Skip to main content
ACS Omega logoLink to ACS Omega
. 2025 Nov 17;10(46):56450–56457. doi: 10.1021/acsomega.5c08318

Understanding Porosity Loss in MOF Pelletization: Intrinsic vs Extrinsic Mechanical Stability in Zn-MOF-74

Ashley L Sutton †,*, Muhammad Munir Sadiq , Michael T Scalzo , Aaron Seeber , Stephanie A Bird §, Kristina Konstas , James I Mardel
PMCID: PMC12658601  PMID: 41322530

Abstract

Metal–organic frameworks (MOFs) are highly studied materials for gas storage and separation, with increasing interest in their application in industrial settings. Industrial processes typically require MOFs to be shaped into robust pellets. Unfortunately, pelletization processing can often result in significant loss in porosity and/or performance. This research examines the mechanical origins of such degradation using Zn-MOF-74 as a model system. A combination of high-pressure single-crystal X-ray diffraction (HP-SCXRD), periodic density functional theory (DFT), powder X-ray diffraction (PXRD), and nitrogen sorption measurements was undertaken to evaluate the mechanical response of Zn-MOF-74 under both hydrostatic (ideal) and uniaxial (practical) stress. SCXRD and DFT show that Zn-MOF-74 remains crystalline under isotropic compression beyond 2.8 GPa, yielding a bulk modulus of ∼15 GPa, confirming significant intrinsic stability under hydrostatic pressure. In contrast, uniaxial pelletization at pressures as low as 0.08–0.77 GPa results in an increasing reduction of BET surface area and microstrain. This difference can be attributed to extrinsic effects, including shear-induced microstructural damage and particle fragmentation. The shear yield strength, estimated empirically from the shear modulus, is only 0.16–0.50 GPa, consistent with the onset of degradation during pressing. These findings reveal the crucial role of nonhydrostatic stress in MOF shaping and demonstrate that bulk modulus alone is insufficient to predict mechanical resilience during pelletization. Strategies such as hydrostatic compaction, binder-assisted shaping, or solvation during pressing have the potential to mitigate porosity loss. This work provides a mechanistic understanding of how to improve MOF processability through both material and engineering solutions.


graphic file with name ao5c08318_0008.jpg


graphic file with name ao5c08318_0006.jpg

Introduction

Metal–organic frameworks (MOFs) are a rapidly growing class of porous materials characterized by their high surface areas, crystallinity, and tunable pore functionality. These features make MOFs attractive for applications such as gas storage and separation, catalysis, and drug delivery. While MOFs are typically synthesized and characterized as fine powders, their translation into real-world applicationsparticularly in industrial fixed-bed reactors- requires shaping into mechanically stable forms. The shaping of powders is commonly encountered in pharmaceutical and industrial separation industries. Shaping benefits include improved handling, reduced dust generation, and potential for increased packing density. Numerous methods for shaping MOF-based powders exist. These include pressed pelletization, granulation, extrusion, 3d printing, and spray drying among others. Pressed pelletization is perhaps the most commonly encountered shaping techniques. This is presumably due to its exceptional cost-effectiveness and the potential for scalable automation which is not currently true of all other shaping methods. Additionally, it has the potential to allow for MOF shaping in a solvent-free manner, without the need for potentially contaminating binders or exposing the material to heat. ,

However, the process of pressed pelletization can introduce mechanical stress that degrades material performance. , Porosity losses in pressed pellets is common. ,, For example, medium-sized crystals (1–2 μm) of Ni-MOF-74 exhibited a substantial decrease in N2 uptake at 77 K after application of pressure associated with pelleting. The exact decrease was pressure dependent, with a pressure of 0.31 GPa resulting in a decrease from ∼265 cc/g to ∼178 cc/g; further losses occurred with an increase to 0.62 GPa, which resulted in a drop to ∼48 cc/g.

It is well-known that pressed pellet formation applies a compression force that may induce framework deformation or collapse. This may arise from either intrinsic limitations (e.g., bulk moduli) of the MOF structure itself (e.g., pressure-induced pore collapse under hydrostatic stress), or extrinsic effects associated with the shaping process (e.g., shear-induced amorphization or crystallite fragmentation). The relative contributions of intrinsic vs extrinsic factors are rarely detailed.

Addressing the gap between intrinsic and extrinsic factors is key to the translation of MOFs into both commercial and industrial applications. A deeper insight into how mechanical forces influence structural and functional stability is needednot only to guide materials selection, but also to improve shaping methods. While many MOFs have substantial bulk moduli, MOFs typically have low shear moduli, rendering them very vulnerable to nonhydrostatic stress during pelletization.

Zn-MOF-74 was selected for this study, as it is one of the most well-studied MOFs with open metal sites. It has previously been shown to have pressure sensitivity, making it an ideal candidate for the examination for mechanical stability under both isostatic and uniaxial pressure. This work provides an insight into the relationship between mechanical stability and porosity retention during pelletization. This will ultimately allow for optimization through both material science and process engineering.

