Next Article in Journal
Bulk Process for Enrichment of Capsinoids from Capsicum Fruit
Previous Article in Journal
Multi-Indicators Decision for Product Design Solutions: A TOPSIS-MOGA Integrated Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress Using Solid-State Materials for Hydrogen Storage: A Short Review

1
Department of Chemistry, Inha University, 100 Inharo, Incheon 22212, Korea
2
Department of Environmental Medical Biology, Wonju College of Medicine, Yonsei University, Wonju 26426, Korea
*
Authors to whom correspondence should be addressed.
Processes 2022, 10(2), 304; https://doi.org/10.3390/pr10020304
Submission received: 4 January 2022 / Revised: 24 January 2022 / Accepted: 30 January 2022 / Published: 3 February 2022
(This article belongs to the Section Biological Processes and Systems)

Abstract

:
With the rapid growth in demand for effective and renewable energy, the hydrogen era has begun. To meet commercial requirements, efficient hydrogen storage techniques are required. So far, four techniques have been suggested for hydrogen storage: compressed storage, hydrogen liquefaction, chemical absorption, and physical adsorption. Currently, high-pressure compressed tanks are used in the industry; however, certain limitations such as high costs, safety concerns, undesirable amounts of occupied space, and low storage capacities are still challenges. Physical hydrogen adsorption is one of the most promising techniques; it uses porous adsorbents, which have material benefits such as low costs, high storage densities, and fast charging–discharging kinetics. During adsorption on material surfaces, hydrogen molecules weakly adsorb at the surface of adsorbents via long-range dispersion forces. The largest challenge in the hydrogen era is the development of progressive materials for efficient hydrogen storage. In designing efficient adsorbents, understanding interfacial interactions between hydrogen molecules and porous material surfaces is important. In this review, we briefly summarize a hydrogen storage technique based on US DOE classifications and examine hydrogen storage targets for feasible commercialization. We also address recent trends in the development of hydrogen storage materials. Lastly, we propose spillover mechanisms for efficient hydrogen storage using solid-state adsorbents.

1. Introduction

Recently, global warming issues surrounding growing populations and increases in the use of fossil fuels compel the replacement of fossil fuels for alternative resources [1,2,3]. Hydrogen is one of the most promising alternative energy sources because of its high energy efficiency, environmental friendliness, and non-toxicity [4,5,6].
The US DOE has announced annual technical targets that it requires to be met for the realistic adoption and expansion of a hydrogen-based society as shown Figure 1 [7,8,9,10]. From the latest study of the annual plan in 2017, We summarize in Table 1 certain important technical targets from the latest study of the annual plan in 2017. Hydrogen storage applications, storage capacity levels, durability/operability, chargingdischarging rate, dormancy, and safety are the primary measurements [11,12]. In the capacity section, targets are set as hydrogen uptake per absolute weight, volume, and cost, all of which are based on their conformity with commercialization. Ultimate targets aligned with the current storage system propose the capability of a minimum of 500 miles, which is the general distance capacity for the light-duty vehicle market in the USA. The resulting capacity target is 6.5 wt%, 2.2 kWh/kg, and 0.065 kg H2/kg system as the gravimetric capacity and 1.7 kWh/L and 0.050 kg H2/L system as the volumetric capacity. The cost set for the final target is 8 USD/kWh net by 3.5 USD/gal on gasoline (~1 kg H2/gge, gge: gasoline gallon equivalent). The target of durability/operability is aligned with the general operating conditions of delivering and storing for fuel cell vehicles. The aim for charging–discharging rates is set to 3–5 min, which is the general filling time for gasoline vehicles. Other parameters on charging–discharging rates, dormancy, and safety set targets based on typical vehicle systems. Among these multiple targets, hydrogen storage capacity is the most exigent factor that importantly affects hydrogen’s deliverability, fuel costs, and the performance of fuel cells.
This review provides a brief summary, with pros and cons, of the following practical hydrogen storage techniques: high-pressure gas storage, hydrogen liquefaction, chemical absorption, and physical adsorption. We address the recent progress of physical adsorption technologies using adsorbents that are divided into carbonaceous and non-carbonaceous materials. Furthermore, this review describes adsorption models for highly efficient hydrogen adsorption behaviors and specific characteristics of hydrogen molecules on spin isomers depending on their temperatures.

2. Hydrogen Storage Techniques

Figure 2 shows the feasible techniques for hydrogen storage, which are classified into four types: high-pressure gas storage, hydrogen liquefaction, chemical absorption, and physical adsorption [13,14,15]. Currently, a method involving physically compressed storage with high-pressure tanks is the only commercialized hydrogen storage technique for mobile systems. The four types of compressed tanks are classified by their cylinder’s materials: Type I comprises entirely metal, Type II comprises metal liner with hoop wrapping, Type III comprises metal liner with full composite wrapping, and Type IV comprises plastic liner with full composite wrapping [16,17]. Both steel and aluminum are used in conventional high-pressure hydrogen storage tanks; however, they are not sufficiently strong. Recently, the development of vessels made of composites, including carbon fiber and epoxy with high mechanical strength, is in progress [18,19,20,21]. However, the high cost and maintenance of high-pressure cylinders have been an impediment to expanding hydrogen fuel cell-based technology. This method requires a considerable amount of energy, and there is a risk of leaking hydrogen at the high pressure of 700 bar. Moreover, compressed hydrogen does not achieve the target value set by the US DOE in 2020 for a gravimetric density of 0.045 kg H2/kg system; the compressed method reached 0.042 kg H2/kg system at a pressure of 700 bar [22,23,24]. In the case of liquefied hydrogen, a large amount of hydrogen is stored in a confined volume compared to high-pressure compression-based storage [25,26]. Cryogenic hydrogen storage at 20 K requires a cooling system and additional energy compared to the high-pressure storage method [27,28,29]. This storage method has a remarkably high storage density compared to other methods. However, the energy cost required for liquefaction is high, and there are hydrogen gas losses because it vaporizes at room temperature. Moreover, it requires double-walled cylinders with good thermal insulation systems. A chemisorption technique is used for storing hydrogen gas as a form of metal hydrides in a solid-state [30,31,32]. This technique has been spotlighted due to its high storage density, remarkable stability, and small space occupancy [33,34,35,36]. A number of metal hydride vessels have been developed [37,38,39]. NaAlH4, LiAlH4, LiBH4, NaBH4, and AlH3 are used as metal hydrogen storage materials; in particular, Mg2NiH4 is in the limelight because of its advantages, such as its high storage capacity, low cost, and light weight [40]. However, the disadvantage is that high temperatures are required for desorption because of the lack of reversibility of hydrogen adsorption–desorption due to its strong bonding force and slow kinetics [41]. Physisorption storage is a process by which hydrogen molecules are weakly adsorbed at the material’s surface via London dispersive forces [42,43,44,45]. Physisorption technology is the most competitive in price since it stores hydrogen molecules at room temperature and at relatively low pressures. In general, materials used for physical adsorption are porous because they require a large specific surface area. This strategy’s advantages are its light weight, high storage density, superior reversibility and cycle stability, and fast charging–discharging speed. Compared to liquid storage at cryogenic temperatures, physisorption reduces hydrogen boil-off loss during storage and consumes relatively low amounts of energy during charging and discharging [23,46]. Porous materials such as zeolites, metal–organic frameworks (MOFs), covalent organic frameworks (COFs), and carbon materials (fullerenes, nanotubes, and graphene) are the most extensively examined materials [47].

3. Hydrogen Storage Materials for Physisorption Methods

Porous materials are mostly used for the physical adsorption of hydrogen because of their high specific surface area, porosity, fluid permeability, regularity, and uniform pore structure [9,48,49,50,51]. Materials such as ACs, porous silica, MOFs, COFs, and zeolites have various characteristics such as shape selectivity, adsorption, stability, and durability [5,52,53]. The high hydrogen storage capacity of the porous materials has been reported as per porosity and uniformity improvement [54,55]. Low energy is required for using hydrogen because of the physical adsorption and desorption of hydrogen molecules [56,57,58,59]. Therefore, to overcome the limitations of existing hydrogen storage methods and reach the target DOE value, physical adsorption storage using porous materials is important (Figure 3c).

3.1. Non-Carbonaceous Materials for Hydrogen Storage

Zeolites are representative silica-based molecular sieves with micropores for catalysts and adsorbents [62,63]. In particular, in addition to cage and channel structures that have high thermal stability and large ion-exchange capacity, they have considerable potential for the physisorption of non-polar gases [64,65]. The maximum hydrogen capacity was up to 2.07 wt% for the zeolite Na-LEV (Figure 3a). In a recent study, an ideal zeolite structure for hydrogen adsorption was estimated from the meta-learning of a zeolites database with Monte Carlo simulations [47]. It reports that the RWY-type zeolite represents a hydrogen uptake of 35 g L1 (around 7 wt%) at 100 bar and 77 K. The AWO-type zeolite had a hydrogen uptake of 10 g L1 (around 7 wt%) at 100 bar and 77 K and 35 g L1 (around 2 wt%) at 100 bar (Figure 4).
Metal–organic frameworks (MOFs), which have a microporous crystalline structure comprising metal ions or clusters, are connected via molecular bridges [66,67,68]. MOFs have good stability, high void volumes, well-defined tailorable cavities of uniform size, high surface areas, and adjustable pore sizes [69,70,71]. In particular, the design flexibility for tuning the porosity of MOFs has attracted attention for their usage as hydrogen adsorbents [72]. Chae et al. synthesized one of the high-surface-area MOFs, known as MOF-177, in 2004 [61]. MOF-177 has a specific surface area of 4500 m2 g−1 when measured at 77 K by N2. Generally, MOFs have high surface areas of >3000 m2 g−1. This is remarkably higher than that of disordered-structure carbons (2030 m2 g−1) and zeolites (904 m2 g−1). Ahmed et al. computationally screened 500,000 reported compounds and demonstrated the record holder of MOF adsorbents (Figure 5). NU-100 shows amounts of 14 and 2 wt% at 100 bar and 1 bar, respectively [73].
Zeolitic imidazolate frameworks (ZIFs) are a class of MOFs that has a similar topology to zeolites [74,75,76,77]. Zhan et al. synthesized the core–shell structure of ZIFs for Matryoshka-type ZIFs via a step-wise liquid-phase epitaxial growth to evaluate the spillover mechanism [78]. They demonstrated that hydrogen atoms migrate through the (100) facets of ZIFs. The dominant factors for the spillover effect are demonstrated also including temperature, hydrogen concentration, and pressure via multi-layer ZIF nanocubes as probes (Figure 6).
Covalent organic frameworks (COFs), a family of crystalline microporous materials that are free of metals and entirely comprise strong covalent bonds, have been considered promising candidates for hydrogen storage because of their high specific surface area and large pore volume [79,80,81]. However, COFs still suffer from high preparation costs, complex synthesis processes, and mechanical instability. Ramirez-Vidal et al. synthesized hyper-crosslinked polymers (HCPs) by employing the Friedel–Crafts reaction using carbazole, anthracene, dibenzothiophene, and benzene as precursors and dimethoxymethane as a crosslinker [82]. The synthesized HCPs (B1FeM2) have a specific surface area of 1137 m2 g−1 and a total pore volume of 0.87 cm3 g−1. The maximum hydrogen capacity was 2.1 wt% at 77 K and 40 bar. After hydrogen adsorption at 140 bar, an irreversible collapse of the texture of the HCPs was observed (Figure 7).

