John Letwabaa,
Uwa Orji Uyor*ab,
Mapula Lucey Mavhungua,
Nwoke Oji Achukac and
Patricia Abimbola Popoolaa
aDepartment of Chemical, Metallurgical & Materials Engineering, Tshwane University of Technology, P.M.B X680, Pretoria, 0001, South Africa. E-mail: uwa.uyor@unn.edu.ng
bDepartment of Metallurgical and Materials Engineering, University of Nigeria, Nsukka, Private Bag 0004, Nsukka, Enugu State, Nigeria
cDepartment of Agricultural and Bioresources Engineering, University of Nigeria, Nsukka, Private Bag 0004, Nsukka, Enugu State, Nigeria
First published on 30th April 2024
Climate change is causing a rise in the need to transition from fossil fuels to renewable and clean energy such as hydrogen as a sustainable energy source. The issue with hydrogen's practical storage, however, prevents it from being widely used as an energy source. Current solutions, such as liquefied and compressed hydrogen storage, are insufficient to meet the U.S. Department of Energy's (US DOE) extensive on-board application requirements. Thus, a backup strategy involving material-based storage is required. Metal organic frameworks (MOFs) belong to the category of crystalline porous materials that have seen rapid interest in the field of energy storage due to their large surface area, high pore volume, and modifiable structure. Therefore, advanced technologies employed in the construction of MOFs, such as solvothermal, mechanochemical, microwave assisted, and sonochemical methods are reviewed. Finally, this review discussed the selected factors and structural characteristics of MOFs, which affect the hydrogen capacity.
Metal organic frameworks (MOF) are promising materials for hydrogen storage technology. They are microporous crystalline structures made up of metal ions connected by organic ligands which produce micropores (<2 nm) and pathways.5 Their unique characteristics namely thermal stability, rigidity, structural flexibility, high void volume, high surface area, adjustable pore size, and tailorable cavities of uniform size made MOFs receive a lot of attention in the research and development for hydrogen storage,6–8 organic pollutant removal,9 rechargeable batteries10,11 sensors application12 and photocatalyst for removal of antibiotics and CO2 photoreduction due to their high active sites and catalytic activity.13 The most common organic linkers like dicarboxylates (malonic acid, succinic acid, oxalic acid, glutaric acid, terephthalic acid), azoles (pyrrodiazole, 1,2,3-triazole) and tricarboxylates (trimesic acid, citric acid) are commonly used in the synthesis of MOFs.2 MOF-5 (Zn4O (BDC)3 where BDC denotes 1,4-benzenedi-carboxylate) is a promising material for hydrogen storage at both cryogenic and ambient conditions. The reported hydrogen adsorption storage capacity at cryogenic conditions and 1 bar is 4.5 wt%, whereas at ambient conditions and 1 bar is 1 wt%.14,15 Furukawa et al.16 prepared MOF-200 (Zn4O(CO2))6 [4,4,4-(benzene-1,3,5-triyl-tris(benzene-4,1-diyl))tribenzoate] and MOF-210 (Zn4O(CO2))6[4,4,4-(benzene-1,3,5-triyl-tris(ethyne-2,1-diyl))tribenzoate]/biphenyl-4,4-dicarboxylate from the solvothermal reaction of organic linkers and zinc nitrate. The hydrogen consumption obtained at 77 K and 80 bar for both MOF-200 and MOF-210 was 15 wt% and 14 wt% respectively. The excess uptake at ambient conditions at 80 bar for MOF-210 was 0.53 wt%. The authors believe that the lower hydrogen uptake was due to the larger average pore diameter recommended by DOE (0.7–1.2 nm to maximize the room temperature hydrogen uptake capacity).
A different intriguing family of MOFs is created by synthesizing copper with carboxylic acids. These MOFs are composed of a paddle wheel-style [Cu2(OOC)4] SBU and a carboxylate linker. Chui et al.17 synthesized MOF which consists of Cu(II) paddlewheel SBU and BTC, with a Langmuir surface area of 917.6 m2 g−1. In general, this kind of MOFs in addition to open metal locations offer large specific surface area and have strong interactions with gas molecules, leading to a higher amount of hydrogen gas being adsorbed. Azolate based MOFs are a class of compounds having high chemical and thermal stability. Their hydrogen storage capacity has been documented.18–20 The highest hydrogen adsorption capacity of one of ZIFs namely Zn(MeIM)2(ZIF-8) reported to date is 3.3 wt% of hydrogen at 77 K and 30 bar.21 This implies that carboxylate-based MOFs are better than ZIFs-based material for the storage of hydrogen. To improve hydrogen storage at ambient conditions, Frost et al.22 carried out calculations to forecast the hydrogen adsorption isotherms of the isoreticular metal organic framework. The findings showed that primarily due to the low adsorption energy between MOF and hydrogen, the connections that already have established for absolute hydrogen adsorption at 77 K are not valid for absolute hydrogen adsorption at 298 K. Also, they found that for large loadings, the excess adsorption of hydrogen at 298 K corresponds better with surface area than with free volume. At fewer loadings, the excess adsorption of hydrogen at 298 K relates well with the heat of adsorption. It was recommended that the MOFs should offer an isosteric heat of 15 kJ mol−1 or greater yet maintain a free volume of 2.5 cm3 g−1 and an empty percentage of 85% to achieve the aim of 9 wt% and 30 g L−1.