Methods

Synthesis of Zn-MOF-74

Crystals of Zn-MOF-74 were prepared using a modified literature procedure. Zinc nitrate hexahydrate (10.0 g, 33.6 mmol) and 2,5-dihydroxyterephthalic acid (H4DOBDC, 2.5 g, 12.6 mmol) were dissolved in a mixture of N,N-dimethylformamide (DMF, 200 mL) and distilled water (25 mL). The solution was sonicated for 10 min to ensure complete dissolution of the reactants, then heated at 100 °C for 22 h in a static (nonstirred) condition. Upon cooling to room temperature, the yellow crystalline product was isolated by decanting the mother liquor. The crystals were soaked with hot DMF (200 mL) for 24 h, then transferred to a Soxhlet apparatus and washed with methanol for 48 h. Finally, the product was dried under vacuum at 60 °C overnight.

Uniaxial Pelleting

Approximately 250 mg of oven-dried Zn-MOF-74 powder was loaded into a 13 mm stainless steel die. The sample was compacted using a manual hydraulic press at the target uniaxial pressure for 1 minute, then carefully released and removed as a self-supporting pellet.

High Pressure Single Crystal X-ray Diffraction

High-pressure single-crystal X-ray diffraction (HP-SCXRD) experiments on Zn-MOF-74 were performed using a micro diamond anvil cell (DAC). Data were collected at ambient temperature on the MX1 beamline at the Australian Synchrotron (λ = 0.71073 Å). , A ruby sphere was loaded alongside the Zn-MOF-74 crystal to enable in situ pressure calibration via ruby fluorescence. The sample chamber was filled with silicone oil, hydrostatic up to ∼3.0 GPa. Diffraction data sets were collected at sequential pressure points to monitor structural evolution under compression.

Powder X-ray Diffraction

X-ray diffraction data were collected using a Bruker D8 Advance A25 X-ray diffractometer equipped with a CuKα source and Lynx Eye XE T-detector. For collection, samples were placed on a silicon zero-background plate within a steel holder.

77 K N2 Isotherms

Nitrogen adsorption isotherms were collected at 77 K using a Micromeritics ASAP 2420 volumetric sorption analyzer. Prior to analysis, samples (including pressed pellets) were activated at 250 °C under dynamic vacuum for 6 hours to remove residual solvent and guest molecules. The Brunauer–Emmett–Teller (BET) surface areas were calculated from the adsorption isotherm using Micromeritics software.

298 K CO2 Isotherms

Carbon dioxide isotherms were collected at 298 K using a Micromeritics Tristar II 3020 analyzer. Prior to analysis, samples (including pressed pellets) were activated at 250 °C under dynamic vacuum for 6 hours to remove residual solvent and guest molecules. Total carbon dioxide uptake was determined from the collected adsorption isotherm.

Density Functional Theory Calculations

Periodic solid-state density functional theory (DFT) calculations were performed using Quantum ESPRESSO version 7.2, , employing periodic boundary conditions. A plane-wave energy cutoff of 1360 eV was chosen based on convergence testing, ensuring total energy accuracy within 1 meV/atom while minimizing Pulay stress artifacts. The Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional was used in conjunction with the DFT-D3 dispersion correction to account for long-range van der Waals interactions.

All calculations employed a Γ-centered k-point mesh of 1 × 1 × 4. Initial geometry optimization was performed at 0 GPa with all solvent molecules removed, including both coordinated and noncoordinated species establishing a fully relaxed, desolvated reference structure. Subsequent pressure points were computed iteratively: the optimized structure at each pressure served as the starting point for the next increment (e.g., the structure optimized at 0.20 GPa was used as input for the 0.40 GPa calculation). At each pressure, both atomic positions and lattice parameters were allowed to fully relax to ensure equilibrium conditions under stress.

This stepwise approach enabled systematic evaluation of pressure-dependent structural changes and ensured mechanical continuity between pressure points. The calculated pressure–volume data were used to extract the bulk modulus.

Results and Discussion

Structural Response of Zn-MOF-74 to Hydrostatic Stress

In this work, each characterization method plays a distinct role in elucidating the mechanism (i.e., intrinsic vs extrinsic) of structural change. HP-SCXRD reveals the framework’s intrinsic response to external pressure through application of isostatic pressure. DFT studies collaborate directly with HP-SCXRD, while providing fine grain detail about atomic changes. PXRD after uniaxial pelleting shows the effects of extrinsic factors (like shear forces). In combination, these techniques allow for a powerful understanding of the mechanisms of porosity loss in Zn-MOF-74. Initially, the intrinsic mechanical stability of Zn-MOF-74 was examined with high-pressure single-crystal X-ray diffraction (HP-SCXRD) using a microdiamond anvil cell (DAC). DAC measurements require a pressure-transmitting fluid for the isotropic application of pressure, and silicone oil was used as the nonpenetrating pressure-transmitting medium, which remains hydrostatic up to ∼3 GPa. Four high-quality diffraction data sets were obtained from two independent single crystals of Zn-MOF-74. For each crystal, structures were resolved at ambient pressure (within the DAC) and following compression to 0.43 GPa (crystal 1) and 1.15 GPa (crystal 2), respectively (see Table ). The retention of crystallinity for diffraction at all pressure points indicates that the crystals remain intact despite significant hydrostatic loading. Based on these experiments, Zn-MOF-74 can withstand at least 1.15 GPa of hydrostatic pressure without loss of long-range order and structural collapse.