3.2. Carbonaceous Materials for Hydrogen Storage

Carbon-based nanomaterials have received considerable attention as potential hydrogen storage materials because of their low costs, low weights, high surface areas, high chemical stabilities, and wide diversities of bulk and pore structures [83,84,85,86,87,88,89]. In particular, many industrial applications are expected because of their moisture-resistant properties. There are many types of carbonaceous materials, including CNTs, fullerenes, graphites, graphene derivates, and activated carbons [90,91,92,93,94,95].
CNTs are carbon macromolecules that have narrow distributions of pore volumes for efficient gas adsorption [96,97,98,99]. Thus, CNTs were obtained in two different primary species characterized by their wall structures as single-walled nanotubes (SWNTs) or multi-walled nanotubes (MWNTs) [100,101,102,103,104]. In 1997, Dillon et al. examined the hydrogen adsorption of CNTs and concluded that they were promising materials for hydrogen storage (Figure 3b) [105]. Fullerene (C60) was predicted in 1991 when it was determined that the hydrogen molecule could be trapped in a C60 cage. Although it is usually unfavorable for hydrogen to be stored in such a space, the high energy barrier required to break the cage open stabilized the hydrogen molecule inside. When fullerene contained 1–29 hydrogen molecules, the C–C bond length increased by 9.3%. The bond was broken at a 14–15% change in length. Changes in the bond’s length caused the fullerenes to depart from their spherical shape [106].
Activated carbons (ACs) are another one of the promising candidates for hydrogen storage materials because of their high specific surface areas, chemical and mechanical stabilities, microporous structures, and low costs [107,108,109,110,111,112,113]. Zhou et al. demonstrated that the hydrogen storage capacity is better for slit-type pores (ACs) than cylinder-type pores (CNTs) [114]. Moreover, they have the characteristic that their pore size can be controlled via conditions of chemical and physical activation [115,116,117,118,119]. One advantage of ACs is that agricultural waste, such as coffee bean dregs, coconut husks, and rice husks, have been used as raw materials for ACs [120,121,122,123]. These characteristics help not only to reduce environmental pollution but also to enable the production of ACs with various properties using various raw materials. Moreover, the advantages of ACs that are the most promising for hydrogen storage materials are that they can be mass-produced, require relatively low costs, and are mostly suitable for commercial use because of their lack of vulnerability to moisture. In one study, microporous carbon derived from melamine and isophthalaldehyde had a 4.0 wt% hydrogen uptake at 77 K and 1 bar [124]. This work recommended one-pot condensation and activation in a process involving a molten salt medium, which helps lower the melting point and leads to the condensation of monomers at a desired polymerization temperature. Molten salt media was selected from two different reagents of eutectic mixture including KOH and NaOH. The prepared melamine- and isophthalaldehyde-based ACs via the KOH-NaOH reagents (MIKN) achieved a specific surface area of 2984 m2/g and a total pore volume of 1.98 cm3/g (Figure 8).
Graphene derivatives such as graphene oxide (GO) and reduced graphene oxide (rGO) are other promising materials for the hydrogen storage system [125,126,127]. Functional groups such as hydroxyls, carbonyls, and carboxyls trap hydrogen molecules via hydrogen bonds that are stronger than the London dispersive force [128]. Moreover, oxygen-containing functional groups not only form gaps between graphene layers but also creates suitable conditions for loading metal nanoparticles in order to create higher hydrogen capacities [10,129]. Graphitic carbons, which are sp2 hybridized and have a sheet-like structure, interact using van der Waals forces. Singh et al. synthesized exfoliated graphite oxide (EGO) with a thermal exfoliation method under various gas conditions [130]. Air-exfoliated EGO represents a specific surface area of 268 m2 g−1, average pore size of 2.9 nm, and total pore volumes of 1.2 cm3 g. The maximum hydrogen uptake of EGOs was 3.34 wt% at 77K and 30 bar. After five cycles, the hydrogen’s capacity decreased by only 9%, indicating a storage amount of 3.02 wt% (Figure 9).
Porous graphite hybridized with nickel (Ni/PG) has been reported as a potential hydrogen storage material [131,132]. Ni/PG provides a specific surface area of 145 m2/g and a total pore volume of 0.056 cm3/g. The optimized samples achieved 4.48 wt% of hydrogen uptake at 298 K and 100 bar (Figure 10).
Because of a large specific surface and high surface activity, MXenes have attracted attention for use as hydrogen storage adsorbents [133,134,135]. MXenes are 2D transition metal carbides or nitrides that have chemical formulations such as Mn+1XnTx (M is an early transition metal, X is a member of group 13 or 14 of the periodic table, and T is a functional group) [136]. MXenes presented the Kubas interaction during hydrogen adsorption, which provides binding energy between chemisorption and physisorption from transition metals in their structures [137,138,139]. Therefore, these unique properties provide superior hydrogen uptakes with adsorption–desorption reversibility. Recently, a partially etched MXene was suggested for high hydrogen uptake performance [140]. The bell-mouth structure from an incompletely etched multilayer provides a fast H2 transport channel for increases in hydrogen uptake. Moreover, additional weak chemisorption via a nanopump effect from a narrow interlayer structure of 7 Å helps to maximize the hydrogen capacity. This work achieved 8.8 wt% H2 uptake at 298 K and 60 bar (Figure 11). Additionally, it was found that the hydrogen storage capacitance of this MXene at the first cycle is 8.07 wt%. After five cycles, the storage amount was 7.56 wt% which represented a retention of 94%, compared to the initial uptake.

4. The Adsorption Models for Hydrogen Storage

Understanding the behaviors of hydrogen adsorption is important for improving hydrogen storage capacity. Moreover, the sorption density on the physisorption of carbonaceous materials is extremely low and has a low hydrogen capacity [141,142,143,144]. Therefore, a progressive strategy is required to create stronger attraction between hydrogen and adsorbents. It is well-known that hydrogen spillover helps obtain highly efficient hydrogen storage capacity by introducing a transition metal component to the surface of porous materials [145,146,147,148,149,150].
The phenomenon of hydrogen spillover can be explained as the dissociative chemisorption of hydrogen molecules to hydrogen atoms on the metal (primary spillover) and migration of atomic hydrogen on the surface of supports (secondary spillover) [151,152]. However, the spillover of hydrogen atoms directly to supports is energetically an unfavorable phenomenon [153]. Therefore, a bridge between metal particles and supports was proposed to allow the surface diffusion of hydrogen atoms from the metals to the supports [154,155]. In this strategy, the energy barriers for the surface diffusion of hydrogen atoms from the metal to the support were overcome via physical bridges with intimate contacts for secondary spillover. However, the secondary spillover from the proposed model is arguable because of a significant energy barrier for the surface migration of hydrogen atoms from metals to supports and the lack of a driving force.
In recently reported studies [131,156], a modified spillover model was proposed to redeem the gap between the real improvement of hydrogen capacities and expected hydrogen spillover effects. The hydrogen molecules adjacent to the adsorbents are classified via three layers in this model (Figure 12). The hydrogen molecules of the first layer dissociatively adsorbed on the metal surfaces via the Kubas interaction from the high oxidative power of the metal [157]. The second layer of hydrogen, physically adsorbed on an adsorbent, was partially dipole-induced because of the strong electron–acceptor characteristics of the metal. The other hydrogen molecules were close to the physically adsorbed adsorbent; however, the equilibrium of the third layer was broken by the attraction of partially acidic carbon surfaces. The carbon surface surrounding the metal lost its electrons (electron acceptor properties) because of the strong electron–acceptor characteristics of the metallic catalysts [158,159]. This environment acts as a driving force for hydrogen affinity properties on carbon surfaces via electron–acceptor–donor interaction behaviors.

5. Conclusions

Modern society is adopting the gradual change to a carbon-free world with the development of novel green energy storage and conversion technologies. Hydrogen has been considered a significant alternative for fossils fuels for a veritably sustainable society. To date, the development of a highly efficient and stable hydrogen storage technique is urgent. To summarize, the US DOE has annually set up detailed targets for hydrogen fuel cell applications. The current state of hydrogen storage techniques is challenging; what is used is a high-pressure compressed tank that has underlying problems because of safety concerns and a lack of energy efficiency. Researchers have examined promising techniques using physical adsorption materials with light weights, high storage densities, and fast charging–discharging kinetics. Regarding highly efficient hydrogen storage capacities, it is important to understand the interfacial interactions between adsorbent surfaces and adsorbates. Here, we briefly addressed the recent progress on adsorbent materials, which can be carbonaceous or non-carbonaceous materials. Furthermore, the spillover effects of hydrogen molecules on solid-state adsorbents are suggested to achieve highly efficient hydrogen storage, which could be an important key point for designing hydrogen adsorbents.