In a different investigation, Pt nanomaterials and carbon black were used to synthesize MOF-5, Zn4O (1,4-benzenedicarboxylate)3 by Kim and colleagues.23 At 298 K and 100 pressure, the composite of carbon black/Pt/MOF-5 adsorbs 0.62 wt% of hydrogen, which is an improvement over MOF-5 (0.44 wt%). It was found that it is challenging for the hydrogen to be kept on the surface of the MOFs if the value of the adsorption enthalpy is smaller, particularly in ambient settings. To enhance the diffusion of dissociated hydrogen to the surface of the adsorbent the loading of a metal catalyst, Pt (1, 4, 7, and 10%) was varied and it was found that 4% is the optimum loading with 0.88 wt% of hydrogen storage capacity. Furthermore, to protect the material from moisture, carbon black was introduced in the optimized formulation. This resulted in the reduction of hydrogen storage capacity from 0.88 wt% to 0.62 wt%. The authors stated that this was attributed to the reduction of surface area and blockage of the pores by carbon black. In this review, we discussed various technologies employed in the construction of MOFs, such as solvothermal, mechanochemical, microwave assisted, and sonochemical methods. Finally, we discuss the selected factors or structural characteristics of MOFs, which affect the hydrogen capacity. This review will form a basis for advanced research and development of novel MOFs for hydrogen energy storage in the promotion of clean and sustainable energy.
Fig. 1 Targeted objectives for hydrogen storage technologies set by the US DOE. Data derived from Energy24 with permission from DOE, copyright 2017. |
As of now, there are no technologies, which complement the DOE hydrogen storage targets. According to several researchers, the best hydrogen storage qualities should be attained to use hydrogen satisfactorily. These qualities include large gravimetric and volume densities, minimal sorption temperatures and heat exchange, excellent reversibility, lightweight, quick reaction kinetics, cycle stability, and reduced costs.27–30 Currently, there is no technology, which is compatible to meet the desired optimum characteristic of hydrogen storage at ambient conditions. Therefore, research and development of solid materials for hydrogen storage are required to fulfil the key role of achieving the set target.
Several methods have been developed for hydrogen storage as illustrated in Fig. 2, which include compression in gas cylinders, liquefaction in cryogenic tanks, storage in metal hydride alloys, physisorption, and chemisorption.31–33 Among these, compression is the most employed technology for hydrogen storage although high pressure has to be used to increase hydrogen density, which results in higher operational costs.2 Liquid hydrogen storage involves a liquefaction process in which the operation temperature is extremely low (−250 °C) which poses a major challenge to be achieved.34 Additionally, up to 40% of energy content can be lost in the process.35 The major setback for physical-based storage (compressed and liquefied) is the failure to achieve the DOE targets24 from energy density and cost perspectives.31
Fig. 2 Hydrogen storage methods. Reproduced from Moradi and Groth34 with permission from Elsevier, copyright 2019. |
Both chemisorption and physisorption are material-based hydrogen storage. In chemical adsorption, hydrogen is chemically bonded to the storage material. There are groups of materials that can be used for chemical sorption such as metal hydrides, ammonia, formic acid, carbohydrates, and many more. Among these groups of materials, metal hydrides are the most attractive candidates due to their ability to achieve relatively high volumetric density.30,36–38 However, the gravimetric capacity does not meet the set target by DOE.24 Moreover, metal hydride suffers from poor kinetics opting for charging time longer than the DOE target,24 and requires high temperature for adsorption/desorption. Adsorption in which gas molecules adhere to the surface of a solid due to weak intermolecular forces of attraction without the creation of a chemical bond between the adsorbate and the adsorbent is known as physical adsorption. For the accumulation of hydrogen, a variety of adsorbates have been created, including carbon-based materials, zeolites, organotransition metal complexes, glass capillary arrays, porous silicas, and metal organic frameworks.39
Physical adsorption on porous materials has the potential to store hydrogen due to their unique characteristics such as fast kinetics at low ambient conditions, large surface area, reversibility of the storage process, and lightweight.40–42 Covalent-organic frameworks (COFs), zeolites, and porous aromatic MOFs are some of the adsorbent materials that have been evaluated for hydrogen storage systems. Ramirez-Vidal et al.43 synthesized one of the COFs namely hyper-crosslinked polymers (HCPs) by utilizing the Friedel–Crafts reaction using carbazole, anthracene, dibenzothiophene, and benzene as raw materials and dimethoxymethane as a crosslinker.43 Physical properties of the synthesized HCPs obtained are a pore volume of 0.87 cm3 g−1 and specific surface area of 1137 m2 g−1. The optimal hydrogen capacity acquired was 2.1 wt% at operating temperature and pressure of 77 K and 40 bar respectively which is below the DOE target.24 Additionally, COFs-based materials suffer from high costs of precursors, poor mechanical properties, and complex synthesis processes.2 Carbon-based materials have been studied as potential hydrogen storage materials due to reduced costs, reduced weights, large surface area, large pore structure, and chemical stability.44–48 Thermally exfoliated graphene oxide with high surface area and pore volume was found to be an attractive material for hydrogen storage due to promising hydrogen uptake. Additionally, the high hydrogen uptakes on exfoliated GO is associated with the distance between the graphene layers, which promotes hydrogen diffusion. In recent work, Singh and De49 used the thermal exfoliation process under various circumstances to synthesize exfoliated graphite oxide. For air-exfoliated EGO with a specific surface area of 286 m2 g−1, mean pore size of 2.9 nm, and total pore volumes of 1.2 cm3 g, the optimal hydrogen absorption was 3.34 wt% at 77 K and 30 bar.