1. Unit Cell Parameters of Single Crystals of Zn-MOF-74 in a Diamond Anvil Cell at Varying Pressures.

Crystal Pressure (GPa) a (Å) Δa (%) c (Å) Δc (%) V 3 )
1 ambient 26.239(8) - 6.6759(13) - 3980.4(18)
1 0.43 26.187(4) –0.20 6.635(2) –0.61 3940.7(15)
2 ambient 26.241(6) - 6.6832(9) - 3985.3(14)
2 1.15 26.145(3) –0.37 6.5414(14) –2.12 3872.2(11)

Increasing the DAC pressure from ambient to 1.15 GPa is associated with a contraction of the a-axis and c-axis by 0.37% and 2.12%, respectively. The total unit cell volume decreases by ∼2.8%. Within MOF-74 frameworks, the c-axis corresponds to the direction of the one-dimensional channels, and compression along this axis is more facile than in the a-b plane. Further, no evidence of crystallographic symmetry breaking or phase transitions was observed. This behavior is expected, with a previous DFT study indicating Zn-MOF-74 has a bulk modulus of between 9.8–12.6 GPa when desolvated. Together, these results indicate that Zn-MOF-74 has bulk moduli significantly higher than the pressure ranges applied during uniaxial pelleting (0.01–0.8 GPa).

It is worth acknowledging that experimental constraints surround the DAC loading process. While the crystals were preactivated by heating under vacuum, the DAC was loaded under ambient laboratory conditions, which allowed for at least partial rehydration of the framework during exposure to air. MOF-74 is well-known to adsorb water readily due to its open metal sites, and even brief exposure to humidity can result in the readsorption of water.

As such, it should be noted that partial solvation or rehydration may improve the apparent mechanical integrity of MOFs during pelletization, thereby influencing both intrinsic and extrinsic responses. While a systematic evaluation of these effects (e.g., humidity-controlled pelletization or pre/posthydration sorption analysis) lies beyond the present scope, this remains an important direction for future work.

Prior studies (both experimental and computational) have shown that both coordinated and pore-filling solvents can lead to significant increases in mechanical resilience in MOFs. , For Zn-MOF-74, Kamencek and Zojer reported that water coordination and pore filling can increase the bulk modulus of Zn-MOF-74 from 12 GPa when fully desolvated to 30 GPa when hydrated. Our belief is that our findings are still valuable, given that during pelleting operations, MOFs are rarely handled under strict anhydrous conditions, and some degree of solvation is expected under real-world pelleting conditions.

Given the limited number of pressure points accessible in the single-crystal experiments, periodic density functional theory (DFT) calculations were used to further investigate the pressure response of Zn-MOF-74 (see Figure ). To model the desolvated framework, all solvent molecules, including those coordinated to the Zn open metal sites, were removed. Previous studies have shown that such desolvation decreases the bulk modulus of Zn-MOF-74, providing a more conservative estimate of its intrinsic mechanical properties under idealized conditions.

1.

1

DFT optimized structure of Zn-MOF-74 at 0.1 GPa (blue), and at 2.8 GPa (orange). Differences in the frameworks highlight pressure-induced structural changes, such as bond distortions, unit cell shifts, or pore contractions.

At each pressure step, both atomic positions and lattice parameters were relaxed while preserving the rhombohedral R3̅ symmetry. The framework exhibited a linear reduction of both a- and c-axes from 0.0 to 2.8 GPa (see SI). No loss of symmetry or amorphization was encountered.

From the DFT-derived pressure–volume relationship (see Figure ), a bulk modulus (K) of 15.0 GPa was extracted by fitting data from 0.1 to 2.8 GPa (eq ).

K=V(ΔP/ΔV) 1

2.

2

Cell volume as a function of pressure calculated from DFT structures, the trend line provides the fit of the bulk modulus.

This value is consistent with prior computational reports (see Figure ). The volume change as a function of pressure shows intrinsic stability for all MOFs examined except NH2-MIL-53­(Al). Nonetheless, pressed pelleting frequently results in reduced porosity/performance, which points to extrinsic losses; however contribution of each has not been thoroughly explored.

3.

3

Normalized volume (V/V 0) as a function of pressure (GPa), illustrating the variability in compressibility across different MOFs. Data shown include Zn-MOF-74 (DFT, this work), MOF-5 (ref. ), and NH2-MIL-53­(Al) (ref. ), PCN-250 (ref. ), Hf-peb (ref. ) and SHF-62-DMF (ref. ). Only data in the range of ambient pressure to 3 GPa is shown.

Examination of the bond lengths and bond angles of Zn-MOF-74 at both 0.0 and 2.8 GPa reveals that most pressure-induced distortions occur around the Zn­(II) coordination environment.

The O–Zn bond lengths contract and expand as much as −0.07 Å and 0.46 Å, respectively. Furthermore, the O–Zn–O angle can deviate as much as −5° to +8° (see Supporting Information for further details). These distortions suggest that the Zn-carboxylate node has some pressure-accommodating capacity. It is plausible that this flexibility allows Zn-MOF-74 to maintain long-range order despite significant hydrostatic loading. These DFT results clearly show that Zn-MOF-74 possesses substantial intrinsic rigidity. Thus, it is unlikely that the intrinsic bulk moduli of Zn-MOF-74 is the determining factor for pore collapse during pelletization.

Porosity Loss upon Pelleting

Pelletization is a common method of shaping MOFs and typically occurs under uniaxial conditions. Pressures used for MOF pelletization span a relatively broad range from 0.01 GPa to beyond 0.8 GPa. Zn-MOF-74 was pelletized under analogous conditions with a series of increasing pressures: 0.08, 0.15, 0.31, 0.46, 0.62, and 0.77 GPa.

Nitrogen isotherms at 77 K were performed to quantify the impact of pelletization on porosity (see Figure ). Prior to compression, the activated powder exhibited a high BET surface area of 884 ± 18 m2/g, consistent with literature for Zn-MOF-74 and indicative of a highly porous framework. Applied pressure of just 0.08 GPa for 1 min resulted in a decrease in BET surface area to 98 ± 2 m2/g (an 89% reduction), highlighting the extreme sensitivity of the material.

4.

4

N2 77 K isotherms of Zn-MOF-74 before and after pelleting under uniaxial conditions at increasing pressures for 1 min.

Such a drastic porosity loss is not unprecedented, as there is significant variability reported for other frameworks under compression. For example, MOF-5 has been documented to display highly divergent BET surface area losses: −48% at 2 MPa, −99% at 10.3 MPa, −42% at 80 MPa, and −87% at 180 MPa. These results underscore the difficulty of drawing direct comparisons between studies and highlight that even drastic porosity losses, such as those we report here for Zn-MOF-74, fall within the broader range of behaviors observed in the literature.

Notably, intermediate pressures showed partial recovery in BET surface area and pore volume, likely due to particle fracture or rearrangement, exposing some residual porosity, supported by t-plot external surface area analysis (see Supporting Information). While both BET area and total pore volume decreased sharply, the external surface area increased at certain pressures (e.g., from 40 ± 1 m2/g at 0.08 GPa to 78 ± 2 m2/g at 0.15 GPa), suggesting the generation of new external surfaces via interparticle fracture or shear-induced delamination. At higher pressures (e.g., 0.62 GPa), a dramatic drop in external surface area to ∼1 m2/g likely reflects particle densification or poor connectivity between residual crystallites. These observations indicate that the dominant cause of porosity loss is not intrinsic framework collapseconsistent with HP-SCXRD and DFT results showing Zn-MOF-74 is stable up to 2.8 GPa hydrostaticallybut rather microstructural damage under uniaxial, nonhydrostatic stress.

Finally, the uptake of CO2 at room temperature as a function of applied uniaxial pressure was examined. Isotherms on a sample of Zn-MOF-74, which had not undergone applied pressure, exhibited an uptake of 2.72 mmol/g. CO2 uptake, however, declined abruptly at just 0.08 GPa, resulting in a reduced capacity of 0.74 mmol/g. Further pressure increases resulted in a continued gradual loss of Zn-MOF-74’s ability to uptake CO2 (see Supporting Information), with a pressure of just 0.31 GPa resulting in a reduction in uptake of just 0.29 mmol/g. These results highlight that beyond BET surface area real-world performance is also affected, which can pose significant challenges, for instance, in CO2 capture.

Follow-up characterization with PXRD (see Figure ) indicates that pelleting is associated with significant peak broadening, reduced peak intensity, and modest loss of crystallinity in the pressed sampleshallmarks of reduced crystallite size and increasing microstrain. The pristine Zn-MOF-74 powder had a crystallite size >2000 nm, which reduced in a clear nonlinear trend that can be approximated by a power-law relationship, with the sample exposed to 0.62 GPa of pressure with a crystallite size of just 60(2) nm, consistent with progressive defect formation under mechanical stress. Concurrently, microstrain rose from 0.021(1)% to 0.088(6)% in a linear fashion, suggesting a significant accumulation of lattice-level distortions, consistent with mechanically induced defect formation. The degree of crystallinity (DOC) showed a modest decline from ∼88% to ∼78% over the same pressure range, suggesting that while long-range order is increasingly disrupted, the overall framework remains largely intact. These trends are consistent with mechanically induced defect accumulation rather than phase transition or amorphization, supported by the absence of new reflections in the diffraction patterns.

5.

5

PXRD of Zn-MOF-74 before and after pelleting under uniaxial conditions at increasing pressures for 1 min.

The BET and PXRD results underscore that extrinsic mechanical effects are an important factor in MOF stability under practical pelleting conditions. While single-crystal and computational studies provide insight into the intrinsic resilience of the framework (under isostatic pressure), bulk powder processingparticularly uniaxial compactionintroduces shear forces and localized stresses that can damage the material at the microstructural level. As shaping of MOFs is essential for industrial applications, understanding the mechanisms that cause performance loss is critical.

Intrinsic vs Extrinsic Framework Stability

Zn-MOF-74 is known to possess a bulk modulus of 10–15 GPa. More relevant to pelletization behavior is the shear modulus, which governs resistance to shape distortion under nonhydrostatic stress. Kamencek and Zojer report a Hill-averaged shear modulus of 4.9 GPa for Zn-MOF-74, derived from DFT-based stress–strain calculations. While this value indicates mechanical softness, it does not directly reflect the stress at which the material begins to fail irreversibly. Instead, shear yield strength is the more relevant measure for structural collapse under applied load, which is typically an order of magnitude smaller than the bulk modulus.

Empirically, shear yield strength can be estimated using the relation τ y G/n, where n ranges from 10 to 30 depending on the material class. For Zn-MOF-74, this yields an estimated τ y of approximately 0.16–0.50 GPa, which is slightly higher than the reported onset of porosity loss under applied uniaxial pressure. The approach used here for shear yield strength assumes elastic–plastic behavior where yield occurs when shear stress exceeds a fraction of the shear modulus. However, for MOFs, this relationship should be treated as approximate, since framework flexibility, anisotropy, and defect distribution can alter the onset of yield. Therefore, the values reported here serve as indicative estimates to enable comparison, rather than an absolute measure of strength.

Further, during uniaxial pressing, failure is often initiated not by uniform stress but by amplified shear gradients along the die wall, local stress concentrations at interparticle contacts or defects, and grain boundary slippage within polycrystalline domains. As such, structural collapse can occur well below the theoretical yield limit derived from bulk moduli. This highlights that in MOF pelletization, the shear yield strength is also a key parameter, not just the bulk modulus. Practically, this suggests that hydrostatic or isostatic shaping techniques, which minimize shear differentials, may offer better preservation of crystallinity and porosity during processing.

Finally, Kamencek and Zojer demonstrated that coordinating water to the open metal sites in Zn-MOF-74 significantly increases the bulk modulus, from ∼12 to ∼30 GPa. It is therefore reasonable to expect that hydration also enhances the shear modulus and, by extension, the shear yield strength. This suggests that maintaining at least partial solvation during pelletization may help to suppress shear failure.

Conclusions

The clear distinction between intrinsic and extrinsic mechanical stability in Zn-MOF-74 has been demonstrated, along with the independent effects they have on MOF stability during pelletization. Bulk moduli would suggest that desolvated Zn-MOF-74 is stable to hydrostatic pressures of 10–15 GPa; however, pelletization typically occurs via uniaxial loads, which leads to shear forces. The shear yield strength is for Zn-MOF-74 is estimated to be ∼0.16–0.50 GPa. This is well within the typically applied pressures for pelleting.

Further, this study demonstrates that the mechanical degradation of Zn-MOF-74 pellets arises predominantly from extrinsic pelletization-induced porosity loss rather than intrinsic collapse of the framework lattice. By integrating high-pressure SCXRD, PXRD, and DFT, we decouple intrinsic and extrinsic contributions, showing how the bulk modulus provides an upper bound for intrinsic crystal stability, while the experimentally derived shear yield strength captures the dominant extrinsic failure mode. This combined approach establishes a systematic framework for assessing MOF mechanical stability under realistic processing conditions and offers a pathway to guide both material selection and shaping strategies for practical applications.

Thus, the shear yield strength should be considered the standard to predict the ability of an MOF to withstand uniaxial pelleting conditions. The literature indicates solvation can significantly improve the bulk modulifuture studies should investigate whether this holds true for shear yield strength. Independent of the bulk modulus, it is apparent that performance loss with pelleting can be reduced by either improving the shear tolerance of the MOF or reducing the shear forces during shaping.

Supplementary Material

ao5c08318_si_001.cif (128.4KB, cif)
ao5c08318_si_002.cif (156KB, cif)
ao5c08318_si_003.cif (156.7KB, cif)
ao5c08318_si_004.cif (153.8KB, cif)
ao5c08318_si_005.pdf (427.1KB, pdf)

Acknowledgments

This project was supported by resources and expertise provided by CSIRO IMT Scientific Computing. Part of this research was undertaken on the MX1 beamline at the Australian Synchrotron.

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c08318.

  • CIF deposited in the CCDC 2475455 (CIF)

  • CIF deposited in the CCDC 2475456 (CIF)

  • CIF deposited in the CCDC 2475457 (CIF)

  • CIF deposited in the CCDC 2475458 (CIF)

  • Additional data relating to HP-SCXRD, DFT calculations, gas sorption, and PXRD (PDF)

The manuscript was written through contributions of all authors.

The authors declare no competing financial interest.

References

  1. Sutton A. L., Melag L., Sadiq M. M., Hill M. R. C.. Capture, Storage, and Release of Oxygen by Metal–Organic Frameworks (MOFs) Angew. Chem., Int. Ed. 2022;61:e202208305. doi: 10.1002/anie.202208305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Sutton A. L., Mardel J. I., Hill M. R.. Metal-Organic Frameworks (MOFs) As Hydrogen Storage Materials At Near-Ambient Temperature. Chem. -Eur. J. 2024;30:e202400717. doi: 10.1002/chem.202400717. [DOI] [PubMed] [Google Scholar]
  3. Jiri, C. Metal-Organic Frameworks; Applications from Catalysis to Gas Storage; Wiley-VCH: Weinheim, 2012. [Google Scholar]
  4. Morris R. E., Wheatley P. S.. Gas Storage in Nanoporous Materials. Angew. Chem., Int. Ed. 2008;47(27):4966–4981. doi: 10.1002/anie.200703934. [DOI] [PubMed] [Google Scholar]
  5. Fan W., Zhang X., Kang Z., Liu X., Sun D.. Isoreticular Chemistry within Metal–Organic Frameworks for Gas Storage and Separation. Coord. Chem. Rev. 2021;443:213968. doi: 10.1016/j.ccr.2021.213968. [DOI] [Google Scholar]
  6. Li H., Wang K., Sun Y., Lollar C. T., Li J., Zhou H. C.. Recent Advances in Gas Storage and Separation Using Metal–Organic Frameworks. Mater. Today. 2018;21(2):108–121. doi: 10.1016/j.mattod.2017.07.006. [DOI] [Google Scholar]
  7. Li H., Li L., Lin R. B., Zhou W., Zhang Z., Xiang S., Chen B.. Porous Metal-Organic Frameworks for Gas Storage and Separation: Status and Challenges. EnergyChem. 2019;1:100006. doi: 10.1016/j.enchem.2019.100006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Dhakshinamoorthy A., Li Z., Garcia H.. Catalysis and Photocatalysis by Metal Organic Frameworks. Chem. Soc. Rev. 2018;47(22):8134–8172. doi: 10.1039/C8CS00256H. [DOI] [PubMed] [Google Scholar]
  9. Bavykina A., Kolobov N., Khan I. S., Bau J. A., Ramirez A., Gascon J.. Metal-Organic Frameworks in Heterogeneous Catalysis: Recent Progress, New Trends, and Future Perspectives. Chem. Rev. 2020;120(16):8468–8535. doi: 10.1021/acs.chemrev.9b00685. [DOI] [PubMed] [Google Scholar]
  10. Li D., Xu H. Q., Jiao L., Jiang H. L.. Metal-Organic Frameworks for Catalysis: State of the Art, Challenges, and Opportunities. EnergyChem. 2019;1(1):100005. doi: 10.1016/j.enchem.2019.100005. [DOI] [Google Scholar]
  11. Zulfiqar A., Miao B., Khan F., Ali N., Ahmed S., Rehman W., Asad M., Nawaz M. A., Mir I. A., Rasheed L.. Metal-Organic Framework (MOF)-Based Catalysts for Sustainable Energy Technologies: A Review. Langmuir. 2025;41(36):24049–24077. doi: 10.1021/acs.langmuir.5c02041. [DOI] [PubMed] [Google Scholar]
  12. Lawson H. D., Walton S. P., Chan C.. Metal-Organic Frameworks for Drug Delivery: A Design Perspective. ACS Appl. Mater. Interfaces. 2021;13(6):7004–7020. doi: 10.1021/acsami.1c01089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Sharabati M. A., Sabouni R., Husseini G. A.. Biomedical Applications of Metal–Organic Frameworks for Disease Diagnosis and Drug Delivery: A Review. Nanomaterials. 2022;12(2):277. doi: 10.3390/nano12020277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Moharramnejad M., Ehsani A., Shahi M., Gharanli S., Saremi H., Malekshah R. E., Basmenj Z. S., Salmani S., Mohammadi M.. MOF as Nanoscale Drug Delivery Devices: Synthesis and Recent Progress in Biomedical Applications. J. Drug Deliv. Sci. Technol. 2023;81:104285. doi: 10.1016/j.jddst.2023.104285. [DOI] [Google Scholar]
  15. Hughes R., Kotamreddy G., Ostace A., Bhattacharyya D., Siegelman R. L., Parker S. T., Didas S. A., Long J. R., Omell B., Matuszewski M.. Isotherm, Kinetic, Process Modeling, and Techno-Economic Analysis of a Diamine-Appended Metal-Organic Framework for CO2Capture Using Fixed Bed Contactors. Energy Fuels. 2021;35(7):6040–6055. doi: 10.1021/acs.energyfuels.0c04359. [DOI] [Google Scholar]
  16. Yeskendir B., Dacquin J. P., Lorgouilloux Y., Courtois C., Royer S., Dhainaut J.. From Metal-Organic Framework Powders to Shaped Solids: Recent Developments and Challenges. Mater. Adv. 2021;2(22):7139–7186. doi: 10.1039/D1MA00630D. [DOI] [Google Scholar]
  17. Zaghloul M. I., Elkady M. F., El-Khouly M. E., Mohamed G. G.. Tailored Zr Bio-MOF Structures for CO2 Adsorption: A Comparative Study of Binderless Pelletization and Chitosan-Based Extrusion. Int. J. Biol. Macromol. 2025;318:145295. doi: 10.1016/j.ijbiomac.2025.145295. [DOI] [PubMed] [Google Scholar]
  18. Finsy V., Verelst H., Alaerts L., De Vos D., Jacobs P. A., Baron G. V., Denayer J. F. M.. Pore-Filling-Dependent Selectivity Effects in the Vapor-Phase Separation of Xylene Isomers on the Metal-Organic Framework MIL-47. J. Am. Chem. Soc. 2008;130(22):7110–7118. doi: 10.1021/ja800686c. [DOI] [PubMed] [Google Scholar]
