. The organic linker for NOTT-112 was synthesized through a four-step procedure, and the procedure for the rht-MOF-7 linker involved only one step56 (see Schemes S1 and S2). Synthesis of the MOFs NU1101, NU-1102, and NU-1103 and experimental details and data were previously reported by Gómez-Gualdrón et al.35 All samples were subject to solvent exchange with acetone or ethanol and thermally activated under dynamic vacuum to remove residual solvent present in the pores of each framework (details in the Supporting Information). For NOTT-112, NU125, and rht-MOF-7, supercritical CO2 activation was performed after the solvent exchange. Gas sorption measurements were performed on the activated samples at the NIST Center for Neutron Research (NCNR) using a computer-controlled Sieverts apparatus. Specifics regarding the instrument and procedure were previously reported.57 N2 isotherm measurements were also performed at Northwestern University (NU) using a Micromeritics Tristar II at 77 K. All gases used for the adsorption measurements were high-purity grade (99.999%).

Paula García-Holley† Benjamin Schweitzer‡ Timur Islamoglu† Yangyang Liu†,∇ Lu Lin† Stephanie Rodriguez§ Mitchell H. Weston§ Joseph T. Hupp† Diego A. Gómez-Gualdrón*,‡ Taner Yildirim*,∥ Omar K. Farha*,†,⊥ †

Department of Chemistry, Northwestern University, Evanston, Illinois 80202, United States ‡ Department of Chemical and Biological Engineering, Colorado School of Mines, Golden, Colorado 80401, United States § NuMat, Skokie, Illinois 60077, United States ∥ NIST Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland 20899, United States ⊥ Department of Chemistry, Faculty of Sciences, King Abdulaziz University, Jeddah 21589, Saudi Arabia



COMPUTATIONAL METHODS Adsorption simulations were performed using Grand Canonical Monte Carlo (GCMC) simulations using the RASPA code.58 751

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters



ASSOCIATED CONTENT

University of Science and Technology; NOTT, University of Nottingham; CSD, Cambridge structural database; LJ, Lennard-Jones; GCMC, grand canonical Monte Carlo; UFF, universal force field

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsenergylett.8b00154. Details on experimental and computational methods as well as adsorption isotherms, isosteric heats of adsorption, and PXRD patterns (PDF)





REFERENCES

(1) EPA. Draft Inventory of U.S. Greenhouse Gas Emissions and Sinks: 1990−2015; U.S. EPA, 2017. (2) Wang, M. Fuel choices for fuel-cell vehicles: well-to-wheels energy and emission impacts. J. Power Sources 2002, 112, 307−321. (3) Target Explanation Document: Onboard Hydrogen Storage for Light-Duty Fuel Cell Vehicles. https://energy.gov/sites/prod/files/ 2015/05/f22/fcto_targets_onboard_hydro_storage_explanation.pdf (accessed May 2017). (4) DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles. https://energy.gov/eere/fuelcells/doe-technicaltargets-onboard-hydrogen-storage-light-duty-vehicles (accessed May 2017). (5) For comparison, the density of hydrogen gas at ambient temperature and 700 bar is 40 g/L, while the density of liquid hydrogen at its atmospheric-pressure boiling point (20 K) is 70 g/L. The density of hydrogen gas at ambient temperature and 350 bar is 25 g/L. (6) Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy 2007, 32, 1121−1140. (7) Gomez-Gualdron, D. A.; Colon, Y. J.; Zhang, X.; Wang, T. C.; Chen, Y.-S.; Hupp, J. T.; Yildirim, T.; Farha, O. K.; Zhang, J.; Snurr, R. Q. Evaluating topologically diverse metal-organic frameworks for cryoadsorbed hydrogen storage. Energy Environ. Sci. 2016, 9, 3279−3289. (8) Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353−358. (9) Langmi, H. W.; Ren, J.; North, B.; Mathe, M.; Bessarabov, D. Hydrogen Storage in Metal-Organic Frameworks: A Review. Electrochim. Acta 2014, 128, 368−392. (10) Thomas, C. E.; James, B. D.; Lomax, F. D., Jr; Kuhn, I. F., Jr Fuel options for the fuel cell vehicle: hydrogen, methanol or gasoline? Int. J. Hydrogen Energy 2000, 25, 551−567. (11) Chahine, R.; Bénard, P. Adsorption Storage of Gaseous Hydrogen at Cryogenic Temperatures. In Advances in Cryogenic Engineering; Kittel, P., Ed.; Springer US: Boston, MA, 1998; pp 1257− 1264. (12) Hydrogen Delivery. https://energy.gov/eere/fuelcells/ hydrogen-delivery (accessed December 2017). (13) Stadie, N. P.; Vajo, J. J.; Cumberland, R. W.; Wilson, A. A.; Ahn, C. C.; Fultz, B. Zeolite-Templated Carbon Materials for High-Pressure Hydrogen Storage. Langmuir 2012, 28, 10057−10063. (14) Vitillo, J. G.; Ricchiardi, G.; Spoto, G.; Zecchina, A. Theoretical maximal storage of hydrogen in zeolitic frameworks. Phys. Chem. Chem. Phys. 2005, 7, 3948−3954. (15) Han, S. S.; Furukawa, H.; Yaghi, O. M.; Goddard, W. A. Covalent Organic Frameworks as Exceptional Hydrogen Storage Materials. J. Am. Chem. Soc. 2008, 130, 11580−11581. (16) Rabbani, M. G.; Sekizkardes, A. K.; Kahveci, Z.; Reich, T. E.; Ding, R.; El-Kaderi, H. M. A 2D Mesoporous Imine-Linked Covalent Organic Framework for High Pressure Gas Storage Applications. Chem. - Eur. J. 2013, 19, 3324−3328. (17) Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J. M.; et al. Targeted Synthesis of a Porous Aromatic Framework with High Stability and Exceptionally High Surface Area. Angew. Chem., Int. Ed. 2009, 48, 9457−9460. (18) Farha, O. K.; Ö zgür Yazaydın, A.; Eryazici, I.; Malliakas, C. D.; Hauser, B. G.; Kanatzidis, M. G.; Nguyen, S. T.; Snurr, R. Q.; Hupp, J. T. De novo synthesis of a metal−organic framework material featuring ultrahigh surface area and gas storage capacities. Nat. Chem. 2010, 2, 944−948. (19) Ferey, G. Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37 (1), 191−214.

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. *E-mail: [email protected]. ORCID

Timur Islamoglu: 0000-0003-3688-9158 Mitchell H. Weston: 0000-0002-4574-5888 Joseph T. Hupp: 0000-0003-3982-9812 Omar K. Farha: 0000-0002-9904-9845 Present Address ∇

Y.L.: Department of Chemistry and Biochemistry, California State University, Los Angeles, CA 90032. Notes

Views expressed in this Energy Focus are those of the authors and not necessarily the views of the ACS. The authors declare the following competing financial interest(s): The co-authors - Omar K. Farha, Joseph T. Hupp, Mitchell H. Weston, and Stephanie Rodriguez - have a financial interest in NuMat Technologies, a company that is commercializing MOFs.



ACKNOWLEDGMENTS The authors are grateful for the financial support from the U.S. Department of Energy, Office of Science: DE-FG02− 08ER15967. This work made use of the J.B. Cohen X-ray Diffraction Facility supported by the MRSEC program of the National Science Foundation (DMR-1720139) at the Materials Research Center of Northwestern University and the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF NNCI-1542205.) This work made use of the IMSERC at Northwestern University, which has received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-1542205); the State of Illinois and International Institute for Nanotechnology (IIN). Molecular simulations were performed in the Mio supercomputer cluster maintained by Colorado School of Mines. D.A.G.-G. thanks Colorado School of Mines for financial support. We gratefully acknowledge Yan Xu and Nicolaas A. Vermeulen for providing UiO-68-Ant ligand as well as helpful advice in the synthesis of NOTT-112 ligand.



ABBREVIATIONS FCV, fuel cell vehicle; NU, Northwestern University; NIST, National Institute of Standards and Technology; NCNR, NIST Center for Neutron Research; DOE, U.S. Department of Energy; HSECoE, Hydrogen Storage Engineering Center of Excellence:; CHG, compressed hydrogen gas; MOF, metal− organic framework; PCN, porous coordination network; CYCU, Chung Yuan Christian University; Ant, anthracene; UiO, University of Oslo; bdc, benzene-1,4-dicarboxylic acid; dabco, 1,4-diazabicyclo[2.2.2]octane; HKUST, Hong Kong 752

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters (20) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal− Organic Frameworks. Chem. Rev. 2012, 112, 673−674. (21) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. (22) An, J.; Rosi, N. L. Tuning MOF CO2 Adsorption Properties via Cation Exchange. J. Am. Chem. Soc. 2010, 132, 5578−5579. (23) Dincă, M.; Long, J. R. Hydrogen Storage in Microporous Metal−Organic Frameworks with Exposed Metal Sites. Angew. Chem., Int. Ed. 2008, 47, 6766−6779. (24) Murray, L. J.; Dinca, M.; Long, J. R. Hydrogen storage in metalorganic frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (25) Wang, Z.; Tanabe, K. K.; Cohen, S. M. Tuning Hydrogen Sorption Properties of Metal−Organic Frameworks by Postsynthetic Covalent Modification. Chem. - Eur. J. 2010, 16, 212−217. (26) Kesanli, B.; Cui, Y.; Smith, M. R.; Bittner, E. W.; Bockrath, B. C.; Lin, W. Highly Interpenetrated Metal−Organic Frameworks for Hydrogen Storage. Angew. Chem., Int. Ed. 2005, 44, 72−75. (27) Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 319, 939−943. (28) Guo, Z.; Wu, H.; Srinivas, G.; Zhou, Y.; Xiang, S.; Chen, Z.; Yang, Y.; Zhou, W.; O’Keeffe, M.; Chen, B. A Metal−Organic Framework with Optimized Open Metal Sites and Pore Spaces for High Methane Storage at Room Temperature. Angew. Chem., Int. Ed. 2011, 50, 3178−3181. (29) 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, 11887−11894. (30) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (31) Adil, K.; Belmabkhout, Y.; Pillai, R. S.; Cadiau, A.; Bhatt, P. M.; Assen, A. H.; Maurin, G.; Eddaoudi, M. Gas/vapour separation using ultra-microporous metal-organic frameworks: insights into the structure/separation relationship. Chem. Soc. Rev. 2017, 46, 3402− 3430. (32) Nakagawa, K.; Tanaka, D.; Horike, S.; Shimomura, S.; Higuchi, M.; Kitagawa, S. Enhanced selectivity of CO2 from a ternary gas mixture in an interdigitated porous framework. Chem. Commun. 2010, 46, 4258−4260. (33) Zhao, D.; Yuan, D.; Zhou, H.-C. The current status of hydrogen storage in metal−organic frameworks. Energy Environ. Sci. 2008, 1, 222−235. (34) Ahmed, A.; Liu, Y.; Purewal, J.; Tran, L. D.; Wong-Foy, A. G.; Veenstra, M.; Matzger, A. J.; Siegel, D. J. Bal-ancing gravimetric and volumetric hydrogen density in MOFs. Energy Environ. Sci. 2017, 10, 2459−2471. (35) Gómez-Gualdrón, D. A.; Wang, T. C.; García-Holley, P.; Sawelewa, R. M.; Argueta, E.; Snurr, R. Q.; Hupp, J. T.; Yildirim, T.; Farha, O. K. Understanding Volumetric and Gravimetric Hydrogen Adsorption Trade-off in Metal−Organic Frameworks. ACS Appl. Mater. Interfaces 2017, 9, 33419−33428. (36) Bobbitt, N. S.; Chen, J.; Snurr, R. Q. High-Throughput Screening of Metal−Organic Frameworks for Hydrogen Storage at Cryogenic Temperature. J. Phys. Chem. C 2016, 120, 27328−27341. (37) Siegel, D. J.; Hardy, B.; HSECoE Team. Engineering an Adsorbent-Based Hydrogen Storage System: What Have We Learned? https://energy.gov/sites/prod/files/2015/02/f19/fcto_h2_storage_ summit_siegel.pdf (accessed June 2017). (38) Gómez-Gualdrón, D. A.; Moghadam, P. Z.; Hupp, J. T.; Farha, O. K.; Snurr, R. Q. Application of consistency criteria to calculate BET areas of micro-and mesoporous metal−organic frameworks. J. Am. Chem. Soc. 2016, 138, 215−224. (39) Some simulation overestimation of measured H2 loadings at 77 K was observed for CYCU-3-Al, UiO-67, rht-MOF-7, and NOTT-112, which are the same MOFs for which the comparison between

simulated and measured N2 loadings at saturation indicate some extent of incomplete solvent removal or pore collapse. (40) Wilmer, C. E.; Farha, O. K.; Yildirim, T.; Eryazici, I.; Krungleviciute, V.; Sarjeant, A. A.; Snurr, R. Q.; Hupp, J. T. Gramscale, high-yield synthesis of a robust metal-organic framework for storing methane and other gases. Energy Environ. Sci. 2013, 6, 1158− 1163. (41) Mondloch, J. E.; Bury, W.; Fairen-Jimenez, D.; Kwon, S.; DeMarco, E. J.; Weston, M. H.; Sarjeant, A. A.; Nguyen, S. T.; Stair, P. C.; Snurr, R. Q.; et al. Vapor-Phase Metalation by Atomic Layer Deposition in a Metal−Organic Framework. J. Am. Chem. Soc. 2013, 135, 10294−10297. (42) Wang, T. C.; Vermeulen, N. A.; Kim, I. S.; Martinson, A. B. F.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K. Scalable synthesis and postmodification of a mesoporous metal-organic framework called NU1000. Nat. Protoc. 2016, 11, 149−162. (43) Li, B.; Zhang, Z.; Li, Y.; Yao, K.; Zhu, Y.; Deng, Z.; Yang, F.; Zhou, X.; Li, G.; Wu, H.; et al. Enhanced Binding Affinity, Remarkable Selectivity, and High Capacity of CO2 by Dual Functionalization of a rht-Type Metal−Organic Framework. Angew. Chem., Int. Ed. 2012, 51, 1412−1415. (44) Chavan, S.; Vitillo, J. G.; Gianolio, D.; Zavorotynska, O.; Civalleri, B.; Jakobsen, S.; Nilsen, M. H.; Valenzano, L.; Lamberti, C.; Lillerud, K. P.; et al. H2 storage in isostructural UiO-67 and UiO-66 MOFs. Phys. Chem. Chem. Phys. 2012, 14, 1614−1626. (45) Metallation of Metal-Organic Frameworks: En Route to Ambient Temperature Storage of Molecular Hydrogen. https:// www.hydrogen.energy.gov/pdfs/progress13/iv_c_5_mondloch_2013. pdf (accessed August 2017). (46) Luebke, R.; Eubank, J. F.; Cairns, A. J.; Belmabkhout, Y.; Wojtas, L.; Eddaoudi, M. The unique rht-MOF platform, ideal for pinpointing the functionalization and CO2 adsorption relationship. Chem. Commun. 2012, 48, 1455−1457. (47) Yan, Y.; Lin, X.; Yang, S.; Blake, A. J.; Dailly, A.; Champness, N. R.; Hubberstey, P.; Schroder, M. Exceptionally high H2 storage by a metal-organic polyhedral framework. Chem. Commun. 2009, 9, 1025− 1027. (48) Alezi, D.; Belmabkhout, Y.; Suyetin, M.; Bhatt, P. M.; Weseliński, Ł. J.; Solovyeva, V.; Adil, K.; Spanopoulos, I.; Trikalitis, P. N.; Emwas, A.-H.; et al. MOF Crystal Chemistry Paving the Way to Gas Storage Needs: Aluminum-Based soc-MOF for CH4, O2, and CO2 Storage. J. Am. Chem. Soc. 2015, 137, 13308−13318. (49) It is important to note that the weight and volume of the tank have not been considered here and would result in lower deliverable capacities. (50) Feng, D.; Wang, K.; Wei, Z.; Chen, Y.-P.; Simon, C. M.; Arvapally, R. K.; Martin, R. L.; Bosch, M.; Liu, T.-F.; Fordham, S.; et al. Kinetically tuned dimensional augmentation as a versatile synthetic route towards robust metal−organic frameworks. Nat. Commun. 2014, 5, 5723. (51) Klet, R. C.; Liu, Y.; Wang, T. C.; Hupp, J. T.; Farha, O. K. Evaluation of Bronsted acidity and proton topology in Zr- and Hfbased metal-organic frameworks using potentiometric acid-base titration. J. Mater. Chem. A 2016, 4, 1479−1485. (52) Liu, Y.; Klet, R. C.; Hupp, J. T.; Farha, O. Probing the correlations between the defects in metal−organic frameworks and their catalytic activity by an epoxide ring-opening reaction. Chem. Commun. 2016, 52, 7806−7809. (53) Wang, C.; Volotskova, O.; Lu, K.; Ahmad, M.; Sun, C.; Xing, L.; Lin, W. Synergistic Assembly of Heavy Metal Clusters and Luminescent Organic Bridging Ligands in Metal−Organic Frameworks for Highly Efficient X-ray Scintillation. J. Am. Chem. Soc. 2014, 136, 6171−6174. (54) Lo, S.-H.; Chien, C.-H.; Lai, Y.-L.; Yang, C.-C.; Lee, J. J.; Raja, D. S.; Lin, C.-H. A mesoporous aluminium metal-organic framework with 3 nm open pores. J. Mater. Chem. A 2013, 1, 324−329. (55) Dybtsev, D. N.; Chun, H.; Kim, K. Rigid and Flexible: A Highly Porous Metal−Organic Framework with Unusual Guest-Dependent Dynamic Behavior. Angew. Chem., Int. Ed. 2004, 43, 5033−5036. 753

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754

Energy Focus

ACS Energy Letters (56) Ma, D.; Li, B.; Zhou, X.; Zhou, Q.; Liu, K.; Zeng, G.; Li, G.; Shi, Z.; Feng, S. A dual functional MOF as a luminescent sensor for quantitatively detecting the concentration of nitrobenzene and temperature. Chem. Commun. 2013, 49, 8964−8966. (57) Zhou, W.; Wu, H.; Hartman, M. R.; Yildirim, T. Hydrogen and Methane Adsorption in Metal−Organic Frameworks: A High-Pressure Volumetric Study. J. Phys. Chem. C 2007, 111, 16131−16137. (58) Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q. RASPA: molecular simulation software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul. 2016, 42, 81−101. (59) Groom, C. R.; Bruno, I. J.; Lightfoot, M. P.; Ward, S. C. The Cambridge structural database. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2016, 72, 171−179. (60) BIOVIA, D. S. Materials Studio; Dassault Systèmes: San Diego, CA, 2017. (61) Feynman, R. P.; Hibbs, A. Quantum Mechanics and Path Integrals; McGraw-Hill: New York, 1965. (62) Levesque, D.; Gicquel, A.; Darkrim, F. L.; Kayiran, S. B. Monte Carlo simulations of hydrogen storage in carbon nanotubes. J. Phys.: Condens. Matter 2002, 14, 9285. (63) Potoff, J. J.; Siepmann, J. I. Vapor−liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676−1682. (64) Rappé, A. K.; Casewit, C. J.; Colwell, K.; Goddard, W.; Skiff, W. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024− 10035. (65) Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309−319.

754

DOI: 10.1021/acsenergylett.8b00154 ACS Energy Lett. 2018, 3, 748−754