The structural features of MOFs include size, porosity, and morphology. Their several factors that affect the structural characteristics of MOFs such as nucleation, additive and synthesis on (size); template, ligand, and additive on (porosity); metal ion, spacer, ligand, capping ligand, hydrotrope, synthesis, and additive on (morphology).51 Recently, Suresh and co-workers52 assessed the effect of an additive consisting of a m-terephenyl-4,4-dicarboxylic acid moiety on the emergence of MOF-5 morphologies (Fig. 3). Fig. 3a shows cubic morphology crystals with {100} crystallographic facets which suggest slow crystal growth. Cuboctahedral-shaped crystals covered by six {100} and eight {111} facets are displayed (Fig. 3b). Octahedral crystal morphology growth along the {111} facet direction and {100} facet. These morphology evolutions demonstrated the effect of additives on interaction with crystallographic facets about the crystal growth rate. Li et al.53 constructed two novel porous MOFs, namely, [Zn4(μ3-OH)2(TPO)2(H2O)2 (product 1)] and [Zn6(μ6-O)(TPO)2](NO3)4·3H2O (product 2)] based on H3TPO ligands via solvothermal methods. These MOFs consist of butterfly-shaped Zn4(μ3-OH)2(CO2)6 and octahedral Zn6(μ6-O)(CO2)6 SBU, which resulted in obtaining a three-dimensional microporous framework with flu and pyr topologies, large cavities and one dimensional channels (see Fig. 4 and 5).53 Generally, the high structural stability of MOFs is associated with SBU as they serve as rigid vertices propagated into a framework by strongly bonding with organic struts. According to Butova et al.54 the choice and deletion of a linker could change the symmetry structure. Whenever a different linker is used, the structure's symmetry is preserved. But as the carbon chain lengthens, the characteristics of the unit cell alter. Moreover, symmetry changes as a result of changes in the mutual arrangement of functionalities.
Fig. 3 Cubic crystal shape (a) is governed by the slower rate of crystal growth in the direction of faces. Additive (m-terephenyl-4,4-dicarboxylate indicated in green) blocking MOF-5 growth along the {111} direction partially or completely at the expense of all {100} facets during crystal growth in creation of cuboctahedral (b) and octahedral (c) crystal morphologies. Reproduced from Suresh et al.52 with permission from American Chemical Society, copyright 2021. |
Fig. 4 (a) Coordination environment for zinc(II) atoms in complex system, (b) cagelike cavity supported by Zn4 cluster and TPO3+ Ligands, (c) packing structure of cagelike cavity, and (d) flu net topology. Reproduced from Li et al.53 with permission from American Chemical Society, copyright 2014. |
Fig. 5 (a) Asymmetric unit of complex system 2, (b) SBU in complex 2, (c) 3D packing stricture of complex system 2 demonstrating the open micropores, and (d) Pry net topography of complex system 2. Reproduced from Li et al.53 with permission from American Chemical Society, copyright 2014. |
Great porosity, great surface area, less density, hardness, bulk modulus, shear modulus, and Young's modulus may be necessary for MOFs utilized in gas adsorption, gas sensing, and gas storage applications.55–57 Suh et al.21 disclosed the BET surface area of MOF-200, MOF-210, and MOF-205 as 4530, 6240, and 4460 m2 g−1, respectively. In terms of crystal density, MOF-200, and MOF-210 recorded the lowest value of 0.25 g cm−3 among MOFs, MOF-205 with a crystal density of 0.38 g cm−3. Mechanical properties are vital properties of MOFs, which among others are affected by the removal of quest molecules and porosity.58 It is reported that a large degree of coordination of the metal center to the organic linkers positively influences the mechanical stability of MOFs.59 Burtch et al.60 improved the young modulus of Zn2(L)2(DABCO) from 2.08 to 6.53 GPa by using N,N-dimethylformamide (DMF) and toluene as a solvent. This improvement was attributed to a change in the geometry of the framework during synthesis.
Also, various techniques and methods such as crystallographic density, helium pycnometry, gas adsorption, and Archimedes' principle are employed to measure the density of metal–organic frameworks (MOFs), each offering its unique advantages and limitations. For instance, crystallographic density derives the MOF's density from its crystal structure, assuming a pristine, defect-free crystal. Often used as a reference, this method is straightforward and omits the need for experimental measurements. However, it overlooks defects and impurities, which can substantially influence MOF density. In contrast, helium pycnometry gauges the skeletal density of MOFs by measuring the volume of displaced helium gas, convertible to the MOF's apparent density. This technique excels in accuracy and reliability, even when defects or impurities are present. Nevertheless, it necessitates specialized equipment and can be time-intensive.62 Gas adsorption takes a different route, focusing on MOF surface area and porosity by adsorbing gas molecules onto the MOF surface. MOF density calculations stem from these surface area and pore volume measurements, critical parameters for many applications. However, this method presumes homogeneity and uniform pore size distribution within the MOF, which may not hold for all MOFs. Meanwhile, Archimedes' principle approaches density measurement by comparing the weight of the MOF in air and a liquid of known density. From this weight difference and the MOF's volume, its density can be determined. Simplicity is its strength, as it requires no specialized equipment. Nonetheless, it can be influenced by factors like air bubbles or surface roughness, potentially impacting measurement accuracy.
It has been shown that the assessment of MOF density relies on a range of techniques, each carrying its own set of benefits and limitations. Helium pycnometry stands out for its reliability and precision, while gas adsorption supplies valuable insights into MOF surface characteristics. Crystallographic density offers a straightforward reference point, albeit without considering defects. Archimedes' principle provides simplicity but can be sensitive to certain environmental factors. Researchers judiciously select these methods based on their research goals and the specific attributes of the MOFs under investigation, ensuring a comprehensive understanding of MOF density and its ramifications in diverse applications. It has been shown that the density of metal–organic frameworks (MOFs) is influenced by several factors, including chemical composition, crystal structure, pore size and framework topology, temperature and pressure conditions, and guest molecules and adsorption.63
The chemical composition of MOFs can affect their density due to differences in atomic mass and packing efficiency. For example, MOFs with heavier metal ions or larger organic ligands tend to have lower densities due to their larger atomic radii and lower packing efficiency. On the other hand, MOFs with lighter metal ions or smaller organic ligands tend to have higher densities due to their smaller atomic radii and higher packing efficiency. For example, MOF-74 is a series of MOFs with different metal ions (e.g., Mg, Co, Ni) and organic ligands (e.g., benzene-1,3,5-tricarboxylate) that have different densities. The density of MOF-74 increases with increasing atomic mass of the metal ion, with Ni-MOF-74 having the highest density.64 The crystal structure of MOFs can affect their density due to differences in packing efficiency and void space. MOFs with higher symmetry and more compact crystal structures tend to have higher densities, while MOFs with lower symmetry and more open crystal structures tend to have lower densities. For example, MOFs with diamondoid or zeolitic imidazolate frameworks (ZIFs) tend to have higher densities due to their compact crystal structures.
Likewise, the pore size and framework topology of MOFs can affect their density due to differences in void space and packing efficiency. MOFs with larger pore sizes and more open framework topologies tend to have lower densities, while MOFs with smaller pore sizes and more compact framework topologies tend to have higher densities. For example, MOFs with isoreticular frameworks (IRFs) tend to have higher densities due to their compact framework topologies.65 Also, Halford was able to demonstrate that using an actinide building block, a material that is unusual, and complex, to create the lowest density of MOFs. He points out that actinides, which are heavy elements near the bottom of the periodic table, are not often thought of as low-density materials. In the research, he prescribed that if the MOF's structure was highly porous and empty, the size of the metal wouldn't matter. Due to this research, it has been recorded in comparison with other known MOF that to date, the complex NU-1301 MOF composed of uranium oxide and tricarboxylate organic linker units has the lowest density66 with 0.124 g cm−3 despite its high surface area and high pore volume. Fig. 6 presents the physical structure of the NU-1301. It is deduced that the NU-1301-unit cell is made up of five cage structures (shown in green and other colours) that are combined to form the cuboctahedron building block (top left), which is composed of uranium oxide nodes and tricarboxylate bridging ligands.
Fig. 6 Simple to complex NU-1301 MOF reproduced from Halford66 with permission from American Chemical Society, copyright 2017. |
The temperature and pressure conditions during MOF synthesis and characterization can affect their density due to changes in crystal structure and guest molecule adsorption. For example, MOFs synthesized at higher temperatures or pressures tend to have higher densities due to increased packing efficiency and decreased void space. However, guest molecules adsorbed in the MOF pores can also affect the density by increasing the overall volume of the MOF. For instance, HKUST-1 is an MOF that undergoes a reversible phase transition at high pressure, resulting in a change in crystal structure and density. At low pressure, HKUST-1 has a lower density due to the presence of guest molecules in the pores, while at high pressure, the guest molecules are expelled, and the density increases. The guest molecules adsorbed in the MOF pores can affect the density by increasing the overall volume of the MOF. For example, MOFs with high guest molecule adsorption tend to have lower densities due to increased void space.
To address these issues, scientists have created a number of densification processes that allow MOFs to be made denser while yet maintaining their internal pore structure. Compaction, pelletization, hot pressing, and solvent-assisted grinding are a few of these methods. Numerous research have shown that it is feasible to improve the mechanical stability and gas adsorption capability of MOFs by adjusting their compact density. For instance, Purewal et al.67 observed that compared to the powder, the volumetric H2 absorption of compacted MOF-5 with a density of 0.51 g cm−3 dramatically increased. MIL-101 compacted at 8 MPa with a density varying between 0.45 and 0.47 g cm−3 was described by Ardelean et al.68 On the other hand, densified Zr-MOFs and HKUST-1 MOFs have a larger volumetric nitrogen adsorption capability than their powder counterparts, according to Dhainaut et al.69 The investigations have demonstrated that densification procedures can influence the surface area, pore volume, and pore size distribution of MOFs. The mechanical stability of MOFs may also be impacted by the densification process. To guarantee that MOFs' storage capacity is not diminished, it is crucial to strike a balance between densification and preserving their natural textural qualities.
Fig. 7 Various methods of synthesizing MOFs and their possible final reaction products. Reproduced from Stock and Biswas71 with permission from American Chemical Society, copyright 2012. |
Fig. 8 MOF-5 SEM-EDX images synthesis at various conditions (a) 105 °C for 144 h, (b) 120 °C for 24 h, and (c) 140 °C for 12 h. Reproduced from Mulyati et al.73 with permission from Indonesian Journal of Chemistry, copyright 2015. |
It is reported that all three samples have cubic morphologies with different particle sizes, which can be categorized as smaller particle size (85–95 μm), and larger particle size (250–300 μm) synthesized at high temperatures (140 °C), shorter time (12 h) and lower temperature and longer time, at 100 °C for 144 and 120 °C for 24 respectively. When the solubility of the precursors is studied, it is discovered that metal salts can only dissolve in strongly polar organic solvents, whereas organic linkers can dissolve in less polar or non-polar organic solvents. As a result, it becomes necessary to combine two or more solvents when conducting synthesis. As a result, the creation of the framework and physical properties can be influenced by the solvent's composition.74 Dietzel et al.75 used 2,5-dihydroxyterephthalic acid as the linker to investigate the impact of solvent change on framework topologies. The system's simplicity was exploited to control how the acidic proton in the side chain dehydrates (hydroxyl group). A three-dimensional nonporous structure with Pts topology [Mg(H2dhtp)(H2O)2] was created using a lower concentration of NaOH (2 mL) in the ethanol/water solvent mixture, and the side chain hydroxyl groups remained protonated. Nevertheless, adding more NaOH (3 mL) to a tetrahydrofuran/water solvent caused the hydroxyl group to dehydrate, which produced MOF-74-Mg, which is made up of 1D helical chains of edge-sharing Mg octahedral. A solvothermal method has been utilized to construct MOFs for hydrogen storage (carboxylate-based, azolate-based, MOFs with mixed ligands and MOFs with metal complexes as building blocks). The setbacks of using the solvothermal method for the construction of MOFs include; lengthy reaction time, large operating temperature, and great cost of the solvent.
(1) |
Microwave development has been used to synthesize various MOFs. Chromium-benzenetricarboxylate (MIL-100-Cr) is produced by Jhung et al.80 using a microwave process from Cr metal, H3BTC, and an aqueous solution of hydrofluoric acid. By using this technique, it was possible to cut the reaction time required to build MIL-100-Cr from 96 to 4 h. Since the reaction time is reduced in MW method, the crystal growth and aggregation rate are limited resulting in obtaining a smaller particle size.81 Hence, it is essential to optimize process parameters such as power input and reaction time to obtain MOFs with satisfactory properties. Using MW-assisted technique, Jhung et al.80 created nanoporous Cr-MIL-101 with diameters ranging from 40 to 90 nm. This was accomplished by changing the heating period from 1 to 40 minutes at 600 W and the energy input from 36 to 1440 kJ. It was discovered that when the crystallization time increases, Cr-dimensions MIL-100's grow and become more uniform. With an increase in energy input, the shape and morphology evolved (from sphere to cubic) (see Fig. 9).
Fig. 9 Impact of crystallization time on the morphology of MIL-101 prepared with microwave irradiation at 210 °C for (a) 1, (b) 2, (c) 10, and (d) 40 min. Reproduced from Jhung et al.80 with permission from John Wiley and Sons, copyright 2006. |
The impact of the synthesizing circumstances, specifically the temperature and heating technique, on the structure of MOFs was investigated by Seo et al.82 The synthesis of MOF at 120 °C and 170 °C generated Cu3(BTC)2(H2O)3 and [Cu2(BTC)(OH)(H2O)]·2H2O phases respectively. Different MOFs can be constructed effectively, rapidly, and with high yield by using the MC method. Vital properties that affect the scalability of MW assisted technique include, heating parameters, dielectric properties, power efficiency, and power density.83 The major setback in the upscaling of this technology is the lack of data on the properties.
Pichon et al.84 were the first to employ mechanochemical free solvent to construct MOF. The reaction was carried out in the steel reactor in which Cu(OAc)2·H2O and isonicotinic acid were ground for 10 minutes with the aid of a steel ball to produce strongly crystalline and one phase result of [Cu(INA)2] with water molecules and acetic acids in pores. The removal of guest molecules can be detached by thermal activation to give the porous compound. The same approach was utilized to synthesize HKUST-1.85 Copper acetate and H3BTC were used as precursors and ball milled for 25 minutes to acquire HKUST-1 with a crystal size of 50 nm in size and a specific surface area of 1713 m2 g−1. Tanaka et al.86 synthesized ZIF-8 by employing mechanosynthesis and conventional processes to assess which method is greatly versatile and effective to attain ZIF-8 with hierarchical superstructure. The use of nanosized ZnO power and zinc acetate made it possible to accelerate the reaction and obtain the best results.
Morphological analysis revealed that rhombic dodecahedral crystals are obtained in the conventional solution (Fig. 10a). The addition of zinc acetate dehydrates in the mechanochemical method affected on the morphology of ZIF-8. Amorphous-looking agglomerates with clearly defined faces and edges can be seen. Zinc acetate was also added, which changed the morphology of the agglomerates from an integrated structure made up of isolated particles to one with dendritic porosity architecture (Fig. 10c). To create nano-sized MOFs, mechanochemical technology provides a quick, easy, and efficient approach to mill a mixture of metal salts and organic ligands. This technology is quicker with high yielding of the product as compared to conventional method. Mechanochemical method does not require additional steps, which are often employed in the solvent method for the removal of entrapped guest molecules that may result in the frame collapsed.
Fig. 10 Morphological analysis of ZIF-8 obtained via different approaches (a) conventional monocrystalline (b) mechanochemical using nanosized ZnO without zinc acetate dehydrate (c) with zinc acetate dehydrate. Reproduced from Tanaka et al.86 with permission from American Chemical Society, copyright 2018. |
The attractiveness of using sonochemical lies in the ability to synthesize MOFs under mild conditions such as at ambient conditions and solvent-free. Qiu et al.90 construct MOF (Zn3(BTC)2) by employing the sonochemical method. As precursors, 20% ethanol was combined with zinc acetate and H3BTC, which were then sonicated for up to 90 minutes. The product had a high yield of 75.3% after only 5 minutes of sonication. In Son et al.91 sonochemical's method was used to create MOF-5. In an inert environment, a solution of zinc nitrate and terephthalic acid in 1-methyl-2-pyrrolidone was combined and sonicated for 10 to 75 minutes. By using this method, the synthesis time was cut from 24 hours (using traditional heating) to 75 minutes. Characteristics of the structure of MOFs (namely the quality of the crystals, particle size, surface area, and pore size) synthesized by sonochemical technique are affected by process parameters such as temperature, solvent type and ratio, reagent concentration, and modulators. Israr et al.89 synthesized Cu-BTC (BTC = 1,3,5-benzene tricarboxylate) using sonochemical technology. They studied the effect of solvent mixtures of water–N,N-dimethylformamide (DMF), and water–ethanol with or without various bases under 20 kHz ultrasonically treated conditions on the morphology of the products (Fig. 11). The deprotonation that leads to crystal formation was promoted by the addition of base in the presence of H2O and ethanol (Fig. 11). Li et al.92 synthesized HKUST-1 by reacting cupric acetate and H3BTC in a mixed solution of DMF/EtOH/H2O under ultrasonic irradiation at ambient conditions for the reaction duration of 5–60 minutes which produced HKUST-1 with a yield of 62.6–85.1%. Among other benefit of using this technique compared to conventional solvothermal is that the obtained nano-crystals have a lower dimensional size in the range of 10–200 nm. Moreover, the sonochemical method for the synthesis of porous MOFs was found to be highly efficient and more environmentally friendly.
Fig. 11 SEM analysis of produce Cu-BTCs under various solvent mixture (a) H2O–DMF, (b) H2O–ethanol–DMF, (c) H2O–ethanol–NaOH, (d) H2O–ethanol–NH4OH, and (e) H2O–ethanol–pridine. Reproduced from Israr et al.89 with permission from Elsevier, copyright 2016. |
In the mixing step, precursors such as organic linkers and metal salts are mixed. The main hazards associated with raw materials, include toxicity, corrosion, and flammability. It is known that metal ion provides the central node for MOFs. These salts (nitrates, sulfates, or chlorides) are carriers of corrosive hazards in the process. In the synthesis of MOFs solvents are utilized e.g., N-methyl-2-pyrrolidone, strong acids, and bases. The hazards associated with the usage of such solvents include toxicity and explosion. Hazards of the synthesis step are related to the flammability and reactivity of the raw materials, being aggravated by process conditions such as operating temperature and pressure. Lee et al.95 indicated that during the mechanochemical synthesis of MOF such as ZIF-8, the use of ammonium nitrate imposes the risk of explosion. The filtration process is common in the synthesis of various MOFs. This process consists of several stages, which use filters and solvent treatment. The major identified hazard associated with this process and washing is the generation of chemical waste from the solvent used.
(2) |
The constant equilibrium of the chemical interaction between the adsorbate molecule and the empty side is KeqA, where ΘA is the fractional occupancy of the adsorption sites, pA is the partial pressure of the adsorbate. The Langmuir adsorption model has some restrictions when applied to MOFs, such as localized adsorption in ultra-micro-pores and multilayer adsorption in supermicropores. Thus, the BET adsorbed model is applicable in these situations. The BET adsorption model is represented by eqn (3).
(3) |
To date, some of the produced MOFs have incredibly high surface area. Chae et al.50 synthesized crystalline Zn4O(1,3,5-benzenetribenzoate)2 known as MOF-177 which is considered as one of the MOFs with a high specific surface area of 4500 m2 g−1 when evaluated at 77 K by nitrogen. This is significantly higher than that of activated carbon (2000 m2 g−1) and zeolite namely zeolite Y (904 m2 g−1). It is known through experimental and calculation results that a linear relationship occurs between the surface area and hydrogen storage capacity measured at 77 K. Preparation of MOF-5 samples, in which the products with different structures were obtained by exploiting the solvent method and drying using three different procedures. The dying conditions were as follows: sample M1 was dried at room temperature for 2 days, sample M2 was dried in the oven at 373 K for 5 h and sample M3 was dried in the vacuum for 2 h at 373 K. The obtained BET surface area of M1, M2 and M3 samples were measured to be 15.8, 143.2 and 516.9 m2 g−1 respectively. The hydrogen capacity of three types of MOFs was measured to be 0.84, 1.17 and 1.57 mmol g−1. The results indicated that different solvent methods have an effect on the structure of MOF-5 and its characteristics namely surface area and porosity. For meso and macro pores, sample M3's heat of adsorption was computed to be 3.68 and 12.45 kJ mol−1, appropriately. These suggest that the interaction between hydrogen molecules and MOF-5 surface is stronger in the macroporous region. Frost et al.22 used simulation on a set of IRMOFs with the same framework topology and surface chemistry to examine the impact of surface area, heat of adsorption, and free volume on hydrogen adsorption capacity. Fig. 11a shows the geometric structure of the assessed IRMOFs.
Fig. 12b presents the effect of surface area on hydrogen adsorption capacity evaluated at intermediate pressure (30 bar). It is evidence that an excellent correlation between surface area and hydrogen uptake is obtainable at the assessed condition. This means IRMOF with higher surface area will have high hydrogen storage capacity. According to Suh et al.21 at high pressure (10–90 bar) at 77 K, there is a favourable link between surface area and hydrogen capacity. At low pressure (1 atm), the hydrogen uptake is correlated with a surface area between 100 and 2000 m2 g−1, though. The interaction between the adsorbent and hydrogen molecules is said to be negatively impacted by a surface area of more than 2000 m2 g−1 at low pressure and 77 K due to the inability to occupy the free space.
Fig. 12 (a) IRMOF chemical structure, building block and geometry and (b) hydrogen adsorption capacity as a function of surface area at 30 bar. Reproduced from Frost et al.22 with permission from American Chemical Society, copyright 2006. |
The optimized synthesis process conditions resulted in obtaining a submicron average crystal size (Fig. 13a). Fig. 13b shows the synthesized samples exhibit a good correlation between nitrogen uptake and surface area. The usage of crystals with a bimodal size distribution or octahedral morphology, which may condense with few structural flaws and efficiently fill space, is a key to success, according to the authors. Saha et al.5 investigate the effects of various synthesis conditions on the performance of hydrogen adsorption, pore textural characteristics, and crystal structure. They demonstrate that MOF-5 materials with higher order crystallinity produce adsorbents with larger crystal sizes, homogeneous pore size distribution, higher specific surface area, greater hydrogen adsorption capacity, and faster hydrogen diffusion rate. The attractiveness of utilizing MOFs in hydrogen storage is due to the modifiable pore size, pore volume, and geometry. Zhang et al.44 examined several MOFs (NU-125, HKUST-1, UiO-68-Ant, NU-1000, Cu-MOF-74 and Zn2(bdc)2(dabco)2) with channel and cage structures to correlate the pore geometry with hydrogen total adsorption.
Fig. 13 (a) Optical micrographs of four different crystal sizes and SEM image for commercial MOF-5 and (b) Nitrogen sorption isotherms (adsorption data are shown with filled symbols while desorption data are shown with empty symbols). Reproduced from Suresh et al.52 with permission from American Chemical Society, copyright 2021. |
The results showed that cage-type MOFs have a higher pore occupancy than channel-type MOFs for the same pore volume (shown in Fig. 14). Yet, as the pore volume grows, the pore occupancy decreases as a result of the reduced interaction potential of adsorption at large pore centers in both scenarios. The total adsorption and pore volume are positively associated up to a pore volume of around 3.3 cm3 g−1. Hydrogen adsorption starts to decrease at pore volumes greater than 3.3 cm3 g−1 as a result of the decreasing pore occupancy at increasing pore volume regions. The authors generated empirical equations to predict total hydrogen adsorption under 100 bar and 77 k for cage-type MOFs as ntot = 0.085 × VP − 0.013 × VP2 and channel-type MOFs as ntot = 0.076 × VP − 0.011 × VP2, where VP denotes the pore volume of the corresponding MOFs. To examine their effects on pore features and hydrogen adsorption, Yang et al.96 produced materials of various qualities (less crystalline, strong crystalline, interwoven, and interwoven with included MWCNTs). The combination of modification and MWCNT insertion in MOF-5 resulted in increased hydrogen adsorption capacity at −196 °C and 1 bar (capacity increased from 1.2 to 2 wt%) and improved thermal stability (decomposition temperature raised from 438 to 510 °C).
Fig. 14 (a) Pore occupancy (100 bar and 77 K) versus pore volume and (b) predicted hydrogen adsorption versus pore volume (100 bar and 77 K) using empirical equations. Reproduced from Zhang et al.97 with permission from John Wiley and Sons, copyright 2020. |
Basically, because spherical pores have more exposed surface atoms and are in closer touch with the sorbate molecule than cylindrical pores,97 the sorbate–sorbent interaction potential in spherical pores is larger. According to Lin et al.30 as pressure rises, pore volume becomes more important and pore size has an impact on hydrogen adsorption.
Fig. 15 H2 adsorption on the Ni site: pictorial representation of two likely Ni2+ ****H2 adducts. The H atoms of the H2 molecule are represented as black sphere. Yellow areas correspond to a −0.015 au value of the electrostatic potential. Reproduced from Vitillo et al.101 with permission from American Chemical Society, copyright 2008. |
By removing the terminal water ligands from the MOF SNU-15 at 220 °C under vacuum, Chae et al.50 produced Co2+ ions containing an open site. It was observed that each Co2+ had an open metal site, which produced a significant isosteric heat of hydrogen adsorption at 0% coverage (15.1 kJ mol−1). According to Gedrich et al.102 positioning of the exposed sites with respect to the hydrogen could maximize hydrogen adsorption. Zeleňák and Saldan103 stated that a method of creating an open metal site by binding substrate without the removal of the ligand should be considered, which involves the alignment of the solvent molecules by the metal aggregates. Kapelewski et al.104 studied the volumetric hydrogen storage capacities of high-performing MOFs: M2 (m-dobdc) (M = Co, Ni; m-dobdc = 4,6-dioxido-1,3-benzenedicarboxylate) and the isomeric frameworks M2 (dobdc) (M = Co, Ni; dobdc = 1,4-dioxido-1,3-benzenedicarboxylate), consisting of an open metal cation site that strongly interacts with hydrogen molecules. The results show that Ni2 (m-dobdc) performs better than other materials with a volumetric capacity of 11 g L−1 and 23 g L−1 measured at the temperature range of −75 and 25 °C and pressure of 100 and 5 bar respectively. Exposed metal cation locations (polarizing Ni2+), which have a significant interaction with hydrogen molecules, were found to be responsible for these outcomes.
This journal is © The Royal Society of Chemistry 2024 |