  19. Wang Z., Liu L., Li Z., Goyal N., Du T., He J., Li G. K.. Shaping of Metal-Organic Frameworks: A Review. Energy Fuels. 2022;36(6):2927–2944. doi: 10.1021/acs.energyfuels.1c03426. [DOI] [Google Scholar]
  20. Li Y., Wen G., Li J., Li Q., Zhang H., Tao B., Zhang J.. Synthesis and Shaping of Metal-Organic Frameworks: A Review. Chem. Commun. 2022;58(82):11488–11506. doi: 10.1039/D2CC04190A. [DOI] [PubMed] [Google Scholar]
  21. Peterson G. W., Decoste J. B., Glover T. G., Huang Y., Jasuja H., Walton K. S.. Effects of Pelletization Pressure on the Physical and Chemical Properties of the Metal-Organic Frameworks Cu3­(BTC)­2 and UiO-66. Microporous Mesoporous Mater. 2013;179:48–53. doi: 10.1016/j.micromeso.2013.02.025. [DOI] [Google Scholar]
  22. Wang T. C., Wright A. M., Hoover W. J., Stoffel K. J., Richardson R. K., Rodriguez S., Flores R. C., Siegfried J. P., Vermeulen N. A., Fuller P. E., Weston M. H., Farha O. K., Morris W.. Surviving under Pressure: The Role of Solvent, Crystal Size, and Morphology during Pelletization of Metal-Organic Frameworks. ACS Appl. Mater. Interfaces. 2021;13(44):52106–52112. doi: 10.1021/acsami.1c09619. [DOI] [PubMed] [Google Scholar]
  23. Serra-Crespo P., Dikhtiarenko A., Stavitski E., Juan-Alcañiz J., Kapteijn F., Coudert F. X., Gascon J.. Experimental Evidence of Negative Linear Compressibility in the MIL-53 Metal-Organic Framework Family. CrystEngComm. 2015;17(2):276–280. doi: 10.1039/C4CE00436A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chapman 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(48):17546–17547. doi: 10.1021/ja908415z. [DOI] [PubMed] [Google Scholar]
  25. Yang K., Zhou G., Xu Q.. The Elasticity of MOFs under Mechanical Pressure. RSC Adv. 2016;6(44):37506–37514. doi: 10.1039/C5RA23149C. [DOI] [Google Scholar]
  26. Wu H., Yildirim T., Zhou W.. Exceptional Mechanical Stability of Highly Porous Zirconium Metal-Organic Framework UiO-66 and Its Important Implications. J. Phys. Chem. Lett. 2013;4(6):925–930. doi: 10.1021/jz4002345. [DOI] [PubMed] [Google Scholar]
  27. Fu Y., Yao Y., Forse A. C., Li J., Mochizuki K., Long J. R., Reimer J. A., De Paëpe G., Kong X.. Solvent-Derived Defects Suppress Adsorption in MOF-74. Nat. Commun. 2023;14(1):2386. doi: 10.1038/s41467-023-38155-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Beamish-Cook J., Shankland K., Murray C. A., Vaqueiro P.. Insights into the Mechanochemical Synthesis of MOF-74. Cryst. Growth Des. 2021;21(5):3047–3055. doi: 10.1021/acs.cgd.1c00213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Quintelier M., Hajizadeh A., Zintler A., Gonçalves B. F., Fernández de Luis R., Esrafili Dizaji L., Vande Velde C. M. L., Wuttke S., Hadermann J.. In Situ Study of the Activation Process of MOF-74 Using Three-Dimensional Electron Diffraction. Chem. Mater. 2024;36(15):7274–7282. doi: 10.1021/acs.chemmater.4c01153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Kim K. J., Culp J. T., Ohodnicki P. R., Thallapally P. K., Tao J.. Synthesis of High-Quality Mg-MOF-74 Thin Films via Vapor-Assisted Crystallization. ACS Appl. Mater. Interfaces. 2021;13(29):35223–35231. doi: 10.1021/acsami.1c12000. [DOI] [PubMed] [Google Scholar]
  31. Adeel M., Ashraf M. W., Sajid M., Hanif S., Shah S. S. A., Nazir M. A.. Exploring MOF-74 Composites: From Novel Synthesis to Cutting-Edge Applications. Inorg. Chem. Commun. 2025;174:114026. doi: 10.1016/j.inoche.2025.114026. [DOI] [Google Scholar]
  32. Nguyen V. T. T., Luu C. L., Nguyen T., Nguyen A. P., Hoang C. T., Ha A. C.. Multifunctional Zn-MOF-74 as the Gas Adsorbent and Photocatalyst. Adv. Nat. Sci.: Nanosci. Nanotechnol. 2020;11(3):035008. doi: 10.1088/2043-6254/ab9d7c. [DOI] [Google Scholar]
  33. Boer S. A., Price J. R., Riboldi-Tunnicliffe A., Williamson R., Rostan R., Summers A., Turner G. F., Jones I., Bond C. S., Vrielink A., Marshall A. C., Hitchings J., Moggach S. A.. High-Pressure Single-Crystal Diffraction at the Australian Synchrotron. J. Synchrotron Radiat. 2023;30:841–846. doi: 10.1107/S160057752300406X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Cowieson N. P., Aragao D., Clift M., Ericsson D. J., Gee C., Harrop S. J., Mudie N., Panjikar S., Price J. R., Riboldi-Tunnicliffe A., Williamson R., Caradoc-Davies T.. MX1: A Bending-Magnet Crystallography Beamline Serving Both Chemical and Macromolecular Crystallography Communities at the Australian Synchrotron. J. Synchrotron Radiat. 2015;22(1):187–190. doi: 10.1107/S1600577514021717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Giannozzi P., Baroni S., Bonini N., Calandra M., Car R., Cavazzoni C., Ceresoli D., Chiarotti G. L., Cococcioni M., Dabo I., Dal Corso A., De Gironcoli S., Fabris S., Fratesi G., Gebauer R., Gerstmann U., Gougoussis C., Kokalj A., Lazzeri M., Martin-Samos L., Marzari N., Mauri F., Mazzarello R., Paolini S., Pasquarello A., Paulatto L., Sbraccia C., Scandolo S., Sclauzero G., Seitsonen A. P., Smogunov A., Umari P., Wentzcovitch R. M.. QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter. 2009;21:395502. doi: 10.1088/0953-8984/21/39/395502. [DOI] [PubMed] [Google Scholar]
  36. Giannozzi P., Andreussi O., Brumme T., Bunau O., Buongiorno Nardelli M., Calandra M., Car R., Cavazzoni C., Ceresoli D., Cococcioni M., Colonna N., Carnimeo I., Dal Corso A., De Gironcoli S., Delugas P., Distasio R. A., Ferretti A., Floris A., Fratesi G., Fugallo G., Gebauer R., Gerstmann U., Giustino F., Gorni T., Jia J., Kawamura M., Ko H. Y., Kokalj A., Kücükbenli E., Lazzeri M., Marsili M., Marzari N., Mauri F., Nguyen N. L., Nguyen H. V., Otero-De-La-Roza A., Paulatto L., Poncé S., Rocca D., Sabatini R., Santra B., Schlipf M., Seitsonen A. P., Smogunov A., Timrov I., Thonhauser T., Umari P., Vast N., Wu X., Baroni S.. Advanced Capabilities for Materials Modelling with Quantum ESPRESSO. J. Phys.: Condens. Matter. 2017;29:465901. doi: 10.1088/1361-648X/aa8f79. [DOI] [PubMed] [Google Scholar]
  37. Perdew J. P., Burke K., Ernzerhof M.. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996;77(18):3865–3868. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
  38. Grimme S., Ehrlich S., Goerigk L.. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011;32(7):1456–1465. doi: 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
  39. Kamencek T., Zojer E.. Understanding the Anisotropic Elastic Properties of Metal-Organic Frameworks at the Nanoscale: The Instructive Example of MOF-74. J. Phys. Chem. C. 2021;125(44):24728–24745. doi: 10.1021/acs.jpcc.1c07882. [DOI] [Google Scholar]
  40. Bazer-Bachi D., Assié L., Lecocq V., Harbuzaru B., Falk V.. Towards Industrial Use of Metal-Organic Framework: Impact of Shaping on the MOF Properties. Powder Technol. 2014;255:52–59. doi: 10.1016/j.powtec.2013.09.013. [DOI] [Google Scholar]
  41. Yuan S., Sun X., Pang J., Lollar C., Qin J. S., Perry Z., Joseph E., Wang X., Fang Y., Bosch M., Sun D., Liu D., Zhou H. C.. PCN-250 under Pressure: Sequential Phase Transformation and the Implications for MOF Densification. Joule. 2017;1(4):806–815. doi: 10.1016/j.joule.2017.09.001. [DOI] [Google Scholar]
  42. Redfern L. R., Farha O. K.. Mechanical Properties of Metal-Organic Frameworks. Chem. Sci. 2019;10(46):10666–10679. doi: 10.1039/C9SC04249K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Tan K., Nijem N., Gao Y., Zuluaga S., Li J., Thonhauser T., Chabal Y. J.. Water Interactions in Metal Organic Frameworks. CrystEngComm. 2015;17(2):247–260. doi: 10.1039/C4CE01406E. [DOI] [Google Scholar]
  44. Baxter S. J., Burtch N. C., Evans J. D., Ready A. D., Wilkinson A. P., Schneemann A.. Recovery of MOF-5 from Extreme High-Pressure Conditions Facilitated by a Modern Pressure Transmitting Medium. Chem. Mater. 2022;34(2):768–776. doi: 10.1021/acs.chemmater.1c03613. [DOI] [Google Scholar]
  45. Sussardi A., Hobday C. L., Marshall R. J., Forgan R. S., Jones A. C., Moggach S. A.. Metal-Organic Frameworks Correlating Pressure-Induced Emission Modulation with Linker Rotation in a Photoluminescent MOF. Angew. Chem., Int. Ed. 2020;59:8118–8122. doi: 10.1002/anie.202000555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Ashworth D. J., Carrington E. J., Roseveare T. M., McMonagle C. J., Ward M. R., Fletcher A. J., Düren T., Warren M. R., Moggach S. A., Oswald I. D. H., Brammer L.. Decoupled MOF Breathing: Pressure-Induced Reversal of Correlation Between Orthogonal Motions in a Diamondoid Framework. Angew. Chem., Int. Ed. 2025;64:e202504297. doi: 10.1002/anie.202504297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Macmillan N. H.. Review: The Theoretical Strength of Solids. J. Mater. Sci. 1972;7:239–254. doi: 10.1007/BF02403513. [DOI] [Google Scholar]
  48. Mackenzie, J. K. A Theory of Sintering and the Theoretical Yield Strength of Solids; University of Bristol: Bristol, 1949. [Google Scholar]
  49. Kelly, A. Strong Solids, 2nd ed.; Clarendon Press: Oxford, 1973. [Google Scholar]
  50. Langdon T. G.. Grain Boundary Sliding Revisited: Developments in Sliding over Four Decades. J. Mater. Sci. 2006;41(3):597–609. doi: 10.1007/s10853-006-6476-0. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao5c08318_si_001.cif (128.4KB, cif)
ao5c08318_si_002.cif (156KB, cif)
ao5c08318_si_003.cif (156.7KB, cif)
ao5c08318_si_004.cif (153.8KB, cif)
ao5c08318_si_005.pdf (427.1KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES