Final4 With Edited Figure
Final4 With Edited Figure
Searching for alternative energy sources has emerged owing to a shortage of fossil fuels, climate changes, etc., and
in this scenario, hydrogen can appear as a very good source of energy. A number of methods have been discovered
for producing hydrogen but the challenges in storing hydrogen are making the limitations of using hydrogen energy
as primary fuel. For practical uses, storage technologies like compressed hydrogen and liquid hydrogen are too
expensive and have some major drawbacks like high energy consumption and cryogenic infrastructure. By
bypassing these drawbacks nanostructured materials are very promising candidates for storing hydrogen taking
advantage of high theoretical capacities and their unique properties. However, these materials also go through
different types of damage during hydrogen cycling. Different damages such as structural damages, sluggish kinetics,
low hydrogen adsorptions, laser exposure issues, defects in multiwalled carbon nanotubes (MWCNT), and poor
reversibility that cause limited storage capacity of the total hydrogen storage system are explored in this review
paper. At last, recent developments were discussed to mitigate these damages for enthusiastic researchers. This
review paper will also offer a critical understanding of the damages associated with nanostructured hydrogen storage
materials for creating appropriate materials for future hydrogen storage technology.
Introduction:
The continuous growth of the world economy and population as well as rapid urbanization has just increased the
energy demand[1]–[5]. We have been using fossil fuels as our principal source of energy since the Industrial
Revolution and that has caused the increment in the levels of greenhouse gases and CO 2 which are the main reasons
for climate change and global warming[6], [7]. In addition, geopolitics plays a very crucial role in not getting access
to crude oils for third-world countries with strict regional policies which has created a need for alternative fuels [8].
Several potential alternatives were proposed before including solar fuels[9], methanol[10], ethanol[11], biomass[12],
hydrogen[13], boron-based materials[14], etc.
Among these fuels, hydrogen holds the highest rank in the index of highest specific energy content and it is
available in large quantities in this universe[15]–[17]. This energy is a secondary form of energy[8]. From these
large quantities, there are several methods to convert the hydrogen to energy. But, here comes the very important
issue regarding storing hydrogen. All energy must be stored safely and efficiently for better energy transfer
throughout the required destinations. A very definite number of hydrogen storage technologies are developed but
they can rarely meet the needs for practical uses[18]–[20]. Fig. 1 summarizes different types of hydrogen storage
technologies and their limitations. The storage systems can be defined in mainly two tracks: physical technology and
material-based technology. At this time, the physical technologies that are compressed hydrogen storage and
liquefied hydrogen storage are mostly used technologies[21]. The main drawbacks of compressed hydrogen storage
are low volumetric density, high energy consumption for compression, and safety concerns for high pressure.
Hydrogen gas is stored in the pressure range of 350-700 bar which concerns very much about safety issues,
especially at the time of loading and unloading the hydrogen[22], [23]. An explosion of a hydrogen transmission
vehicle occurred at a coal-fired power plant on 8 January 2007 which killed one person and injured 10 others. The
investigation report pointed out that the premature failure of a pressure relief device at the time of unloading
hydrogen was the main reason for that accident[24]. The main drawbacks of liquefied hydrogen storage are high
energy consumption for liquefaction and requirements of cryogenic infrastructure. Hydrogen is cooled at -253C at
ambient pressure(1 atm) to liquefy and stored in cryogenic tanks[25]–[27]. However, it is very costly to arrange the
refrigeration system and the complexity of cryogenic infrastructure[28], [29]. On the other hand, the hydrogen atoms
can penetrate the inner part of the containing vessel which could be the main reason for hydrogen embrittlement. In
this process, the inner part of the containing vessel loses its ductility and strength as a result it becomes more brittle.
This can create structural damage which leads to the catastrophic disruption of the containing vessel [30]–[32]. The
material-based technology is also divided into two parts: physisorption and chemisorption[21]. The process in which
hydrogen is bonded with the storage systems by weak attraction forces like the van der Waals force is known as
physisorption and the process in which hydrogen is bonded with the storage systems by strong attraction forces like
ionic and covalent force is known as chemisorption[33], [34]. Metal-organic frameworks (MOFs), covalent organic
frameworks (COFs), carbon nanotubes (CNTs), zeolite, graphene, etc. are examples of material-based technology
for the physisorption process and compounds like LiAlH 4, LiBH4, MgH2, LaNi5, AlH3, liquid organic hydrogen
carrier (LOHC), etc. are examples of material-based technology for the chemisorption process[21].
Several limitations are found while using nanostructured hydrogen storage systems. The structure of LiAlH 4 is
totally changed at the time of releasing hydrogen[38]. Pores have a symbiotic relation with storage capacity but
connective pores can be a great threat because these can damage the structure of the storage walls and that can lead
to catastrophic disruption[39]. Poor kinetics is a very challenging issue, especially for the Mg-based nanoparticles
that form cracks after a number of cycles and these also affect the capacity of the storage systems [40]–[42]. The
reduced adsorption rate in certain conditions and the synthesis process are also a very serious issue to concern [36],
[43]. For the use of MWCNT, laser treatments reduced the entrapment of hydrogen which leads to limited storage
capacity[44], [45]. Again the negative impact on the characteristics of storage is caused by temperature differences,
poor reversibility while selecting inappropriate materials, and complicated synthesis process[46]–[48]. This review
finds and summarizes the details of the number of damages of nanostructured hydrogen storages to help the
researchers address the solutions to these damages.
Physisorption:
Selected components of a fluid phase known as solutes are transferred to the surface of an insoluble solid, this
process is known as adsorption. Where the absorbed solute is specified as adsorbate and the solid material is
adsorbent. Molecules of fluid diffuse into the adsorbent’s surface if the fluid phase is revealed to the adsorbent. If
these molecules are held by physical or intermolecular force then this type of adsorption is stated as physisorption or
physical adsorption.[49]
Physical adsorption of gas can be defined as the accumulation of molecules at the conjunction of the surface of the
solid and the gas phase. This interaction creates an attraction between fluctuating and induced dipoles between these
molecules (gas phase and the surface of the solid). London Dispersive force helps hydrogen molecules to be
adsorbed at the surface of materials.[50],[51],[52],[53] As the surfaces are the main dominating factor, highly
porous materials with large surface areas are sought after for physisorption.[54] Moreover, these are favorable in the
case of physisorption because of uniform pore construction, regularity, and fluid permeability [55][56][57][58][59].
Materials such as porous silica, MOFs, activated carbons (AC), and zeolites are some of the candidates for
physisorption.[60][61][62].To meet the target of DOE values, the physical adsorption process should use these
effective materials.[63]The physisorption storage process is an attractive and one of the most viable options because
of its storage capability in low enthalpy (room temperature) and relatively low pressures. In recent times the
materials used in the physisorption process are porous (zeolites, metal-organic frameworks (MOFs), carbon
materials like fullerenes, etc.) and have a large specific surface area. Which is the reason they become lightweight,
have high storage density, and Cycle stability. Because of the low Van der Wales force it is far easier for desorption
than other processes[63].
With low bonding energy (Van der Waals) the H 2 atoms get adsorbed onto the surface of the storage material as
shown in the Fig.2. [64]
The physisorption process/technique has shown promising results by being producible in 298K temperature and 1.5-
30 bar pressure. [65] Since no energy barrier remains between the gas phase and the absorbed state, it is guaranteed
that there will be fast adsorption and desorption kinetics. In addition, there is no need for extra thermal control
systems or any small heat management.
While comparing the physisorption process and liquid storage at cryogenic temperature, physisorption reduces
hydrogen boil-off loss during storing hydrogen and needs low energy while charging and discharging. [66][67].
Some of the improvements in the physisorption sector have been made so that it can meet the expected criteria. The
binding energy of neutral aromatic molecules such as coronene has been doubled because of the anionic charged
systems in addition to the van der Waals force. Layered anionic structural advantages include a quadrupled total
binding energy which is an astounding result in the case of physisorption.[65]
MOFs considerably show very high internal surface areas of 3000 to 7000 m 2 g−1. [54] Also, the improvement of
MOF monoliths showed total volumetric capacities of over 40 g L−1 at 77 K. [54] A recent development suggests
that the usage of adsorbents at subcritical temperature of hydrogen compared to liquid hydrogen storage showed a
reduced boil-off (extended dormancy). A new future to high-density storage/transportation by physisorption can be
imminent as fundamental investigations on subcritical hydrogen adsorption in silica showed accumulation of super
dense hydrogen monolayer. It can be a new possibility for large-scale energy storage by cryogenics. [54]
Chemisorption:
The process in which H2 molecules bonded by chemical reaction with the storage materials, is called chemisorption.
Unlike physisorption, chemisorption forms strong chemical bonds like ionic, covalent, or metallic [34], [68], [69].
The nanostructure storage materials itself have several factors that can affect the chemisorption mechanism. The size
and shape of the nanomaterial play an essential role in this regard. Smaller nanoparticles, which have more surface
atoms, will provide more hydrogen adsorption sites ultimately increasing storage capacity [70]. Also, the type of
elements used in nanostructures affects the nature and strength of bonds with hydrogen for example some transition
metals like platinum or palladium show a strong affinity towards hydrogen resulting in efficient chemisorption [71].
Additionally, having defects or dopants within the nanostructure further modulates the chemical environment
leading to interaction with hydrogen.
Figure 3: Schematic diagram of chemisorption on metal surface[72].
In the case of chemisorption, the potential of high energy storage, easy usage, and most significantly liquid
integrated systems used in the same infrastructure of today's resources is fairly noticeable. However, the release
process is comparatively difficult due to irreversible storage approaches. Also, a couple of significant variables
make it harder to incorporate with recent storage units[73]. For example, ammonia and other compounds are
unsuitable because they produce gaseous pollutants, and metal hydrides have poor kinetics and insufficient stability
[74].
Spillover Mechanism:
Small amounts of transition metals (TMs) nanoparticles such as platinum (Pt), palladium (Pd), and nickel (Ni) are
incorporated into carbon nanostructures to enhance the hydrogen storage capacity. This can be achieved by
physically mixing the metals with carbon or directly doping through wet chemistry methods. The improved
hydrogen storage capacity in metal-doped carbons is attributed to the spillover mechanism. This involves the
transfer of an active species, specifically atomic hydrogen, from a surface where it is absorbed or formed (such as a
transition metal nanoparticle) to another surface that typically does not adsorb or form the active species (like the
carbon nanostructure) [75], [76]. There are three steps included in the hydrogen spillover mechanism which are
shown in Fig 4: 1) Dissociate molecular hydrogen to atomic hydrogen and bind to the surface of transition metal
(TM) nanoparticles; 2) Migrate atomic hydrogen from the TM nanoparticles to the carbon surface and 3) Diffuse
atomic hydrogen along the carbon surface, forming a stable C-H bond after traveling some distance [75].
Figure 4: Hydrogen spillover mechanism: Step 1) Dissociation of molecular hydrogen and formation of TM-H
bonds; Step 2) Migration of atomic hydrogen to the carbon surface; Step 3) diffusion of atomic hydrogen along the
carbon surface [75].
The carbon nanostructure’s characteristics are vital for hydrogen spillover, with the majority of hydrogen being
stored on the carbon substrate and only a minor portion on the metal nanoparticles. Researchers have experimentally
studied hydrogen spillover to various carbon materials, such as activated carbons, carbon nanotubes, MWCNTs,
carbon nanospheres, carbon nanofibers, fullerenes, and graphene-like materials [75], [77]–[80].
In the spillover mechanism, hydrogen molecules (H 2) are dissociated into atomic hydrogen (H) on the surface of
transition metal nanoparticles such as Pt or Pd which act as catalysts. Then the dissociated hydrogen atoms migrate
from the transition metal nanoparticle surface to the receptor material, usually a high-surface area like activated
carbon or other carbon nanostructured material [75]. The atomic hydrogen adsorbs onto the carbon nanostructured
material and may further diffuse into its bulk or bind to specific sites, thus storing hydrogen in a more stable form.
The hydrogen atoms can either be physisorbed (weakly bind via van der Waals forces) or chemisorbed (from
stronger chemical bonds) to the receptor material [75]. The balance between these interactions influences the overall
storage capacity and stability of the storage system.
Adhesion kinetics, which describes the rate at which hydrogen atoms adhere to and migrate across the surface, plays
a crucial role in the spillover mechanism for hydrogen storage. The kinetics of the dissociation process are
influenced by the catalytic properties of the metal, available surface area, and the required bond dissociation. On the
other hand, structural damage can lead to mechanical degradation, reducing the material’s ability to withstand
multiple hydrogen cycles and reducing the spillover mechanism kinetics. Density functional theory (DFT)
calculations are often used to predict and understand the behavior of hydrogen on nanomaterials [75]. These
theoretical models help identify favorably binding sites, migration pathways, and energy barriers associated with the
spillover process [75], [81].
To ensure carbon nanostructures effectively store hydrogen, atomic hydrogen must be supplied in sufficient
quantities, and with reasonable kinetics to enable surface diffusion. TM nanoparticles must dissociate molecular
hydrogen and bind atomic hydrogen at low enough energies to allow migration from the nanoparticles to the carbon
nanostructure at ambient temperatures. The efficiency of hydrogen dissociation and atomic hydrogen bonding is
influenced by the type of TM nanoparticles, their size, hydrogen surface coverage, and how the nanoparticle is
anchored to the carbon substrate [75].
Krasheninnikov et al., [82] through DFT calculations, demonstrated that vacancy defects in graphene create a strain
field extending approximately 2 nm from the defect site. The strain makes the carbon atoms in the pristine graphene
lattice more reactive. They speculated that these strain fields attract TM atoms to the defect sites. The impact of
these strain fields on the activation energy for atomic hydrogen migration from a TM cluster or small nanoparticles
bound to a vacancy defect site is of particular interest for hydrogen storage via spillover in carbon nanostructures
[75].
Nanomaterials with pore optimal pore sizes have a larger surface area available for hydrogen adsorption. Larger
surface areas provide more sites for hydrogen to adhere to, enhancing the overall storage capacity in the spillover
mechanism. Also, pores must be of an appropriate size to allow hydrogen molecules to enter and diffuse through the
material. If the pore size is too small, hydrogen molecules may not enter effectively, reducing storage capacity. On
the other hand, if the pores are too large, the surface area might be insufficient for maximizing adsorption sites. [83]
Recent experimental studies have shown that adding oxygen functional groups and lattice dopants can increase
hydrogen adsorption capacity to over 3 wt% at 298K, 40 bar in the spillover mechanism [84]. Although introducing
oxygen functionalization and lattice dopants can create more reactive sites by disrupting π-conjugation, it does not
necessarily increase the total number of reactive sites and may even reduce them. Also, there are significant
concerns regarding the reversibility of these materials. Contradictory experimental results have been reported, and
theoretical studies suggest that sample degradation is more favored over H 2 desorption in the spillover mechanism
[75].
Hydrogen storage technologies rely heavily on nanostructured materials due to their high surface area and unique
properties. However, these materials are susceptible to various damages and defects that can significantly impair
their performance. Issues such as high-temperature activation, continuous cycling, and improper catalyst coatings
can lead to the collapse of structural cavities, formation of defective sites, and degradation of the material’s
integrity. Understanding and mitigating these damages is crucial for optimizing the efficiency, reversibility, and
longevity of hydrogen storage systems.
Structural Defects and Damages:
Structural damage is defined in terms of the material microstructure or nanostructure by continuum damage
mechanics (CDM) and the damage variable is transmitted to the macroscopic strength or stiffness of the structure.
This allows the state of damage to be predicted before the initiation of macroscopic flow and permits the estimation
of residual strength or service life of an existing structure. So, structural damage is any kind of damage that has an
impact on its structural integrity [85].
Yurum et al. [43] found the enormous chemical and structural complexity and heterogeneity of activated carbons
between theoretical and experimental. This kind of difference or not correlating between the experimental and
theoretical is the major disadvantage [43]. Because it reduces the overall hydrogen storage efficiency.
As an example, on the surface of a multi-walled nanotube (MWNT) for hydrogen storage, numerous defective
cavities collapse at high temperatures and MWNT may have thinned. Because at a high temperature of 1100 C
specific surface area and pore volume become decreased when CO2 activation is extreme [48]. Extreme CO2
activation can be harmful due to the formation of less porous carbon as a result of decreased specific surface area
and pore volume. Also, the graphitic structure of MWNT becomes destroyed for the defective cavities’
decomposition of the external surface of activation temperature at 1100C (Acti-1100) [48]. At 1100C temperature
CO2 reacts with these defect sites, breaking the carbon-carbon bonds and removing carbon from the structure. The
removal of carbon atoms leads to the breakdown of the hexagonal lattice structure which is defined as a graphitic
structure. Another kind of structural damage is found when a catalyst is coated in nanomaterials where a new cluster
forms which is structurally different from the previous one. Fig 5 shows the coating process of the Ti catalyst in
LiAlH4 nanomaterials.
Figure 5: Illustration of the coating process of Ti at the surface of LiAlH4 nanoparticles [86].
Pratthana et al.[86] found that in LiAlH4 lattice Ti can be sited at the void of LiAlH4@Ti for this reason the bond of
Al-H bonds weakening hydrogen released and moveable Al-H creates, Ti-Al clusters formed [86]. LiAlH4@Ti
displayed only one broad Al-H stretching vibration, which could be due to alterations in the Al-H environment
caused by the formation of the Ti shell. The slight shift towards higher wavenumbers in all the Al-H bending
vibrations suggests a modification in the Al-H environment [86]. The mobility allows the Al-H bonds to reconfigure
and interact with the Ti atoms, leading to the formation of Al-Ti clusters. These statements indicate the adverse
effect of structural damage. So, a material should be selected for hydrogen storage when the reversibility of
hydrogen uptake is optimal. In the case of LiAlH 4, with a core-shell structure even if the material is a good one
when it comes to reversibility the uptake of hydrogen with a high storage capacity fails to do the job, and this
attribute is considered to be a loss of core-shell structured LiAlH 4[86]. The reversibility in LiAlH4@Ti decreased
because of the loss of the core-shell structure or the formation of stable products that require higher temperatures
and hydrogen pressures to regenerate the initial LiAlH 4 phase. The sintering of LiAlH 4 nanoparticles during heat
treatment results in a loss of their nanoscale properties. This is another reason for the almost complete lack of
reversibility in LiAlH4@Ti [86]. There is another kind of structural damage which is called continuous cycling.
During continuous cycling, internal stress is created in the structure of the metal hydride which is capable of
fracturing the metal hydride as a result reduction of the lifetime of the hydrogen storage. Also, some submicron
particles are created. This damage was found by Liu et al. [42].
For hydrogen storage in multi-walled carbon nanotubes (MWCNTs), Palladium (Pd) is a very good and effective
catalyst as revealed by recent articles. By the nanosecond laser irradiation, MWCNTs furnished with palladium
nanoparticles are fabricated to improve the property of hydrogen storage. Laser-induced Pd- Pd-furnished MWCNTs
are analyzed for various laser doses [39]. The hydrogen desorption radically reduces for the long laser exposure time
of 120 min which is noticed by Mortazavi et al.[39]. This happened because, in the MWCNT structure, total
destruction creates pores whose diameter is larger than the micro-pores. On the other hand, significant structural
damage occurs because of meso-pores which are formed by the connection of each micro-pores for higher laser
doses.
Thermogravimetric analysis (TGA) reveals that extended exposure time to laser irradiation increases MWCNTs'
damage. Specifically, higher laser doses lead to the formation of amorphous carbons due to induced weakly
graphitic structure [39]. Also, at elevated laser doses, there are notable reduction-induced micro-pores, characterized
as defective nanoscale cavities on the surface area of MWCNTs. These cavities act as barriers, impeding the
enhancement of hydrogen storage capacity [39]. The collision and adsorption of hydrogen with high incident kinetic
energies on a single-walled boron nitride (BN) nanotube were investigated by Han et al. [87]. All of the H
(Hydrogen) atoms separated at the outer wall can go inside the nanotube and form H2 by recombination in the range
of 23-26 eV at higher incident energies [88]. These overall interactions can lead to changes in the structural
properties of the BN nanotubes. However, the tubular structure may automatically be damaged by higher incident
energy[88]. At high incident energies, the kinetic energy of incoming hydrogen atoms can be sufficient to break the
strong B-N bonds in the nanotube structure. This bond breaking can create defects such as vacancies, where boron
or nitrogen atoms are knocked out of their lattice position. This process degrades the structural integrity and can
eventually lead to the destruction of the nanotube’s tubular structure.
Figure 6: TEM (a) and HRTEM (b, c, d) characterization of the TiS2 nanotubes. The rectangle in c marks the area
magnified in d. The arrows in d display the defects, and the inset is the corresponding electron diffraction pattern
[89].
Chen et al. [89] observed various types of defects in TiS 2 nanotubes which are revealed by the defective deformation
of the nanotubes [89]. To understand the connection between TiS2 nanotube structures and their hydrogen uptake
properties, the sample was examined using a transmission electron microscope (TEM) and high-resolution TEM
(HRTEM) as shown in Fig 6. Through the nanotube walls, defects are situated, and in the nanotips planner defects
are present. This action is shown for continuous hydrogen absorption-desorption cycles when interlayers of the
structure expand irregularly[89]. The point defects cause local distortions in the nanotube lattice. The immediate
area around a point defect may experience compressive or tensile stresses, leading to localized bending. On the other
hand, at the nanotube tips, planar defects can cause more extensive deformation. The cyclic nature of hydrogen
absorption and desorption causes the layers within the nanotube structure to expand and contract. This irregular
expansion can increase existing defects or create new ones.
Another damage is found in palladium hollow nanoparticles (Pd HNPs) at the time of hydrogen adsorption. When
hydrogen is formed only inside the cavity of Pd hollow nanoparticles the ratio of H/Pd was found maximum which
was 1.21 [90]. But when hydrogen is absorbed on the surface both inside and outside the ratio was found 1.70,
which is 25% larger than the results for Pd NPs observed by Crespo et al.[91]. This indicates that the hollow
structure of Pd HNPs allows for a higher hydrogen storage capacity than Pd NPs but the Pd HNPs break up or are
structurally compromised when the ratio exceeds 1.70 [90]. Excessive hydrogen adsorption leads to significant
internal stress and pressure, causing the nanoparticles to break apart. So, this also suggests a threshold beyond
which the structural integrity cannot be maintained.
The kinetic mechanism in the hydrogen storage system includes hydrogen absorption and desorption rates. These
two are important parameters for the use of hydrogen storage systems. A good kinetic process increases the
performance level of the hydrogen storage system by affecting the hydrogen uptake and release. For chemisorption
materials enhancing the charge-discharge kinetics is one of the main challenges to face. In physisorption, hydrogen
stays in its molecular form and loosely attaches to the surface of the storage material because of the weak van der
Waals force. So, hydrogen can generally be stored only at a cryogenic temperature under high pressure. Thus, this is
another significant challenge for physisorption to minimize the physical forces like temperature and pressure and get
better hydrogen storage capacity. On the other hand, thermodynamics is also important to know the behavior like
chemical reactions, and mechanical deformation of materials at different temperatures and pressures.
The study by Chen et al. [92] describes the synthesis of Mg nanowires using a simple vapor transport method. The
study highlights the enhanced kinetics of hydrogen absorption and desorption reactions exhibited by Mg nanowires.
Thinner Mg/MgH2 nanowires (diameter < 30 nm) require significantly less desorption energy as a result of higher
desorption kinetics.[41] This is because there is size dependency of desorption kinetics, the smaller the size of
particles, the faster the desorption of the content of hydrogen. For nanoparticles, desorption is faster due to their
short diffusion lengths and large surface area for H recombination which is investigated by Liu et al. [37]. On the
other hand, Gross et al. [93] synthesized nanoparticulate MgH2 samples incorporated in a porous carbon host with a
pore size of 2 to 30 nm in diameter, due to the small nanoparticle size and the catalytic effect of Cu and Ni from the
wetting layers, they found that the MgH2 sample in a nano-porous carbon host could release about 5 wt.% hydrogen
at 250°C [41]. However, despite the small crystal size of a few nanometers and the introduction of catalysts in the
storage material, they could not observe significant thermodynamic changes in their system. Therefore, it can be
concluded that thermodynamic properties remain unchanged when using catalysts and nanostructures within the size
range of 5 to 300 nm [41]. Wagemans et al. [94] proposed that changes in thermodynamic values could be detected
during hydrogen sorption when Mg nanoparticles are smaller than 1.3 nm, due to surface energy alterations and
other nanoscale effects. Their quantum chemical study revealed that desorption energy significantly decreases when
the crystalline size drops below 1.3 nm [94]. The stability of thermodynamics is a big issue for hydrogen storage
systems. Thus, developing Mg-based nanomaterials for onboard hydrogen storage remains challenging [41].
The kinetics becomes sluggish for the continuous activity of charging and discharging of hydrogen in nanostructured
metal hydrides for hydrogen storage[95]. Shao et al.[96] found that in nanostructured Mg-based hydrogen storage
material in the absorption and desorption cycle, a crack has been generated on the surface of Mg. For this reason, the
surface becomes spongier and there appear to be channels within the particles due to the movement of hydrogen in
and out of them which affects the kinetics [96]. Developing spongier surfaces with channels due to crack formation
generally enhances hydrogen absorption and desorption kinetics by increasing the surface area, improving hydrogen
diffusion, and reducing the diffusion path length [96]. However, the long-term mechanical stability decreases for
continuous cycling, as a result, kinetics become sluggish.
Figure 7: Hydrogen sorption kinetics of polyaniline nanospheres at room temperature showing good reversibility in the initial
run.[40]
Niemann et al.[40] observed that at room temperature the hydrogen sorption kinetics of polyaniline nanospheres are
represented by Fig. 7 [40]. The hydrogen uptake and release of ∼4.0wt.% occurs in the initial run which is
noticeable from this graph. However, the hydrogen storage capacity and kinetics were decreased during continuous
cycling [40]. This is because during continuous cycling internal structure of storage materials is harmed by breaking
bonds of the storage materials which affects its storing capacity and absorption-desorption kinetics due to the
reduction of active reaction sites [42], [96]. On the other hand, too high plateau pressure at relatively low
temperatures for LiAlH4 which makes the hydride comparatively slow hydrogen desorption kinetics found by Rafi-
Ud-Din et al.[47]. The storage material with high thermodynamic stability requires higher temperature and pressure
to release stored hydrogen and may also exhibit slow hydrogen kinetics. This can be impractical and energy-
intensive, making the storage system less efficient.
Thus, kinetics and thermodynamics mechanisms play an important role in hydrogen storage systems and they should
be improved to get more output. Kinetics can be improved by decreasing particle size to get more surface area and
short diffusion distance and introducing catalysts to get more active sites. On the other hand, thermodynamic
stability can be changed when the particle size is smaller than the 1.3 nm range. Also, one effective method to
decrease the thermodynamic stability of Mg-based materials is to create alloys by combining Mg with other metals,
such as Ni. When Mg is alloyed with Ni, Mg 2Ni is formed, which can absorb hydrogen to produce the hydride
Mg2NiH4. This hydride releases hydrogen with a lower desorption enthalpy of 65 kjmol-1 H2 [96].
There are a couple of circumstances affecting the hydrogen adsorption rate. One of the cases is that Gas-phase
persulfate-oxidized activated carbons increased the hydrogen adsorption rate at a certain temperature (400 ℃-
700℃) but on the other hand liquid-phase persulfate-oxidized activated carbons displayed a loss of hydrogen
adsorption capacity. The reason behind this is that liquid phase persulfate oxidized activated carbons fill up some
adsorption sites ultimately lowering the effective surface area and porosity of the adsorbent. [43]
Then comes the observing problem of measuring adsorption capacity. This is a difficult task because this parameter
depends on multiple variables like usable surface area, size of the pores, surface topology, chemical composition of
the surface, also pressure and temperature. Moreover, the lack of non-standardized test conditions used by many
research groups and individuals brings out the complexity of adsorption dependency in a clear view. Also, in many
cases of experiments, to introduce experimental conditional variations, many have added impurities, contaminants,
or high-pressure isothermal variabilities. These are only a couple of problematic situations among others. Even
though theoretical calculations should have focused on examining multiple parameters and their interrelation,
generally most theoretical calculations examined only one characteristic of the adsorption. [43]
There are considerations to make in the case of the synthesis process too. The vapor phase deposition or templating
process is one of the synthesis processes that increases the surface area of carbon materials and ultimately improves
the adsorption of hydrogen but this process is costly for mass production which makes it ineffective for hydrogen
storage. This is where electrospinning comes into play, it is a synthesis process where with the help of electric force,
sub-micron-sized uninterrupted fiber from polymer solution/melt is produced. Its cost-effectiveness and versatility
defer it from the vapor phase deposition process. However, Kim et al.[97] stated that in this process, nanoporous
carbon fiber while being coated by Pd, shows slow adsorption kinetics at 77k and decreased micropore volume as Pd
nanoparticles are still attached to carbon surfaces. Also because of the Pd presence, hydrogen gets dissociated over
Pd because of incomplete occupation of d orbitals. These chemisorbed hydrogens sometimes travel to the Pd-C
interface for carbon support and form stable bonds which in return makes the adsorption irreversible. [97]
In the case of surface modification of CNT, hydrogen storage measurements are generally done at high pressure and
room temperature. Even though the Functional Carbon nanotube (F-CNT) showed promising results, we noticed a
low amount of hydrogen adsorption of 0.65 wt% at 80 bar pressure.[98]
Figure 8: Isotherm curves at (a) 77 K and (b) 300 K for graphene with different vacancy defects[99].
Fig 8a. presents the difference in hydrogen adsorption of different defected graphene in 77k. Where monovacancy
(MV) shows the maximum adsorption rate with the pristine graphene being the lowest.
Fig 8b. presents the difference in hydrogen adsorption of different defected graphene in 300k. Where the values
stand at 1.17 wt%,2.2083 wt%, 2.1043 wt%, 2.3098 wt%, 3.006 wt%, and 3.096 wt% in 100 bar pressures of
pristine, Monovacancy (MV), Stone Wales (SW),5-8-5 double vacancy (DV),555-777 DV and 5555-6-7777 DV
defects, respectively.
By Fig. 8(a),(b) Kag et al.[99] stated that defective graphene has an increased hydrogen adsorption rate compared to
the pristine graphene hydrogen adsorption rate. These defects have dangling bonds because of their vacancies,
ultimately storing more hydrogen than the pristine one. [99].
As an example of defects caused by the wrong-selected material, p-CNT’s (Pristine multiwalled carbon nanotubes)
hydrogen adsorption is at room temperature and high pressure is less than <0.5 wt%, which is much lower.
Nanocomposites conducting polymers show less hydrogen adsorption (1.4% -1.7%) which is less than expected.
[40]
Figure 9: Variation of wt% of hydrogen adsorption on graphene with strain at 77K and 10 bar[99].
Fig 9 shows the wt% of hydrogen adsorption with strained pristine graphene. Kag et al.[99] state the lack of strain
decreases the chemical activity and makes the graphene chemically less active for hydrogen adsorption. With that,
the C-C bond length also decreases, because the total surface area decreases to make the physisorption hard (as it is
a surface phenomenon), and the wt% decreases with the decrease of strain.[99]
A common finding is that only hydrogen physisorption or adsorption cannot meet the DOE (Department of Energy)
requirements at atmospheric temperature and under high-pressure conditions. The urgency of focusing on multiple
variables is very apparent/obvious.[43]
The reactions begin at 150C, 180C, and >350C respectively, and release 5.3 wt.%, 2.6 wt.%, and 2.6 wt.% of
hydrogen respectively[38], [104]–[106].From this, it is clear that at lower, more practical temperatures, LiAlH 4
stubbornly clings to its hydrogen payload. However certain issues like weak reversibility and sluggish kinetics are
the key problems for using LiAlH4 as a practical use of hydrogen storage[107], [108].
Almost the same types of problems were found by Rafi-ud-din et al.[47] and stated that the highest hydrogen can be
stored in LiAlH4 but the significance of this capacity can be demoralized by the presence of an exothermic reaction
and at high pressure and low temperature the release of hydrogen process can be totally stopped which means the
hydride will be completely irreversible[47]. For the exothermic reaction, the temperature is bounded in the range of
350C. So, it is very hard to draw out the hydrogen from LiAlH 4 though it stores maximum hydrogen. This unused
hydrogen will reduce the storage of hydrogen and that’s how the significance of capacity is demoralized.
One of the major problems in storing hydrogen is managing heat. The repeated uptake and release process can create
heat, which requires strategies like physical absorption to control[111]. Moradi et al.[112] stated that the mitigation
of thermal management by physical absorption occurs only because of the continuous uptake and release process of
hydrogen in the storage unit[112]. Unfortunately, the amount of stored usable hydrogen is very low for these
methods. Here the main challenge is a storage unit using more energy for dealing with heat rather than holding
hydrogen. Almost similar properties are found in nanostructured magnesium-based storage systems by Zhang et al.
[113] and they noticed that these magnesium-based storage systems often struggle with slow hydrogen release[113].
The rate of hydrogen release is hampered by poor thermodynamics which results in low usable hydrogen. The slow
rate of releasing hydrogen is a very big problem for real-world uses. Catalyst materials can speed up the process of
releasing hydrogen, acting like a jumpstart[114]. Integrating porous structures inside the storage materials could also
help the flow of hydrogen, further boosting efficiency[115].
Porous materials can absorb more hydrogen like a sponge can soak up more water. The more surface area the
sponge has, the more water it can soak up. This same idea applies to storing hydrogen. The more surface area the
storage material has, the more hydrogen can be stored. Porous materials specially graphene stand out here. But it is
not as simple as possible. Here temperature and pressure play a very crucial role. Srinivas et al.[116] stated that the
amount of hydrogen is much lower in porous carbon materials and the number is almost 0.5 wt.% at ambient
temperature and high pressure[116]–[118]. But later they found a solution in graphene and stated disrupting the
tightly stacked layers of graphene at normal temperature creates extra surface area, boosting its ability to store
hydrogen[116].
Mortazavi et al.[39] worked on Pd-nanoparticles decorated with MWCNTs and observed the drastic disruption of
hydrogen storage with the increases in laser exposure time. This reveals how laser exposure, sometimes used during
cycling processes, can disrupt the MWCNT structure. This disruption creates pores in the material, effectively
reducing the available space for hydrogen atoms. As a result, the storage capacity drops significantly [39]. When the
time is independent the exposure to higher laser doses has the same phenomenon to reduce hydrogen storage
capacity[39], [45].
It is discussed before that graphene-like storage can store relatively large amounts of hydrogen because of its porous
structure which increases graphene's surface area. But for storage like Graphene Oxide Polyaniline (GO-PANI), this
scenario has totally changed. It may look promising on paper, but its capacity to hold hydrogen is just too low.
Storages like Graphene Oxide (GO) have relatively higher storage capacity than GO-PANI[98]. This is caused by
the polymerization of aniline at the surface of GO (Fig. 11b)[119]. It is anticipated that aniline is absorbed on GO
surfaces and forms the PANI chains [120], [121]. The effective utilization of GO layers for uptake storage can be
potentially limited by this polymerization process which leads to a reduced storage capacity[119].
Fig 12 represents hydrogen adsorption via PLA (Pulsed Laser Ablation) treatment in Ni-MWCNT (Ni-NPs).
The aim was to analyze the Ni-MWCNT conditions on associating with Ni-NP which was done by PLA (Pulsed
Laser Ablation) in water. Mehrabi et al.[45] found that the structure of MWCNTs when exposed to higher laser
doses leads to structural damage with larger pores. Which ultimately hampers the hydrogen storage capacity by
reducing it. It also reduced the desorption temperature as well. The report shows that faulty sites of MWCNT (with
pore size <2nm) strongly affect hydrogen trapping. Even though the density of Ni NPs in CNT is increased, the
hydrogen uptake drops with the higher dose of the laser. [45]
By CO2 activation, we can improve the defective cavities and improve the porosity enhancements of MWCNT. But
in case of excess activation, we have noticed a lower specific area and pore volume, and the reasons are deduced to
be thinned MWCNTs and a huge number of defects on the surface area. In the end, this resulted in a collapse of the
MWCNT (Acti-1100) surface.[48]
In the case of laser treatment of MWCNT, when the MWCNT is exposed for 120 mins the hydrogen storage ability
reduces drastically. It is theorized that the destruction of the MWCNT structure and the formation of large pores that
are incapable of hydrogen and resulting in lower hydrogen storage.[39]
As defects and damages influence the hydrogen nano storage capabilities (hydrogen storage capacity, desorption,
etc.), there are a couple of extrinsic/intrinsic factors come into play while producing the most beneficial hydrogen
storage system. Moderating the factors for a highly capable system (enhancement of temperature, optimized pore
distribution, etc.), will bring the expected results.
Temperature:
Temperature plays an important role in hydrogen storage systems. Temperature can regulate the rate of hydrogen
storage. Also, temperature creates an impact on the structure of nanomaterials. For chemical sorption materials
metal hydrides can be used which are the most famous group of materials. However chemical sorption materials
face main challenges which are operating temperature, and enhancing the charge-discharge kinetics [112].
The gap widths of smaller nanoparticles become smaller by enhancing the temperature. At a temperature which is
above 303K, the size dependence may be shown by lattice expansion found by Yamauchi et al.[122]. That means
with enhancing temperature the difference in the lattice expansion for the nanoparticles becomes smaller than that
for the bulk. There is a complex dependency on temperature for adsorption capacity [43].
Temperature can impact the stability of the hydrogen storage nanomaterial. Extremely high temperatures may lead
to structural changes or degradation of the material and also affect its hydrogen storage capacity. Also, by enhancing
the activation temperature storage capacity of activated MWNTs increased. Activation temperature increased to
900C, about 0.78wt.% storage capacity improved in activated MWNTs as a result of the defective structure
generated in the activated MWNTs which is observed by Lee et al. [48].
On the other hand, the hydrogen uptake properties of thermally exfoliated graphene oxide (EGR) are significantly
influenced by the exfoliation temperature which is investigated by Singh et al.[123]. Graphene oxide (GO) was
exfoliated at various temperatures ranging from 200C to 500C for hydrogen storing shown in Fig 13. These
temperature variations were found to affect the surface substantially and the structural characteristics of the resulting
samples [123].
Figure 13: Schematic of hydrogen uptake over exfoliated GO samples prepared at different temperatures[124].
The exfoliated graphene oxide samples exhibited a fluffy layered structure. As the exfoliation temperature increased,
the distance between graphene layers also increased resulting in a higher surface area and pore volume. The
maximum uptake of 3.12 wt% was observed for the GO sample exfoliated at 300C ( at -196C and 30 bar ). This is
because the GO sample exfoliated at 300C had the highest surface area, measuring 248 m 2/g, and the largest total
pore volume, at 1.64 cm3/g. The higher surface area offered the most active sites for monolayer adsorption, whereas
a larger pore volume attributed to the mesoporous structure, likely facilitated multilayer adsorption. The sample
exfoliated at the lowest temperature of 200C had the least effective layer separation, resulting in the lowest surface
area and pore volume. At a higher exfoliation temperature of 500C, partial damage to the layers likely contributed
to a reduction in both surface area and pore volume. [123]
Sheng et al. [46] found that different temperature shows different impacts on the specific surface area of double-
walled carbon nanotubes (DWCNTs) with Pd loading [46]. The largest surface area was found in DWCNTs with
2wt.% Pd at 400C (2%Pd/DWCNTs-400C). These results indicate that during the H 2 reduction process at high
temperatures, defect sites can be created near the metal particles. It is observed that the surface area of the
2%Pd/DWCNTs-500C sample is lower than the 2%Pd/DWCNTs-400C sample. This is because some of the
effective defects may undergo recrystallization during annealing processes at high temperatures.[46] For this reason,
effective defects are reduced resulting in a decrease in surface area. The effective surface area of the
2%Pd/DWCNTs-300C sample is also less than the 2%Pd/DWCNTs-400C sample. Because at lower temperatures
the catalytic activity may not be as better as compared to higher temperatures, resulting in less carbon gasification
and lower surface area. So, based on the surface area of DWCNTs samples are in the order : 2%Pd/DWCNTs-400°C
> 2%Pd/DWCNTs-500°C > 2%Pd/DWCNTs-300°C [46]. The specific surface area of 2%Pd/DWCNTs-400°C,
2%Pd/DWCNTs-500°C and 2%Pd/DWCNTs-300°C samples found respectively 209.1, 205.4 and 201.2 m2/g [46].
The order of the specific surface of the samples creates a similar order in hydrogen adsorption capacity because
hydrogen adsorption is enhanced by increasing the specific surface area. Hydrogen adsorption capacity found 1.85,
2.00, and 1.93 wt.% for sample 2%Pd/DWCNTs-300°C, 2%Pd/DWCNTs-400°C and 2%Pd/DWCNTs-500°C
respectively [46].
Pores:
The hydrogen storage system is very important for porous materials. They provide a formwork that allows hydrogen
molecules to be retained and released. There are various bodies of porous materials being pursued in the search for
ideal conditions for hydrogen storage such as activated carbons, Metal-Organic Frameworks (MOFs), Porous
Organic Polymers (POPs), porous carbon materials, etc.[83], [125], [126]. There are a few advantages and
disadvantages of pores which are directly connected to the amount of stored hydrogen.
Gogotsi et al.[127] presented the relationship between hydrogen uptake and pore size in Fig. 15, as the pores
become smaller, the surface area of any material increases. The maximum capacity is provided by the smallest
pores, with larger pores contributing only slightly. An increase in surface area means more areas for hydrogen
molecules to bond hence more storage[43], [127], [128]. Conversely, Yurum et al.[43] have found that there was a
high volumetric capacity associated with micro and nano-sized pores compared to meso-pores because hydrogen can
easily make bonds with micro and nano-sized pores rather than meso-pores[43]. In plainer terms, these smaller pores
can pack in more hydrogen per unit of volume. Kim et al.[97] some pore size distribution in Fig. 14. There were
some pores networked with different diameters. This variation had a huge impact on hydrogen storage capacity. As
discussed earlier, smaller pores give greater surface area for binding of hydrogen molecules hence they may lead to
increased capacity to store it[43], [127], [128]. However, the presence of a few bigger pores within the network that
can be seen from the figure might decrease the overall surface area and may finally hinder its storage capability.
Figure 15: Effect of pore size on hydrogen sorption[127], [128].
Pore size and distribution are critical even though they are necessary. This work is done by Mehrabi et al.[45] and
illustrates how dangerous large pores are. For instance, they discovered that many MWCNTs had bigger holes than
those produced after using high laser doses during material processing. These larger empty spaces, however, have
their own price, since they reduce the total surface area available for reaction on the catalysts' which reduces the
efficiency of catalysis per gram molecule or atom present in it[45]. Thus, a reduction in the surface area makes
hydrogen-binding sites fewer thereby decreasing storage ability as well. Another threat to MWCNTs has been
identified by Lee et al.[48]. Excessive activation is another danger that can take place besides high doses of laser
that also lead to the creation of extra pores and reduced surface area[48]. This is not only a mere contraction in
space, these abnormally larger pores could even crush the thinned layer of MWCNTs, which may result in
significant structural damage, hence affecting storage capacity.
Mortazavi et al.[39] highlighted more risk factors about pore continuity. Large mesopores are formed as a result of
microporous interconnectivity that can be caused by using high laser doses according to their research findings[39].
These newly created mesopores impose considerable harm on the storage material, increasing its structural
instability, and thereby diminishing hydrogen storage capacity.
Pressure:
The effect of pressure on nanomaterials in hydrogen storage systems is pivotal in determining their storage capacity,
efficiency, and operational performance. It is a critical factor influencing adsorption-desorption kinetics.
Generally, increasing the pressure enhances the amount of hydrogen that nanomaterials can absorb. At 77 K, Talyzin
et al. [129] showed that graphene samples with different surface areas enhanced gravimetric capacity by increasing
pressure, as shown in Fig 16. This is due to the higher concentration of hydrogen molecules at high pressures, which
drives more hydrogen onto the material’s pores and surface [129].
Figure 16: H2 adsorption isotherms for graphene samples with various surface areas at 77K [129].
On the other hand, at high pressure, nanomaterials may undergo structural changes that can either increase or
decrease their storage capacity. High pressure can also lead to the deformation of the nanomaterial’s lattice
structure. Excessive high pressure might cause pore collapse in nanomaterials, leading to reduced surface area and
decreased hydrogen storage capacity. So, an optimal pressure is required to make the best outcome. Also, Repeated
exposure to high pressure can lead to mechanical fatigue or degradation of the nanomaterials [130]. When
nanomaterials are subjected to high pressure repeatedly, the material experiences cycling loading, where the
pressure is applied and then released multiple times. This repeated pressure can cause the accumulation of internal
stresses within the material’s structure. These accumulated stresses can lead to the initiation and propagation of
microcracks. Habib et al. [130] found that when the high pressure is abruptly released, a significant phenomenon
occurs. The rapid decrease in pressure causes the internal gas volume to expand, initiating the desorption process.
As desorption progresses, bubbles form, leading to the create microcracks [130]. After this microcracks grow and
cracks, defects, or dislocations are generated. These structural defects can reduce the material’s ability to store
hydrogen efficiently over multiple cycles.
Also, nanomaterials might experience phase transition at high pressure, which can alter their hydrogen storage
properties [131]. Metal hydrides might transition to different crystalline phases under pressure, affecting how
hydrogen is absorbed and released. Palladium hydride, along with other metal hydrides, serves as an effective model
system for solute intercalation due to its rapid hydrogenation and dehydrogenation kinetics, which occur at easily
attainable transition pressure [131].
Baldi et al. [131] found that Pd nanocrystals exhibit size-dependent loading pressures when subjected to varying
hydrogen pressures. Smaller particles (below 15 nm) demonstrate more complex thermodynamics with the sloped
transition region, whereas larger particles (above 15 nm) exhibit abrupt phase transition, indicating distinct behavior
based on particle size [131]. Also, surface stress, due to hydrogen concentration at the surface of nanomaterials,
impacts the equilibrium loading pressure. Smaller particles tend to load at lower pressures compared to larger ones,
which might be attributed to the increased surface-to-volume ratio in smaller nanocrystals, affecting their
thermodynamics [131].
So, the relationship between pressure, particle size, and surface effects leads to varying behavior in loading and
unloading. Pressure not only influences thermodynamics but also interacts with size-dependent properties, leading to
different phase transition behaviors across different sizes.
Synergistic:
Two or more different nanomaterials or components often work together to form a synergistic effect that enhances
the performance of hydrogen storage beyond their individual contributions. This can be done by pairing metal
nanoparticles with porous supports and altering the nanostructure through alloying, doping, or strain engineering.
Figure 17: Hydrogen absorption and desorption curves of (a,c)Mg-Ni-V and (b,d) Mg-Ni nanocomposites[132].
Xie et al.[132] compared Mg-Ni-V nanocomposite doped hydrogen storage and Mg-Ni nanocomposite doped
hydrogen storage to highlight the synergistic effect on the storage system. At 423 K, the Mg-Ni-V nanocomposite
absorbed 1 and 2 wt% H2 within 10 and 60 min respectively which was two times greater than the Mg-Ni
nanocomposite[133]. At the temperature of 473 K within the span of 30 min, the Mg-Ni-V nanocomposite uptake
4.6 wt% H2 while 3.7, 3.5, and 1.3 wt% H2 can be adsorbed by individual catalysts like Mg-V[134], Mg-Ni, and Mg-
Co[135] respectively. After 60 min of absorption, the capacity of H 2 reached 5.1 wt% for the Mg-Ni-V
nanocomposite which is superior to those of Mg-Ni and Mg-V nanocomposites with the capacity of 4.8 and 3.8 wt%
respectively[134]. The same scenario was also found at the temperature of 523 K, 573 K, 623 K, and 673 K. For the
desorption process, the synergistic effect plays a significant role. At the temperature of 673 K within the span of 5
min, the desorption from the Mg-Ni-V nanocomposite was 5.9 wt% H 2 which was higher than the desorption from
Mg-Ni nanocomposite that was 5.5 wt% H 2. At 623 K, the desorption reached 5 and 3.5 wt% H 2 for the Mg-Ni-V
and Mg-Ni nanocomposite within the sorption time of 10 min. When the temperature decreased to 573 K, the
desorption amount was 3 and 2.6 wt% H 2 for the time of 30 min and 3.8 and 2.6 wt% H 2 for the time of 60 min for
the nanocomposite of Mg-Ni-V and Mg-Ni respectively[136]. The same went for other temperatures. All of this
information is graphically presented in Fig. 17[132]. Yang et al.[137] also found this synergistic effect when
incorporating Pt-MWCNT into the MOF-5. At 298 K and 100 bar pressure, the absorption amount for the total
system of Pt-MWCNT@MOF-5 was 0.55 wt% H 2 which was greater than the individual absorption amount of the
parent (0.3 wt%) and incorporated nanocomposite (0.2 wt%) of H 2. Musyoka et al.[138] used Cr-based MOF (MIL-
101) to incorporate zeolite templated caron (ZTC) to highlight this effect. Their findings are presented in Fig. 18.
The absorption amount for the system of ZTC@MIL-101(Cr) was 2.55 et% H 2 at the temperature and pressure of 77
K and 1 bar. However, the individual performances of MIL-101 (Cr) and ZTC were 1.91 and 2.39 wt% H 2 at the
same temperature and pressure.
Figure 18: H2-sorption isotherms of MIL-101 (Cr), ZT, and ZTC@MIL-101 samples at 77 K[138].
The long-term usage of storage materials in industry which is expected, has repeated hydrogenation and
dehydrogenation cycles for a long period. Which causes effects (reduced hydrogen storage capacity, reduced rate of
hydration or dehydration, etc.) in some nano-materials, and in others, it isn’t noticeable.
While testing the mechanism of capacity reduction in long exposure to cycling, the hydrogen storage capacity is
classified as effective capacity and absolute capacity. Absolute capacity is the maximum amount of hydrogen that a
system can preserve and release while being in a long dehydrogenation process, determined by thermal factors. On
the other hand, the capability to store and release hydrogen in a short operation time is an effective capacity. It is
determined by thermodynamic and kinetic factors. A system may hold high absolute capacity but low effective
capacity, resulting in poor de/hydrogenation kinetics. Moreover, cyclic stability relies on many intrinsic and
extrinsic factors (phase composition, crystal structure disorders, sintering and pulverization, grain size, purity of H 2,
etc)[139].
Cycling can pulverize the bulk of the alloy and reduce the particle size distribution of the metal composition as can
be seen in LaNi4.25Al0.75. The absorption kinetics and adsorption/desorption were tested at 363K,383 K, and 403K for
1000 cycles. The storage capacity absorption kinetics were slightly altered after long exposure to the system, on the
other hand, degradation of absorption kinetics was huge. Particle size distribution was narrower (0.36–103 m)
compared to primary activation (0.78–301 m), with volume mean diameter reducing from 34.81 to 11.34 m. This
phenomenon was explained by the pulverization and the impurities from charging hydrogen that were subjected to
cycling.[140]
Cycling may also cause agglomeration, weakened hydrogen spillover effect, and decreased storage capacity in some
cases. Palladium nanoparticle decorated nitrogen-doped graphene (Pd/N-rGO) got a lot of attention because of its
high absorption capacity at moderate conditions. While testing its hydrogen absorption-desorption cycle
performance, which is essential for their practical application, it was found that the storage capacity dropped from
2.9% to 0.8% at 25℃ and 4 MPa pressure. The palladium nanoparticles were shed from the composition sheet after
the test and agglomerated. Which resulted in a weak hydrogen spillover effect and reduced hydrogen storage
capacity. The hydrogen storage capacity is reduced by 72% due to the agglomeration, and shedding weakens the
hydrogen spillover effect. Also, the nitrogen content reduced the hydrogen absorption sites. These phenomena
happened in only three cycles of adsorption/ desorption.[141]
While comparing V, Fe, and Ti with ferrovanadium (FeV) and Ti sponge, under the long-term cyclic experiment of
1000 de-/hydrogenation, both groups of the material showed almost similar activation and degradation behavior. A
1.5 mass% H2 was recorded for the pure/pristine material and 1.4 mass% H 2 was stored with FeV and Ti sponge
when examined after 1000 cycles. This process shows that by replacing the raw material with similar quality
materials (V and Ti substituted by FeV and Ti sponge) we can reduce cost by up to 83%.[142]
Hydrogen absorption-desorption cycle performance, which is essential for their practical application in the industry
shows the varied results of selection materials, with mixed outcomes (some drastic negative impact and others being
fruitful). The moderation of materials in long cyclic exposure ensures the benefits of industrial use.
Future Outlook:
The performance of a hydrogen storage system is strongly dependent on the internal structure of nanomaterials.
However structural damage affects its performance and it reduces the lifetime of the storage system. The damage in
structure like breaking bonds, forming defects, forming pores, etc. can occur due to high temperature, pressure, or
continuous cycling of the system. In continuous cycling, nano-structured materials experience internal stress leading
to potential fracturing. Structural damage can be protected during continuous cycling by developing protective
coatings and surface treatments. Structural integrity can be enhanced by integrating nanomaterials into a matrix of
another material. Investigating materials capable of self-repairing to extend the operational life of hydrogen storage
systems. Also, designing the storage system in a way that distributes mechanical stress more evenly across the
material can reduce localized damage. On the other hand, thermodynamic stability is a big issue for hydrogen
storage systems. Thermodynamic stability reduces the performance of hydrogen storage systems. Creating
nanostructured materials with tailored particle sizes and morphologies can influence thermodynamic stability. Also,
developing composite materials that combine different nanomaterials can leverage their individual thermodynamic
properties to achieve balanced stability. The kinetics of hydrogen charging and discharging in nanostructured metal
hydrides for hydrogen storage becomes sluggish with continuous activity which reduces the hydrogen storage
performance. Incorporating catalytic additives into nanostructured metal hydrides can significantly improve
hydrogen adsorption and desorption rates. Catalysts such as Pt, Pd, or transition metal oxides can facilitate faster
adsorption and desorption by lowering activation energy barriers. Also, utilizing advanced synthesis techniques to
control the size, shape, and distribution of nanomaterials can improve kinetic properties. Methods like chemical
vapor deposition (CVD), atomic layer deposition (ALD), and electrochemical deposition can create nanomaterials
with superior kinetic behavior. Researchers are using different types of catalysts like magnesium, platinum, rare
earth, etc.-based materials to improve the uptake rate of hydrogen in storage systems. However, there is a massive
gap in finding a suitable way to release 100% or close to 100% hydrogen from the storage system because current
nanostructured storage systems don’t release near about 100% hydrogen. Researchers can use computational
theories, DFT, molecular dynamics, data-driven technology, etc. to discover the solutions to these release-based
problems and apply the proper computational findings in the laboratory. After several cycles of using porous storage
systems, the pores started to increase their radius and reduce their storage capacity. Researchers can use different
self-healing materials in the porous storage systems to observe how they behave on the targeted sites. As Hydrogen
adsorption depends on multiple variables, the non-standardized test conditions used in many research groups and
individuals made it more complicated to locate the hydrogen adsorption dependency and characteristics. Also, a
couple of approaches from researchers caused problematic criteria/situations. Generally, most of the research was
examined focusing on one characteristic. This is why it is much needed to have research following multiple
characteristics and a standard environment for examining and discovering hydrogen adsorption’s dependency and
characteristics. Moving on to the importance of combining two nanostructures, it is acquired that MgH 2 and LiBH4
need to be inserted following orders then the lack of suitable solvents shown to achieve 2LiBH 4-MgH2 composite.
An inhomogeneous distribution of LiBH4 restricts the mix of LiBH4 and MgH2 resulting in blocking the pores from
the primary infiltration of MgH2 and in the end resulting in inferior effective results like low dehydrogenation, low
hydrogen kinetics, and weak reversibility. This is the reason a new and improved synthesis process is greatly desired
and much needed but there is still a lack of solution for this MgH 2, LiBH4 synthesis problem. Another developing
area is managing the pore size and distribution within nanomaterials to prevent issues like large pores that reduce
hydrogen storage capacity. Utilizing controlled activation processes (like CO 2 activation) to create a uniform pore
size distribution, avoiding the formation of excessively large pores that decrease the surface area and storage
capacity. Techniques to repair or mitigate the effects of over-activation, such as post-synthesis treatments, could also
be beneficial.
Conducting comprehensive experimental studies to validate theoretical predictions and understand the real-world
performance of nanomaterial-based hydrogen storage systems. Experiments should focus on long-term cycling
stability, thermodynamic behavior, kinetic performance, hydrogen adsorption-desorption rate, and its dependency
and characteristics.
Conclusion:
Nanostructured storage materials represent a significant advancement in addressing hydrogen storage challenges.
These materials offer substantially larger surface areas compared to bulk materials, enabling higher hydrogen
storage capacities and faster absorption-desorption rates. Additionally, they provide enhanced durability over
conventional storage systems. However, despite these advantages, nanostructured hydrogen storage systems also
face limitations and disadvantages that must be addressed to fully realize their potential. Structural damage is a
significant challenge in hydrogen storage systems, affecting the performance and efficiency of materials like multi-
walled carbon nanotubes (MWCNTs) and palladium nanoparticles, LiAlH 4 lattice. High temperatures, laser doses,
and continuous cycling can create defects, reduce surface area, and disrupt the material structure, leading to
decreased hydrogen storage capacity and durability. Minimizing these damages is essential for optimizing hydrogen
uptake and enhancing the reversibility of storage materials. On the other hand, optimizing hydrogen storage systems
involves balancing kinetics, thermodynamics, and materials stability. Enhancing kinetics through nanostructuring
and appropriate synthesis methods is crucial, but maintaining thermodynamic stability remains challenging.
Effective hydrogen storage relies on maximizing surface era and managing pore size, yet practical issues like heat
management, structural degradation, and high-release temperature hinder performance. Innovative approaches, such
as catalyst doping and optimized nanostructure engineering, are essential to improve reversibility and efficiency.
Despite promising materials like MWCNTs and LiAlH 4, achieving practical, high-capacity hydrogen storage
requires overcoming these multifaceted challenges. Temperature and pore characteristics are crucial for optimizing
hydrogen storage systems. Proper temperature management can enhance adsorption rates and create beneficial
defect sites, but excessively high temperatures can degrade nanomaterial structures and reduce storage capacity.
Optimal temperatures maximize storage efficiency and minimize material damage. Similarly, smaller pore sizes in
porous materials enhance storage capacity by increasing surface area, while larger pores from excessive activation
or higher laser doses reduce it. Balancing pore size and distribution is essential for achieving efficient hydrogen
storage in materials like activated carbons, MOFs, POPs, and MWCNTs. The performance and lifespan of hydrogen
storage systems are heavily influenced by the internal structure of nanomaterials, which can be damaged by higher
temperatures, pressures, and continuous cycling. Structural damage, such as breaking bonds and forming defects or
pores, deteriorates storage capacity. Research must focus on integrating self-healing materials to address this issue.
Increased laser exposure results in more damage and amorphous carbons, affecting storage efficiency. Standardized
testing conditions are crucial to understanding hydrogen adsorption characteristics. Combining MgH 2 and LiBH4 is
challenging due to inhomogeneous distribution, necessitating improved synthesis processes to enhance hydrogen
kinetics and reversibility.
References:
[1] U. Nations, “World population prospects 2019,” Vol (ST/ESA/SE. A/424) Dep. Econ. Soc. Aff. Popul. Div.,
2019.
[2] K. Alanne and S. Cao, “An overview of the concept and technology of ubiquitous energy,” Appl. Energy,
vol. 238, no. November 2018, pp. 284–302, 2019, doi: 10.1016/j.apenergy.2019.01.100.
[4] J. Zhang, Techno-economic analysis and optimization of distributed energy systems. Mississippi State
University, 2018.
[5] J. C. Radcliffe, “The water energy nexus in Australia – The outcome of two crises,” Water-Energy Nexus,
vol. 1, no. 1, pp. 66–85, 2018, doi: 10.1016/j.wen.2018.07.003.
[6] A. L. Dicks and D. A. J. Rand, Fuel cell systems explained. John Wiley & Sons, 2018.
[7] M. Boudellal, Power-to-gas: Renewable hydrogen economy for the energy transition. Walter de Gruyter
GmbH & Co KG, 2023.
[8] A. Campen, K. Mondal, and T. Wiltowski, “Separation of hydrogen from syngas using a regenerative
system,” Int. J. Hydrogen Energy, vol. 33, pp. 332–339, Jan. 2008, doi: 10.1016/j.ijhydene.2007.07.016.
[9] A. Shahsavari and M. Akbari, “Potential of solar energy in developing countries for reducing energy-related
emissions,” Renew. Sustain. Energy Rev., vol. 90, pp. 275–291, Jul. 2018, doi: 10.1016/j.rser.2018.03.065.
[11] D. Andjelkovic, B. Antić, M. Vujanić, M. Subotić, and L. Radovanovic, “The perspectives of applying
ethanol as an alternate fuel,” Energy Sources, Part B Econ. Planning, Policy, vol. 12, pp. 1–10, May 2017,
doi: 10.1080/15567249.2012.683930.
[12] M. Balat and G. Ayar, “Biomass Energy in the World, Use of Biomass and Potential Trends,” Energy
Sources, vol. 27, no. 10, pp. 931–940, Jul. 2005, doi: 10.1080/00908310490449045.
[13] M. Balat, “Potential importance of hydrogen as a future solution to environmental and transportation
problems,” Int. J. Hydrogen Energy, vol. 33, no. 15, pp. 4013–4029, 2008, doi:
10.1016/j.ijhydene.2008.05.047.
[14] J. KANDASAMY, I. Gokalp, F. San, H. Ustunel, A. Yontar, and E. Eroğlu, Exploring the Potential of
Boron Based Materials for Energy and Hydrogen Generation: An Overview of Some Studies in Türkiye.
2022.
[15] S. Yolcular, “Hydrogen Production for Energy Use in European Union Countries and Turkey,” Energy
Sources, Part A Recover. Util. Environ. Eff., vol. 31, no. 15, pp. 1329–1337, Aug. 2009, doi:
10.1080/15567030802089615.
[16] C. Barboza, “Towards a Renewable Energy Decision Making Model,” Procedia Comput. Sci., vol. 44, pp.
568–577, 2015, doi: https://doi.org/10.1016/j.procs.2015.03.017.
[17] S. E. Hosseini and M. A. Wahid, “Hydrogen production from renewable and sustainable energy resources:
Promising green energy carrier for clean development,” Renew. Sustain. Energy Rev., vol. 57, pp. 850–866,
2016, doi: https://doi.org/10.1016/j.rser.2015.12.112.
[18] M. Chen et al., “Synergy between metallic components of MoNi alloy for catalyzing highly efficient
hydrogen storage of MgH2,” Nano Res., vol. 13, no. 8, pp. 2063–2071, 2020, doi: 10.1007/s12274-020-
2808-7.
[19] C. Duan et al., “Novel core–shell structured MgH2/AlH3@CNT nanocomposites with extremely high
dehydriding–rehydriding properties derived from nanoconfinement,” J. Mater. Chem. A, vol. 9, no. 17, pp.
10921–10932, 2021, doi: 10.1039/D1TA01938D.
[20] J. Kang et al., “Realizing Two-Electron Transfer in Ni(OH)(2) Nanosheets for Energy Storage.,” J. Am.
Chem. Soc., vol. 144, no. 20, pp. 8969–8976, May 2022, doi: 10.1021/jacs.1c13523.
[21] S. Liu and J. Shui, “Mechanism and properties of emerging nanostructured hydrogen storage materials,”
Batter. Energy, vol. 1, no. 4, 2022, doi: 10.1002/bte2.20220033.
[22] P. Ferreira-Aparicio, J. J. Conde, and A. M. Chaparro, “3 - Hydrogen storage options for portable fuel-cell
systems,” P. Ferreira-Aparicio and A. M. B. T.-P. H. E. S. Chaparro, Eds., Academic Press, 2018, pp. 41–
50. doi: https://doi.org/10.1016/B978-0-12-813128-2.00003-6.
[23] B. G. Pollet, I. Staffell, J. L. Shang, and V. Molkov, “22 - Fuel-cell (hydrogen) electric hybrid vehicles,” R.
B. T.-A. F. and A. V. T. for I. E. P. Folkson, Ed., Woodhead Publishing, 2014, pp. 685–735. doi:
https://doi.org/10.1533/9780857097422.3.685.
[24] PNNL-29731, “Hydrogen Incident Examples - Select Summaries of Hydrogen Incidents from the
H2tools.org Lessons Learned Database,” Hydrog. Saf. Tools, vol. PNNL-29731, no. March, 2020.
[25] M. Aziz, “Liquid Hydrogen: A Review on Liquefaction, Storage, Transportation, and Safety,” Energies, vol.
14, p. 5917, Sep. 2021, doi: 10.3390/en14185917.
[26] T. Zhang, J. Uratani, Y. Huang, L. Xu, S. Griffiths, and Y. Ding, “Hydrogen liquefaction and storage:
Recent progress and perspectives,” Renew. Sustain. Energy Rev., vol. 176, p. 113204, 2023, doi:
https://doi.org/10.1016/j.rser.2023.113204.
[27] R. Krishna et al., “Hydrogen Storage for Energy Application,” J. Liu, Ed., Rijeka: IntechOpen, 2012, p. Ch.
10. doi: 10.5772/51238.
[28] R. K. Ahluwalia, J. K. Peng, H. S. Roh, T. Q. Hua, C. Houchins, and B. D. James, “Supercritical cryo-
compressed hydrogen storage for fuel cell electric buses,” Int. J. Hydrogen Energy, vol. 43, no. 22, pp.
10215–10231, 2018, doi: https://doi.org/10.1016/j.ijhydene.2018.04.113.
[29] Z. Yanxing, M. Gong, Z. Yuan, X. Dong, and S. Jun, “Thermodynamics analysis of hydrogen storage based
on compressed gaseous hydrogen, liquid hydrogen and cryo-compressed hydrogen,” Int. J. Hydrogen
Energy, vol. 44, May 2019, doi: 10.1016/j.ijhydene.2019.04.207.
[30] A. Behvar, M. Haghshenas, and M. B. Djukic, “Hydrogen embrittlement and hydrogen-induced crack
initiation in additively manufactured metals: A critical review on mechanical and cyclic loading,” Int. J.
Hydrogen Energy, vol. 58, pp. 1214–1239, Mar. 2024, doi: 10.1016/j.ijhydene.2024.01.232.
[31] P. C. Okonkwo et al., “A focused review of the hydrogen storage tank embrittlement mechanism process,”
Int. J. Hydrogen Energy, vol. 48, no. 35, pp. 12935–12948, 2023, doi:
https://doi.org/10.1016/j.ijhydene.2022.12.252.
[32] S. K. Dwivedi and M. Vishwakarma, “Hydrogen embrittlement in different materials: A review,” Int. J.
Hydrogen Energy, vol. 43, no. 46, pp. 21603–21616, 2018, doi:
https://doi.org/10.1016/j.ijhydene.2018.09.201.
[33] M. P. Mudoi, P. Sharma, and A. S. Khichi, “A review of gas adsorption on shale and the influencing factors
of CH4 and CO2 adsorption,” J. Pet. Sci. Eng., vol. 217, p. 110897, 2022, doi:
https://doi.org/10.1016/j.petrol.2022.110897.
[34] M. Atif, H. Z. Haider, R. Bongiovanni, M. Fayyaz, T. Razzaq, and S. Gul, “Physisorption and chemisorption
trends in surface modification of carbon black,” Surfaces and Interfaces, vol. 31, p. 102080, 2022, doi:
https://doi.org/10.1016/j.surfin.2022.102080.
[35] M. Hirscher and B. Panella, “Nanostructures with High Surface Area for Hydrogen Storage.,” J. Alloys
Compd., vol. 404–406, pp. 399–401, Dec. 2005, doi: 10.1016/j.jallcom.2004.11.109.
[36] F. Ding and B. I. Yakobson, “Challenges in hydrogen adsorptions: From physisorption to chemisorption,”
Front. Phys., vol. 6, no. 2, pp. 142–150, 2011, doi: 10.1007/s11467-011-0171-6.
[37] W. Liu et al., “Size-dependent hydrogen trapping in palladium nanoparticles,” J. Mater. Chem. A, vol. 9, no.
16, pp. 10354–10363, 2021, doi: 10.1039/d0ta12174f.
[38] N. A. Sazelee and M. Ismail, “Recent advances in catalyst-enhanced LiAlH4 for solid-state hydrogen
storage: A review,” Int. J. Hydrogen Energy, vol. 46, no. 13, pp. 9123–9141, 2021, doi:
10.1016/j.ijhydene.2020.12.208.
[39] S. Z. Mortazavi, P. Parvin, A. Reyhani, R. Malekfar, and S. Mirershadi, “Hydrogen storage property of laser
induced Pd-nanoparticle decorated multi-walled carbon nanotubes,” RSC Adv., vol. 3, no. 5, pp. 1397–1409,
2013, doi: 10.1039/c2ra22224h.
[41] H. Shao, G. Xin, J. Zheng, X. Li, and E. Akiba, “Nanotechnology in Mg-based materials for hydrogen
storage,” Nano Energy, vol. 1, no. 4, pp. 590–601, 2012, doi: 10.1016/j.nanoen.2012.05.005.
[42] L. Liu et al., “Metal Hydride Composite Structures for Improved Heat Transfer and Stability for Hydrogen
Storage and Compression Applications,” Inorganics, vol. 11, no. 5, pp. 1–17, 2023, doi:
10.3390/inorganics11050181.
[43] Y. Yürüm, A. Taralp, and T. N. Veziroglu, “Storage of hydrogen in nanostructured carbon materials,” Int. J.
Hydrogen Energy, vol. 34, no. 9, pp. 3784–3798, 2009, doi: 10.1016/j.ijhydene.2009.03.001.
[44] M. Aghababaei, A. A. Ghoreyshi, and K. Esfandiari, “Optimizing the conditions of multi-walled carbon
nanotubes surface activation and loading metal nanoparticles for enhanced hydrogen storage,” Int. J.
Hydrogen Energy, vol. 45, no. 43, pp. 23112–23121, 2020, doi: 10.1016/j.ijhydene.2020.06.201.
[45] M. Mehrabi, A. Reyhani, P. Parvin, and S. Z. Mortazavi, “Surface structural alteration of multi-walled
carbon nanotubes decorated by nickel nanoparticles based on laser ablation/chemical reduction methods to
enhance hydrogen storage properties,” Int. J. Hydrogen Energy, vol. 44, no. 7, pp. 3812–3823, 2019, doi:
10.1016/j.ijhydene.2018.12.122.
[46] Q. Sheng, H. Wu, D. Wexler, and H. Liu, “Effects of reducing temperatures on the hydrogen storage
capacity of double-walled carbon nanotubes with Pd loading,” J. Nanosci. Nanotechnol., vol. 14, no. 6, pp.
4706–4709, 2014, doi: 10.1166/jnn.2014.8251.
[47] Rafi-Ud-Din, L. Zhang, L. Ping, and Q. Xuanhui, “Catalytic effects of nano-sized TiC additions on the
hydrogen storage properties of LiAlH4,” J. Alloys Compd., vol. 508, no. 1, pp. 119–128, 2010, doi:
10.1016/j.jallcom.2010.08.008.
[48] S. Y. Lee and S. J. Park, “Effect of temperature on activated carbon nanotubes for hydrogen storage
behaviors,” Int. J. Hydrogen Energy, vol. 35, no. 13, pp. 6757–6762, 2010, doi:
10.1016/j.ijhydene.2010.03.114.
[49] C. Erkey and M. Türk, “Thermodynamics and kinetics of adsorption of metal complexes on surfaces from
supercritical solutions,” Supercrit. Fluid Sci. Technol., vol. 8, pp. 73–127, Jan. 2021, doi: 10.1016/B978-0-
444-64089-5.00047-0.
[50] X. Zhang et al., “Optimization of the Pore Structures of MOFs for Record High Hydrogen Volumetric
Working Capacity,” Adv. Mater., vol. 32, no. 17, pp. 1–6, 2020, doi: 10.1002/adma.201907995.
[51] Y. J. Heo, S. H. Yeon, and S. J. Park, “Defining contribution of micropore size to hydrogen physisorption
behaviors: A new approach based on DFT pore volumes,” Carbon N. Y., vol. 143, pp. 288–293, 2019, doi:
10.1016/j.carbon.2018.11.019.
[52] K. Suresh, D. Aulakh, J. Purewal, D. J. Siegel, M. Veenstra, and A. J. Matzger, “Optimizing Hydrogen
Storage in MOFs through Engineering of Crystal Morphology and Control of Crystal Size,” J. Am. Chem.
Soc., vol. 143, no. 28, pp. 10727–10734, 2021, doi: 10.1021/jacs.1c04926.
[53] U. Eberle, M. Felderhoff, and F. Schüth, “Chemical and physical solutions for hydrogen storage,” Angew.
Chemie - Int. Ed., vol. 48, no. 36, pp. 6608–6630, 2009, doi: 10.1002/anie.200806293.
[54] M. Hirscher, L. Zhang, and H. Oh, “Nanoporous adsorbents for hydrogen storage,” Appl. Phys. A Mater.
Sci. Process., vol. 129, no. 2, pp. 1–10, 2023, doi: 10.1007/s00339-023-06397-4.
[55] D. Czarna-Juszkiewicz, J. Cader, and M. Wdowin, “From coal ashes to solid sorbents for hydrogen storage,”
J. Clean. Prod., vol. 270, p. 122355, 2020, doi: 10.1016/j.jclepro.2020.122355.
[56] Q. Weng et al., “Hydrogen Storage in Carbon and Oxygen Co-Doped Porous Boron Nitrides,” Adv. Funct.
Mater., vol. 31, no. 4, pp. 7–9, 2021, doi: 10.1002/adfm.202007381.
[58] E. Rozzi, F. D. Minuto, and A. Lanzini, “Dynamic modeling and thermal management of a Power-to-Power
system with hydrogen storage in microporous adsorbent materials,” J. Energy Storage, vol. 41, no. April, p.
102953, 2021, doi: 10.1016/j.est.2021.102953.
[59] J. Wang, Y. Chen, L. Yuan, M. Zhang, and C. Zhang, “Scandium decoration of boron doped porous
graphene for high-capacity hydrogen storage,” Molecules, vol. 24, no. 13, 2019, doi:
10.3390/molecules24132382.
[60] M. D. Allendorf et al., “An assessment of strategies for the development of solid-state adsorbents for
vehicular hydrogen storage,” Energy Environ. Sci., vol. 11, no. 10, pp. 2784–2812, 2018, doi:
10.1039/c8ee01085d.
[61] G. Li et al., “Hydrogen storage in Pd nanocrystals covered with a metal-organic framework,” Nat. Mater.,
vol. 13, no. 8, pp. 802–806, 2014, doi: 10.1038/nmat4030.
[63] S. Y. Lee, J. H. Lee, Y. H. Kim, J. W. Kim, K. J. Lee, and S. J. Park, “Recent Progress Using Solid-State
Materials for Hydrogen Storage: A Short Review,” Processes, vol. 10, no. 2, pp. 1–19, 2022, doi:
10.3390/pr10020304.
[64] M. M. Rampai, C. B. Mtshali, N. S. Seroka, and L. Khotseng, “Hydrogen production, storage, and
transportation: recent advances,” RSC Adv., vol. 14, no. 10, pp. 6699–6718, 2024, doi: 10.1039/d3ra08305e.
[65] L. G. Scanlon et al., “Hydrogen storage based on physisorption,” J. Phys. Chem. B, vol. 113, no. 14, pp.
4708–4717, 2009, doi: 10.1021/jp809097v.
[66] E. Rivard, M. Trudeau, and K. Zaghib, “Hydrogen storage for mobility: A review,” Materials (Basel)., vol.
12, no. 12, 2019, doi: 10.3390/ma12121973.
[67] Z. Q. Zuo, P. J. Sun, W. B. Jiang, X. J. Qin, P. Li, and Y. H. Huang, “Thermal stratification suppression in
reduced or zero boil-off hydrogen tank by self-spinning spray bar,” Int. J. Hydrogen Energy, vol. 44, no. 36,
pp. 20158–20172, 2019, doi: 10.1016/j.ijhydene.2019.05.241.
[68] Y. Zhang, S. Wu, L. Wang, and X. Zhang, “Chemisorption solid materials for hydrogen storage near
ambient temperature: a review,” Front. Energy, vol. 17, no. 1, pp. 72–101, 2023, doi: 10.1007/s11708-022-
0835-7.
[69] S. Schaefer, V. Fierro, A. Szczurek, M. T. Izquierdo, and A. Celzard, “Physisorption, chemisorption and
spill-over contributions to hydrogen storage,” Int. J. Hydrogen Energy, vol. 41, no. 39, pp. 17442–17452,
2016, doi: https://doi.org/10.1016/j.ijhydene.2016.07.262.
[70] Y. Xu, Y. Li, L. Gao, Y. Liu, and Z. Ding, “Advances and Prospects of Nanomaterials for Solid-State
Hydrogen Storage,” Nanomaterials, vol. 14, no. 12. 2024. doi: 10.3390/nano14121036.
[71] B. W. J. Chen and M. Mavrikakis, “Effects of composition and morphology on the hydrogen storage
properties of transition metal hydrides: Insights from PtPd nanoclusters,” Nano Energy, vol. 63, p. 103858,
2019, doi: https://doi.org/10.1016/j.nanoen.2019.103858.
[72] M. R. Ajayakumar et al., “Neutral Organic Radical Formation by Chemisorption on Metal Surfaces,” J.
Phys. Chem. Lett., vol. 11, no. 10, pp. 3897–3904, 2020, doi: 10.1021/acs.jpclett.0c00269.
[73] T. H. Le, M. P. Kim, C. H. Park, and Q. N. Tran, “Recent Developments in Materials for Physical Hydrogen
Storage: A Review,” Materials (Basel)., vol. 17, no. 3, 2024, doi: 10.3390/ma17030666.
[74] Y. Kojima, “Hydrogen storage materials for hydrogen and energy carriers,” Int. J. Hydrogen Energy, vol.
44, no. 33, pp. 18179–18192, 2019, doi: 10.1016/j.ijhydene.2019.05.119.
[75] D. S. Pyle, E. M. A. Gray, and C. J. Webb, “Hydrogen storage in carbon nanostructures via spillover,” Int.
J. Hydrogen Energy, vol. 41, no. 42, pp. 19098–19113, 2016, doi: 10.1016/j.ijhydene.2016.08.061.
[76] W. C. Conner and J. L. Falconer, “Spillover in Heterogeneous Catalysis,” Chem. Rev., vol. 95, no. 3, pp.
759–788, 1995, doi: 10.1021/cr00035a014.
[77] C. C. Huang, N. W. Pu, C. A. Wang, J. C. Huang, Y. Sung, and M. Der Ger, “Hydrogen storage in graphene
decorated with Pd and Pt nano-particles using an electroless deposition technique,” Sep. Purif. Technol., vol.
82, no. 1, pp. 210–215, 2011, doi: 10.1016/j.seppur.2011.09.020.
[78] N. R. Stuckert, L. Wang, and R. T. Yang, “Characteristics of hydrogen storage by spillover on Pt-doped
carbon and catalyst-bridged metal organic framework,” Langmuir, vol. 26, no. 14, pp. 11963–11971, 2010,
doi: 10.1021/la101377u.
[79] A. Ansón et al., “Hydrogen capacity of palladium-loaded carbon materials,” J. Phys. Chem. B, vol. 110, no.
13, pp. 6643–6648, 2006, doi: 10.1021/jp057206c.
[81] H. Cheng, L. Chen, A. C. Cooper, X. Sha, and G. P. Pez, “Hydrogen spillover in the context of hydrogen
storage using solid-state materials,” Energy Environ. Sci., vol. 1, no. 3, pp. 338–354, 2008, doi:
10.1039/b807618a.
[82] A. V. Krasheninnikov and R. M. Nieminen, “Attractive interaction between transition-metal atom impurities
and vacancies in graphene: A first-principles study,” Theor. Chem. Acc., vol. 129, no. 3–5, pp. 625–630,
2011, doi: 10.1007/s00214-011-0910-3.
[83] J. Huang et al., “Revealing contribution of pore size to high hydrogen storage capacity,” Int. J. Hydrogen
Energy, vol. 43, no. 39, pp. 18077–18082, 2018, doi: 10.1016/j.ijhydene.2018.08.027.
[86] C. Pratthana and K. F. Aguey-Zinsou, “LiAlH4Nanoparticles Encapsulated within Metallic Titanium Shells
for Enhanced Hydrogen Storage,” ACS Appl. Nano Mater., vol. 5, no. 11, pp. 16413–16422, 2022, doi:
10.1021/acsanm.2c03483.
[87] S. S. Han, J. K. Kang, H. M. Lee, A. C. T. Van Duin, and W. A. Goddard, “The theoretical study on
interaction of hydrogen with single-walled boron nitride nanotubes. I. The reactive force field ReaxFF HBN
development,” J. Chem. Phys., vol. 123, no. 11, 2005, doi: 10.1063/1.1999628.
[88] F. Chen and J. Chen, “Storage of hydrogen and lithium in inorganic nanotubes and nanowires,” J. Mater.
Res., vol. 21, no. 11, pp. 2744–2757, 2006, doi: 10.1557/jmr.2006.0337.
[89] J. Chen, S. L. Li, Z. L. Tao, Y. T. Shen, and C. X. Cui, “Titanium disulfide nanotubes as hydrogen-storage
materials,” J. Am. Chem. Soc., vol. 125, no. 18, pp. 5284–5285, 2003, doi: 10.1021/ja034601c.
[90] M. Zarezadeh Mehrizi, J. Abdi, M. Rezakazemi, and E. Salehi, “A review on recent advances in hollow
spheres for hydrogen storage,” Int. J. Hydrogen Energy, vol. 45, no. 35, pp. 17583–17604, 2020, doi:
10.1016/j.ijhydene.2020.04.201.
[91] E. A. Crespo, M. Ruda, S. Ramos De Debiaggi, E. M. Bringa, F. U. Braschi, and G. Bertolino, “Hydrogen
absorption in Pd nanoparticles of different shapes,” Int. J. Hydrogen Energy, vol. 37, no. 19, pp. 14831–
14837, 2012, doi: 10.1016/j.ijhydene.2011.12.075.
[92] W. Li, C. Li, H. Ma, and J. Chen, “Magnesium nanowires: Enhanced kinetics for hydrogen absorption and
desorption,” J. Am. Chem. Soc., vol. 129, no. 21, pp. 6710–6711, 2007, doi: 10.1021/ja071323z.
[93] A. F. Gross, C. C. Ahn, S. L. Van Atta, P. Liu, and J. J. Vajo, “Fabrication and hydrogen sorption behaviour
of nanoparticulate MgH 2 incorporated in a porous carbon host,” Nanotechnology, vol. 20, no. 20, 2009, doi:
10.1088/0957-4484/20/20/204005.
[94] R. W. P. Wagemans, J. H. Van Lenthe, P. E. De Jongh, A. J. Van Dillen, and K. P. De Jong, “Hydrogen
storage in magnesium clusters: Quantum chemical study,” J. Am. Chem. Soc., vol. 127, no. 47, pp. 16675–
16680, 2005, doi: 10.1021/ja054569h.
[95] A. Schneemann et al., “Nanostructured Metal Hydrides for Hydrogen Storage,” Chem. Rev., vol. 118, no.
22, pp. 10775–10839, 2018, doi: 10.1021/acs.chemrev.8b00313.
[96] H. Y. Shao and X. G. Li, “Kinetics and thermodynamics of nanostructured Mg-based hydrogen storage
materials synthesized from metal nanoparticles,” Adv. Mater. Res., vol. 924, pp. 189–192, 2014, doi:
10.4028/www.scientific.net/AMR.924.189.
[97] H. Kim, D. Lee, and J. Moon, “Co-electrospun Pd-coated porous carbon nanofibers for hydrogen storage
applications,” Int. J. Hydrogen Energy, vol. 36, no. 5, pp. 3566–3573, 2011, doi:
10.1016/j.ijhydene.2010.12.041.
[98] R. S. Rajaura et al., “Structural and surface modification of carbon nanotubes for enhanced hydrogen
storage density,” Nano-Structures and Nano-Objects, vol. 14, pp. 57–65, 2018, doi:
10.1016/j.nanoso.2018.01.005.
[99] D. Kag, N. Luhadiya, N. D. Patil, and S. I. Kundalwal, “Strain and defect engineering of graphene for
hydrogen storage via atomistic modelling,” Int. J. Hydrogen Energy, vol. 46, no. 43, pp. 22599–22610,
2021, doi: 10.1016/j.ijhydene.2021.04.098.
[100] K. Kumar, M. Alam, D. Rakshit, and V. Dutta, “Operational characteristics of metal hydride energy storage
system in microgrid,” Energy Convers. Manag., vol. 187, no. March, pp. 176–190, 2019, doi:
10.1016/j.enconman.2019.03.019.
[101] Y. Zhang, B. Bijeljic, Y. Gao, S. Goodarzi, S. Foroughi, and M. J. Blunt, “Pore-Scale Observations of
Hydrogen Trapping and Migration in Porous Rock: Demonstrating the Effect of Ostwald Ripening,”
Geophys. Res. Lett., vol. 50, no. 7, pp. 1–8, 2023, doi: 10.1029/2022GL102383.
[102] J. Fu, L. Röntzsch, T. Schmidt, T. Weißgärber, and B. Kieback, “Improved dehydrogenation properties of
lithium alanate (LiAlH 4) doped by low energy grinding,” J. Alloys Compd., vol. 525, pp. 73–77, 2012, doi:
10.1016/j.jallcom.2012.02.076.
[103] S. Liu et al., “Study on hydrogen release capacity of LiAlH4 doped with CeO2,” Xiyou Jinshu Cailiao Yu
Gongcheng/Rare Met. Mater. Eng., vol. 43, no. 3, pp. 544–547, 2014, doi: 10.1016/s1875-5372(14)60073-4.
[104] V. P. Balema, V. K. Pecharsky, and K. W. Dennis, “Solid state phase transformations in LiAlH4 during
high-energy ball-milling,” J. Alloys Compd., vol. 313, no. 1–2, pp. 69–74, 2000, doi: 10.1016/S0925-
8388(00)01201-9.
[105] D. S. Easton, J. H. Schneibel, and S. A. Speakman, “Factors affecting hydrogen release from lithium alanate
(LiAlH4),” J. Alloys Compd., vol. 398, no. 1–2, pp. 245–248, 2005, doi: 10.1016/j.jallcom.2005.02.017.
[106] H. W. Brinks, B. C. Hauback, P. Norby, and H. Fjellvåg, “The decomposition of LiAlD4 studied by in-situ
X-ray and neutron diffraction,” J. Alloys Compd., vol. 351, no. 1–2, pp. 222–227, 2003, doi: 10.1016/S0925-
8388(02)01021-6.
[107] S. Li et al., “Synergistic hydrogen desorption properties of the 4LiAlH4+ Mg2NiH4composite,” J. Alloys
Compd., vol. 697, pp. 80–85, 2017, doi: 10.1016/j.jallcom.2016.12.137.
[108] Y. Yang et al., “Multi-hydride systems with enhanced hydrogen storage properties derived from Mg(BH 4)
2 and LiAlH 4,” Int. J. Hydrogen Energy, vol. 37, no. 14, pp. 10733–10742, 2012, doi:
10.1016/j.ijhydene.2012.04.068.
[109] M. Sterner and I. Stadler, Handbook of Energy Storage: Demand, Technologies, Integration. 2019. doi:
10.1007/978-3-662-55504-0.
[110] S. Niaz, T. Manzoor, and A. H. Pandith, “Hydrogen storage: Materials, methods and perspectives,” Renew.
Sustain. Energy Rev., vol. 50, pp. 457–469, 2015, doi: 10.1016/j.rser.2015.05.011.
[111] S. Jana, P. Muthukumar, and N. N. Raju, “Hydrogen charging and discharging studies on embedded cooling
tube metal hydride reactor designed for fuel cell applications,” Int. J. Hydrogen Energy, vol. 48, no. 96, pp.
37847–37859, 2023, doi: 10.1016/j.ijhydene.2022.07.118.
[112] R. Moradi and K. M. Groth, “Hydrogen storage and delivery: Review of the state of the art technologies and
risk and reliability analysis,” Int. J. Hydrogen Energy, vol. 44, no. 23, pp. 12254–12269, 2019, doi:
10.1016/j.ijhydene.2019.03.041.
[113] J. Zhang et al., “Metal Hydride Nanoparticles with Ultrahigh Structural Stability and Hydrogen Storage
Activity Derived from Microencapsulated Nanoconfinement,” Adv. Mater., vol. 29, no. 24, pp. 1–6, 2017,
doi: 10.1002/adma.201700760.
[114] Z. Li, S. Li, Z. Yuan, Y. Zhang, and J. Zhang, “Effect of catalysts on microstructure, hydrogen storage
thermodynamics, and kinetics performance of La5Mg85Ni10 alloy,” Int. J. Hydrogen Energy, vol. 44, no.
45, pp. 24839–24848, 2019, doi: 10.1016/j.ijhydene.2019.07.225.
[116] G. Srinivas, Y. Zhu, R. Piner, N. Skipper, M. Ellerby, and R. Ruoff, “Synthesis of graphene-like nanosheets
and their hydrogen adsorption capacity,” Carbon N. Y., vol. 48, no. 3, pp. 630–635, 2010, doi:
10.1016/j.carbon.2009.10.003.
[117] B. Panella, M. Hirscher, and S. Roth, “Hydrogen adsorption in different carbon nanostructures,” Carbon N.
Y., vol. 43, no. 10, pp. 2209–2214, 2005, doi: 10.1016/j.carbon.2005.03.037.
[118] K. M. Thomas, “Hydrogen adsorption and storage on porous materials,” Catal. Today, vol. 120, no. 3-4
SPEC. ISS., pp. 389–398, 2007, doi: 10.1016/j.cattod.2006.09.015.
[119] H. Gul, A. U. H. A. Shah, U. Krewer, and S. Bilal, “Study on direct synthesis of energy efficient
multifunctional polyaniline–graphene oxide nanocomposite and its application in aqueous symmetric
supercapacitor devices,” Nanomaterials, vol. 10, no. 1, 2020, doi: 10.3390/nano10010118.
[120] A. G. Tabrizi, N. Arsalani, A. Mohammadi, L. S. Ghadimi, I. Ahadzadeh, and H. Namazi, “A new route for
the synthesis of polyaniline nanoarrays on graphene oxide for high-performance supercapacitors,”
Electrochim. Acta, vol. 265, pp. 379–390, 2018, doi: https://doi.org/10.1016/j.electacta.2018.01.166.
[121] L. R. Vargas, A. K. Poli, R. de C. L. Dutra, C. B. de Souza, M. R. Baldan, and E. S. Gonçalves, “Formation
of composite polyaniline and graphene oxide by physical mixture method,” J. Aerosp. Technol. Manag., vol.
9, no. 1, pp. 29–38, 2017, doi: 10.5028/jatm.v9i1.697.
[122] M. Yamauchi, R. Ikeda, H. Kitagawa, and M. Takata, “Nanosize effects on hydrogen storage in palladium,”
J. Phys. Chem. C, vol. 112, no. 9, pp. 3294–3299, 2008, doi: 10.1021/jp710447j.
[123] S. B. Singh and M. De, “Thermally exfoliated graphene oxide for hydrogen storage,” Mater. Chem. Phys.,
vol. 239, no. April 2019, p. 122102, 2020, doi: 10.1016/j.matchemphys.2019.122102.
[124] S. B. Singh and M. De, “Thermally exfoliated graphene oxide for hydrogen storage,” Mater. Chem. Phys.,
vol. 239, no. April 2019, p. 122102, 2020, doi: 10.1016/j.matchemphys.2019.122102.
[125] F. Paquin, J. Rivnay, A. Salleo, N. Stingelin, and C. Silva, “Multi-phase semicrystalline microstructures
drive exciton dissociation in neat plastic semiconductors,” J. Mater. Chem. C, vol. 3, no. 207890, pp.
10715–10722, 2015, doi: 10.1039/b000000x.
[126] P. García-Holley et al., “Benchmark Study of Hydrogen Storage in Metal-Organic Frameworks under
Temperature and Pressure Swing Conditions,” ACS Energy Lett., vol. 3, no. 3, pp. 748–754, 2018, doi:
10.1021/acsenergylett.8b00154.
[127] Y. Gogotsi, R. K. Dash, G. Yushin, T. Yildirim, G. Laudisio, and J. E. Fischer, “Tailoring of nanoscale
porosity in carbide-derived carbons for hydrogen storage,” J. Am. Chem. Soc., vol. 127, no. 46, pp. 16006–
16007, 2005, doi: 10.1021/ja0550529.
[128] X. Yu, Z. Tang, D. Sun, L. Ouyang, and M. Zhu, “Recent advances and remaining challenges of
nanostructured materials for hydrogen storage applications,” Prog. Mater. Sci., vol. 88, pp. 1–48, 2017, doi:
10.1016/j.pmatsci.2017.03.001.
[129] E. Boateng and A. Chen, “Recent advances in nanomaterial-based solid-state hydrogen storage,” Mater.
Today Adv., vol. 6, p. 100022, 2020, doi: 10.1016/j.mtadv.2019.100022.
[131] A. Baldi, T. C. Narayan, A. L. Koh, and J. A. Dionne, “In situ detection of hydrogen-induced phase
transitions in individual palladium nanocrystals,” Nat. Mater., vol. 13, no. 12, pp. 1143–1148, 2014, doi:
10.1038/nmat4086.
[132] X. Xie, M. Chen, P. Liu, J. Shang, and T. Liu, “Synergistic catalytic effects of the Ni and V nanoparticles on
the hydrogen storage properties of Mg-Ni-V nanocomposite,” Chem. Eng. J., vol. 347, no. July 2017, pp.
145–155, 2018, doi: 10.1016/j.cej.2018.04.084.
[133] J. Zou, S. Long, X. Chen, X. Zeng, and W. Ding, “Preparation and hydrogen sorption properties of a Ni
decorated Mg based Mg@Ni nano-composite,” Int. J. Hydrogen Energy, vol. 40, no. 4, pp. 1820–1828,
2015, doi: https://doi.org/10.1016/j.ijhydene.2014.11.113.
[134] T. Liu, T. Zhang, C. Qin, M. Zhu, and X. Li, “Improved hydrogen storage properties of Mg–V nanoparticles
prepared by hydrogen plasma–metal reaction,” J. Power Sources, vol. 196, no. 22, pp. 9599–9604, 2011,
doi: https://doi.org/10.1016/j.jpowsour.2011.07.078.
[135] C. Lu, J. Zou, X. Zeng, and W. Ding, “Hydrogen storage properties of core-shell structured Mg@TM
(TM = Co, V) composites,” Int. J. Hydrogen Energy, vol. 42, no. 22, pp. 15246–15255, 2017, doi:
https://doi.org/10.1016/j.ijhydene.2017.04.063.
[136] T. Liu, C. Chen, H. Wang, and Y. Wu, “Enhanced Hydrogen Storage Properties of Mg–Ti–V
Nanocomposite at Moderate Temperatures,” J. Phys. Chem. C, vol. 118, no. 39, pp. 22419–22425, Oct.
2014, doi: 10.1021/jp5061073.
[137] J. Yang, P. Li, L. Wang, X. Guo, J. Guo, and S. Liu, “In-situ synthesis of Ni-MOF@CNT on graphene/Ni
foam substrate as a novel self-supporting hybrid structure for all-solid-state supercapacitors with a high
energy density,” J. Electroanal. Chem., vol. 848, p. 113301, 2019, doi:
https://doi.org/10.1016/j.jelechem.2019.113301.
[138] N. M. Musyoka et al., “Synthesis of a hybrid MIL-101(Cr)/ZTC composite for hydrogen storage
applications,” Res. Chem. Intermed., vol. 42, no. 6, pp. 5299–5307, 2016, doi: 10.1007/s11164-015-2361-2.
[139] Q. Li, Y. Li, B. Liu, X. Lu, T. Zhang, and Q. Gu, “The cycling stability of the: In situ formed Mg-based
nanocomposite catalyzed by YH2,” J. Mater. Chem. A, vol. 5, no. 33, pp. 17532–17543, 2017, doi:
10.1039/c7ta04551d.
[140] H. H. Cheng, H. G. Yang, S. L. Li, X. X. Deng, D. M. Chen, and K. Yang, “Effect of hydrogen
absorption/desorption cycling on hydrogen storage performance of LaNi4.25Al0.75,” J. Alloys Compd., vol.
453, no. 1–2, pp. 448–452, 2008, doi: 10.1016/j.jallcom.2006.11.112.
[141] J. Li et al., “Hydrogen absorption-desorption cycle decay mechanism of palladium nanoparticle decorated
nitrogen doped graphene,” Prog. Nat. Sci. Mater. Int., vol. 31, no. 4, pp. 514–520, 2021, doi:
10.1016/j.pnsc.2021.06.009.
[142] U. Ulmer, M. Dieterich, A. Pohl, R. Dittmeyer, M. Linder, and M. Fichtner, “Study of the structural,
thermodynamic and cyclic effects of vanadium and titanium substitution in laves-phase AB2 hydrogen
storage alloys,” Int. J. Hydrogen Energy, vol. 42, no. 31, pp. 20103–20110, Aug. 2017, doi:
10.1016/J.IJHYDENE.2017.06.137.