Author Contributions

Conceptualization, S.-Y.L., J.-H.L., Y.-H.K. and J.-W.K.; Writing—original draft preparation, S.-Y.L., J.-H.L., Y.-H.K. and J.-W.K.; Writing—review and editing, K.-J.L. and S.-J.P.; Supervision, K.-J.L. and S.-J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (or Industrial Strategic Technology Development Program) (20012373, Innovative technology for CO2 free hydrogen pro-duction using molten catalysts) funded By the Ministry of Trade, Industry and Energy (MOTIE, Korea). This work was also supported by Korea Evaluation institute of Industrial Technology (KEIT) through the Carbon Cluster Construction project [10083586, Development of petroleum based graphite fibers with ultra-high thermal conductivity] funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schlapbach, L.; Zuttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  2. Schoedel, A.; Ji, Z.; Yaghi, O.M. The role of metal-organic frameworks in a carbon-neutral energy cycle. Nat. Energy 2016, 1, 16034. [Google Scholar] [CrossRef]
  3. Schneemann, A.; White, J.L.; Kang, S.; Jeong, S.; Wan, L.W.F.; Cho, E.S.; Heo, T.W.; Prendergast, D.; Urban, J.J.; Wood, B.C.; et al. Nanostructured Metal Hydrides for Hydrogen Storage. Chem. Rev. 2018, 118, 10775–10839. [Google Scholar] [CrossRef]
  4. Staffell, I.; Scamman, D.; Abad, A.V.; Balcombe, P.; Dodds, P.E.; Ekins, P.; Shah, N.; Ward, K.R. The role of hydrogen and fuel cells in the global energy system. Energy Environ. Sci. 2019, 12, 463–491. [Google Scholar] [CrossRef] [Green Version]
  5. Allendorf, M.D.; Hulvey, Z.; Gennett, T.; Ahmed, A.; Autrey, T.; Camp, J.; Cho, E.S.; Furukawa, H.; Haranczyk, M.; Head-Gordon, M.; et al. An assessment of strategies for the development of solid-state adsorbents for vehicular hydrogen storage. Energy Environ. Sci. 2018, 11, 2784–2812. [Google Scholar] [CrossRef] [Green Version]
  6. Ueckerdt, F.; Bauer, C.; Dirnaichner, A.; Everall, J.; Sacchi, R.; Luderer, G. Potential and risks of hydrogen-based e-fuels in climate change mitigation. Nat. Clim. Chang. 2021, 11, 384–393. [Google Scholar] [CrossRef]
  7. Doe, U. Target explanation document: Onboard hydrogen storage for light-duty fuel cell vehicles. US Drive 2017, 1, 1–29. [Google Scholar]
  8. Morris, L.; Hales, J.J.; Trudeau, M.L.; Georgiev, P.; Embs, J.P.; Eckert, J.; Kaltsoyannis, N.; Antonelli, D.M. A manganese hydride molecular sieve for practical hydrogen storage under ambient conditions. Energy Environ. Sci. 2019, 12, 1580–1591. [Google Scholar] [CrossRef] [Green Version]
  9. Czarna-Juszkiewicz, D.; Cader, J.; Wdowin, M. From coal ashes to solid sorbents for hydrogen storage. J. Clean. Prod. 2020, 270, 122355. [Google Scholar] [CrossRef]
  10. Shiraz, H.G.; Tavakoli, O. Investigation of graphene-based systems for hydrogen storage. Renew. Sustain. Energy Rev. 2017, 74, 104–109. [Google Scholar] [CrossRef]
  11. Hassan, I.A.; Ramadan, H.S.; Saleh, M.A.; Hissel, D. Hydrogen storage technologies for stationary and mobile applications: Review, analysis and perspectives. Renew. Sustain. Energy Rev. 2021, 149, 111311. [Google Scholar] [CrossRef]
  12. Xiao, R.F.; Tian, G.; Hou, Y.; Chen, S.T.; Cheng, C.; Chen, L. Effects of cooling-recovery venting on the performance of cryo-compressed hydrogen storage for automotive applications. Appl. Energy 2020, 269, 115143. [Google Scholar] [CrossRef]
  13. Ramirez-Vidal, P.; Canevesi, R.L.S.; Sdanghi, G.; Schaefer, S.; Maranzana, G.; Celzard, A.; Fierro, V. A Step Forward in Understanding the Hydrogen Adsorption and Compression on Activated Carbons. ACS Appl. Mater. Interfaces 2021, 13, 12562–12574. [Google Scholar] [CrossRef] [PubMed]
  14. Zhou, L. Progress and problems in hydrogen storage methods. Renew. Sustain. Energy Rev. 2005, 9, 395–408. [Google Scholar] [CrossRef]
  15. Thornton, A.W.; Simon, C.M.; Kim, J.; Kwon, O.; Deeg, K.S.; Konstas, K.; Pas, S.J.; Hill, M.R.; Winkler, D.A.; Haranczyk, M.; et al. Materials Genome in Action: Identifying the Performance Limits of Physical Hydrogen Storage. Chem. Mater. 2017, 29, 2844–2854. [Google Scholar] [CrossRef]
  16. Li, M.X.; Bai, Y.F.; Zhang, C.Z.; Song, Y.X.; Jiang, S.F.; Grouset, D.; Zhang, M.J. Review on the research of hydrogen storage system fast refueling in fuel cell vehicle. Int. J. Hydrogen Energy 2019, 44, 10677–10693. [Google Scholar] [CrossRef] [Green Version]
  17. Su, Y.; Lv, H.; Zhou, W.; Zhang, C. Review of the Hydrogen Permeability of the Liner Material of Type IV On-Board Hydrogen Storage Tank. World Electr. Veh. J. 2021, 12, 130. [Google Scholar] [CrossRef]
  18. Benitez, A.; Wulf, C.; de Palmenaer, A.; Lengersdorf, M.; Roding, T.; Grube, T.; Robinius, M.; Stolten, D.; Kuckshinrichs, W. Ecological assessment of fuel cell electric vehicles with special focus on type IV carbon fiber hydrogen tank. J. Clean. Prod. 2021, 278, 123277. [Google Scholar] [CrossRef]
  19. Liu, H.R.; Wang, C.J.; Chen, B.; Zhang, Z. A further study of pyrolysis of carbon fibre-epoxy composite from hydrogen tank: Search optimization for kinetic parameters via a Shuffled Complex Evolution. J. Hazard. Mater. 2019, 374, 20–25. [Google Scholar] [CrossRef]
  20. Zhang, Z.; Wang, C.J.; Huang, G.; Liu, H.R.; Yang, S.L.; Zhang, A.F. Thermal degradation behaviors and reaction mechanism of carbon fibre-epoxy composite from hydrogen tank by TG-FTIR. J. Hazard. Mater. 2018, 357, 73–80. [Google Scholar] [CrossRef]
  21. Kim, W.J.; Heo, Y.J.; Lee, J.H.; Rhee, K.Y.; Park, S.J. Effect of Atmospheric-Pressure Plasma Treatments on Fracture Toughness of Carbon Fibers-Reinforced Composites. Molecules 2021, 26, 3698. [Google Scholar] [CrossRef] [PubMed]
  22. Hua, T.Q.; Roh, H.S.; Ahluwalia, R.K. Performance assessment of 700-bar compressed hydrogen storage for light duty fuel cell vehicles. Int. J. Hydrogen Energy 2017, 42, 25121–25129. [Google Scholar] [CrossRef]
  23. Rivard, E.; Trudeau, M.; Zaghib, K. Hydrogen Storage for Mobility: A Review. Materials 2019, 12, 1973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sdanghi, G.; Maranzana, G.; Celzard, A.; Fierro, V. Review of the current technologies and performances of hydrogen compression for stationary and automotive applications. Renew. Sustain. Energy Rev. 2019, 102, 150–170. [Google Scholar] [CrossRef]
  25. Reuss, M.; Grube, T.; Robinius, M.; Preuster, P.; Wasserscheid, P.; Stolten, D. Seasonal storage and alternative carriers: A flexible hydrogen supply chain model. Appl. Energy 2017, 200, 290–302. [Google Scholar] [CrossRef]
  26. Suh, M.P.; Park, H.J.; Prasad, T.K.; Lim, D.-W. Hydrogen storage in metal–organic frameworks. Chem. Rev. 2012, 112, 782–835. [Google Scholar] [CrossRef]
  27. Zheng, J.; Zhou, H.; Wang, C.G.; Ye, E.Y.; Xu, J.W.; Loh, X.J.; Li, Z.B. Current research progress and perspectives on liquid hydrogen rich molecules in sustainable hydrogen storage. Energy Storage Mater. 2021, 35, 695–722. [Google Scholar] [CrossRef]
  28. Zhao, Y.X.; Gong, M.Q.; Zhou, Y.; Dong, X.Q.; Shen, J. Thermodynamics analysis of hydrogen storage based on compressed gaseous hydrogen, liquid hydrogen and cryo-compressed hydrogen. Int. J. Hydrogen Energy 2019, 44, 16833–16840. [Google Scholar] [CrossRef]
  29. Wei, G.M.; Zhang, J.F. Numerical Study of the Filling Process of a Liquid Hydrogen Storage Tank under Different Sloshing Conditions. Processes 2020, 8, 1020. [Google Scholar] [CrossRef]
  30. Bai, X.S.; Yang, W.W.; Tang, X.Y.; Yang, F.S.; Jiao, Y.H.; Yang, Y. Optimization of tree-shaped fin structures towards enhanced absorption performance of metal hydride hydrogen storage device: A numerical study. Energy 2021, 220, 119738. [Google Scholar] [CrossRef]
  31. Gao, S.C.; Liu, H.Z.; Xu, L.; Li, S.Q.; Wang, X.H.; Yan, M. Hydrogen storage properties of nano-CoB/CNTs catalyzed MgH2. J. Alloy. Compd. 2018, 735, 635–642. [Google Scholar] [CrossRef]
  32. Lotoskyy, M.; Denys, R.; Yartys, V.A.; Eriksen, J.; Goh, J.; Nyamsi, S.N.; Sita, C.; Cummings, F. An outstanding effect of graphite in nano-MgH2-TiH2 on hydrogen storage performance. J. Mater. Chem. A 2018, 6, 10740–10754. [Google Scholar] [CrossRef]
  33. Cho, E.S.; Ruminski, A.M.; Liu, Y.S.; Shea, P.T.; Kang, S.Y.; Zaia, E.W.; Park, J.Y.; Chuang, Y.D.; Yuk, J.M.; Zhou, X.W.; et al. Hierarchically Controlled Inside-Out Doping of Mg Nanocomposites for Moderate Temperature Hydrogen Storage. Adv. Funct. Mater. 2017, 27, 1704316. [Google Scholar] [CrossRef]
  34. Zhang, J.G.; Zhu, Y.F.; Lin, H.J.; Liu, Y.N.; Zhang, Y.; Li, S.Y.; Ma, Z.L.; Li, L.Q. Metal Hydride Nanoparticles with Ultrahigh Structural Stability and Hydrogen Storage Activity Derived from Microencapsulated Nanoconfinement. Adv. Mater. 2017, 29, 1700760. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Q.; Luo, Q.; Gu, Q.F. Insights into the composition exploration of novel hydrogen storage alloys: Evaluation of the Mg-Ni-Nd-H phase diagram. J. Mater. Chem. A 2017, 5, 3848–3864. [Google Scholar] [CrossRef]
  36. Liu, Y.S.; Wang, S.; Li, Z.L.; Gao, M.X.; Liu, Y.F.; Sun, W.P.; Pan, H.G. Enhanced Hydrogen Storage Performance of MgH2 by the Catalysis of a Novel Intersected Y2O3/NiO Hybrid. Processes 2021, 9, 892. [Google Scholar] [CrossRef]
  37. Nguyen, H.Q.; Shabani, B. Review of metal hydride hydrogen storage thermal management for use in the fuel cell systems. Int. J. Hydrogen Energy 2021, 46, 31699–31726. [Google Scholar] [CrossRef]
  38. Hang, Z.M.; Hu, Z.C.; Xiao, X.Z.; Jiang, R.C.; Zhang, M. Enhancing Hydrogen Storage Kinetics and Cycling Properties of NaMgH3 by 2D Transition Metal Carbide MXene Ti3C2. Processes 2021, 9, 1690. [Google Scholar] [CrossRef]
  39. Lin, X.; Xie, W.; Zhu, Q.; Yang, H.G.; Li, Q. Rational optimization of metal hydride tank with LaNi4.25Al0.75 as hydrogen storage medium. Chem. Eng. J. 2021, 421, 127844. [Google Scholar] [CrossRef]
  40. Rizo-Acosta, P.; Cuevas, F.; Latroche, M. Hydrides of early transition metals as catalysts and grain growth inhibitors for enhanced reversible hydrogen storage in nanostructured magnesium. J. Mater. Chem. A 2019, 7, 23064–23075. [Google Scholar] [CrossRef]
  41. Wang, H.; Wu, G.; Cao, H.; Pistidda, C.; Chaudhary, A.L.; Garroni, S.; Dornheim, M.; Chen, P. Near ambient condition hydrogen storage in a synergized tricomponent hydride system. Adv. Energy Mater. 2017, 7, 1602456. [Google Scholar] [CrossRef]
  42. Zhang, X.; Lin, R.B.; Wang, J.; Wang, B.; Liang, B.; Yildirim, T.; Zhang, J.; Zhou, W.; Chen, B.L. Optimization of the Pore Structures of MOFs for Record High Hydrogen Volumetric Working Capacity. Adv. Mater. 2020, 32, 1907995. [Google Scholar] [CrossRef] [PubMed]
  43. Eberle, U.; Felderhoff, M.; Schuth, F. Chemical and Physical Solutions for Hydrogen Storage. Angew. Chem. Int. Edit 2009, 48, 6608–6630. [Google Scholar] [CrossRef] [PubMed]
  44. Suresh, K.; Aulakh, D.; Purewal, J.; Siegel, D.J.; Veenstra, M.; Matzger, A.J. Optimizing Hydrogen Storage in MOFs through Engineering of Crystal Morphology and Control of Crystal Size. J. Am. Chem. Soc. 2021, 143, 10727–10734. [Google Scholar] [CrossRef] [PubMed]
  45. Heo, Y.J.; Yeon, S.H.; Park, S.J. Defining contribution of micropore size to hydrogen physisorption behaviors: A new approach based on DFT pore volumes. Carbon 2019, 143, 288–293. [Google Scholar] [CrossRef]
  46. Zuo, Z.Q.; Sun, P.J.; Jiang, W.B.; Qin, X.J.; Li, P.; Huang, Y.H. Thermal stratification suppression in reduced or zero boil-off hydrogen tank by self-spinning spray bar. Int. J. Hydrogen Energy 2019, 44, 20158–20172. [Google Scholar] [CrossRef]
  47. Sun, Y.Z.S.; DeJaco, R.F.; Li, Z.; Tang, D.; Glante, S.; Sholl, D.S.; Colina, C.M.; Snurr, R.Q.; Thommes, M.; Hartmann, M.; et al. Fingerprinting diverse nanoporous materials for optimal hydrogen storage conditions using meta-learning. Sci. Adv. 2021, 7, eabg3983. [Google Scholar] [CrossRef]
  48. Weng, Q.H.; Zeng, L.L.; Chen, Z.W.; Han, Y.X.; Jiang, K.; Bando, Y.; Golberg, D. Hydrogen Storage in Carbon and Oxygen Co-Doped Porous Boron Nitrides. Adv. Funct. Mater. 2021, 31, 2007381. [Google Scholar] [CrossRef]
  49. Balderas-Xicohtencatl, R.; Schmieder, P.; Denysenko, D.; Volkmer, D.; Hirscher, M. High Volumetric Hydrogen Storage Capacity using Interpenetrated Metal-Organic Frameworks. Energy Technol.-Ger. 2018, 6, 510–512. [Google Scholar] [CrossRef] [Green Version]
  50. Rozzi, E.; Minuto, F.D.; Lanzini, A. Dynamic modeling and thermal management of a Power-to-Power system with hydrogen storage in microporous adsorbent materials. J. Energy Storage 2021, 41, 102953. [Google Scholar] [CrossRef]
  51. Wang, J.; Chen, Y.H.; Yuan, L.H.; Zhang, M.L.; Zhang, C.R. Scandium Decoration of Boron Doped Porous Graphene for High-Capacity Hydrogen Storage. Molecules 2019, 24, 2382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Li, G.Q.; Kobayashi, H.; Taylor, J.M.; Ikeda, R.; Kubota, Y.; Kato, K.; Takata, M.; Yamamoto, T.; Toh, S.; Matsumura, S.; et al. Hydrogen storage in Pd nanocrystals covered with a metal-organic framework. Nat. Mater. 2014, 13, 802–806. [Google Scholar] [CrossRef] [PubMed]
  53. Reardon, H.; Hanlon, J.M.; Hughes, R.W.; Godula-Jopek, A.; Mandal, T.K.; Gregory, D.H. Emerging concepts in solid-state hydrogen storage: The role of nanomaterials design. Energy Environ. Sci. 2012, 5, 5951–5979. [Google Scholar] [CrossRef]
  54. Lee, S.Y.; Park, S.J. Preparation and characterization of ordered porous carbons for increasing hydrogen storage behaviors. J. Solid State Chem. 2011, 184, 2655–2660. [Google Scholar] [CrossRef]
  55. Lee, S.Y.; Park, S.J. Influence of oxygen-functional groups on carbon replicas for hydrogen adsorption. Phys Status Solidi A 2012, 209, 694–697. [Google Scholar] [CrossRef]
  56. Liang, X.Y.; Ng, S.P.; Ding, N.; Wu, C.M.L. Strain-induced switch for hydrogen storage in cobalt-decorated nitrogen-doped graphene. Appl. Surf. Sci. 2019, 473, 174–181. [Google Scholar] [CrossRef]
  57. Kumar, S.; Kumar, T.J.D. Electronic Structure Calculations of Hydrogen Storage in Lithium Decorated Metal-Graphyne Framework. ACS Appl. Mater. Interfaces 2017, 9, 28659–28666. [Google Scholar] [CrossRef]
  58. Makowski, P.; Thomas, A.; Kuhn, P.; Goettmann, F. Organic materials for hydrogen storage applications: From physisorption on organic solids to chemisorption in organic molecules. Energy Environ. Sci. 2009, 2, 480–490. [Google Scholar] [CrossRef]
  59. Guo, C.X.; Wang, Y.; Li, C.M. Hierarchical Graphene-Based Material for Over 4.0 Wt % Physisorption Hydrogen Storage Capacity. ACS Sustain. Chem. Eng. 2013, 1, 14–18. [Google Scholar] [CrossRef]
  60. Dong, J.X.; Wang, X.Y.; Xu, H.; Zhao, Q.; Li, J.P. Hydrogen storage in several microporous zeolites. Int. J. Hydrogen Energy 2007, 32, 4998–5004. [Google Scholar] [CrossRef]
  61. Chae, H.K.; Siberio-Perez, D.Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A.J.; O’Keeffe, M.; Yaghi, O.M. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427, 523–527. [Google Scholar] [CrossRef] [PubMed]
  62. Langmi, H.W.; Book, D.; Walton, A.; Johnson, S.R.; Al-Mamouri, M.M.; Speight, J.D.; Edwards, P.P.; Harris, I.R.; Anderson, P.A. Hydrogen storage in ion-exchanged zeolites. J. Alloy. Compd. 2005, 404, 637–642. [Google Scholar] [CrossRef]
  63. Forster, P.M.; Eckert, J.; Chang, J.S.; Park, S.E.; Ferey, G.; Cheetham, A.K. Hydrogen adsorption in nanoporous Nickel(II) phosphates. J. Am. Chem. Soc. 2003, 125, 1309–1312. [Google Scholar] [CrossRef] [PubMed]
  64. Arean, C.O.; Palomino, G.T.; Carayol, M.R.L. Variable temperature FT-IR studies on hydrogen adsorption on the zeolite (Mg,Na)-Y. Appl. Surf. Sci. 2007, 253, 5701–5704. [Google Scholar] [CrossRef]
  65. Lee, S.Y.; Park, S.J. Synthesis of zeolite-casted microporous carbons and their hydrogen storage capacity. J. Colloid Interface Sci. 2012, 384, 116–120. [Google Scholar] [CrossRef]
  66. Kapelewski, M.T.; Runcevski, T.; Tarver, J.D.; Jiang, H.Z.H.; Hurst, K.E.; Parilla, P.A.; Ayala, A.; Gennett, T.; FitzGerald, S.A.; Brown, C.M.; et al. Record High Hydrogen Storage Capacity in the Metal-Organic Framework Ni-2(m-dobdc) at Near-Ambient Temperatures. Chem. Mater. 2018, 30, 8179–8189. [Google Scholar] [CrossRef]
  67. Rostami, S.; Pour, A.N.; Salimi, A.; Abolghasempour, A. Hydrogen adsorption in metal- organic frameworks (MOFs): Effects of adsorbent architecture. Int. J. Hydrogen Energy 2018, 43, 7072–7080. [Google Scholar] [CrossRef]
  68. Lee, S.Y.; Park, S.J. Effect of platinum doping of activated carbon on hydrogen storage behaviors of metal-organic frameworks-5. Int. J. Hydrogen Energy 2011, 36, 8381–8387. [Google Scholar] [CrossRef]
  69. Daglar, H.; Gulbalkan, H.C.; Avci, G.; Aksu, G.O.; Altundal, O.F.; Altintas, C.; Erucar, I.; Keskin, S. Effect of Metal-Organic Framework (MOF) Database Selection on the Assessment of Gas Storage and Separation Potentials of MOFs. Angew. Chem. Int. Edit 2021, 60, 7828–7837. [Google Scholar] [CrossRef]
  70. Camp, J.; Stavila, V.; Allendorf, M.D.; Prendergast, D.; Haranczyk, M. Critical Factors in Computational Characterization of Hydrogen Storage in Metal-Organic Frameworks. J. Phys. Chem. C 2018, 122, 18957–18967. [Google Scholar] [CrossRef]
  71. Jaramillo, D.E.; Jiang, H.Z.H.; Evans, H.A.; Chakraborty, R.; Furukawa, H.; Brown, C.M.; Head-Gordon, M.; Long, J.R. Ambient-Temperature Hydrogen Storage via Vanadium(II)-Dihydrogen Complexation in a Metal-Organic Framework. J. Am. Chem. Soc. 2021, 143, 6248–6256. [Google Scholar] [CrossRef] [PubMed]
  72. Witman, M.; Ling, S.L.; Gladysiak, A.; Stylianou, K.C.; Smit, B.; Slater, B.; Haranczyk, M. Rational Design of a Low-Cost, High-Performance Metal-Organic Framework for Hydrogen Storage and Carbon Capture. J. Phys. Chem. C 2017, 121, 1171–1181. [Google Scholar] [CrossRef] [PubMed]
  73. Ahmed, A.; Seth, S.; Purewal, J.; Wong-Foy, A.G.; Veenstra, M.; Matzger, A.J.; Siegel, D.J. Exceptional hydrogen storage achieved by screening nearly half a million metal-organic frameworks. Nat. Commun. 2019, 10, 1568. [Google Scholar] [CrossRef] [Green Version]
  74. Klein, R.A.; Shulda, S.; Parilla, P.A.; Le Magueres, P.; Richardson, R.K.; Morris, W.; Brown, C.M.; McGuirk, C.M. Structural resolution and mechanistic insight into hydrogen adsorption in flexible ZIF-7. Chem. Sci. 2021, 12, 15620–15631. [Google Scholar] [CrossRef] [PubMed]
  75. Panchariya, D.K.; Rai, R.K.; Kumar, E.A.; Singh, S.K. Core-Shell Zeolitic Imidazolate Frameworks for Enhanced Hydrogen Storage. ACS Omega 2018, 3, 167–175. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, J.; Ji, D.; Zhou, H.; Yan, X.F.; Yuan, A.H. Nickel-Platinum Nanoparticles Supported on Zeolitic Imidazolate Framework/Graphene Oxide as High-Performance Adsorbents for Ambient-Temperature Hydrogen Storage. J. Nanosci. Nanotechnol. 2017, 17, 1400–1406. [Google Scholar] [CrossRef]
  77. Ethiraj, J.; Palla, S.; Reinsch, H. Insights into high pressure gas adsorption properties of ZIF-67: Experimental and theoretical studies. Microporous Mesoporous Mater. 2020, 294, 109867. [Google Scholar] [CrossRef]
  78. Zhan, G.W.; Zeng, H.C. Hydrogen spillover through Matryoshka-type (ZIFs@)(n-1)ZIFs nanocubes. Nat. Commun. 2018, 9, 3778. [Google Scholar] [CrossRef] [Green Version]
  79. Ke, Z.P.; Cheng, Y.Y.; Yang, S.Y.; Li, F.; Ding, L.F. Modification of COF-108 via impregnation/functionalization and Li-doping for hydrogen storage at ambient temperature. Int. J. Hydrogen Energy 2017, 42, 11461–11468. [Google Scholar] [CrossRef]
  80. Tong, M.M.; Zhu, W.C.; Li, J.; Long, Z.Y.; Zhao, S.; Chen, G.J.; Lan, Y.S. An easy way to identify high performing covalent organic frameworks for hydrogen storage. Chem. Commun. 2020, 56, 6376–6379. [Google Scholar] [CrossRef]
  81. Braunecker, W.A.; Shulda, S.; Martinez, M.B.; Hurst, K.E.; Koubek, J.T.; Zaccarine, S.; Mow, R.E.; Pylypenko, S.; Sellinger, A.; Gennett, T.; et al. Thermal Activation of a Copper-Loaded Covalent Organic Framework for Near-Ambient Temperature Hydrogen Storage and Delivery. ACS Mater. Lett. 2020, 2, 227–232. [Google Scholar] [CrossRef]
  82. Ramirez-Vidal, P.; Suarez-Garcia, F.; Canevesi, R.L.S.; Castro-Muniz, A.; Gadonneix, P.; Paredes, J.I.; Celzard, A.; Fierro, V. Irreversible deformation of hyper-crosslinked polymers after hydrogen adsorption. J. Colloid Interface Sci. 2022, 605, 513–527. [Google Scholar] [CrossRef]
  83. Zhang, H.Y.; Zhu, Y.W.; Liu, Q.Y.; Li, X.W. Preparation of porous carbon materials from biomass pyrolysis vapors for hydrogen storage. Appl. Energy 2022, 306, 118131. [Google Scholar] [CrossRef]
  84. Naheed, L.; Koppel, M.; Paalo, M.; Alsabawi, K.; Lamb, K.E.; Gray, E.M.; Janes, A.; Webb, C.J. Hydrogen adsorption properties of carbide-derived carbons at ambient temperature and high pressure. Int. J. Hydrogen Energy 2021, 46, 15761–15772. [Google Scholar] [CrossRef]
  85. Lee, J.H.; Park, S.J. Recent advances in preparations and applications of carbon aerogels: A review. Carbon 2020, 163, 1–18. [Google Scholar] [CrossRef]
  86. Choi, Y.K.; Park, S.J. Hydrogen storage capacity of highly porous carbons synthesized from biomass-derived aerogels. Carbon Lett. 2015, 16, 127–131. [Google Scholar] [CrossRef] [Green Version]
  87. Choi, Y.K.; Park, S.J. Preparation and characterization of sucrose-based microporous carbons for increasing hydrogen storage. J. Ind. Eng. Chem. 2015, 28, 32–36. [Google Scholar] [CrossRef]
  88. Im, J.S.; Kwon, O.; Kim, Y.H.; Park, S.J.; Lee, Y.S. The effect of embedded vanadium catalyst on activated electrospun CFs for hydrogen storage. Microporous Mesoporous Mater. 2008, 115, 514–521. [Google Scholar] [CrossRef]
  89. Jung, M.J.; Kim, J.W.; Im, J.S.; Park, S.J.; Lee, Y.S. Nitrogen and hydrogen adsorption of activated carbon fibers modified by fluorination. J. Ind. Eng. Chem. 2009, 15, 410–414. [Google Scholar] [CrossRef]
  90. Ustinov, E.A.; Gavrilov, V.Y.; Mel’gunov, S.; Sokolov, V.V.; Berveno, V.P.; Berveno, A.V. Characterization of activated carbons with low-temperature hydrogen adsorption. Carbon 2017, 121, 563–573. [Google Scholar] [CrossRef]
  91. Gaboardi, M.; Amade, N.S.; Aramini, M.; Milanese, C.; Magnani, G.; Sanna, S.; Ricco, M.; Pontiroli, D. Extending the hydrogen storage limit in fullerene. Carbon 2017, 120, 77–82. [Google Scholar] [CrossRef] [Green Version]
  92. Alhameedi, K.; Hussain, T.; Bae, H.; Jayatilaka, D.; Lee, H.; Karton, A. Reversible hydrogen storage properties of defect-engineered C4N nanosheets under ambient conditions. Carbon 2019, 152, 344–353. [Google Scholar] [CrossRef]
  93. Kim, B.J.; Bae, K.M.; An, K.H.; Park, S.J. Effects of Mechanical Ball Milling with Active Gases on Hydrogen Adsorption Behaviors of Graphite Flakes. J. Nanosci. Nanotechnol. 2012, 12, 5713–5718. [Google Scholar] [CrossRef] [PubMed]
  94. Kim, C.K.; Park, B.H.; Park, S.J.; Kim, C.K. Modeling studies on the uptake of hydrogen molecules by graphene. J. Mol. Modeling 2015, 21, 240. [Google Scholar] [CrossRef]
  95. Lee, H.I.; Kim, W.J.; Heo, Y.J.; Son, Y.R.; Park, S.J. Control of interlayer spacing of expanded graphite for improved hydrogen storage capacity. Carbon Lett. 2018, 27, 117–120. [Google Scholar] [CrossRef]
  96. Kaskun, S.; Akinay, Y.; Kayfeci, M. Improved hydrogen adsorption of ZnO doped multi-walled carbon nanotubes. Int. J. Hydrogen Energy 2020, 45, 34949–34955. [Google Scholar] [CrossRef]
  97. Wang, Y.X.; Hu, X.D.; Guo, T.; Tian, W.G.; Hao, J.; Guo, Q.J. The competitive adsorption mechanism of CO2, H2O and O2 on a solid amine adsorbent. Chem. Eng. J. 2021, 416, 129007. [Google Scholar] [CrossRef]
  98. Lee, J.H.; Rhee, K.Y.; Park, S.J. Effects of cryomilling on the structures and hydrogen storage characteristics of multi-walled carbon nanotubes. Int. J. Hydrogen Energy 2010, 35, 7850–7857. [Google Scholar] [CrossRef]
  99. Lee, S.Y.; Park, S.J. Effect of temperature on activated carbon nanotubes for hydrogen storage behaviors. Int. J. Hydrogen Energy 2010, 35, 6757–6762. [Google Scholar] [CrossRef]
  100. Lee, S.Y.; Rhee, K.Y.; Nahm, S.H.; Park, S.J. Effect of p-type multi-walled carbon nanotubes for improving hydrogen storage behaviors. J. Solid State Chem. 2014, 210, 256–260. [Google Scholar] [CrossRef]
  101. Han, Y.J.; Park, S.J. Influence of nickel nanoparticles on hydrogen storage behaviors of MWCNTs. Appl. Surf. Sci. 2017, 415, 85–89. [Google Scholar] [CrossRef]
  102. Lee, S.Y.; Park, S.J. Effect of chemical treatments on hydrogen storage behaviors of multi-walled carbon nanotubes. Mater. Chem. Phys. 2010, 124, 1011–1014. [Google Scholar] [CrossRef]
  103. Lee, S.Y.; Park, S.J. Hydrogen Adsorption of Acid-treated Multi-walled Carbon Nanotubes at Low Temperature. Bull. Korean Chem. Soc. 2010, 31, 1596–1600. [Google Scholar] [CrossRef] [Green Version]
  104. Lee, S.Y.; Park, S.J. Influence of the pore size in multi-walled carbon nanotubes on the hydrogen storage behaviors. J. Solid State Chem. 2012, 194, 307–312. [Google Scholar] [CrossRef]
  105. Dillon, A.C.; Jones, K.M.; Bekkedahl, T.A.; Kiang, C.H.; Bethune, D.S.; Heben, M.J. Storage of hydrogen in single-walled carbon nanotubes. Nature 1997, 386, 377–379. [Google Scholar] [CrossRef]
  106. Pupysheva, O.V.; Farajian, A.A.; Yakobson, B.I. Fullerene nanocage capacity for hydrogen storage. Nano Lett. 2008, 8, 767–774. [Google Scholar] [CrossRef]
  107. Blankenship, T.S.; Blankenship, T.S.; Balahmar, N.; Mokaya, R. Oxygen-rich microporous carbons with exceptional hydrogen storage capacity. Nat. Commun. 2017, 8, 1545. [Google Scholar] [CrossRef] [Green Version]
  108. Sdanghi, G.; Nicolas, V.; Mozet, K.; Schaefer, S.; Maranzana, G.; Celzard, A.; Fierro, V. A 70 MPa hydrogen thermally driven compressor based on cyclic adsorption-desorption on activated carbon. Carbon 2020, 161, 466–478. [Google Scholar] [CrossRef]
  109. Kim, C.H.; Lee, S.Y.; Park, S.J. Efficient micropore sizes for carbon dioxide physisorption of pine cone-based carbonaceous materials at different temperatures. J. CO2 Util. 2021, 54, 101770. [Google Scholar] [CrossRef]
  110. Han, Y.J.; Park, S.J. Effect of nickel on hydrogen storage behaviors of carbon aerogel hybrid. Carbon Lett. 2015, 16, 281–285. [Google Scholar] [CrossRef] [Green Version]
  111. Han, Y.J.; Park, S.J. Hydrogen Storage Behaviors of Porous Carbons Derived from Poly(vinylidene fluoride). J. Nanosci. Nanotechnol. 2017, 17, 8075–8080. [Google Scholar] [CrossRef]
  112. Kim, B.J.; An, K.H.; Park, S.J. Preparation and Characterization of Highly Porous Carbons for Hydrogen Storage. J. Nanosci. Nanotechnol. 2011, 11, 860–864. [Google Scholar] [CrossRef] [PubMed]
  113. Kim, B.J.; Park, S.J. Influence of surface treatments on micropore structure and hydrogen adsorption behavior of nanoporous carbons. J. Colloid Interface Sci. 2007, 311, 619–621. [Google Scholar] [CrossRef] [PubMed]
  114. Zhou, L.; Zhou, Y.P.; Sun, Y. A comparative study of hydrogen adsorption on superactivated carbon versus carbon nanotubes. Int. J. Hydrogen Energy 2004, 29, 475–479. [Google Scholar] [CrossRef]
  115. Im, J.S.; Park, S.J.; Kim, T.J.; Kim, Y.H.; Lee, Y.S. The study of controlling pore size on electrospun carbon nanofibers for hydrogen adsorption. J. Colloid Interface Sci. 2008, 318, 42–49. [Google Scholar] [CrossRef]
  116. Im, J.S.; Park, S.J.; Lee, Y.S. Superior prospect of chemically activated electrospun carbon fibers for hydrogen storage. Mater. Res. Bull. 2009, 44, 1871–1878. [Google Scholar] [CrossRef]
  117. Im, J.S.; Park, S.J.; Lee, Y.S. The metal-carbon-fluorine system for improving hydrogen storage by using metal and fluorine with different levels of electronegativity. Int. J. Hydrogen Energy 2009, 34, 1423–1428. [Google Scholar] [CrossRef]
  118. Lee, H.M.; Heo, Y.J.; An, K.H.; Jung, S.C.; Chung, D.C.; Park, S.J.; Kim, B.J. A study on optimal pore range for high pressure hydrogen storage behaviors by porous hard carbon materials prepared from a polymeric precursor. Int. J. Hydrogen Energy 2018, 43, 5894–5902. [Google Scholar] [CrossRef]
  119. Park, J.H.; Park, S.J. Expansion of effective pore size on hydrogen physisorption of porous carbons at low temperatures with high pressures. Carbon 2020, 158, 364–371. [Google Scholar] [CrossRef]
  120. Sevilla, M.; Mokaya, R. Energy storage applications of activated carbons: Supercapacitors and hydrogen storage. Energy Environ. Sci. 2014, 7, 1250–1280. [Google Scholar] [CrossRef] [Green Version]
  121. Bastos-Neto, M.; Patzschke, C.; Lange, M.; Mollmer, J.; Moller, A.; Fichtner, S.; Schrage, C.; Lassig, D.; Lincke, J.; Staudt, R.; et al. Assessment of hydrogen storage by physisorption in porous materials. Energy Environ. Sci. 2012, 5, 8294–8303. [Google Scholar] [CrossRef]
  122. Heo, Y.J.; Park, S.J. Synthesis of activated carbon derived from rice husks for improving hydrogen storage capacity. J. Ind. Eng. Chem. 2015, 31, 330–334. [Google Scholar] [CrossRef]
  123. Lee, J.H.; Park, S.J. Potassium Oxalate as an Alternative Activating Reagent of Corn Starch-Derived Porous Carbons for Methane Storage. J. Nanosci. Nanotechnol. 2020, 20, 7124–7129. [Google Scholar] [CrossRef] [PubMed]
  124. Rehman, A.; Park, S.J. Microporous carbons derived from melamine and isophthalaldehyde: One-pot condensation and activation in a molten salt medium for efficient gas adsorption. Sci. Rep. 2018, 8, 6092. [Google Scholar] [CrossRef] [Green Version]
  125. Koh, J.; Choi, E.; Sakaki, K.; Kim, D.; Han, S.M.; Kim, S.; Cho, E.S. Uncovering the encapsulation effect of reduced graphene oxide sheets on the hydrogen storage properties of palladium nanocubes. Nanoscale 2021, 13, 16942–16951. [Google Scholar] [CrossRef]
  126. Dhar, P.; Gaur, S.S.; Kumar, A.; Katiyar, V. Cellulose Nanocrystal Templated Graphene Nanoscrolls for High Performance Supercapacitors and Hydrogen Storage: An Experimental and Molecular Simulation Study. Sci. Rep. 2018, 8, 3886. [Google Scholar] [CrossRef] [Green Version]
  127. Ruse, E.; Buzaglo, M.; Pri-Bar, I.; Shunak, L.; Nadiv, R.; Pevzner, S.; Siton-Mendelson, O.; Skripnyuk, V.M.; Rabkin, E.; Regev, O. Hydrogen storage kinetics: The graphene nanoplatelet size effect. Carbon 2018, 130, 369–376. [Google Scholar] [CrossRef]
  128. Mun, S.J.; Park, S.J. Graphitic Carbon Nitride Materials for Photocatalytic Hydrogen Production via Water Splitting: A Short Review. Catalysts 2019, 9, 805. [Google Scholar] [CrossRef] [Green Version]
  129. Wang, L.; Lee, K.; Sun, Y.Y.; Lucking, M.; Chen, Z.F.; Zhao, J.J.; Zhang, S.B.B. Graphene Oxide as an Ideal Substrate for Hydrogen Storage. ACS Nano 2009, 3, 2995–3000. [Google Scholar] [CrossRef]
  130. Singh, S.B.; De, M. Effects of gaseous environments on physicochemical properties of thermally exfoliated graphene oxides for hydrogen storage: A comparative study. J. Porous Mater. 2021, 28, 875–888. [Google Scholar] [CrossRef]
  131. Kim, B.J.; Park, S.J. Optimization of the pore structure of nickel/graphite hybrid materials for hydrogen storage. Int. J. Hydrogen Energy 2011, 36, 648–653. [Google Scholar] [CrossRef]
  132. Lee, S.Y.; Park, S.J. Hydrogen Storage Behaviors of Ni-Doped Graphene Oxide/MIL-101 Hybrid Composites. J. Nanosci. Nanotechnol. 2013, 13, 443–447. [Google Scholar] [CrossRef] [PubMed]
  133. Kumar, P.; Singh, S.; Hashmi, S.A.R.; Kim, K.H. MXenes: Emerging 2D materials for hydrogen storage. Nano Energy 2021, 85, 105989. [Google Scholar] [CrossRef]
  134. Jiang, R.C.; Xiao, X.Z.; Zheng, J.G.; Chen, M.; Chen, L.X. Remarkable hydrogen absorption/desorption behaviors and mechanism of sodium alanates in-situ doped with Ti-based 2D MXene. Mater. Chem. Phys. 2020, 242, 122529. [Google Scholar] [CrossRef]
  135. Kannan, K.; Sadasivuni, K.K.; Abdullah, A.M.; Kumar, B. Current Trends in MXene-Based Nanomaterials for Energy Storage and Conversion System: A Mini Review. Catalysts 2020, 10, 495. [Google Scholar] [CrossRef]
  136. Murali, G.; Rawal, J.; Modigunta, J.K.R.; Park, Y.H.; Lee, J.H.; Lee, S.Y.; Park, S.J.; In, I. A review on MXenes: New-generation 2D materials for supercapacitors. Sustain. Energy Fuels 2021, 5, 5672–5693. [Google Scholar] [CrossRef]
  137. Hu, Q.K.; Sun, D.D.; Wu, Q.H.; Wang, H.Y.; Wang, L.B.; Liu, B.Z.; Zhou, A.G.; He, J.L. MXene: A New Family of Promising Hydrogen Storage Medium. J. Phys. Chem. A 2013, 117, 14253–14260. [Google Scholar] [CrossRef]
  138. Yadav, A.; Dashora, A.; Patel, N.; Miotello, A.; Press, M.; Kothari, D.C. Study of 2D MXene Cr2C material for hydrogen storage using density functional theory. Appl. Surf. Sci. 2016, 389, 88–95. [Google Scholar] [CrossRef]
  139. Sun, S.J.; Liao, C.; Hafez, A.M.; Zhu, H.L.; Wu, S.P. Two-dimensional MXenes for energy storage. Chem. Eng. J. 2018, 338, 27–45. [Google Scholar] [CrossRef]
  140. Liu, S.Y.; Liu, J.Y.; Liu, X.F.; Shang, J.X.; Xu, L.; Yu, R.H.; Shui, J.L. Hydrogen storage in incompletely etched multilayer Ti2CTx at room temperature. Nat. Nanotechnol. 2021, 16, 331–336. [Google Scholar] [CrossRef]
  141. Shi, R.; Yan, H.X.; Zhang, J.G.; Gao, H.G.; Zhu, Y.F.; Liu, Y.N.; Hu, X.H.; Zhang, Y.; Li, L.Q. Vacancy-Mediated Hydrogen Spillover Improving Hydrogen Storage Properties and Air Stability of Metal Hydrides. Small 2021, 17, 2100852. [Google Scholar] [CrossRef] [PubMed]
  142. Kang, P.C.; Ou, Y.S.; Li, G.L.; Chang, J.K.; Wang, C.Y. Room-Temperature Hydrogen Adsorption via Spillover in Pt Nanoparticle-Decorated UiO-66 Nanoparticles: Implications for Hydrogen Storage. ACS Appl. Nano Mater. 2021, 4, 11269–11280. [Google Scholar] [CrossRef]
  143. Kim, B.J.; Lee, Y.S.; Park, S.J. A study on the hydrogen storage capacity of Ni-plated porous carbon nanofibers. Int. J. Hydrogen Energy 2008, 33, 4112–4115. [Google Scholar] [CrossRef]
  144. Kim, B.J.; Lee, Y.S.; Park, S.J. Novel porous carbons synthesized from polymeric precursors for hydrogen storage. Int. J. Hydrogen Energy 2008, 33, 2254–2259. [Google Scholar] [CrossRef]
  145. Karim, W.; Spreafico, C.; Kleibert, A.; Gobrecht, J.; VandeVondele, J.; Ekinci, Y.; van Bokhoven, J.A. Catalyst support effects on hydrogen spillover. Nature 2017, 541, 68–71. [Google Scholar] [CrossRef]
  146. Plerdsranoy, P.; Thaweelap, N.; Poo-arporn, Y.; Khajondetchairit, P.; Suthirakun, S.; Fongkaew, I.; Chanlek, N.; Utke, O.; Pangon, A.; Utke, R. Hydrogen adsorption of O/N-rich hierarchical carbon scaffold decorated with Ni nanoparticles: Experimental and computational studies. Int. J. Hydrogen Energy 2021, 46, 5427–5440. [Google Scholar] [CrossRef]
  147. Kim, B.J.; Lee, Y.S.E.; Park, S.J. Preparation of platinum-decorated porous graphite nanofibers, and their hydrogen storage behaviors. J. Colloid Interface Sci. 2008, 318, 530–533. [Google Scholar] [CrossRef]
  148. Lee, S.Y.; Park, S.J. Influence of CO2 activation on hydrogen storage behaviors of platinum-loaded activated carbon nanotubes. J. Solid State Chem. 2010, 183, 2951–2956. [Google Scholar] [CrossRef]
  149. Park, S.J.; Lee, S.Y. A study on hydrogen-storage behaviors of nickel-loaded mesoporous MCM-41. J. Colloid Interface Sci. 2010, 346, 194–198. [Google Scholar] [CrossRef]
  150. Pyle, D.S.; Gray, E.M.; Webb, C.J. Hydrogen storage in carbon nanostructures via spillover. Int. J. Hydrogen Energy 2016, 41, 19098–19113. [Google Scholar] [CrossRef]
  151. Wang, L.F.; Yang, R.T. New sorbents for hydrogen storage by hydrogen spillover—A review. Energy Environ. Sci. 2008, 1, 268–279. [Google Scholar] [CrossRef]
  152. Park, S.J.; Lee, S.Y. Hydrogen storage behaviors of platinum-supported multi-walled carbon nanotubes. Int. J. Hydrogen Energy 2010, 35, 13048–13054. [Google Scholar] [CrossRef]
  153. Prins, R. Hydrogen Spillover. Facts and Fiction. Chem. Rev. 2012, 112, 2714–2738. [Google Scholar] [CrossRef] [PubMed]
  154. Lachawiec, A.J.; Qi, G.S.; Yang, R.T. Hydrogen storage in nanostructured carbons by spillover: Bridge-building enhancement. Langmuir 2005, 21, 11418–11424. [Google Scholar] [CrossRef]
  155. Li, Y.W.; Yang, R.T. Hydrogen storage in metal-organic frameworks by bridged hydrogen spillover. J. Am. Chem. Soc. 2006, 128, 8136–8137. [Google Scholar] [CrossRef]
  156. Park, S.J.; Kim, B.J.; Lee, Y.S.; Cho, M.J. Influence of copper electroplating on high pressure hydrogen-storage behaviors of activated carbon fibers. Int. J. Hydrogen Energy 2008, 33, 1706–1710. [Google Scholar] [CrossRef]
  157. Hoang, T.K.A.; Antonelli, D.M. Exploiting the Kubas Interaction in the Design of Hydrogen Storage Materials. Adv. Mater. 2009, 21, 1787–1800. [Google Scholar] [CrossRef]
  158. Lee, S.Y.; Lee, J.H.; Kim, Y.H.; Rhee, K.Y.; Park, S.J. Roles of London Dispersive and Polar Components of Nano-Metal-Coated Activated Carbons for Improving Carbon Dioxide Uptake. Coatings 2021, 11, 691. [Google Scholar] [CrossRef]
  159. Im, J.S.; Park, S.J.; Kim, T.; Lee, Y.S. Hydrogen storage evaluation based on investigations of the catalytic properties of metal/metal oxides in electrospun carbon fibers. Int. J. Hydrogen Energy 2009, 34, 3382–3388. [Google Scholar] [CrossRef]
Figure 1. The US DOE targets for hydrogen storage technology. Reproduced from ref. [2]; copyright (2019) with permission from Royal Society of Chemistry, Reproduced from ref. [7]; copyright (2017) with permission from US DOE.
Figure 1. The US DOE targets for hydrogen storage technology. Reproduced from ref. [2]; copyright (2019) with permission from Royal Society of Chemistry, Reproduced from ref. [7]; copyright (2017) with permission from US DOE.
Processes 10 00304 g001
Figure 2. Schematic illustrations of various hydrogen storage techniques.
Figure 2. Schematic illustrations of various hydrogen storage techniques.
Processes 10 00304 g002
Figure 3. (a) Three-dimensional structural information, SEM image, and hydrogen uptake on 77 K and 16 bar of LEV-type zeolite for hydrogen adsorption, (b) TEM image and hydrogen storage uptakes on various temperatures (233–318 K) and 110 bar, and (c) various candidates for hydrogen adsorbents. Reproduced from ref. [60]; copyright (2007) with permission from Elsevier. Reproduced from ref. [61]; copyright (2004) with permission from Nature Research.
Figure 3. (a) Three-dimensional structural information, SEM image, and hydrogen uptake on 77 K and 16 bar of LEV-type zeolite for hydrogen adsorption, (b) TEM image and hydrogen storage uptakes on various temperatures (233–318 K) and 110 bar, and (c) various candidates for hydrogen adsorbents. Reproduced from ref. [60]; copyright (2007) with permission from Elsevier. Reproduced from ref. [61]; copyright (2004) with permission from Nature Research.
Processes 10 00304 g003
Figure 4. (a) Hydrogen adsorption behaviors of various types of zeolites, (b,c) 3D illustrations for hydrogen adsorbed in AWO-type zeolite, and (d) hydrogen adsorption isotherms of AWO-zeolite. Reproduced from ref. [47]; copyright (2021) with permission from AAAS.
Figure 4. (a) Hydrogen adsorption behaviors of various types of zeolites, (b,c) 3D illustrations for hydrogen adsorbed in AWO-type zeolite, and (d) hydrogen adsorption isotherms of AWO-zeolite. Reproduced from ref. [47]; copyright (2021) with permission from AAAS.
Processes 10 00304 g004
Figure 5. (a) Structural illustration of various types of MOF and (b) their hydrogen capacities at 77 K and 100 bar. Reproduced from ref. [73]; copyright (2019) with permission from Nature Research.
Figure 5. (a) Structural illustration of various types of MOF and (b) their hydrogen capacities at 77 K and 100 bar. Reproduced from ref. [73]; copyright (2019) with permission from Nature Research.
Processes 10 00304 g005
Figure 6. (a) Schematic illustration of the hydrogen spillover and diffusion mechanism on Matryoshka-type ZIFs, (b) designs of Matryoshka-type ZIFs, and (c) TEM image of metal loaded-ZIF for hydrogen spillover. Reproduced from ref. [78]; copyright (2018) with permission from Nature Research.
Figure 6. (a) Schematic illustration of the hydrogen spillover and diffusion mechanism on Matryoshka-type ZIFs, (b) designs of Matryoshka-type ZIFs, and (c) TEM image of metal loaded-ZIF for hydrogen spillover. Reproduced from ref. [78]; copyright (2018) with permission from Nature Research.
Processes 10 00304 g006
Figure 7. (a) Schematic illustration of the preparation process for hyper-crosslinked polymers (HCPs); the monomers used were (A) anthracene, (B) benzene, (C) carbazole, and (D) dibenzothiophene. (b) Adsorption–desorption isotherms for N2 at 77 K. (c) Hydrogen adsorption curves at 77 K and 14 MPa for COF (B1Fe1M2) and COF-GO composites (B1Fe1M2-GO) after the first 12 h of outgassing. Reproduced from ref. [82]; copyright (2022) with permission from Elsevier.
Figure 7. (a) Schematic illustration of the preparation process for hyper-crosslinked polymers (HCPs); the monomers used were (A) anthracene, (B) benzene, (C) carbazole, and (D) dibenzothiophene. (b) Adsorption–desorption isotherms for N2 at 77 K. (c) Hydrogen adsorption curves at 77 K and 14 MPa for COF (B1Fe1M2) and COF-GO composites (B1Fe1M2-GO) after the first 12 h of outgassing. Reproduced from ref. [82]; copyright (2022) with permission from Elsevier.
Processes 10 00304 g007
Figure 8. (a) Schematic illustration for the one-pot synthesis of microporous carbon derived from melamine and isophthalaldehyde with molten salt as reagents, (b) 77 K/N2 adsorption–desorption isotherms for prepared samples, (c) pore size distributions calculated by the NLDFT model, and (d) hydrogen uptakes at 77 K and 1 bar. Reproduced from ref. [124]; copyright (2018) with permission from Nature Research.
Figure 8. (a) Schematic illustration for the one-pot synthesis of microporous carbon derived from melamine and isophthalaldehyde with molten salt as reagents, (b) 77 K/N2 adsorption–desorption isotherms for prepared samples, (c) pore size distributions calculated by the NLDFT model, and (d) hydrogen uptakes at 77 K and 1 bar. Reproduced from ref. [124]; copyright (2018) with permission from Nature Research.
Processes 10 00304 g008
Figure 9. (a) Preparation process of thermally exfoliated graphene oxide (EGO) under various gas conditions, (b) N2 adsorption–desorption isotherms, (c) pore size distributions, (d) hydrogen uptakes at 77 K and 30 bar, and (e) cycle stability for hydrogen uptake under 5 cycles. Reproduced from ref. [130]; copyright (2021) with permission from Springer.
Figure 9. (a) Preparation process of thermally exfoliated graphene oxide (EGO) under various gas conditions, (b) N2 adsorption–desorption isotherms, (c) pore size distributions, (d) hydrogen uptakes at 77 K and 30 bar, and (e) cycle stability for hydrogen uptake under 5 cycles. Reproduced from ref. [130]; copyright (2021) with permission from Springer.
Processes 10 00304 g009
Figure 10. (a) Illustration of hydrogen adsorption sites of graphitic surfaces: natural graphite (NPG, left) and porous graphite (PG, right). (b) XRD of NPG and nickel/PG hybrids. (c) Hydrogen adsorption curves at 298 K and 100 bar. Reproduced from ref. [131]; copyright (2011) with permission from Elsevier.
Figure 10. (a) Illustration of hydrogen adsorption sites of graphitic surfaces: natural graphite (NPG, left) and porous graphite (PG, right). (b) XRD of NPG and nickel/PG hybrids. (c) Hydrogen adsorption curves at 298 K and 100 bar. Reproduced from ref. [131]; copyright (2011) with permission from Elsevier.
Processes 10 00304 g010
Figure 11. (a,b) TEM images of an MXene for hydrogen storage with a bell-mouth structure, (c) the hydrogen storage mechanism on an incompletely etched MXene, (d,e) hydrogen uptakes of MXene samples at 298 K and 60 bar, (f) XRD patterns of pristine, hydrogenated, and dehydrogenated MXene, and (g) hydrogen uptake cycles. Reproduced from ref. [140]; copyright (2021) with permission from Nature Research.
Figure 11. (a,b) TEM images of an MXene for hydrogen storage with a bell-mouth structure, (c) the hydrogen storage mechanism on an incompletely etched MXene, (d,e) hydrogen uptakes of MXene samples at 298 K and 60 bar, (f) XRD patterns of pristine, hydrogenated, and dehydrogenated MXene, and (g) hydrogen uptake cycles. Reproduced from ref. [140]; copyright (2021) with permission from Nature Research.
Processes 10 00304 g011
Figure 12. Comparison of two representatives: Yang’s spillover model and Park’s modified spillover model in H2 storage (dn: distance between the metal and the number of the hydrogen layers; BEdn: binding energy between two dns).
Figure 12. Comparison of two representatives: Yang’s spillover model and Park’s modified spillover model in H2 storage (dn: distance between the metal and the number of the hydrogen layers; BEdn: binding energy between two dns).
Processes 10 00304 g012
Table 1. Technical system targets for onboard hydrogen storage for light-duty fuel cell vehicles. Reproduced with permission from ref. [7]; copyright (2017), the US DOE.
Table 1. Technical system targets for onboard hydrogen storage for light-duty fuel cell vehicles. Reproduced with permission from ref. [7]; copyright (2017), the US DOE.
Units20202025Ultimate
Storage capacity
System gravimetric capacity
Usable, specific energy from H2
(net useful energy/max system mass)
kWh/kg
(kg H2/kg system)
1.5
(0.045)
1.8
(0.055)
2.2
(0.065)
System volumetric capacity
Usable energy density from H2
(net useful energy/max system volume)
kWh/L
(kg H2/L system)
1.0
(0.030)
1.3
(0.040)
1.7
(0.050)
Storage system cost
Storage system costUSD/kWh net
(USD/kg H2)
10
(333)
9
(300)
8
(266)
Fuel costUSD/gge at pump444
Durability/operability
Operating ambient temperature°C−40/60 (Sun)−40/60 (Sun)−40/60 (Sun)
Min/max delivery temperature°C−40/85−40/85−40/85
Operational cycle life (1/4 tank to full)cycles150015001500
Min/max delivery pressure from storage systembar (abs)5/125/125/12
Onboard efficiency%909090
“Well” to power plant efficiency%606060
Charging–discharging rates
System fill timemin3–53–53–5
Minimum/average full flow rate(g/s)/kW0.02/0.0040.02/0.0040.02/0.004
Start time to full flow (−20 °C/20 °C)s15/515/515/5
Transient response at operating temperatures0.750.750.75
Dormancy
Dormancy time targetdays71014
Boil-off loss target%101010
Environmental health and safety
Permeation and leakage-Meet or exceed SAE J2579 for system safety
Toxicity-Meet or exceed applicable standards
Safety-Conduct and evaluate failure analysis
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, S.-Y.; Lee, J.-H.; Kim, Y.-H.; Kim, J.-W.; Lee, K.-J.; Park, S.-J. Recent Progress Using Solid-State Materials for Hydrogen Storage: A Short Review. Processes 2022, 10, 304. https://doi.org/10.3390/pr10020304

AMA Style

Lee S-Y, Lee J-H, Kim Y-H, Kim J-W, Lee K-J, Park S-J. Recent Progress Using Solid-State Materials for Hydrogen Storage: A Short Review. Processes. 2022; 10(2):304. https://doi.org/10.3390/pr10020304

Chicago/Turabian Style

Lee, Seul-Yi, Jong-Hoon Lee, Yeong-Hun Kim, Jong-Woo Kim, Kyu-Jae Lee, and Soo-Jin Park. 2022. "Recent Progress Using Solid-State Materials for Hydrogen Storage: A Short Review" Processes 10, no. 2: 304. https://doi.org/10.3390/pr10020304

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop