0% found this document useful (0 votes)
17 views29 pages

Zhang 2022 Prog. Energy 4 042013

This document reviews the fundamentals of hydrogen storage in nanoporous materials, highlighting the efficiency of physisorption as an alternative to traditional storage methods. It discusses the performance of various materials, including zeolites and metal-organic frameworks, in adsorbing hydrogen at low temperatures and moderate pressures. The review also addresses the use of advanced computational methods, such as machine learning, for optimizing hydrogen storage solutions and emphasizes the importance of reproducibility in experimental data.

Uploaded by

Roberto G. Silva
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
17 views29 pages

Zhang 2022 Prog. Energy 4 042013

This document reviews the fundamentals of hydrogen storage in nanoporous materials, highlighting the efficiency of physisorption as an alternative to traditional storage methods. It discusses the performance of various materials, including zeolites and metal-organic frameworks, in adsorbing hydrogen at low temperatures and moderate pressures. The review also addresses the use of advanced computational methods, such as machine learning, for optimizing hydrogen storage solutions and emphasizes the importance of reproducibility in experimental data.

Uploaded by

Roberto G. Silva
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

Progress in Energy

TOPICAL REVIEW • OPEN ACCESS You may also like


- Advances in thermal conductivity for
Fundamentals of hydrogen storage in nanoporous energy applications: a review
Qiye Zheng, Menglong Hao, Ruijiao Miao
materials et al.

- A continuum of physics-based lithium-ion


battery models reviewed
To cite this article: Linda Zhang et al 2022 Prog. Energy 4 042013 F Brosa Planella, W Ai, A M Boyce et al.

- Review of parameterisation and a novel


database (LiionDB) for continuum Li-ion
battery models
A A Wang, S E J O’Kane, F Brosa Planella
View the article online for updates and enhancements. et al.

This content was downloaded from IP address 190.131.28.18 on 06/02/2023 at 23:30


Prog. Energy 4 (2022) 042013 https://doi.org/10.1088/2516-1083/ac8d44

Progress in Energy

TOPICAL REVIEW

Fundamentals of hydrogen storage in nanoporous materials


OPEN ACCESS
Linda Zhang1, Mark D Allendorf 2, Rafael Balderas-Xicohténcatl1,3, Darren P Broom4,
RECEIVED George S Fanourgakis5, George E Froudakis5, Thomas Gennett6,7, Katherine E Hurst6,
12 May 2022
Sanliang Ling8, Chiara Milanese9, Philip A Parilla6, Daniele Pontiroli10, Mauro Riccò10,
REVISED
15 August 2022 Sarah Shulda6, Vitalie Stavila2, Theodore A Steriotis11, Colin J Webb12, Matthew Witman2
ACCEPTED FOR PUBLICATION
and Michael Hirscher1,∗
26 August 2022 1
Max-Planck-Institut für Intelligente Systeme, Heisenbergstrasse 3, 70569 Stuttgart, Germany
2
PUBLISHED Sandia National Laboratories, Livermore, CA 94551, United States of America
14 September 2022 3
Neutron Scattering Division, Neutron Sciences Directorate, Oak Ridge National Laboratory, Oak Ridge, TN 37831,
United States of America
4
Original content from Hiden Isochema, 422 Europa Boulevard, Warrington WA5 7TS, United Kingdom
5
this work may be used Department of Chemistry, University of Crete, Voutes Campus, GR-70013 Heraklion, Crete, Greece
under the terms of the 6
National Renewable Energy Laboratory, Golden, CO 80401, United States of America
Creative Commons 7
Department of Chemistry, Colorado School of Mines, Golden, CO 80401, United States of America
Attribution 4.0 licence. 8
Advanced Materials Research Group, Faculty of Engineering, University of Nottingham, University Park, Nottingham NG7 2RD,
Any further distribution United Kingdom
of this work must 9
maintain attribution to Pavia Hydrogen Lab, Chemistry Department, Physical Chemistry Section, University of Pavia and C.S.G.I., Viale Taramelli 16,
the author(s) and the title I-27100 Pavia, Italy
of the work, journal 10
Nanocarbon Laboratory, Dipartimento di Scienze Matematiche, Fisiche e Informatiche, Università di Parma, Parco Area delle
citation and DOI. Scienze, 7/a, I-43124 Parma, Italy
11
National Center for Scientific Research ‘Demokritos’, 15341 Ag. ParaskeviAttikis, Athens, Greece
12
Queensland Micro- and Nanotechnology Centre, Griffith University, Brisbane, QLD 4111, Australia

Author to whom any correspondence should be addressed.
E-mail: [email protected]

Keywords: energy storage, porous materials, adsorption, machine learning

Abstract
Physisorption of hydrogen in nanoporous materials offers an efficient and competitive alternative
for hydrogen storage. At low temperatures (e.g. 77 K) and moderate pressures (below 100 bar)
molecular H2 adsorbs reversibly, with very fast kinetics, at high density on the inner surfaces of
materials such as zeolites, activated carbons and metal–organic frameworks (MOFs). This review,
by experts of Task 40 ‘Energy Storage and Conversion based on Hydrogen’ of the Hydrogen
Technology Collaboration Programme of the International Energy Agency, covers the
fundamentals of H2 adsorption in nanoporous materials and assessment of their storage
performance. The discussion includes recent work on H2 adsorption at both low temperature and
high pressure, new findings on the assessment of the hydrogen storage performance of materials,
the correlation of volumetric and gravimetric H2 storage capacities, usable capacity, and optimum
operating temperature. The application of neutron scattering as an ideal tool for characterising H2
adsorption is summarised and state-of-the-art computational methods, such as machine learning,
are considered for the discovery of new MOFs for H2 storage applications, as well as the modelling
of flexible porous networks for optimised H2 delivery. The discussion focuses moreover on
additional important issues, such as sustainable materials synthesis and improved reproducibility
of experimental H2 adsorption isotherm data by interlaboratory exercises and reference materials.

1. Introduction

Developing a safe, affordable and efficient way of storing H2 is a key priority in hydrogen energy research.
Current fuel cell vehicles, such as the Toyota Mirai, use 700 bar compressed H2 , which provides a gravimetric
H2 capacity of approximately 5.7 wt% and a volumetric capacity of 40 g H2 l−1 [1]. Compressed H2 storage
offers quick refill times and provides long ranges for fuel cell vehicles, but it also has some disadvantages.
Compressing H2 to 700 bar, for example, consumes energy, and further gains in volumetric and gravimetric

© 2022 The Author(s). Published by IOP Publishing Ltd


Prog. Energy 4 (2022) 042013 L Zhang et al

capacity can only be achieved by using yet higher pressures. The need for high pressures limits tank shape to
non-conformable carbon fibre-reinforced composite cylinders, which are also expensive, while isenthalpic
expansion of H2 during charging leads to temperature increases, so H2 must be pre-cooled to 233 K to avoid
overheating [2]. H2 can also be stored as a liquid, but this requires very low temperatures, below 30 K [3, 4].
Cooling H2 to this extent consumes more energy than 700 bar compression and long-term storage requires
highly insulated tanks, which are, again, rather expensive. A third option, known as cryo-compression,
combines cooling and compression, to achieve high volumetric densities close to liquid H2 in the gaseous
state [5, 6].
A general alternative to storing H2 in compressed, liquid or cryo-compressed form is to use materials that
absorb or adsorb hydrogen, to provide high gravimetric and volumetric capacities at more practical
pressures and temperatures [7, 8]. Various options exist, as discussed throughout this special issue, but each
one has disadvantages, as well as advantages. Metal and complex hydrides bind hydrogen in their structure,
and can therefore be categorised as a form of chemical hydrogen storage [4]. Complex hydrides can provide
high gravimetric and volumetric capacities, but they are not always reversible, often require high
temperatures for operation, and, when reversible, usually have sluggish hydrogen absorption and desorption
kinetics. Interstitial metal hydrides that can operate at near ambient temperature, meanwhile, can possess
impressive volumetric capacities, exceeding that of liquid H2 [9], but they usually exhibit low gravimetric
capacities, below 2 wt%. Moreover, in all cases, since hydrogen is chemically bound via ionic, metallic or
covalent bonds, the respective sorption and desorption enthalpies are considerable, and heat management
then becomes a key issue for optimal operation [10].
Hydrogen adsorption in nanoporous materials is an alternative physical, rather than chemical, solution
to the hydrogen storage problem [4]. In this case, molecular H2 is physically adsorbed, or physisorbed, in the
pores of materials with very high internal surface areas and hence extended gas–solid interfaces, such as
zeolites [11], activated carbons, and metal–organic frameworks (MOFs) [12–16]. Using this approach, H2
can be stored at higher densities than compressed gas, at pressures below 100 bar, and at higher temperatures
than those required for liquid storage. The practical problem is that low temperatures, in the region of 77 K,
and pressures up to 100 bar, are still required to achieve high capacities; however, there are also advantages,
compared to chemical storage in hydrides, which include rapid sorption kinetics and the low heat of
adsorption. The key point is that H2 storage in nanoporous materials may offer an intermediate solution that
provides both high gravimetric and volumetric storage densities at pressures below 100 bar, thus reducing
both compression or liquefaction losses and the challenges of handling and distributing H2 as high pressure
(700 bar) gas or cryogenic liquid.
This article covers the fundamentals of H2 storage in nanoporous materials, and recent work
investigating the adsorption behaviour of H2 at low temperatures and high pressures. We also consider the
correlation of gravimetric and volumetric H2 capacities, enhancing deliverable capacities through the use of
flexible materials, machine learning (ML) for predicting storage performance, and assessing the limits of H2
storage in flexible materials. Other important issues, such as reproducibility testing using interlaboratory
exercises and the need for reference materials, are also discussed.

2. Fundamentals

2.1. Gas adsorption in porous materials


The critical point of H2 is around 33 K and 13 bar, and its boiling point is just over 20 K [17]. Very low
temperatures are therefore required to condense hydrogen. These physical properties are due to weak H2 –H2
interactions, which originate from H2 molecules having no charge, no dipole moment, a relatively weak
quadrupole moment, and a low polarisability. These same characteristics lead to the weak physical interaction
of molecular H2 with surfaces [18–20], but it is nevertheless possible to adsorb fairly large quantities of H2 at
supercritical temperatures, in materials possessing high enough surface areas and pores of proper size.
Various factors affect the adsorption of gases by porous materials, including the strength of the
adsorbate–adsorbent interactions, the specific surface area (SSA) of the material, typical pore widths or
diameters, and total pore volume. The strength of adsorbate–adsorbent interactions depends on the
properties of both the gas and the solid. SSA, meanwhile, defines the extent of the gas–solid interface and
thus the quantity of surface sites available for adsorption, per unit mass of material; although the concept
breaks down somewhat when pores are sufficiently narrow. At low pressures, the amount of adsorption
depends mainly on pore width or diameter. In very narrow pores, the overlapping potentials of opposing
pore walls, or of neighbouring framework atoms, lead to stronger solid–fluid interactions compared to open,
flat surfaces and therefore to larger amounts of adsorption at relatively low pressure. The saturation uptake
of a material, however, at high pressure, tends to scale with the total pore volume up to pore sizes beyond
which adsorption is not significant [21].

2
Prog. Energy 4 (2022) 042013 L Zhang et al

Adsorption behaviour is typically characterised by the shape of isotherms—plots of uptake against


pressure—and varies with temperature. It also correlates with the physical properties of the adsorbate and
the nature of the adsorbent. Adsorption at subcritical temperatures differs to supercritical adsorption.
Nitrogen is one of the most common adsorbates, due to its widespread use for determining
Brunauer–Emmett–Teller (BET) areas, a common approximation of the accessible SSA of a solid. The critical
temperature of N2 is 126 K, while measurements are often performed at its boiling point, 77 K, because of
the practical convenience of using liquid N2 as a cryogen. At such subcritical temperatures, monolayer
adsorption can occur on relatively flat or open surfaces, followed by multilayer formation at higher relative
pressures, P/P0 , where P is absolute pressure and P0 is the saturation pressure of the adsorptive. This is the
case for non-porous or macroporous materials, and identifying the point at which a statistical monolayer is
formed is the principle behind BET surface area determination.
Materials of interest for adsorptive hydrogen storage, however, are usually microporous or mesoporous.
Terminology associated with both porous materials and gas adsorption has been defined in widely accepted
International Union of Pure and Applied Chemistry (IUPAC) guidelines [22]. Microporous materials are
defined as having pore sizes <2 nm; mesoporous materials, between 2 and 50 nm; while the term nanoporous
refers to any material with a pore size <100 nm [22]. Most porous hydrogen storage materials are therefore
nanoporous. In mesoporous materials, at subcritical temperatures, monolayer and multilayer adsorption can
occur, followed by capillary condensation, which in most cases gives a characteristic (Type IV) isotherm
shape with a sudden increase of the amount adsorbed coupled many times with hysteresis between
adsorption and desorption isotherms [22–24]. In micropores, however, the adsorption process differs
because adsorption only occurs via pore-filling and there is no capillary condensation. This leads to a
so-called Type I isotherm, which is concave to the pressure axis, with no observable hysteresis. This difference
in adsorption behaviour—capillary condensation versus pore filling or else Type IV versus I—is the
historical origin of the otherwise rather arbitrary 2 nm threshold between micropores and mesopores, as
defined in the IUPAC guidelines [22].
At supercritical temperatures, even in mesopores, capillary condensation, which is associated with a
phase transition from a gas to liquid-like state [23], does not occur. The presence of the surface, however,
affects the density of the adsorbate in the pores, to a varying degree, depending on pore size and the strength
of the adsorbate–adsorbent interactions. The magnitude of the interaction between an adsorbate and either a
surface or a pore of a given size is usually characterised by the heat or enthalpy of adsorption, ∆H, which is
typically expressed as isosteric enthalpy of adsorption, ∆H st —the differential enthalpy at constant surface
coverage or loading [25]. For H2 on an open carbon surface, for example, ∆H st is approximately 4 kJ mol−1
[26], which is rather low considering that the thermal energy at 77 K is around 1 kJ mol−1 . ∆H st , however,
increases under confinement, while heterogeneities also alter the strength of interaction [27]. Regardless of
the chemistry of the adsorbent, however, there is a general trend of increasing enthalpy with decreasing pore
size up to the point of molecular sieving [13, 14]. This is a result of the enhanced interaction within small
pores due to potential overlap. In some materials, H2 can also adsorb more strongly on certain types of sites,
due to electrostatic interactions [18, 19]. Examples include cations in zeolites [28–30] and open metal sites
(OMSs) in MOFs [16, 31]. In these cases, ∆H st for H2 can exceed 10 kJ mol−1 at low loadings.
Isosteric enthalpies are usually determined experimentally by measuring adsorption isotherms at two, or
preferably more, closely spaced temperatures, and then applying the Clausius–Clapeyron (CC) equation or
van ‘t Hoff relation [8, 32], while a virial-type analysis is also possible [33, 34]. This results in a plot of ∆H st
versus coverage or loading, θ. For H2 , most of the times ∆H st tends to decrease as a function of coverage
[14, 25]. This differs to species that may exhibit strong adsorbate–adsorbate interactions at higher coverages.
For H2 , the highest values of ∆H st —for strongly interacting sites or the narrowest pores—are found at low
coverages. ∆H st at zero coverage can also be calculated by extrapolation or direct calculation. This is, in
principle, the energy release due to the adsorption of a single test molecule (θ → 0). The magnitude of ∆H st
has several practical consequences. Firstly, adsorption can be realised at higher temperatures for higher
values of ∆H st . Secondly, more heat is generated upon adsorption when ∆H st is higher, potentially raising
heat management issues. Thirdly, from a hydrogen storage perspective, the loading dependence of ∆H st
affects the deliverable amount of H2 at any given pressure, because a high value of ∆H st at low loading,
regardless of the value at high loading, will lead to more H2 being trapped in the material at low pressure,
below the delivery pressure of the store [35]. This point is discussed in more detail in section 3.1.2.
Adsorbent heterogeneity due to different surface sites, such as heteroatoms, functional groups, surface
defects and so forth, as well as due to different pore sizes within the material also affects ∆H st , as well as
isotherm shape. The simplest case is described by the Langmuir model, in which each adsorption site is
energetically equivalent and there are no adsorbate–adsorbate interactions. In this case, ∆H st is constant as a
function of loading and the isotherm is described mathematically by the Langmuir equation. Most real

3
Prog. Energy 4 (2022) 042013 L Zhang et al

adsorbents, however, are not so homogeneous, and the Langmuir equation rarely provides an adequate fit to
experimental H2 adsorption data for nanoporous materials; except perhaps over a limited pressure range.

2.2. Intermolecular interactions


Physical adsorption is generally based on van der Waals interactions, which include long-range attractive
forces, between fluctuating, induced or permanent electric moments, but also short-range strong repulsive
forces, due to the interaction of overlapping atomic or molecular orbitals (Pauli’s exclusion principle)
[36, 37]. London dispersion forces are due to electron density fluctuations within atoms, which induce
electrical moments in neighbouring atoms and thus attraction [38, 39]. The potential energy of such
attractive (negative) interactions between two atoms is given by [40],

εd (r) = −A1 r−6 − A2 r−8 − A3 r−10 ∼


= −Ar−6 (1)

where r is the distance between the centres of the atoms and A1 (=A), A2 , A3 are the dispersion constants for
instantaneous dipole–dipole, dipole–quadrupole and quadrupole–quadrupole interactions, respectively; the
r −8 and r −10 terms are usually negligible. Short range repulsion (positive) potential energy, meanwhile, can
be described by [41],

εR (r) = Br−m (2)

where B is an empirical constant and m is usually set to 12. The total potential energy between two atoms as a
function of their distance can then be approximated by the Lennard-Jones (LJ) expression [42],

εLJ (r) = εd (r) + εR (r) = −Ar−6 + Br−12 (3)

which has the general shape shown in figure 1(a). This can be recast in a more commonly used form after
considering a van der Waals-type diameter, σ, with εLJ (σ) = 0 and an interatomic distance, r 0 , where the
interaction is strongest (i.e. the energy, ε0 , is at a minimum and thus dεLJ (r) /dr|r0 = 0). Then,
[ ]
−6 −12
εLJ (r) = −4ε0 (σ/r) + (σ/r) (4)

√ √ √
where ε0 = εLJ (r0 ) = − (A/4) σ −6 and σ = − 6 B/A (r0 = 6 2B/A = 6 2σ).
For two
( different
) atoms, i and j, the Lorentz–Berthelot mixing rules can be used, giving

σij = 1/2 σi + σj , while εij = εi εj . The dispersion constant Aij for the atoms is directly related to their
properties, for instance, through the Kirkwood–Müller relationship [43, 44],

6mc2 αi αj
Aij = ( ) (5)
(αi /χi ) + αj /χj

where α is polarisability, χ is magnetic susceptibility, m is electron mass and c is the speed of light. As an
example, the α and χ values for H2 are 0.79 × 10−24 cm3 and 4.0 × 10−4 cm3 , respectively, while those for
N2 are 1.76 × 10−24 cm3 and 12.0 × 10−4 cm3 [45]. The effect of these different adsorbate properties can be
seen in the different calculated potentials for H2 and N2 in figure 1(a).
The LJ expression is a ‘generic’ potential and is thus not necessarily limited to atoms; it can in principle
be used for any pair of LJ particles (e.g. figure 1 pertains to H2 and N2 spherical molecular models). For
instance, the position-dependent energy landscape, U LJ (r), that a molecule i experiences when close to an
open surface, or inside a porous framework comprising j atoms, can be presented as the following
summation of all the pairwise LJ potentials,
∑ [( )−6 ( )−12 ]
ULJ (r) = −4ε0 σij /rij + σij /rij . (6)
j

For homogeneous flat surfaces the gas-solid interaction potential is more or less similar to the pairwise
interaction curve of figure 1(a). The variable in this case, however, is the distance from the surface, z. Upon
confinement in pores, such as slits and cylinders, the interaction of neighbouring atoms of the solid sum up
in a constructive manner and produce much deeper potential wells compared to the open surface, as shown
in figure 1(b). In the case of simple pore models (e.g. slit-shaped or cylindrical pores), the energy landscape
can be easily described, as shown in figures 1(b) and (c). Confinement of a gas molecule, however, in a 3D
porous solid—for instance, a zeolite or MOF—is far more complicated, as the energy landscape varies in the
x, y and z directions, as shown in figure 1(d) for the case of the MOF, HKUST-1.

4
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 1. (a) LJ potential energy between two molecules (H2 or N2 ) versus their distance, (b) position dependent potential energy
of a single H2 molecule confined in carbon slit pores of varying width (z = 0 is the centre of the pore), (c) LJ zero potential
isosurface (U LJ = 0) for H2 in a slit pore with graphene walls. Calculations are based on the visualisation method of [46]. (d) H2
zero potential isosurface in a typical MOF (HKUST-1). Constructed with i-RASPA visualisation software [47].

For the case of H2 the interactions become more complicated as the adsorption temperature is reduced
since it is a very light gas and therefore has a considerable de Broglie wavelength. For this reason, nuclear
quantum effects are expected to contribute significantly to the adsorption process at temperatures below
100 K. For such processes, the quadratic term of the Feynman–Hibbs (FH) effective potential can be used
[48], in order to ‘smear out’ the deepest part of the classical potential curve,

( )
ij βℏ2
Uall = ULJ + ∇2 (ULJ ) (7)
24µm

−1
where ℏ = h/2π, h is the Planck constant, β = (kT) and µm is the reduced mass, given by,

( )−1
1 1
µm = + (8)
mi mj

where mi is the mass of atom i. The FH expression adequately describes the quantum spreading of H2
( )1/
molecules when λ∗ = 2πβℏ2 /mσH2 2 2 ⩽ 0.5. Here λ∗ is the reduced de Broglie wavelength, and m the

mass of H2 . The reduced thermal wavelength λ∗ at temperature T is given by the expression: 4.17/ T and
its numerical value at 77 K is 0.47. At lower temperatures, quantum effects become increasingly important
and more elaborate approaches, such as the path integral formalism, are required to describe H2
interactions [49].

5
Prog. Energy 4 (2022) 042013 L Zhang et al

2.3. H2 adsorption at subcritical temperatures


As discussed above, the low molecular mass of H2 and its weak intermolecular potential result in fascinating
properties at cryogenic temperatures of the adsorbed, liquid, and solid phases, comparable only with He.
Similarly to other gases below their critical points (e.g. N2 at 77 K or argon at 87 K) adsorption of H2 at
subcritical temperatures (<33 K) is usually associated, depending on the pore structure and morphology,
with processes such as pore-filling and capillary condensation that are not observed at supercritical
temperatures. However, in addition to this typical behaviour, H2 adsorbed at subcritical temperatures, near
its boiling point (20.3 K), exhibits unique quantum properties that must be considered for the study of
adsorption at such conditions, as well as for the design of cryo-storage systems.
One example that depends on adsorption potential and pore size is the difference in kinetics and
adsorption energy upon isotopic (H2 /D2 ) exchange. The small difference in mass between the two isotopes
(equations (7) and (8)) can lead to quantum sieving which may be kinetic in ultra-microporous (<0.7 nm)
materials [50, 51] due to the larger de Broglie wavelength of the lighter isotope (D2 diffuses faster than H2 in
small pores) or enhanced chemical affinity for the heavier isotope on strong adsorption sites, such as OMSs
in MOFs [51–54]. Furthermore, the hydrogen molecule, H–H, possesses two possible proton spin state
configurations, ortho-hydrogen (parallel, ↑↑) and para-hydrogen (anti-parallel, ↑↓). Because of the
symmetry of the wavefunction, para-hydrogen only has even rotational numbers (J = 0, 2, 4,…), while
ortho-hydrogen only has odd numbers (J = 1, 3, 5,…). The lowest rotational state of ortho-hydrogen is
J = 1, which implies a quadrupole moment, leading to a stronger interaction when it is physisorbed.
At ambient temperature, the para/ortho ratio is 1:3, whereas at near-boiling temperatures (20 K) it is
nearly 1:0. If a H2 molecule is in the presence of a magnetic centre there may be spin interchange, resulting in
ortho-para conversion. This conversion has important practical implications due to the large latent heat of
conversion of 1.42 kJ mol−1 at 20 K, which is higher than the latent heat of vaporisation of 0.89 kJ mol−1 at
the same temperature [55]. This effect is often neglected, due to the experimental complication of
controlling the para-ortho conversion, and the low temperature at which it is observed.
The quantum nature of molecular H2 can be also observed in the high compressibility of the bulk solid
phase due to the absence of multiple electron shells. Applying a pressure of 100 bar to hydrogen in the solid
state results in a volume decrease of 5% [56], whereas other solids are barely compressible. Argon, for
example, changes its volume by only 0.8% [57]. This large compressibility also occurs in the adsorbed phase
of H2 , which can possess a density higher than the bulk at subcritical temperatures and near ambient
pressure, due to short H2 –H2 intermolecular distances.

2.3.1. High density of H2 inside the pore system


At 77 K, H2 physisorption is mainly governed by the surface of the solid, with a linear correlation observed
between the SSA (area per unit of mass) and the H2 gravimetric capacity at elevated pressures, as specified by
Chahine’s rule [58]. An analogous relation for the volumetric H2 capacity was recently verified
experimentally on a series of MOFs [59]. Gravimetric and volumetric uptake can therefore be correlated to
the surface area of the adsorbent [59–61], with an average surface H2 density of 1.9 × 10−2 mg m−2 ,
measured for many MOFs [59]. This surface density is lower than the bulk density of H2 , and can be
understood as the number of H2 molecules per unit area, equivalent to an intermolecular (H2 –H2 ) distance
of 4.74 Å at 77 K. This is larger than the intermolecular distance in the solid (3.76 Å) and liquid (4.05 Å)
state. Increasing the adsorbed layer density (or reducing the intermolecular H2 –H2 distance) is important, as
it relates directly to an increase in both volumetric and gravimetric H2 storage capacity of a material.
Measurements of N2 adsorption at 77 K are routinely used to assess the monolayer capacity of a material;
although using Ar at 87 K is preferable, due to its lack of quadrupole moment, as recommended in the
IUPAC guidelines [22]. The monolayer capacity, combined with the cross-sectional area of the probe
molecule—the area a single molecule occupies on the surface—can be converted into the SSA. Assuming
knowledge of the H2 cross-sectional area, a subcritical H2 adsorption experiment (<33 K) can therefore, in
principle, be used to assess surface area, while simultaneously measuring the maximum capacity of an
adsorbent [62]. Despite its importance, the H2 cross-sectional area cannot be estimated a priori. H2
molecules exhibit weak intermolecular interactions (∼0.3–0.5 kJ mol−1 ) [63], while the H2 -surface
interaction potential for conventional materials such as carbon or silica is ∼4 kJ mol−1 [26]. The density of
the adsorbed layer or cross-sectional area at low temperatures therefore depends strongly on the H2 -surface
interaction, which will be dictated by both the surface chemistry and morphology of the adsorbent.
Subcritical adsorption of light gases, such as He and H2 , has been studied since the 1940s, with some
reports revealing unusually high He and H2 capacities, compared to other gases such as Ar, N2 or CO2 , in the
first adsorbed layer (monolayer) [65, 66]. In 1956, Steele [67] introduced a phenomenological model using
the concept of double-layer adsorption (a bilayer) to explain high He monolayer capacities on carbon. Pace
and Siebert [68] reported a large discrepancy between the monolayer capacities on carbon of N2 at 77 K and

6
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 2. Hydrogen and argon adsorption isotherms of porous silica at boiling temperature, respectively. The monolayer
capacities calculated by BET show a difference of almost two times more for H2 , indicating that the two times more molecules
cover the same surface [64]. The relative pressure refers to absolute pressure (P) divided by the condensation pressure (P0 ).

H2 at 20 K, suggesting a short intermolecular H2 –H2 distance (2.95 Å). Such experimental evidence
indicated an H2 adsorbed layer with approximately double the density in graphite-like carbon at 20 K
compared to H2 bulk liquid density. Similar observations have been reported since for H2 on silica in 1990
[69, 70], 1997 [71] and 2014 [72], for H2 on carbon in 2004 [73], and for He on zeolites in 1994 [74]. Despite
these experiments showing a high adsorption capacity of H2 (and He) at subcritical temperatures, however,
the microscopic nature is still under discussion, with some reports ascribing it to a high-density,
monolayer-like H2 phase, and others assuming the formation of a bilayer.
In recent work, Balderas-Xicohténcatl et al [64, 75] studied the density of a single layer of H2 adsorbed at
20 K on a series of micro-, meso- and non-porous silicas and carbons, as well as the model material KIT-6, a
mesoporous silica [76]. High-resolution gas adsorption experiments and inelastic neutron scattering (INS)
were used to independently demonstrate layer formation of H2 with an intermolecular distance of 2.9 Å.
Figure 2 shows an example of the comparison of H2 (20 K) and Ar (87 K) adsorption isotherms for porous
silica. The isotherms show that the H2 BET monolayer capacity is almost double that of Ar for the same
surface. The intermolecular H2 –H2 distance was calculated by comparing the H2 monolayer capacity and the
surface area, and equates to a volumetric density of 201 g cm−3 , almost three times the bulk-solid density of
H2 (80.0 g cm−3 ). These experimental results were supported by path integral Monte Carlo simulations and
ab initio calculations, including nuclear quantum effects, allowing rationalisation of the high-density phase
by the relatively small intermolecular repulsion of the compressed H2 compared to the surface-adsorbate
attraction [64, 75]. This high density, or short intermolecular H2 –H2 distance, still requires further study,
but could potentially be used to increase both volumetric and gravimetric H2 storage capacities of materials
for cryogenic storage systems.

2.3.2. H2 adsorption near its boiling temperature


Storing H2 as a liquid in a cryogenic tank has been proposed as an economically viable option for large-scale
storage and transport applications, due to the higher energy density and better area/volume ratio of a large
tank, which translates into better thermal isolation [77]. Hydrogen liquefaction requires temperatures as low
as 33.1 K and a moderate pressure of 12.8 bar. In this context, H2 adsorption in nanoporous materials near
the boiling point could be used to optimise such cryogenic tanks.
One of the first subcritical H2 adsorption isotherms measured on a MOF, MIL-101, was reported by
Streppel et al [62]. This material possesses a trimodal pore size distribution, which offered the opportunity to

7
Prog. Energy 4 (2022) 042013 L Zhang et al

explore the pore filling effect that occurs at different pressures. Near the condensation pressure, the material
is saturated due to the filling of all pores. At this point, the loading reaches an upper limit that defines the
saturation capacity of the material. Saturation capacity is technologically relevant since it marks the upper
physical limit of an adsorbent-filled tank.
Oh et al [78] used an isochoric (constant volume) adsorption experiment to measure the capacity of an
empty and a MIL-101 filled sample cell. The isochoric experiment is a direct measurement of the pressure
increase with temperature. For the empty tank, the pressure rapidly increases at 20 K, which corresponds to
the boiling temperature of H2 . In the presence of the adsorbent, the pressure starts to increase at ca. 40 K
over wider temperature range. Hence, H2 adsorption data near boiling temperature for high-capacity
sorbents is also required to assess the potential of achieving the gravimetric and volumetric capacity
requirements for on-board H2 storage systems based on cryo-adsorption.

2.4. Adsorption at high pressure


When measuring gas adsorption at high pressure it is important to distinguish between the excess and
absolute adsorbed quantities [15, 74]. Techniques for measuring adsorption isotherms, including both the
gravimetric and volumetric/manometric methods, determine excess adsorption, which is the amount
adsorbed over and above the molar quantity that would be present in the absence of gas–solid interactions.
Absolute adsorption, meanwhile, is the common term for the total quantity present in the adsorbed phase. An
additional term, total adsorption, is sometimes used to refer to the sum of adsorbate molecules in the
accessible pore volume of the adsorbent. Calculation of absolute uptake from excess uptake, determined
from experimentally measured values of pressure, temperature and calibrated volumes, requires knowledge
of either the volume occupied by the adsorbate or its density [79]. Experimental determination of the density
of H2 adsorbed in the pores of a material is, however, very difficult [15, 80]. One way to achieve this is to
continue to increase the pressure in an experimental isotherm past the point at which adsorption saturates.
Provided there is no further adsorption for higher pressure steps and the adsorbate density does not change,
the calculated isotherm uptake becomes linear with respect to the bulk gas density, and the slope of this
linear portion of the isotherm gives the volume of the adsorbate [81, 82].
As H2 adsorbs relatively weakly in most porous materials, achieving saturation requires low
temperatures, high pressure, or both. Due to difficulties with handling and measuring high pressure H2
adsorption, low temperatures were first to be investigated. Poirier and Dailly analysed H2 adsorption
measurements on IRMOF-1 [83] and a range of MOFs [84–86] at 50–100 K up to 40 bar, determining
densities of the adsorbed H2 of 0.068 g cm−3 and 0.05–0.06 g cm−3 respectively. This technique has also been
applied to isotherms for other gases, including CO2 [87–89], and N2 and CH4 [89]. In all the H2 studies
mentioned above, pressures less than 100 bar were employed, necessitating temperatures below 100 K in
order to reach saturation. At higher temperatures, much higher pressures are needed, as demonstrated by
measurements made by Voskuilen et al [90] who found that saturation was not reached at ambient
temperature for pressures up to 500 bar, for various nanoporous materials, including porous carbons, five
different MOFs, and a hyper-crosslinked polymer.
More recently, H2 adsorption isotherms have been measured on a commercial activated carbon,
Filtrasorb 400, at ambient temperature [91]. In order to achieve saturation, measurements were made up to
2000 bar. A typical excess isotherm is shown in figure 3, in which the calculated uptake in wt% is plotted
against gas pressure (Plot A). It should be noted that performing adsorption measurements up to this
pressure is not routine. It requires H2 compression, as typical supply cylinder pressures are of the order of
200 bar, as well as the use of specialist high pressure components, including valves, tubes and fittings. Most
commercial H2 adsorption instruments operate up to a maximum pressure of only 200 bar, while the
majority of gas adsorption instruments, more generally, for use, for example, for N2 adsorption, only operate
up to ambient pressure (1 bar).
Assuming the adsorbed phase volume, which is often taken as the pore volume of the sample [82, 89, 92],
is constant for all points on the isotherm and it is the adsorbate density that changes with uptake, the
absolute isotherm can be constructed from the excess isotherm by adding an uptake amount equivalent to
the bulk gas density times the adsorbate volume [91]. The absolute isotherm calculated in this way is shown
in figure 3, plot B. Again, assuming the adsorbate volume is constant, the density of the adsorbate can be
calculated from the absolute mass adsorbed divided by the adsorbate volume, as shown in Plot C.
A linear fit to the last six points of the excess uptake versus gas density isotherm had a regression
coefficient of R = 0.9987 and gave a volume for the adsorbate of 0.625 cm3 for the 0.480 g sample. From the
maximum uptake of 5.50 wt%, this gives a maximum density of 0.0447 g cm−3 for the adsorbed H2 averaged
over the sample. Compared to the H2 adsorbate density at 50 K on activated carbon of 0.06 g cm−3 [93], this
ambient temperature adsorbate density is slightly lower, as might be expected for a higher temperature.

8
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 3. Hydrogen uptake isotherms and adsorbate density at ambient temperature to 2000 bar: plot A, excess isotherm
determined from the experimental measurements; B, absolute isotherm calculated from the excess isotherm and the volume of
the adsorbate; C, calculated hydrogen adsorbate density. Reproduced from [91], with permission from Springer Nature.

Densities in the range 0.05–0.07 g cm−3 have been reported on MOFs at cryogenic temperatures. All these
values are lower than the density of liquid H2 at the boiling point and 1 bar, which is 0.071 g cm−3 .

3. Assessing hydrogen storage performance of nanoporous materials

Various measures of the hydrogen storage capabilities of a material have been employed over many years. The
primary technique is calculation of the uptake of H2 as a function of pressure, made by measurements of
pressure and temperature over the range of applied pressure. Values for the uptake can be used to determine
the gravimetric and, together with knowledge of the volume occupied by the sample, the volumetric,
capacities. Another tool for probing hydrogen storage properties is neutron scattering which can yield
additional information, such as the location of the adsorbed H2 .
More recently, computational simulations, such as force field and first-principles density functional
theory (DFT)-based atomistic modelling, as well as ML-based high throughput screening, are delivering new
information and guidance for experimental approaches to new potential materials for hydrogen storage.
Finally, hydrogen storage measurements are only useful if they are accurate and interlaboratory
comparison studies indicate that improvements are required, particularly with regard to using consistent
terminology and publishing sample preparation details.

3.1. Volumetric and gravimetric capacities


While the amount of hydrogen stored in or on a material is a useful measure of the capacity of a potential
hydrogen storage material, the space occupied by the material is also of importance. This is particularly so for
practical applications, such as passenger vehicles where there is limited space for fuel tanks. The former
capacity, the gravimetric capacity, is usually described in terms of the weight of the hydrogen compared to
the weight of the material, whereas the volumetric capacity represents the mass of hydrogen per unit volume
occupied by the material.
These capacities and the correlation between the two are discussed in detail below, together with the
usable capacity, which is the capacity after practical constraints such as maximum tank pressures and
minimum fuel-cell feed pressures are taken into account.

9
Prog. Energy 4 (2022) 042013 L Zhang et al

3.1.1. Correlation of volumetric and gravimetric H2 storage


From a gravimetric perspective, H2 storage capacity refers to the amount of H2 adsorbed per unit mass,
expressed for instance as g H2 per g of adsorbent or wt% (g H2 per 100 g of H2 loaded adsorbent). At lower
temperatures (77 K) and higher pressures (>20 bar), the maximum storage capacity has been found to be
directly related to the SSA accessible to H2 molecules, showing a linear correlation, known as Chahine’s rule
[58]. This results in 1 wt% H2 uptake per 500 m2 g−1 of BET area at 77 K. Most of the early development of
porous materials for H2 storage focused on optimising surface area, with the highest excess gravimetric H2
adsorption capacity of 9.95 wt% measured at 77 K for a material, NU-100, with a SSA of 6143 m2 g−1 [94].
Porous materials with very high SSAs, however, tend to have large pores, which increases the free volume in
their open frameworks and decreases their volumetric H2 storage capacity. A trade-off between total
volumetric and gravimetric H2 capacity has since been identified [60], indicating that MOFs with the best
gravimetric performance will generally exhibit relatively modest volumetric capacities.
Volumetric capacity is a primary consideration when evaluating porous materials for H2 storage and
becomes an issue of increasing importance in applications where economy of space is crucial. For instance in
the transport sector, and especially for light-duty vehicles, the volume of the H2 storage tank will be limited.
It is therefore the limiting factor in determining the driving range of a vehicle. Volumetric capacity refers to
the amount of H2 adsorbed per unit volume, as g H2 l−1 , in a volume of the tank filled with adsorbent. For
an adsorbent, storage capacity is governed by the accessible surface area for the gas; therefore, volumetric
capacity will be determined by the surface area per unit volume, i.e. volumetric surface area, rather than the
gravimetric SSA. This results in an analogous relation to Chahine’s rule, between volumetric uptake and
volumetric surface area, which was recently verified with experimental data using the packing density and the
single crystal density of MOFs [59]. In the same paper, a phenomenological model was developed, based on
experimental data that show a direct correlation between volumetric and gravimetric uptake. This suggests
that only increasing SSA will not produce a significant increase in the volumetric capacity.
To increase the available volumetric surface area, interpenetrated MOFs have been investigated.
Interpenetration, meaning the intergrowth of two or more frameworks, is often viewed as a problem when
trying to synthesise MOFs [97]; however, in certain cases, it has been found to reduce pore volume, thus
increasing the volumetric surface area. CFA-7, the interpenetrated network of the MFU-4 family, has been
reported to have a volumetric surface area of 2697 m2 ml−1 , while the non-interpenetrated MFU-4l possesses
a volumetric surface area of only 1670 m2 ml−1 . The absolute volumetric H2 storage capacity at 77 K
therefore increases from 25 g H2 l−1 to 50 g H2 l−1 at 20 bar in MFU-4l and CFA-7, respectively [98].
In recent years, although many nanoporous materials, with different pore sizes, structures and SSAs, have
been studied for H2 storage, the packing density of adsorbent beds has often been overlooked. Typically, the
single-crystal density based on crystallographic analysis is used to calculate the volumetric capacity of
crystalline materials such as MOFs. However, packing density is important for practical applications. Packing
density is commonly a factor of two lower than the single-crystal density. Zacharia et al [95] showed that by
compacting powder samples, packing density can be increased and the specific volume and interparticle
voids reduced. In turn, by eliminating the interparticle voids, the surface area per volume will increase and
result in higher volumetric H2 uptake. Balderas-Xicohténcatl et al [59] reported a linear relation of the
inverse of the packing density versus SSA for different porous materials. As shown in figure 4, the specific
volume shows a linear correlation with SSA calculated for many MOFs using the single crystal volume (blue)
and the packing volume of the powder (red). In addition, the similar slopes of the plots of both packing and
single crystal volume indicate no significant loss of SSA. For the single crystal case, the intercept corresponds
to the skeletal volume, while for powders, the intercept corresponds to the skeletal volume plus the
interparticle void volume. This gap between the loose powder packing and single crystal densities can be
closed by compacting powder to form monoliths or pellets and reducing the interparticle void volume. Two
examples, using MIL-101 (green stars) [96] and MOF-177 (black crosses) [95], show that mechanical
compaction of powders reduces the specific volume down to the single crystal volume. For the case of
MOF-177, compacting the material further results in the loss of surface area by destruction of the pores and
the points follow the linear correlation between the single crystal volume and SSA.
Many attempts have been made to increase volumetric capacity by increasing packing density.
Pelletised/compacted SNU-70 exhibits the highest H2 capacity at 77 K and 100 bar of 33 g ml−1 [99]. The
current record improvement at 298 K is for zeolite-templated carbon/reduced graphene oxide (ZTC/rGO)
monoliths, which have been reported by Gabe et al [100] to exhibit a volumetric H2 storage capacity of
11.2 g l−1 , increasing ∼8% over pure compression. Using co-ordinatively unsaturated metal sites (or OMSs),
which exhibit strong interactions with H2 , also offers a way of increasing volumetric H2 storage capacity. The
adsorption strength of the positive charge density at the metal cation site increases the amount of gas
bound at the working temperature. The best performing material was Ni2 (m-dobdc) (m-dobdc4− =
4,6-dioxido-1,3-benzenedicarboxylate), with a record of 12 g H2 l−1 and 0.9 wt% at 25 ◦ C and 100 bar.

10
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 4. Specific volume using the packing density and single-crystal density (red circles and blue triangles, respectively) as a
function of the SSA for porous materials. Data from Zacharia et al [95] obtained for a mechanically densified MOF-177 are
included as black crosses and data from Blanita et al [96] for MIL-101 are included as green stars. The green arrow symbolises the
gap for improving the volumetric storage capacity, that can be closed by densification, pelletising or forming monoliths.

However, this approach is limited by the surface density of OMSs in the pores of the framework. Volumetric
and gravimetric H2 capacities are both key factors determining practical system performance in fuel cell
vehicles. It is therefore crucial to optimise the volumetric and gravimetric capacities as concurrent
parameters.

3.1.2. Usable capacity


Despite the importance of maximum gravimetric and volumetric capacities, usable capacity is another key
practical consideration for evaluating adsorbents for H2 storage. Maximum storage capacities typically
reported for adsorbents are defined as the amount adsorbed between vacuum and a maximum storage
pressure. From a practical perspective, however, a minimum pressure, in the range 1.5–5 bar, is required to
supply a back pressure to the fuel cell stack. Any H2 adsorbed at pressures below 1.5–5 bar will remain in the
material and will therefore be unusable. Usable or deliverable capacity is thus defined as the amount of H2
released from the adsorbent between full tank conditions and the fuel cell stack back pressure [35].
As mentioned before, surface area, commonly measured by Ar/N2 adsorption and calculated using the
BET method [22], defines the accessible area available for H2 adsorption (see section 2.1). Storage
performance, however, also depends on other adsorbent properties, such as total pore volume, which
determines saturation capacities at high pressure, and the magnitude and variation of the gas–solid
interaction potential. The latter depends not only on pore size, but also on surface chemistry. The right
balance of these properties is required, but deconvoluting the individual contributions is difficult. So,
although the search for effective H2 adsorbents has been ongoing for at least two decades, there are still
significant hurdles to developing and deploying nanoporous materials for H2 storage [16]. Regardless of the
impressive maximum gravimetric H2 capacities reported so far, optimum solutions, in terms of usable or
deliverable gravimetric and volumetric capacities, under different temperature and pressure operating
conditions, are yet to be achieved.
For a given material, the optimum operating temperature is the temperature at which the maximum
usable capacity is obtained. In principle, adsorption on materials exhibiting low ∆H st requires low
temperatures/high pressures, while stronger gas-solid interactions and therefore increased ∆H st may allow
significant adsorption at higher temperatures and/or lower pressures. In this respect, attempts have been
made to develop materials optimised for ∆H st , in order to elevate the operating temperature. The optimum
∆H st was estimated by Bhatia and Myers [101] to be 15–20 kJ mol−1 , to achieve H2 adsorption at ambient
temperature at the charging pressure. One way of increasing ∆H st is to introduce strong adsorption centres,

11
Prog. Energy 4 (2022) 042013 L Zhang et al

such as undercoordinated metals, to enhance the interaction strength, which will increase the adsorption
temperature. However, the surface chemistry is not the only property to be considered; pore size is also
important since in small pores, ∆H st increases due to overlap of van der Waals forces [13]. A higher ∆H st
increases the temperature at which H2 can be adsorbed, but it also leads to a weaker temperature dependence
of the maximum H2 uptake at a given pressure [102].
In this respect, Kapelewski et al [103], for example, employed a family of structural isomers of
M2 (m-dobdc), featuring M2+ cation sites with a higher apparent charge density, to increase H2 binding
enthalpies to 8.8–12.3 kJ mol−1 . The H2 capacity was ∼12 g H2 l−1 at 298 K. An even higher H2 binding
enthalpy (−21 kJ mol−1 ) has been reported by Jaramillo et al [104] for the V2 Cl2.8 (btdd) framework, which
contains a high density of exposed vanadium (II) sites. However, even though the total H2 capacity at
ambient temperature was enhanced, both materials showed no increase in usable capacity. To understand
this observation, the Langmuir–Freundlich model is applied, in which the usable capacity depends not
explicitly on temperature, but on the saturation adsorption and the ratio of the charging pressure to
discharge pressure of the tank. The optimum usable capacity can be described as follows:
( )η/
pmax 2
( ) pmin −1
kopt = k pmax , pmin , Topt = nm ( )η/ (9)
pmax 2
pmin +1

where kopt is the usable H2 storage capacity at optimum operating temperature, pmax is the maximum
permissible pressure of a tank, pmin is the required minimum pressure of a fuel cell, nm is the saturation
uptake of a sample in a monolayer, and η is Freundlich exponent.
The parameter η is a temperature-independent material constant. Nevertheless, for several materials with
variable enthalpies of adsorption, the following correlation between η and the enthalpy of adsorption is
suggested,

−4.63 kJ mol−1
η≈ . (10)
∆Hst

Based on the above equations, a lower enthalpy, which decreases the amount of H2 adsorbed at low
pressures, can increase the usable capacity at a particular temperature. In 2012, Schlichtenmayer and
Hirscher [13] studied a series of nanoporous materials and found a correlation between the average enthalpy
of adsorption and the excess H2 uptake at 77 K and 20 bar. However, the different enthalpies of adsorption
were mainly achieved by varying pore size, rather than by including strong adsorption sites, e.g. OMSs. Since
smaller pore size typically coincides with a reduction in SSA, a tendency of lower saturation uptake with
increasing enthalpy was observed. In 2016, the same authors evaluated the usable capacity, between 2 and
20 bar, and found a higher optimal operating temperature for materials with higher enthalpies of adsorption
[35]. Glante et al [105] recently reported a correlation between optimal operating temperature and usable
capacity, using the same analysis method. A series of MOFs were investigated and compared to zeolite Ca–A.
The optimal operating temperature for most of the MOFs was below 90 K, while for zeolite Ca–A it was
∼120 K. In addition, the usable fraction decreases if one is using a material like zeolite A or a carbon
molecular sieve, which have small pore diameters and thus a higher enthalpy of adsorption. This
phenomenon has also been confirmed by a computational study by Sun et al [106], in which the maximum
working capacity was predicted to decrease with increasing optimal temperatures, after screening 64 state
points.
Figure 5 summarises the reported data on the correlation of usable fraction and the optimum operating
temperature for different porous materials. As the materials have different SSAs (and thus uptakes) usable
fraction, defined as the usable capacity at the optimum operating temperature normalised to the uptake at
77 K, has been chosen for comparison. Overall, the higher enthalpy of adsorption associated with small pores
or strong adsorption sites increases working temperature, but at the expense of usable capacity. The trade-off
between optimal operating temperature and high working capacity is a major roadblock for adsorptive H2
storage, and needs to be addressed further.
In an ideal scenario, for practical applications, all H2 uptake/release would occur above the minimum
operating pressure of the storage unit, and usable capacity would equal total capacity; although this is usually
not the case. Typical rigid materials exhibit classical Langmuir-type absolute adsorption isotherms, where the
amount of gas adsorbed increases significantly at relatively low pressure, before reaching a plateau. It is
therefore difficult to improve the usable capacity of such adsorbents. To optimise usable capacity, an
adsorbent with an ‘S-shaped’ or ‘stepped’ adsorption isotherm is desired, in which the adsorbed amount

12
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 5. The usable fraction (usable capacity at the optimum operating temperature normalised to the uptake at 77 K and
25 bar) of all materials versus their optimum operating temperature. The materials PAF-1, DUT-6, DUT-8(Cu), DUT-9, and
IRMOF-1 are inserted at 77 K, since their optimum temperature could not be identified within the measured temperature range.

would be small at low pressure but would rise sharply just above the delivery pressure. Such stepped
isotherms have been reported for the flexible compounds Co(bdq) and Fe(bdq) by Mason et al [107] and
were attributed to a structural phase transition, enabling higher usable CH4 storage capacities than rigid
adsorbents. Flexible MOFs that exhibit ‘gate-opening’ behaviour, in which the non-porous structure expands
to a porous framework above a certain pressure, have shown hysteretic H2 adsorption behaviour, resulting in
a higher usable capacity. For example MIL-53(Al) exhibited flexibility during H2 adsorption, revealing an
increase in usable capacity [108].
Another option to increase usable capacity is to apply a temperature-pressure swing (TPS) when
emptying the tank, by warming the pressure vessel in its depleted state to a higher final temperature. This was
first considered by comparing MOF-177 and AX-21 in the pressure range up to 20 bar and at temperatures
from 77 K to 125 K and at room temperature [12]. AX-21_33 shows a usable capacity of 3.5 wt% in the case
of isothermal operation at 77 K, but 5.6 wt% when the tank is loaded at 77 K and then increased by 40 K
during unloading, with a pressure drop from 20 bar to 2 bar. Under the same condition, the usable capacity
of MOF-177 increases from 6.1 wt% to 7.4 wt%. As another example, H2 deliverable capacities under
conditions corresponding to charging at 100 bar and 77 K and discharging at 5 bar and 160 K, were evaluated
for 14 MOFs [109]. Among the MOFs studied, the gravimetric and volumetric deliverable capacities for
NU-125 (49 g l−1 , 8.5 wt%), NU-1000 (48 g l−1 , 8.3 wt%), and UiO-68-Ant (47 g l−1 , 7.8 wt%) are
promising for applications in H2 storage and delivery. Moreover, a recent experimental investigation was
carried out on NU-1501-Al, which shows one of the highest deliverable hydrogen capacities (14.0 wt%,
46.2 g H2 l−1 ) under a combined TPS from 77 K/100 bar to 160 K/5 bar [110].
The challenge for enhancing usable H2 storage capacity is thus twofold: identify materials whereby H2
drives structural phase transitions above a certain pressure at a practical temperature but also perform
systematic studies of different materials over a range of pressures and temperature. In both fields,
computational methods have become an increasingly powerful tool, both for explaining and interpreting
experimental results and for guiding experimental work.

13
Prog. Energy 4 (2022) 042013 L Zhang et al

3.2. Neutron scattering characterisation of adsorption systems


Understanding the properties of molecular H2 confined in nanopores is critical to designing and developing
new materials and/or processes towards improving current H2 storage technologies. The microscopic nature
of adsorbed molecular H2 depends on the interaction with the adsorbent surface and the geometry of the
pores [72], while confinement can result in phenomena such as phase transitions and hysteresis [111].
Moreover, an adsorbent can also undergo structural changes, due to flexibility, breathing, gate-opening, and
so forth, upon gas adsorption or other external stimuli that modify the local environment of the adsorbed H2
molecules [112].
The aforementioned widely used adsorption methodologies are bulk experimental approaches based on
statistically averaged observations. They are key techniques for assessing the performance of hydrogen
storage materials but fail to provide direct information on the atomic-molecular level. Scattering and
diffraction techniques either in powdered or single crystal samples has proven pivotal in resolving the
structure of new crystalline porous sorbents, while synchrotron radiation is in several cases essential due to
the low density of the materials. Neutron scattering on the other hand is an ideal tool for the in-situ
characterisation of H2 storage materials due to the large cross-section of hydrogen (1 H and 2 H). This strong
neutron scattering power means that atomic positions and motion can easily be detected [113, 114].
However, neutron scattering experiments require a neutron source, typically either a nuclear reactor or a
particle accelerator-based spallation source, which are generally large national or international facilities. For
interested readers, an earlier discussion of neutron scattering studies of H2 in nanoporous materials and
existing challenges was provided by Broom et al [15].
Elastic scattering involves no change in energy of the scattered neutrons and results in diffraction, which
provides atomic and/or magnetic structural information, similar to x-ray diffraction (XRD), but with the
difference that neutrons interact with atomic nuclei or magnetic moments, rather than the electrons of
atoms in a sample. Neutron powder diffraction (NPD) has been successfully applied to understand the
structural response of porous frameworks upon gas adsorption, such as MIL-53 [115, 116] and ZIF-7 [117],
and to determine adsorbate positions in crystalline porous frameworks [118–120].
INS is especially relevant for H2 storage in nanoporous materials because it provides a direct
spectroscopic probe of the dynamics and local environment of physisorbed H2 , which can be difficult to
detect using other techniques. It therefore provides valuable microscopic information on the effects of both
surface interactions and confinement. Typically, this characterisation is performed at low temperatures,
below the melting point (∼5 K), with H2 showing quantum properties such as rotational transitions. The
lowest possible rotational transition (J = 1 to J = 0) has a characteristic energy of 14.6 meV, known as the
free rotor transition. Such a transition is directly visible in an INS experiment and is affected by the local
environment of the H2 molecule. Confinement and adsorption potential symmetry can create a rotational
barrier that hinders the free rotor transition [121], and so the hindered rotor transition energies and
intensities can be studied to understand the local environment of H2 . Typically, only the molecules in direct
contact with the surface are affected by the rotational barrier and the hindering effect can also be used to
distinguish the different adsorbed positions in a complex interaction, such as in porous materials with
different pore sizes or strong adsorption sites [54]. INS has been used to study molecular H2 confined within
carbon-based materials [72, 122], MOFs [123, 124] and covalent organic frameworks [125]. In combination
with structural characterisation methods, such as XRD or NPD, INS can help explain complex adsorbate or
adsorbent behaviour, such as flexibility, phase transitions, or quantum effects.

3.3. Computational methods


In parallel to experimental studies, computer simulations and quantum chemical calculations have provided
key insights into the hydrogen storage properties of nanoporous materials. Grand canonical Monte Carlo
(GCMC) simulations have been widely used to calculate H2 adsorption capacities at specific temperatures
and pressures, while binding sites and energies can be determined using first principles methods, such as
electronic DFT and ab initio calculations. An earlier discussion of these methods and the challenges involved
in performing accurate simulations of H2 adsorption was provided by Broom et al [15], while Allendorf et al
[16] compared different theoretical methods of calculating H2 physisorption, including first principles
methods, ab initio and classical molecular dynamics, and GCMC simulations. The most common theoretical
methods used to compute H2 interactions with adsorbent materials have been summarised in [16], see table
3 therein. In this section, we will now consider the screening of MOFs using ML, and recent work assessing
the limits of the deliverable H2 capacity of flexible materials.

3.3.1. Screening MOFs for H2 storage using ML


A huge number of nanoporous materials have now been synthesised experimentally, with new materials
being reported almost on a daily basis. Some of these materials have been tested for their H2 storage capacity;

14
Prog. Energy 4 (2022) 042013 L Zhang et al

however, due to the expense and time requirements associated with physical experiments, a large proportion
of these nanoporous materials are yet to be evaluated experimentally for H2 storage. Over the past few
decades, with advances in computing hardware and software, as well as the development of accurate
nanoporous materials databases [126, 127] and efficient ML algorithms [128–130], it has become feasible to
run high-throughput computational screening on hundreds of thousands of experimentally synthesised and
hypothetical nanoporous materials to assess their suitability for H2 storage. This approach is much cheaper,
and candidate materials with targeted gas storage properties can be identified in a much shorter time scale,
and indeed this has led to the discovery of a number of nanoporous materials with excellent gas storage
properties.
ML has been used extensively over the last decade to identify materials with specific properties, in a range
of different fields. Once an ML model has been constructed, it can provide almost instantaneous predictions
for unknown materials. Traditionally, to develop predictive ML models, data from the literature
(experimental or theoretical), or data constructed for the purpose of the specific study, are employed. Several
ML algorithms are then trained and validated using the data before arriving at the best-performing
predictive model. In principle, assuming the ML descriptors are appropriate and the amount of data is
sufficient, the ML algorithm can provide predictions of unprecedented accuracy. The development and
evaluation of ML descriptors, as well as the accuracy of various ML algorithms for predicting H2 and other
gas adsorption capacities of MOFs, have been studied extensively over the last few years. For H2 adsorption
by MOFs, structural features, such as void fraction, surface area, pore volume, and so forth, have been used
as descriptors leading to accurate ML predictions; however, improved performance has been recorded when
energy-based descriptors were also employed [129, 131].
In one of the first applications of ML in this area, Borboudakis et al [130] constructed a database of 100
experimentally studied MOFs. The metal corners, linkers and functional groups were used as descriptors.
A number of different ML algorithms were capable of providing reasonable predictions for the gravimetric
capacity of these materials. A combination of ML and molecular simulations was used by Thornton et al
[132] to screen a library of ∼850 000 materials. Neural networks were trained to predict H2 adsorption by
the materials, using their structural features as descriptors and data generated using GCMC simulations.
Candidates with the most promising volumetric working capacities between 100 and 1 bar were identified.
More recently, Ahmed and Siegel [128] employed a diverse set of 918 734 MOFs. A sub-set comprising
24 674 MOFs was used to train the ML algorithm, while seven structural features were used as descriptors.
The extremely randomised trees algorithm [133] identified more than 8000 MOFs appropriate for pressure
swing (PS) and 95 materials for TPS which exceed, within the accuracy of the modelling parameterisation,
the gravimetric and volumetric capacities of state-of-the-art materials.
Fanourgakis et al [134] recently introduced an iterative self-consistent (SC) approach aimed at rapidly
identifying the top-performing materials from a large database of candidates, using a minimum amount of
information. The procedure is illustrated in figure 6 and is briefly described as follows: initially a data set is
created using information for a small number of materials. These materials could be selected from the
database either randomly or using semi-empirical models [60, 132]. An initial ML model is trained on this
data and is used for predictions for the remaining materials in the database. A predefined number of ML
predicted top-performing MOFs enriches the previous data set which is used, in turn, to construct a second
ML model. This procedure is repeated until the predefined number of the predicted top-performing
materials has been included in the training data set. Even though the ML predictions are not very accurate
during the first iterations, due to the small training set sizes, the majority of materials included in the
training data during the next iterations will have high capacities. As a result, successive ML models gradually
improve in the region of interest, namely for materials having large capacities. The final ML model provided
significantly higher accuracy for materials with large capacities compared to the materials with lower
capacities.
Application of the above method to CH4 adsorption by nanoporous materials [134] showed that more
than 70 of the 100 top-performing materials could be identified with only a small amount of information
(260–390 MOFs). It is important to mention that, while the previous approach was applied to two databases
with sizes differing by more than an order of magnitude (∼5000 MOFs and ∼67 000 covalent organic
frameworks (COFs)), the amount of information finally required was similar in both cases. Secondly, the
accuracy of the present approach (number of identified top-performing MOFs) is significantly higher than
when data sets of similar sizes, containing randomly selected materials, are used instead. For example, in an
application of the approach it was found [134] that the SC procedure converged requiring information for
only 306 of the 4763 CoRE MOFs. Among them were 76 of the top-100 performing MOFs for CH4 storage at
P = 5.8 bar. Instead, the ML model trained using a dataset of 306 randomly selected MOFs was capable of
successfully identifying only 50 of the top-100 materials. For constructing efficient ML schemes, aimed at
identifying top-performing materials, the proposed iterative procedure can therefore significantly reduce the

15
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 6. Flowchart of the iterative SC approach used for the identification of the top-performing MOFs for H2 storage.

number of required simulations. While in the previous application of the SC approach, CH4 adsorption
capacities of materials were computed using GCMC simulations, in principle, experimentally determined
adsorption capacities may be used as well.
The same methodology for identifying top-performing materials was employed for H2 storage.
Candidates were selected from a collection of experimentally synthesised and hypothetical MOFs created by
Ahmed et al [135]. The gravimetric capacity of 98 694 materials for PS between P = 100 bar and 5 bar at
T = 77 K, as well as for TPS between P = 100 bar at T = 77 K and P = 5 bar at T = 160 K was also computed
by GCMC simulations by the same researchers. Usable capacities, along with several structural features of the
MOFs, including void fraction, mass density, and pore limited diameter, were freely distributed by the
authors. These results were used to evaluate the approach in terms of accuracy (number of identified
top-performing materials) and efficiency (total amount of reference information required). During the
application of the SC approach, it was assumed that the adsorption capacities of 100 randomly selected
materials are known. Also, after each iteration, the top-100 materials predicted by the ML algorithm were
examined. Those not included in the training set were considered during the next iteration. After convergence
of the procedure, the results were evaluated by examining the number of materials that were among the
top-100 performing ones. In order to avoid any bias from the initial choice of the randomly selected MOFs,
the procedure was repeated 100 times and the average results were computed. The final results of the two
data sets are shown in figure 7. It is easily seen that under both conditions (PS and TPS) the SC approach was
capable of identifying 98 of the top-100 performing materials. On average, information for less than 300
materials was required (298 for PS and 253 for TPS). Since the number of candidates that were examined is
∼3 orders of magnitude larger than the materials for which GCMC simulations were needed (i.e. ∼100 000
versus ∼300) it can be concluded that the proposed SC approach combines both accuracy and efficiency.

3.3.2. Probing limits of deliverable H2 capacity by modelling intrinsically flexible materials


To achieve technical targets for on-board vehicular storage of H2 , it is important to understand the
fundamental limits of deliverable or usable capacity of nanoporous materials, as discussed in section 3.1.2.
The deliverable or usable capacity of a Langmuirian material is maximised by an optimal free energy of
adsorption ∆G represented by the orange isotherm in figure 8(a) and is decreased when adsorption is too
strong (blue) or too weak (green) [136]. Hydrogen’s adsorption enthalpy in porous materials via
physisorption is almost exclusively too weak to provide the deliverable capacity required by technical targets
(at non-cryogenic temperatures) [137]. Rare exceptions have been demonstrated where materials containing
OMSs approach the optimal [104], or even too strong [138], binding energetics; nonetheless, analogues of
these materials with a sufficiently high density of such adsorption sites have yet to be discovered [139]. Even
if the optimal ∆G of adsorption can be achieved in a high OMS density framework, rigid materials with

16
Prog. Energy 4 (2022) 042013 L Zhang et al

Figure 7. Evaluation of the performance of the SC approach on the identification of performing materials for H2 storage. In each
bar, the number of materials identified as top-performing (blue region) and the total number of materials used for the training of
the ML algorithm (grey region) are denoted. The percentage of successfully identified materials over the number of materials for
which accurate GCMC simulations were required is given in parentheses. The thermodynamic conditions and the database used
are denoted next to each bar. The red error bars at the top of each bar show the minimum and maximum values found during the
100 individual runs, while the green error bars the corresponding standard deviation.

Figure 8. (a) Schematic isotherms of three different Langmuirian materials where deliverable capacity (DC) is maximised from an
optimal ∆G of adsorption (orange) or reduced via too strong (blue) or too weak (green) binding of the adsorbate. An
intrinsically flexible material with a non-porous (black) isotherm can further maximise deliverable capacity over the optimal
Langmuirian material. (b) Examples of intrinsically flexible (left), Sr(NDC) [145], vs extrinsically flexible (right), Co(BDP) [141],
MOFs with visualisation of the non-porous and porous states.

Langmuir adsorption behaviour will still suffer a drop in H2 deliverable capacity (as with any gas) due to
unremovable capacity at the discharge pressure, as discussed in section 3.1.2.
One solution is to perform a temperature swing, whereby charging occurs at low temperatures and
discharging at high temperatures (also represented by blue and green isotherms, respectively) [140]. Another
way to circumvent this problem is to exploit the structural flexibility that many nanoporous materials
possess, which can be either intrinsic flexibility (e.g. linker rotation/vibration, see figure 8(b) left) or extrinsic
flexibility (e.g. ‘breathing’ behaviour, see figure 8(b) right) or both. Extrinsically flexible nanoporous
materials have been shown to deliver H2 via an S-shaped isotherm through a large volume contraction from
an open-pore phase (charged) to a narrow-pore phase (discharged), therefore increasing the H2 deliverable
capacity, represented by the black isotherm in figure 8(a) [141]. However, the weak interactions of H2 in this
non-OMS framework limited such a phase transition at feasible H2 pressures to cryogenic temperatures;
furthermore, such large volume contraction/expansion would also pose a significant challenge in practical
applications, as high mechanical stress associated with large volume changes in some of these materials may

17
Prog. Energy 4 (2022) 042013 L Zhang et al

affect long-term structural integrity and therefore adsorption capacity after successive cycles of
charging/discharging [142]. Significant attention is therefore placed on intrinsically flexible nanoporous
materials which do not experience significant volume change after repeated charging/discharging cycles, and
can maintain their structural integrity and mechanical stability under realistic operating conditions [143].
Computational modelling and simulation provide significant insights into such materials and elucidate how
and when pore geometry, H2 binding energetics, and host energetics can produce this desirable
non-Langmuirian adsorption profile [144].
In large-scale high-throughput screening studies of nanoporous materials, as described in the previous
section, a widely used approximation is the rigid structure assumption. This makes such studies feasible, but
also biases the results when the adsorbates of interest have size/shape commensurate with the pores of the
adsorbent [146]. In reality, many nanoporous materials have some degree of structural flexibility. One of the
consequences is that some of the flexible adsorbents with small pores under the rigid structure
approximation (e.g. close to or smaller than the kinetic diameter of a target adsorbate) can be nominally
non-porous; however, a slightly higher energy open pore configuration could be stabilised by uptake of H2
molecules at sufficiently high chemical potential, if H2 adsorption is sufficiently strong in the open state and
the penalty for framework distortion to the open state is sufficiently small. Such materials are destined to be
missed by these high-throughput screening studies.
To understand the limit of deliverable capacity in intrinsically flexible nanoporous materials, a statistical
adsorption model has been developed by Witman et al [144]. Taking CH4 as an example, it was
demonstrated that a perfectly designed nanoporous material with intrinsically flexible slit-pores could
achieve higher deliverable CH4 capacity than the best benchmark systems known to date, with little to no
total volume change [144]. Inspired by this flexible slit pore model, electronic DFT calculations and GCMC
simulations were performed, from which a known MOF (see figure 8(b) left) was identified that validates key
features of the statistical adsorption model. It was also demonstrated that the adsorption thermodynamics,
including the energy penalty associated with intrinsic adsorbent linker rotation/vibration and adsorbate
binding energy, can be isoreticularly tuned by modifying the linker as well as the metal species of an existing
intrinsically flexible MOF. While this study was initially focused on CH4 , it is envisaged that the same
computational framework can also be used to study H2 adsorption and estimate the limit of H2 deliverable
capacity in intrinsically flexible nanoporous materials.
The next challenge would be developing efficient computational approaches to identify intrinsically
flexible nanoporous materials with the right pore parameters that lead to optimal H2 deliverable capacity.
Ultimately, a large-scale high-throughput screening study, with carefully chosen and physically informed
materials descriptors, relies on a better fundamental understanding of the origin of such intrinsic structural
flexibility. An alternative and viable approach, however, would be looking into an existing nanoporous
materials database [126], which contains ‘cleaned’ MOF structures of which guest or solvent molecules were
removed from the pores, i.e. these MOF structures resemble the porous (‘charged’) state. Computational
geometry optimisations of these MOF structures in the absence of guest/solvent molecules, using DFT, for
example, may result in MOF structures that resemble the nonporous (‘discharged’) state, providing (a) the
open-pore to narrow-pore transition is enthalpy driven, and (b) there is no significant kinetic barrier for this
transition. Comparing the two MOF structures in different states and taking into account the prerequisites
on negligible or small overall volume change, as well as appropriate pore parameters, it may be possible to
identify some, if not all, of the intrinsically flexible nanoporous materials that have excellent H2 deliverable
capacity.
To conclude, effort should be made towards rationally designing nanoporous materials analogous to the
above flexible slit pore adsorption model, and we call for continued discovery of intrinsically flexible
nanoporous materials with high H2 deliverable capacity, where such materials remain hidden from rigid
structure screening studies due to their nominal non- or low-porosity.

3.4. Reproducibility and interlaboratory exercises


Experimental reproducibility has been an issue in hydrogen storage material research, due to the difficulties
of accurately characterising the hydrogen storage properties of materials [8, 147]. Problems have mostly
affected nanostructured and nanoporous materials, for which it can be challenging to comprehensively
characterise samples, particularly in terms of purity. Experimental errors when measuring H2 sorption can
also be particularly severe for low density materials and increase with increasing pressure [79, 148, 149].
Nanoporous materials of interest for H2 storage, including carbons and MOFs, are almost exclusively low
density, and nanoporosity can lead to difficulties in accurately determining sample volume, which is typically
required to perform accurate measurements. In light of these difficulties, Broom and Hirscher [150] recently
presented measurement and reporting guidelines, aimed at improving reproducibility in hydrogen storage
material research.

18
Prog. Energy 4 (2022) 042013 L Zhang et al

Table 1. The relative standard deviation for excess gravimetric hydrogen capacity for 2009 [151] and 2019 [154] for measurements at
liquid N2 and ambient temperatures.

Comparison of the relative standard deviation for excess gravimetric capacity measurements on different
carbon samples

Liquid N2 temperature Ambient temperature


2019 study 2019 study
Pressure (bar) 2009 study Sample 1 Sample 2 2009 study Sample 1 Sample 2

5 10.5 5.6 2.2 30 26.8 15.5


10 9.5 3.1 1.1 27.4 12.3 11.1
15 11.1 4.3 1.2 26.5 10.0 11.0
25 — — — 22.5 8.1 9.3
50 — — — 21.5 6.2 7.8

The challenges of obtaining agreement between H2 sorption measurements made in different


laboratories, which are at the heart of the above problems, were demonstrated by a round robin study
published in 2009. Zlotea et al [151] reported measurements of H2 adsorption on a porous carbon up to
200 bar, and found large variations between laboratories in the reported gravimetric capacities. A later, 2013
study involving 14 laboratories subsequently reported measurements of hydrogen absorption by a ball-milled
MgH2 sample [152]. Isotherms were measured at 553 and 593 K and both sets of results had a roughly 7%
relative standard deviation in the gravimetric capacity at the plateau equilibrium pressure. In 2016, a
comparison of measurements of gravimetric excess capacities for two different carbon samples was
published, involving four different research groups who each had over 10 years of experience in the hydrogen
storage field [153]. This study showed good measurement agreement and described explicit details of sample
preparation and calculations used, providing benchmark H2 adsorption data for the research community.
Nevertheless, these studies clearly illustrated the inherent difficulty in achieving agreement between H2
sorption measurements made in different laboratories.
Subsequently, in 2019, a study with 13 participating laboratories focused on both gravimetric and
volumetric H2 storage capacities [154]. This investigation involved measuring H2 adsorption by two porous
carbon samples, a pelletised and powder material with surface areas of 1270 m2 g−1 and 2400 m2 g−1 ,
respectively, at liquid N2 and ambient temperatures. In an attempt to compare the 2019 results to the 2009
study, the data from the 2009 work was interpolated to a common pressure and the relative standard
deviation was calculated based on the data in the supplementary information. Table 1 lists the relative
standard deviation of the 2019 and the 2009 [151] studies for the excess gravimetric capacities at various
pressures. It is important to note that data in table 1 represents measurements of different carbon materials,
and therefore different magnitudes of excess adsorption. There is also a difference in the methodology for the
statistical analysis of the two studies; the 2019 study includes all the reported data, while some outlying data
are removed from the analysis from the 2009 work. At both liquid N2 and ambient temperatures, the relative
standard deviation for Sample 1 and 2 in the 2019 study are roughly 50% lower than that of the 2009 data.
In addition to reporting gravimetric capacities for porous carbon samples, the most recent study
highlighted the importance of a consistent methodology for determining volumetric H2 adsorption capacity
(see section 3.1.1). Both excess and total volumetric capacities are normalised by a characteristic volume,
where the volume can be defined in different ways, for example, crystalline volume, packing volume and
envelope volume. A complete discussion of various volumes associated with H2 storage measurements can be
found elsewhere [155–157]. This study used packing volume. One striking finding was the difference in the
relative standard deviation of the packing volumes of the pelletised material (4.6%) compared to that of the
powder material (27.8%). This uncertainty was reflected in the relative standard deviations of the excess and
total volumetric capacities and was attributed to uncertainty in the packing volume measurement and/or real
differences in the sample packing densities. This study showed the need for robust and universally accepted
definitions, and consistent measurement protocols for reporting volumetric capacities. A material’s
volumetric capacity is an important performance metric for applications, and measurement protocols must
be explicit in order to provide meaningful comparison of a material’s hydrogen storage performance.

4. Future developments

The future progress of nanoporous materials for H2 storage depends on the one side on a large-scale and
environmentally-friendly production of affordable materials and on the other side on a reliable and highly

19
Prog. Energy 4 (2022) 042013 L Zhang et al

reproducible characterisation of materials properties. The production of materials from sustainable sources
will gain importance in a world with limited resources and reduced energy consumption. New
interlaboratory tests on hydrogen adsorption properties over a wide range of temperatures and pressures will
be needed to get reliable enthalpies of H2 adsorption, which are required for designing large storage systems.
This brings up the demand of reference materials for validation of high pressure H2 adsorption
measurements on porous materials and special challenges in characterising materials under extreme
pressures.

4.1. Sustainable material production


Worldwide population growth has led to increasing demand for food and other basic resources, resulting in
the intensification of agricultural and industrial activities, and hence in the amount of generated waste.
Global annual production of biomass waste and of municipal solid waste (MSW) are of the order of 140 Gt
[158, 159] and over 2.1 Bt (0.74 kg/person/day) [160], respectively. Global waste production is expected to
grow to 3.40 billion tonnes by 2050, resulting from more than double population growth over the same
period. Waste management is therefore critical, and the availability of rapid, low-cost, and economically
feasible methods of recovering by-products and turning them into new sources is of considerable academic
and industrial interest. Globally, organic (food and green) waste constitutes 44% of total MSW, followed by
paper and cardboard (17%), and plastic (12%) [161]. All these materials can be considered, from a circular
economy perspective, as a rich and underutilised renewable source of carbon-based materials.
Biochar, a carbon-rich solid residue (C content from 40% to 90%) formed by the pyrolysis of biomass at
moderate temperatures (from 673 K to 873 K), has been attracting growing attention in recent years, due to
abundant surface functional groups that can have significant effects on the thermodynamics of
heterogeneous reactions occurring at the interface, and the possibility of easily tuning porosity by varying the
thermal treatment temperature and duration. The application of biochar-based materials as functional
composites in energy storage and conversion field is very promising [162]. Concerning H2 storage, biochar
must be activated using, for example, KOH, ZnCl2 or steam [163] in order to create the microporosity
required for H2 adsorption [164]. The biochar material prepared by Xiao et al [165], by pyrolysis at 1073 K
for 4 h under N2 atmosphere and subsequent KOH activation of melaleuca bark, for example, shows a surface
area up to 3170 m2 g−1 , a micropore volume up to 0.86 cm3 g−1 , and has a H2 storage capacity of 4.08 wt%
at 77 K and 10 bar. Zhang et al [166] prepared microporous carbon with surface area of 3200 m2 g−1 and
pore volume of 1.44 cm3 g−1 via pyrolysis and KOH activation of cornstalks and obtained a maximum H2
adsorption capacity of 4.4 wt% at 77 K and 40 bar. Biochar was prepared from rice bran pyrolysis at 923 K
for 12 h and subsequent activation with KOH, obtaining an activated material with a BET area of
2270 m2 g−1 and a pore volume of 1.22 cm3 g−1 , able to reversibly store up to 4.2 wt% H2 in less than 1 min
at 77 K and 8 bar (manuscript in preparation). One drawback of these materials is the low H2 adsorption at
room temperature (maximum 0.85 wt%), requiring high pressure (up to 100 bar) [167]. Intercalation with
alkaline and alkaline earth metals and decoration with transition metals will be attempted on biochar from
cereals and vegetable peels to verify their effect in inducing/improving the absorption performance close to
room temperature, potentially making these materials appealing for practical applications.
A rather young field is ‘green’ synthesis of MOFs by either new routes at room-temperature and in water,
avoiding energy demanding solvothermal processes [168], or the use of sustainable precursors based on
waste materials [169, 170]. MOFs can also be prepared from waste: polyethylene terephthalate bottles, for
example, were used to obtain the acid linker terephthalic acid, that served to prepare MOF-5 and UiO-66(Zr)
[171, 172]. This last material showed a SSA of 814 m2 g−1 , micropore volume of 0.28 cm3 g−1 , and adsorb
1.2 wt% H2 at 77 K and 1 bar (to be compared with 1.5 wt% for the commercial cage).

4.2. New inter-laboratory exercises


The interlaboratory exercises performed to date have focussed on measuring H2 adsorption capacities at
77 K and ambient temperature. Another important parameter, however, is the enthalpy of adsorption, ∆H st ,
or the coverage dependent isosteric heat of adsorption, qst , which is a metric for the strength of the
adsorbent-adsorbate interaction, often used to assess adsorbent properties and compare materials. ∆H st can
be measured directly via calorimetry [25, 173] or indirectly through applying the CC approximation to
adsorption isotherms taken at different temperatures. Derivation of the CC approximation requires
assumptions that may not hold under experimental conditions e.g. the gas in the bulk phase behaves ideally,
and the molar volume of the bulk gas is much greater than the molar volume of the gas in the adsorbed
phase. In addition, the CC approximation is based on absolute adsorption while isotherm measurements
typically measure excess adsorption. Since absolute adsorption cannot be directly measured experimentally,
conversion of excess to absolute is required.

20
Prog. Energy 4 (2022) 042013 L Zhang et al

The CC approximation is given by,

∆Hst 1
ln (P) = − +C (11)
R T
( )
ln (p1 /p2 )
∆Hst = RT1 T2 . (12)
T2 − T1 n

The enthalpy of adsorption can be calculated using equation (11) by plotting the natural log of the
equilibrium pressure, ln(P), for a constant coverage, n, as a function of inverse temperature, 1/T, for various
measured isotherms. The enthalpy of adsorption is determined from the slope of the line multiplied by the
ideal gas constant, R. The method results in ∆H st as a function of moles adsorbed. Equation (12) is a
discretised form of equation (11) that can be used with isotherms measured only at two temperatures. Using
either equation requires fitting the experimental data with a model isotherm to determine
pressure-temperature pairs corresponding to constant moles adsorbed at each isotherm temperature. Details
of how the methods are carried out can be found elsewhere [34, 157].
Several reports in the literature have highlighted discrepancies in calculated enthalpies of adsorption via
the CC approximation arising from differences in how the equations are applied and how the experimental
data is processed [25, 174]. Examples include: whether the CC equation is corrected for the non-ideality of
the gas; if and how the data is converted from excess to absolute sorption and the associating assumptions;
which isotherm model is used to fit the data; the number of isotherms included in the analysis; the isotherm
temperature range; and finally if each isotherm is fit individually or if a single temperature dependent model
is used to fit all the isotherms. These variation in analyses make it inherently challenging to judiciously
compare competing materials.
An interlaboratory study for determining the enthalpy of adsorption and the associated range of
variability would elucidate if and how, and to what extent, different variables and approaches affects the
reported results. The objective would be to develop a consensus protocol that allows researchers to converge
on best practices for determining the isosteric heat of adsorption. This will enable the hydrogen storage
community to better define and compare adsorbate–adsorbent interaction strengths.
An effective study would require identification of a reference material which is uniform, abundant, stable,
with a well-defined activation protocol, and an appropriate and measurable adsorption capacity. The
material should be well-characterised and have a limited number of homogeneous adsorption sites. Specific
parameters for all experiments would be clear and well-defined, including the degassing procedure, sample
size, equilibrium time, and pressure range.
The focus would first be to obtain isothermal adsorption measurements at the most experimentally
accessible temperatures, for example, 77 K, 89 K, 273 K, and 303 K. Laboratories with more extensive
temperature control capabilities could provide isotherm data at additional temperatures. Each research
group would be asked to calculate the enthalpy of adsorption using the method they deem most accurate to
characterise the extent of deviation within the reported values. In addition, using their same data, researchers
would be asked to calculate the enthalpy of adsorption with a common method to determine if consistent
results are achievable for material evaluation across different laboratories. The relative standard deviation of
the isosteric heat of adsorption will provide a measure of the sensitivity of the calculation method and how
differences in measurements are reflected in the calculated data.

4.3. Reference materials


Reference materials are invaluable for validating measurements of any physical or chemical property.
Organisations such as the US National Institute of Standards and Technology (NIST) and the Bundesanstalt
für Materialforschung und -prüfung, in Germany, provide reference materials for a vast array of purposes. As
discussed above, problems have occurred in hydrogen storage material research due to difficulties in
obtaining agreement between high pressure H2 sorption measurements made on the same material in
different laboratories. However, there are currently no reference materials available for H2 adsorption in
nanoporous materials, although a EURAMET project, MefHySto, is seeking to rectify this situation.
Reference measurements of high pressure CO2 and CH4 adsorption on the NIST reference materials, RM
8852 and RM 8850, respectively, have been published recently [175, 176], but there is a clear need for
equivalent work to be performed for H2 adsorption. RM 8852 is ammonium ZSM-5 zeolite and RM 8850 is
zeolite Y. The key point is that the availability of a stable nanoporous material, for which the high pressure
H2 adsorption capacity has been determined reproducibly in different laboratories using different apparatus
and techniques, would allow independent researchers and laboratories to test the accuracy of their
measurement apparatus and protocols. This is a crucial precursor to performing accurate and reproducible

21
Prog. Energy 4 (2022) 042013 L Zhang et al

H2 adsorption measurements on new materials, and so identification of a suitable reference material would
have a significant impact on the field.

4.4. Measurement challenges


A considerable amount of computational and experimental research has been conducted on H2 storage in
nanoporous materials over the last two decades or so, as outlined in this article and elsewhere [13, 15, 16, 21,
31, 155]. However, there are perhaps surprising gaps in our knowledge of H2 adsorption by different
materials over a wide range of temperatures and pressures. This information is of fundamental interest, but
the H2 adsorption behaviour of materials, as a function of temperature and pressure, is also required to
accurately model the performance of adsorptive H2 storage units, particularly under different TPS
conditions.
The majority of H2 adsorption data in the literature has been measured at 77 K, due to the relative ease of
performing adsorption experiments at this temperature using liquid N2 . Variable low temperature
measurements are far more scarce, as are measurements performed beyond 200 bar, at any temperature. It
could be argued that H2 adsorption beyond 200 bar is of limited technological relevance, as the aim of using
adsorption to store H2 is to lower the upper storage pressure to 100 bar or so. However, this does not mean it
is not worthy of further study. For some materials, for example, framework flexibility could lead to changes
in adsorption behaviour or adsorbed phase density under different temperature and pressure conditions.
With regard to the challenges of such work, it is important to note the variability and irreproducibility of
high pressure H2 adsorption data. It has proven difficult to achieve good agreement between data measured
at a single temperature, 77 K, for various reasons [79, 150], so it seems likely that data measured at variable
temperatures could be subject to even greater variation. The practical significance of variable temperature
data is the possible need to use TPS conditions for storage units, in order to maximise usable capacity. As
noted recently by Humayun and Tomasko [88], there has been only a limited number of experimental studies
on the usable capacity of different materials under temperature swing conditions.
Two of the fundamental issues with measuring H2 adsorption at high pressure and variable temperature
are accurately describing the density of high pressure H2 , as a function of temperature and pressure, and the
density or volume of a material with pores of molecular dimensions. Small errors in the gas density and the
volume or density of the solid can lead to large errors in the calculated excess adsorption isotherms. Practical
issues include the challenge of operating high accuracy apparatus with H2 , which can be particularly
susceptible to leakage. Leakage must be avoided when making H2 adsorption measurements [79, 150].
Commercial apparatus for such measurements is not widely available and so measurement systems must be
custom built [91]. It is therefore likely that this will remain a rather specialist area. To the best of our
knowledge, no instruments for making H2 adsorption measurements to pressures significantly greater than
200 bar, but at variable low temperatures, have been reported. It is possible that in-situ neutron scattering
experiments could be used to probe the behaviour of H2 in nanoporous materials at high pressure and
variable low temperatures, but the development of appropriate in-situ cells for such work brings its own
challenges.

5. Conclusion

This article has addressed several important aspects of the fundamentals of hydrogen storage in nanoporous
materials, as well as assessment of the performance of different materials, particularly in terms of usable
volumetric and gravimetric capacity. Computational techniques, such as ML, have also been discussed, in the
context of searching for new rigid and flexible MOFs, together with other important issues, including
sustainable production of sorbents, and the use of interlaboratory exercises and the need for standard
reference materials, to help improve the reproducibility of experimental H2 adsorption data.
Research into the use of adsorption to store H2 in porous materials has evolved significantly in recent
years. Although adsorptive H2 storage was first reported over 40 years ago, using activated carbons, the
emergence of carbon nanotubes as a possible storage medium in the late 1990s and MOFs, in the early 2000s,
greatly increased the focus on H2 adsorption as a possible solution to the hydrogen storage problem,
alongside hydrides, which have a far longer history. Much of the earlier strategy and work, however, focussed
on maximising gravimetric SSA and thus increasing gravimetric capacities, while performance assessment
was carried out mainly by adsorption measurements at 77 K. Although this approach is still reasonable to
some extent, there has been a notable shift towards considering alternative working conditions, such as
temperature swings and adsorption at higher temperatures, and also assessing additional parameters such as
usable capacity and volumetric capacity, which have been discussed extensively in this article. Another
significant change in recent years has been the introduction of ML methods to rapidly test the H2 storage
capacity of newly synthesised materials or screen hypothetical structures for new candidates with high H2

22
Prog. Energy 4 (2022) 042013 L Zhang et al

uptake. This has allowed a more strategic approach, compared to simply synthesising a material or several
materials in a trial-and-error fashion and testing them for H2 storage experimentally or using GCMC
simulation to estimate H2 uptakes. Nevertheless, current ML approaches can certainly be improved by
considering variable working conditions, for example, pressure-temperature swings or avoiding missing
flexible frameworks, by assuming that all structures are rigid.
At the start of this article, different aspects of the adsorption of H2 by nanoporous materials were
discussed, including some of the physics behind intermolecular interactions and the adsorption of H2 in
porous materials at different temperatures and pressures, while densification of the nano-confined adsorbed
phase under diverse conditions was also detailed. As noted later, there is much that remains unknown about
the behaviour of H2 in nanoporous materials, under a wider range of temperatures and pressures than is
typically studied. Further work in this area may lead to a greater understanding of H2 adsorption, more
generally, and this could take us closer to the ultimate goal, which is the development of affordable hydrogen
storage tanks that can satisfy the technical requirements for practical applications, in terms of usable
volumetric and gravimetric capacity, refill time, safety and energy consumption.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Acknowledgments

This paper was realised within the framework of the Hydrogen Technology Collaboration Programme
(Hydrogen TCP) of the International Energy Agency (IEA) in Task 40 ‘Energy storage and conversion based
on hydrogen’.
RBX and MH kindly acknowledge funding from the EMPIR programme co-financed by the Participating
States and from the European Union’s Horizon 2020 research and innovation programme (Project Number:
19ENG03).
RBX gratefully acknowledges research funding from the Hydrogen Materials—Advanced Research
Consortium (HyMARC), established as part of the Energy Materials Network under the US Department of
Energy, Office of Energy Efficiency and Renewable Energy, Hydrogen and Fuel Cell Technologies Office,
under Contract Number DE-AC05-00OR22725.
Sandia National Laboratories is a multimission laboratory managed and operated by National
Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell
International, Inc., for the US Department of Energy’s National Nuclear Security Administration under
Contract DE-NA-0003525. M W, M A, and V S gratefully acknowledge funding from the US Department of
Energy, Office of Energy Efficiency and Renewable Energy, Hydrogen and Fuel Cell Technologies Office,
through the Hydrogen Storage Materials Advanced Research Consortium (HyMARC). This paper describes
objective technical results and analysis. Any subjective views or opinions that might be expressed in the paper
do not necessarily represent the views of the US Department of Energy or the United States Government.
S L acknowledges the use of the Sulis supercomputer through the HPC Midlands+ Consortium and the
ARCHER2 supercomputer through membership of the UK’s HPC Materials Chemistry Consortium, which
are funded by EPSRC Grant Nos. EP/T022108/1 and EP/R029431/1, respectively.
This work was authored in part by the National Renewable Energy Laboratory, operated by Alliance for
Sustainable Energy, LLC, for the US Department of Energy (DOE) under Contract No.
DE-AC36-08GO28308. Funding provided by Hydrogen Materials—Advanced Research Consortium
(HyMARC), established as part of the Energy Materials Network under the US Department of Energy, Office
of Energy Efficiency and Renewable Energy, Fuel Cell Technologies Office, under Contract Number
DEAC36-08-GO28308. The views expressed in the article do not necessarily represent the views of the DOE
or the US Government. The US Government retains and the publisher, by accepting the article for
publication, acknowledges that the US Government retains a nonexclusive, paid-up, irrevocable, worldwide
license to publish or reproduce the published form of this work, or allow others to do so, for US Government
purposes.
C M, D P and M R gratefully acknowledge research funding from Fondazione Cariplo, GHELF project
(Gaining Health and Energy from Lombard Agrifood Waste; n◦ 2019–2152).

ORCID iDs

Linda Zhang  https://orcid.org/0000-0003-3841-544X


Mark D Allendorf  https://orcid.org/0000-0001-5645-8246

23
Prog. Energy 4 (2022) 042013 L Zhang et al

Rafael Balderas-Xicohténcatl  https://orcid.org/0000-0002-5514-412X


Darren P Broom  https://orcid.org/0000-0002-1328-7376
George S Fanourgakis  https://orcid.org/0000-0001-6158-6824
George E Froudakis  https://orcid.org/0000-0002-6907-1822
Thomas Gennett  https://orcid.org/0000-0003-1259-6588
Katherine E Hurst  https://orcid.org/0000-0003-4596-9504
Sanliang Ling  https://orcid.org/0000-0003-1574-7476
Chiara Milanese  https://orcid.org/0000-0002-3763-6657
Philip A Parilla  https://orcid.org/0000-0002-2056-0289
Sarah Shulda  https://orcid.org/0000-0001-8946-1249
Vitalie Stavila  https://orcid.org/0000-0003-0981-0432
Theodore A Steriotis  https://orcid.org/0000-0003-4978-7513
Colin J Webb  https://orcid.org/0000-0001-6659-0726
Matthew Witman  https://orcid.org/0000-0001-6263-5114
Michael Hirscher  https://orcid.org/0000-0002-3143-2119

References
[1] Toyota 2017 Mirai product information (available at: https://ssl.toyota.com/mirai/assets/core/Docs/Mirai%20Specs.pdf)
[2] Zheng J, Liu X, Xu P, Liu P, Zhao Y and Yang J 2012 Development of high pressure gaseous hydrogen storage technologies Int. J.
Hydrog. Energy 37 1048–57
[3] Züttel A 2003 Materials for hydrogen storage Mater. Today 6 24–33
[4] Eberle U, Felderhoff M and Schueth F 2009 Chemical and physical solutions for hydrogen storage Angew. Chem., Int. Ed.
48 6608–30
[5] Aceves S M, Espinosa-Loza F, Ledesma-Orozco E, Ross T O, Weisberg A H, Brunner T C and Kircher O 2010 High-density
automotive hydrogen storage with cryogenic capable pressure vessels Int. J. Hydrog. Energy 35 1219–26
[6] Moreno-Blanco J, Petitpas G, Espinosa-Loza F, Elizalde-Blancas F, Martinez-Frias J and Aceves S M 2019 The storage
performance of automotive cryo-compressed hydrogen vessels Int. J. Hydrog. Energy 44 16841–51
[7] M Hirscher (ed) 2010 Handbook of Hydrogen Storage: New Materials for Future Energy Storage (New York: Wiley)
[8] Broom D P 2011 Hydrogen Storage Materials: The Characterisation of Their Storage Properties (London: Springer)
[9] Buschow K, Bouten P and Miedema A 1982 Hydrides formed from intermetallic compounds of two transition metals: a special
class of ternary alloys Rep. Prog. Phys. 45 937
[10] Jorgensen S W 2011 Hydrogen storage tanks for vehicles: recent progress and current status Curr. Opin. Solid State Mater. Sci.
15 39–43
[11] Langmi H, Walton A, Al-Mamouria M M, Johnson S R, Book D, Speight J D, Edwards P P, Gameson I, Anderson P A and
Harris I R 2003 Hydrogen adsorption in zeolites A, X, Y and rho J. Alloys Compd. 356 710–5
[12] Schlichtenmayer M, Streppel B and Hirscher M 2011 Hydrogen physisorption in high SSA microporous materials–a comparison
between AX-21_33 and MOF-177 at cryogenic conditions Int. J. Hydrog. Energy 36 586–91
[13] Schlichtenmayer M and Hirscher M 2012 Nanosponges for hydrogen storage J. Mater. Chem. 22 10134–43
[14] Tedds S, Walton A, Broom D P and Book D 2011 Characterisation of porous hydrogen storage materials: carbons, zeolites, MOFs
and PIMs Faraday Discuss. 151 75–94
[15] Broom D P et al 2016 Outlook and challenges for hydrogen storage in nanoporous materials Appl. Phys. A 122 151
[16] Allendorf M D et al 2018 An assessment of strategies for the development of solid-state adsorbents for vehicular hydrogen storage
Energy Environ. Sci. 11 2784–812
[17] Ebeling W, Kraeft W D and Kremp D 1986 Hydrogen plasma—phase diagram and properties Europhys. News 17 52–55
[18] Lochan R C and Head-Gordon M 2006 Computational studies of molecular hydrogen binding affinities: the role of dispersion
forces, electrostatics, and orbital interactions Phys. Chem. Chem. Phys. 8 1357–70
[19] Klontzas E, Tylianakis E and Froudakis G E 2011 On the enhancement of molecular hydrogen interactions in nanoporous solids
for improved hydrogen storage J. Phys. Chem. Lett. 2 1824–30
[20] Christmann K 1988 Interaction of hydrogen with solid surfaces Surf. Sci. Rep. 9 1–163
[21] Thomas K M 2009 Adsorption and desorption of hydrogen on metal–organic framework materials for storage applications:
comparison with other nanoporous materials Dalton Trans. 9 1487–505
[22] Thommes M, Kaneko K, Neimark A V, Olivier J P, Rodriguez-Reinoso F, Rouquerol J and Sing K S W 2015 Physisorption of gases,
with special reference to the evaluation of surface area and pore size distribution (IUPAC technical report) Pure Appl. Chem.
87 1051–69
[23] Gelb L D, Gubbins K, Radhakrishnan R and Sliwinska-Bartkowiak M 1999 Phase separation in confined systems Rep. Prog. Phys.
62 1573
[24] Thommes M 2010 Physical adsorption characterization of nanoporous materials Chem. Ing. Technol. 82 1059–73
[25] Kloutse A, Zacharia R, Cossement D, Chahine R, Balderas-Xicohténcatl R, Oh H, Streppel B, Schlichtenmayer M and Hirscher M
2015 Isosteric heat of hydrogen adsorption on MOFs: comparison between adsorption calorimetry, sorption isosteric method,
and analytical models Appl. Phys. A 121 1417–24
[26] Wang Q and Johnson J K 1998 Hydrogen adsorption on graphite and in carbon slit pores from path integral simulations Mol.
Phys. 95 299–309
[27] Gotzias A, Tylianakis E, Froudakis G and Steriotis T 2012 Theoretical study of hydrogen adsorption in oxygen functionalized
carbon slit pores Microporous Mesoporous Mater. 154 38–44
[28] van den Berg A W and Areán C O 2008 Materials for hydrogen storage: current research trends and perspectives Chem. Commun.
6 668–81
[29] Garrone E, Bonelli B and Areán C O 2008 Enthalpy–entropy correlation for hydrogen adsorption on zeolites Chem. Phys. Lett.
456 68–70

24
Prog. Energy 4 (2022) 042013 L Zhang et al

[30] Palomino G T, Carayol M L and Areán C O 2006 Hydrogen adsorption on magnesium-exchanged zeolites J. Mater. Chem.
16 2884–5
[31] Murray L J, Dincă M and Long J R 2009 Hydrogen storage in metal–organic frameworks Chem. Soc. Rev. 38 1294–314
[32] Rouquerol J, Rouquerol F, Llewellyn P, Maurin G and Sing K S W 2014 Adsorption by Powders and Porous Solids: Principles,
Methodology and Applications (Amsterdam: Elsevier) pp 467–527
[33] Czepirski L and Jagiełło J 1989 Virial-type thermal equation of gas–solid adsorption Chem. Eng. Sci. 44 797–801
[34] Nuhnen A and Janiak C 2020 A practical guide to calculate the isosteric heat/enthalpy of adsorption via adsorption isotherms in
metal–organic frameworks, MOFs Dalton Trans. 49 10295–307
[35] Schlichtenmayer M and Hirscher M 2016 The usable capacity of porous materials for hydrogen storage Appl. Phys. A 122 379
[36] Gregg S J and Sing K S W 1982 Adsorption, Surface Area and Porosity 2nd edn (London: Academic)
[37] Margenau H 1939 van der Waals forces Rev. Mod. Phys. 11 1–35
[38] London F 1930 Zur Theorie Und Systematik Der Molekularkräfte Z. Phys. 63 245–79
[39] London F 1930 Über einige Eigenschaften und Anwendungen der Molekularkräfte Z. Phys. Chem. B 11 222–51
[40] Young D M and Crowell A D 1962 Physical Adsorption of Gases (London: Butterworths Scientific Publications)
[41] Abrahamson A A 1963 Repulsive interaction potentials between rare-gas atoms. Homonuclear two-center systems Phys. Rev.
130 693
[42] Lennard-Jones J E 1931 Cohesion Proc. Phys. Soc. 43 461
[43] Müller A 1936 The van der Waals potential and the lattice energy of a n-CH2 chain molecule in a paraffin crystal Proc. R. Soc. A
154 624–39
[44] Kirkwood J G 1932 Polarisierbarkeiten, suszeptibilitaten und van der waalssche krafte der atome mit mehreren elektronen Phys.
Z. 33 57
[45] Olney T N, Cann N, Cooper G and Brion C 1997 Absolute scale determination for photoabsorption spectra and the calculation of
molecular properties using dipole sum-rules Chem. Phys. 223 59–98
[46] Karozis S, Charalambopoulou G, Steriotis T, Stubos A and Kainourgiakis M 2017 Determining the specific surface area of
metal-organic frameworks based on a computational approach Colloids Surf. A 526 14
[47] Dubbeldam D, Calero S and Vlugt T J 2018 iRASPA: GPU-accelerated visualization software for materials scientists Mol. Simul.
44 653–76
[48] Feynman R 1972 Statistical Mechanics: A Set of Lectures ed D Pines (New York: W A Benjamin)
[49] Sesé L M 1995 Feynman-Hibbs potentials and path integrals for quantum Lennard-Jones systems: theory and Monte Carlo
simulations Mol. Phys. 85 931–47
[50] Beenakker J, Borman V and Krylov S Y 1995 Molecular transport in subnanometer pores: zero-point energy, reduced
dimensionality and quantum sieving Chem. Phys. Lett. 232 379–82
[51] Teufel J, Oh H, Hirscher M, Wahiduzzaman M, Zhechkov L, Kuc A, Heine T, Denysenko D and Volkmer D 2013 MFU-4–a
metal-organic framework for highly effective H2 /D2 separation Adv. Mater. 25 635–9
[52] Li W et al 2016 Transformation of metal-organic frameworks for molecular sieving membranes Nat. Commun. 7 1–9
[53] Savchenko I, Gu B, Heine T, Jakowski J and Garashchuk S 2017 Nuclear quantum effects on adsorption of H2 and isotopologues
on metal ions Chem. Phys. Lett. 670 64–70
[54] Weinrauch I et al 2017 Capture of heavy hydrogen isotopes in a metal-organic framework with active Cu (I) sites Nat. Commun.
8 1–7
[55] Al Ghafri S et al 2022 Hydrogen liquefaction: a review of the fundamental physics, engineering practice and future opportunities
Energy Environ. Sci. 15 2690–731
[56] Anderson M and Swenson C 1974 Experimental compressions for normal hydrogen and normal deuterium to 25 kbar at 4.2 K
Phys. Rev. B 10 5184
[57] Silvera I F 1980 The solid molecular hydrogens in the condensed phase: fundamentals and static properties Rev. Mod. Phys. 52 393
[58] Chahine R and Bose T (International Association for Hydrogen Energy) 1996 Characterization and optimization of adsorbents
for hydrogen storage Hydrogen Energy Progress XI: Proc. 11th World Hydrogen Energy Conf. vol 2 pp 1259–64
[59] Balderas-Xicohténcatl R, Schlichtenmayer M and Hirscher M 2018 Volumetric hydrogen storage capacity in metal–organic
frameworks Energy Technol. 6 578–82
[60] Goldsmith J, Wong-Foy A G, Cafarella M J and Siegel D J 2013 Theoretical limits of hydrogen storage in metal–organic
frameworks: opportunities and trade-offs Chem. Mater. 25 3373–82
[61] Gómez-Gualdrón D A, Wang T C, García-Holley P, Sawelewa R M, Argueta E, Snurr R Q, Hupp J T, Yildirim T and Farha O K
2017 Understanding volumetric and gravimetric hydrogen adsorption trade-off in metal–organic frameworks ACS Appl. Mater.
Interfaces 9 33419–28
[62] Streppel B and Hirscher M 2011 Bet specific surface area and pore structure of MOFs determined by hydrogen adsorption at 20 K
Phys. Chem. Chem. Phys. 13 3220–2
[63] Diep P and Johnson J K 2000 An accurate H2 –H2 interaction potential from first principles J. Chem. Phys. 112 4465–73
[64] Balderas-Xicohténcatl R 2019 High-density hydrogen monolayer formation and isotope diffusion in porous media PhD Thesis
(University of Stuttgart Germany) (https://doi.org/10.18419/opus-10792)
[65] Schaeffer W, Smith W and Wendell C 1949 The adsorption of helium on carbon black at liquid helium temperatures J. Am. Chem.
Soc. 71 863–7
[66] Singh R P and Band W 1955 The anomalous monolayer adsorption of helium J. Phys. Chem. 59 663–5
[67] Steele W A 1956 Concerning a theory of multilayer adsorption, with particular reference to adsorbed helium J. Chem. Phys.
25 819–23
[68] Pace E and Siebert A 1959 Heat of adsorption of parahydrogen and orthodeuterium on graphon J. Phys. Chem. 63 1398–400
[69] Huber T and Huber C 1990 Adsorption of hydrogen on porous vycor glass J. Low Temp. Phys. 80 315–23
[70] Huber T and Huber C 1990 Vibrational spectroscopy of porous vycor glass: surface hydroxyl perturbations upon adsorption of
hydrogen J. Phys. Chem. 94 2505–11
[71] Edler K J, Reynolds P A, Branton P J, Trouw F R and White J W 1997 Structure and dynamics of hydrogen sorption in
mesoporous MCM-41 J. Chem. Soc. Faraday Trans. 93 1667–74
[72] Prisk T R, Bryan M and Sokol P E 2014 Diffusive and rotational dynamics of condensed n-H2 confined in MCM-41 Phys. Chem.
Chem. Phys. 16 17960–74
[73] Tanaka H, Kanoh H, El-Merraoui M, Steele W A, Yudasaka M, Iijima S and Kaneko K 2004 Quantum effects on hydrogen
adsorption in internal nanospaces of single-wall carbon nanohorns J. Phys. Chem. B 108 17457–65

25
Prog. Energy 4 (2022) 042013 L Zhang et al

[74] Setoyama N and Kaneko K 1995 Density of He adsorbed in micropores at 4.2 K Adsorption 1 165–73
[75] Balderas-Xicohténcatl R et al 2022 Formation of a super-dense hydrogen monolayer on mesoporous silica Nat. Chem. accepted
(https://doi.org/10.1038/s41557-022-01019-7)
[76] Kleitz F, Choi S H and Ryoo R 2003 Cubic Ia3d large mesoporous silica: synthesis and replication to platinum nanowires, carbon
nanorods and carbon nanotubes Chem. Commun. 17 2136–7
[77] Zohuri B 2019 Hydrogen Energy: Challenges and Solutions for a Cleaner Future (Cham: Springer) pp 121–39
[78] Oh H, Lupu D, Blanita G and Hirscher M 2014 Experimental assessment of physical upper limit for hydrogen storage capacity at
20 K in densified MIL-101 monoliths RSC Adv. 4 2648–51
[79] Broom D P and Webb C J 2017 Pitfalls in the characterisation of the hydrogen sorption properties of materials Int. J. Hydrog.
Energy 42 29320–43
[80] Sircar S 2001 Measurement of gibbsian surface excess AlChE J. 47 1169–76
[81] Myers A L and Monson P A 2002 Adsorption in porous materials at high pressure: theory and experiment Langmuir 18 10261–73
[82] Myers A L and Monson P A 2014 Physical adsorption of gases: the case for absolute adsorption as the basis for thermodynamic
analysis Adsorption 20 591–622
[83] Poirier E and Dailly A 2008 Investigation of the hydrogen state in IRMOF-1 from measurements and modeling of adsorption
isotherms at high gas densities J. Phys. Chem. C 112 13047–52
[84] Poirier E and Dailly A 2009 Thermodynamic study of the adsorbed hydrogen phase in Cu-based metal-organic frameworks at
cryogenic temperatures Energy Environ. Sci. 2 420–5
[85] Poirier E and Dailly A 2009 On the nature of the adsorbed hydrogen phase in microporous metal–organic frameworks at
supercritical temperatures Langmuir 25 12169–76
[86] Poirier E and Dailly A 2009 Thermodynamics of hydrogen adsorption in MOF-177 at low temperatures: measurements and
modelling Nanotechnology 20 204006
[87] Pini R 2014 Interpretation of net and excess adsorption isotherms in microporous adsorbents Microporous Mesoporous Mater.
187 40–52
[88] Humayun R and Tomasko D L 2000 High-resolution adsorption isotherms of supercritical carbon dioxide on activated carbon
AlChE J. 46 2065–75
[89] Möllmer J, Möller A, Dreisbach F, Gläser R and Staudt R 2011 High pressure adsorption of hydrogen, nitrogen, carbon dioxide
and methane on the metal–organic framework HKUST-1 Microporous Mesoporous Mater. 138 140–8
[90] Voskuilen T G, Pourpoint T L and Dailly A M 2012 Hydrogen adsorption on microporous materials at ambient temperatures and
pressures up to 50 MPa Adsorption 18 239–49
[91] Naheed L, Lamb K E, Gray E M and Webb C J 2021 Extracting adsorbate information from manometric uptake measurements of
hydrogen at high pressure and ambient temperature Adsorption 27 1–11
[92] Sircar S 1999 Gibbsian surface excess for gas adsorption revisited Ind. Eng. Chem. Res. 38 3670–82
[93] Poirier E 2014 Ultimate H2 and CH4 adsorption in slit-like carbon nanopores at 298 K: a molecular dynamics study RSC Adv.
4 22848–55
[94] Farha O K, Yazaydın A Ö, Eryazici I, Malliakas C D, Hauser B G, Kanatzidis M G, Nguyen S T, Snurr R Q and Hupp J T 2010 De
novo synthesis of a metal–organic framework material featuring ultrahigh surface area and gas storage capacities Nat. Chem.
2 944–8
[95] Zacharia R, Cossement D, Lafi L and Chahine R 2010 Volumetric hydrogen sorption capacity of monoliths prepared by
mechanical densification of MOF-177 J. Mater. Chem. 20 2145–51
[96] Blanita G, Coldea I, Misan I and Lupu D 2014 Hydrogen cryo-adsorption by hexagonal prism monoliths of MIL-101 Int. J.
Hydrog. Energy 39 17040–6
[97] Jiang H-L, Makal T A and Zhou H-C 2013 Interpenetration control in metal–organic frameworks for functional applications
Coord. Chem. Rev. 257 2232–49
[98] Balderas-Xicohténcatl R, Schmieder P, Denysenko D, Volkmer D and Hirscher M 2018 High volumetric hydrogen storage
capacity using interpenetrated metal-organic frameworks Energy Technol. 6 510–2
[99] Purewal J, Veenstra M, Tamburello D, Ahmed A, Matzger A J, Wong-Foy A G, Seth S, Liu Y and Siegel D J 2019 Estimation of
system-level hydrogen storage for metal-organic frameworks with high volumetric storage density Int. J. Hydrog. Energy
44 15135–45
[100] Gabe A et al 2021 High-density monolithic pellets of double-sided graphene fragments based on zeolite-templated carbon J.
Mater. Chem. A 9 7503–7
[101] Bhatia S K and Myers A L 2006 Optimum conditions for adsorptive storage Langmuir 22 1688–700
[102] Schmitz B, Müller U, Trukhan N, Schubert M, Férey G and Hirscher M 2008 Heat of adsorption for hydrogen in microporous
high-surface-area materials ChemPhysChem 9 2181–4
[103] Kapelewski M T et al 2014 M2 (m-dobdc)(M = Mg, Mn, Fe, Co, Ni) metal–organic frameworks exhibiting increased charge
density and enhanced H2 binding at the open metal sites J. Am. Chem. Soc. 136 12119–29
[104] Jaramillo D E, Jiang H Z, Evans H A, Chakraborty R, Furukawa H, Brown C M, Head-Gordon M and Long J R 2021
Ambient-temperature hydrogen storage via vanadium(II)-dihydrogen complexation in a metal–organic framework J. Am. Chem.
Soc. 143 6248–56
[105] Glante S, Fischer M and Hartmann M 2021 Investigation of the optimum conditions for adsorptive hydrogen storage Mergent
Mater. 4 1295–303
[106] Sun Y et al 2021 Fingerprinting diverse nanoporous materials for optimal hydrogen storage conditions using meta-learning Sci.
Adv. 7 eabg3983
[107] Mason J A et al 2015 Methane storage in flexible metal–organic frameworks with intrinsic thermal management Nature
527 357–61
[108] Mota J P, Martins D, Lopes D, Catarino I and Bonfait G G 2017 Structural transitions in the MIL-53 (Al) metal–organic
framework upon cryogenic hydrogen adsorption J. Phys. Chem. C 121 24252–63
[109] García-Holley P et al 2018 Benchmark study of hydrogen storage in metal–organic frameworks under temperature and pressure
swing conditions ACS Energy Lett. 3 748–54
[110] Chen Z et al 2020 Balancing volumetric and gravimetric uptake in highly porous materials for clean energy Science 368 297–303
[111] Narehood D, Grube N, Dimeo R, Brown D and Sokol P 2003 Inelastic neutron scattering of H2 in xerogel J. Low Temp. Phys.
132 223–37

26
Prog. Energy 4 (2022) 042013 L Zhang et al

[112] Schneemann A, Bon V, Schwedler I, Senkovska I, Kaskel S and Fischer R A 2014 Flexible metal–organic frameworks Chem. Soc.
Rev. 43 6062–96
[113] Chen Y-P, Bashir S and Liu J L 2017 Nanostructured Materials for Next-Generation Energy Storage and Conversion: Hydrogen
Production, Storage, and Utilization (London: Springer)
[114] Mitchell P C H, Parker S F, Ramirez-Cuesta A J and Tomkinson J 2005 Vibrational Spectroscopy with Neutrons: With Applications
in Chemistry, Biology, Materials Science and Catalysis vol 3 (Singapore: World Scientific)
[115] Pollock R A, Her J-H, Brown C M, Liu Y and Dailly A 2014 Kinetic trapping of D2 in MIL-53 (Al) observed using neutron
scattering J. Phys. Chem. C 118 18197–206
[116] Kim J Y et al 2020 Specific isotope-responsive breathing transition in flexible metal–organic frameworks J. Am. Chem. Soc.
142 13278–82
[117] Klein R A, Shulda S, Parilla P A, Le Magueres P, Richardson R K, Morris W, Brown C M and McGuirk C M 2021 Structural
resolution and mechanistic insight into hydrogen adsorption in flexible ZIF-7 Chem. Sci. 12 15620–31
[118] Dinca M, Dailly A, Liu Y, Brown C M, Neumann D A and Long J R 2006 Hydrogen storage in a microporous metal-organic
framework with exposed Mn2+ coordination sites J. Am. Chem. Soc. 128 16876–83
[119] Peterson V K, Liu Y, Brown C M and Kepert C J 2006 Neutron powder diffraction study of D2 sorption in Cu3 (1, 3,
5-benzenetricarboxylate)2 J. Am. Chem. Soc. 128 15578–9
[120] Yildirim T and Hartman M 2005 Direct observation of hydrogen adsorption sites and nanocage formation in metal-organic
frameworks Phys. Rev. Lett. 95 215504
[121] Requena A, Peña R and Serna A 1982 Perturbation for a rigid rotator in an electric field Int. J. Quantum Chem. 22 1263–70
[122] Georgiev P A, Ross D K, Albers P and Ramirez-Cuesta A J 2006 The rotational and translational dynamics of molecular hydrogen
physisorbed in activated carbon: a direct probe of microporosity and hydrogen storage performance Carbon 44 2724–38
[123] Rowsell J L C, Eckert J and Yaghi O M 2005 Characterization of H2 binding sites in prototypical metal-organic frameworks by
inelastic neutron scattering J. Am. Chem. Soc. 127 14904–10
[124] Liu Y, Brown C, Neumann D, Peterson V and Kepert C 2007 Inelastic neutron scattering of H2 adsorbed in HKUST-1 J. Alloys
Compd. 446 385–8
[125] Pham T, Forrest K A, Mostrom M, Hunt J R, Furukawa H, Eckert J and Space B 2017 The rotational dynamics of H2 adsorbed in
covalent organic frameworks Phys. Chem. Chem. Phys. 19 13075–82
[126] Chung Y G et al 2019 Advances, updates, and analytics for the computation-ready, experimental metal–organic framework
database: core MOF 2019 J. Chem. Eng. Data 64 5985–98
[127] Moghadam P Z, Li A, Wiggin S B, Tao A, Maloney A G, Wood P A, Ward S C and Fairen-Jimenez D 2017 Development of a
Cambridge Structural Database subset: a collection of metal–organic frameworks for past, present, and future Chem. Mater.
29 2618–25
[128] Ahmed A and Siegel D J 2021 Predicting hydrogen storage in MOFs via machine learning Patterns 2 100291
[129] Bucior B J, Bobbitt N S, Islamoglu T, Goswami S, Gopalan A, Yildirim T, Farha O K, Bagheri N and Snurr R Q 2019 Energy-based
descriptors to rapidly predict hydrogen storage in metal–organic frameworks Mol. Syst. Des. Eng. 4 162–74
[130] Borboudakis G, Stergiannakos T, Frysali M, Klontzas E, Tsamardinos I and Froudakis G E 2017 Chemically intuited, large-scale
screening of MOFs by machine learning techniques npj Comput. Mater. 3 1–7
[131] Fanourgakis G S, Gkagkas K, Tylianakis E and Froudakis G 2020 A generic machine learning algorithm for the prediction of gas
adsorption in nanoporous materials J. Phys. Chem. C 124 7117–26
[132] Thornton A W et al 2017 Materials genome in action: identifying the performance limits of physical hydrogen storage Chem.
Mater. 29 2844–54
[133] Geurts P, Ernst D and Wehenkel L 2006 Extremely randomized trees Mach. Learn. 63 3–42
[134] Fanourgakis G S, Gkagkas K, Tylianakis E and Froudakis G 2020 Fast screening of large databases for top performing
nanomaterials using a self-consistent, machine learning based approach J. Phys. Chem. C 124 19639–48
[135] Ahmed A, Seth S, Purewal J, Wong-Foy A G, Veenstra M, Matzger A J and Siegel D J 2019 Exceptional hydrogen storage achieved
by screening nearly half a million metal-organic frameworks Nat. Commun. 10 1–9
[136] Simon C M et al 2015 The materials genome in action: identifying the performance limits for methane storage Energy Environ.
Sci. 8 1190–9
[137] Llewellyn P L et al 2008 High uptakes of CO2 and CH4 in mesoporous metal-organic frameworks MIL-100 and MIL-101
Langmuir 24 7245–50
[138] Denysenko D, Grzywa M, Jelic J, Reuter K and Volkmer D 2014 Scorpionate-type coordination in MFU-4l metal–organic
frameworks: small-molecule binding and activation upon the thermally activated formation of open metal sites Angew. Chem.,
Int. Ed. 53 5832–6
[139] Witman M, Ling S, Gladysiak A, Stylianou K C, Smit B, Slater B and Haranczyk M 2017 Rational design of a low-cost,
high-performance metal–organic framework for hydrogen storage and carbon capture J. Phys. Chem. C 121 1171–81
[140] Kapelewski M T et al 2018 Record high hydrogen storage capacity in the metal–organic framework Ni2 (m-dobdc) at
near-ambient temperatures Chem. Mater. 30 8179–89
[141] Choi H J, Dinca M and Long J R 2008 Broadly hysteretic H2 adsorption in the microporous metal–organic framework Co(1,
4-benzenedipyrazolate) J. Am. Chem. Soc. 130 7848–50
[142] Bon V, Kavoosi N, Senkovska I and Kaskel S 2015 Tolerance of flexible MOFs toward repeated adsorption stress ACS Appl. Mater.
Interfaces 7 22292–300
[143] Kitagawa S, Kitaura R and Noro S I 2004 Functional porous coordination polymers Angew. Chem., Int. Ed. 43 2334–75
[144] Witman M, Ling S, Stavila V, Wijeratne P, Furukawa H and Allendorf M D 2020 Design principles for the ultimate gas deliverable
capacity material: nonporous to porous deformations without volume change Mol. Syst. Des. Eng. 5 1491–503
[145] Raja D S, Luo J-H, Yeh C-T, Jiang Y-C, Hsu K-F and Lin C-H 2014 Novel alkali and alkaline earth metal coordination polymers
based on 1, 4-naphthalenedicarboxylic acid: synthesis, structural characterization and properties CrystEngComm 16 1985–94
[146] Witman M, Ling S, Jawahery S, Boyd P G, Haranczyk M, Slater B and Smit B 2017 The influence of intrinsic framework flexibility
on adsorption in nanoporous materials J. Am. Chem. Soc. 139 5547–57
[147] Broom D P and Hirscher M 2016 Irreproducibility in hydrogen storage material research Energy Environ. Sci. 9 3368–80
[148] Blach T and Gray E M 2007 Sieverts apparatus and methodology for accurate determination of hydrogen uptake by light-atom
hosts J. Alloys Compd. 446 692–7
[149] Webb C J and Gray E M 2014 Analysis of the uncertainties in gas uptake measurements using the Sieverts method Int. J. Hydrog.
Energy 39 366–75

27
Prog. Energy 4 (2022) 042013 L Zhang et al

[150] Broom D P and Hirscher M 2021 Improving reproducibility in hydrogen storage material research ChemPhysChem 22 2141–57
[151] Zlotea C, Moretto P and Steriotis T 2009 A round robin characterisation of the hydrogen sorption properties of a carbon based
material Int. J. Hydrog. Energy 34 3044–57
[152] Moretto P et al 2013 A round robin test exercise on hydrogen absorption/desorption properties of a magnesium hydride based
material Int. J. Hydrog. Energy 38 6704–17
[153] Hurst K E, Parilla P A, O’Neill K J and Gennett T 2016 An international multi-laboratory investigation of carbon-based hydrogen
sorbent materials Appl. Phys. A 122 42
[154] Hurst K E et al 2019 An international laboratory comparison study of volumetric and gravimetric hydrogen adsorption
measurements ChemPhysChem 20 1997–2009
[155] Broom D P, Webb C J, Fanourgakis G S, Froudakis G E, Trikalitis P N and Hirscher M 2019 Concepts for improving hydrogen
storage in nanoporous materials Int. J. Hydrog. Energy 44 7768–79
[156] Parilla P A, Gross K, Hurst K and Gennett T 2016 Recommended volumetric capacity definitions and protocols for accurate,
standardized and unambiguous metrics for hydrogen storage materials Appl. Phys. A 122 201
[157] Gross K J et al 2016 Recommended Best Practices for the Characterization of Storage Properties of Hydrogen Storage Materials
(Golden, CO: EMN-HYMARC)
[158] Tripathi N, Hills C D, Singh R S and Atkinson C J 2019 Biomass waste utilisation in low-carbon products: harnessing a major
potential resource npj Clim. Atmos. Sci. 2 1–10
[159] UNEP 2015 Converting waste agricultural biomass into a resource (United Nations Environment Programme Division of
Technology, Industry and Economics International Environmental Technology Centre) (available at: www.unep.org/ietc/Portals/
136/Publications/Waste%20Management/WasteAgriculturalBiomassEST_Compendium.pdf)
[160] Sharma K D and Jain S 2020 Municipal solid waste generation, composition, and management: the global scenario Soc. Responsib.
J. 16 917–48
[161] World Energy Council, World Energy Resources 2016 Summary (available at: www.worldenergy.org/assets/images/imported/
2016/10/World-Energy-Resources-Full-report-2016.10.03.pdf)
[162] Liu W-J, Jiang H and Yu H-Q 2019 Emerging applications of biochar-based materials for energy storage and conversion Energy
Environ. Sci. 12 1751–79
[163] Liu W-J, Jiang H and Yu H-Q 2015 Development of biochar-based functional materials: toward a sustainable platform carbon
material Chem. Rev. 115 12251–85
[164] Züttel A, Sudan P, Mauron P and Wenger P 2004 Model for the hydrogen adsorption on carbon nanostructures Appl. Phys. A
78 941–6
[165] Xiao Y, Dong H, Long C, Zheng M, Lei B, Zhang H and Liu Y 2014 Melaleuca bark based porous carbons for hydrogen storage
Int. J. Hydrog. Energy 39 11661–7
[166] Zhang F, Ma H, Chen J, Li G-D, Zhang Y and Chen J-S 2008 Preparation and gas storage of high surface area microporous carbon
derived from biomass source cornstalks Bioresour. Technol. 99 4803–8
[167] Sevilla M and Mokaya R 2014 Energy storage applications of activated carbons: supercapacitors and hydrogen storage Energy
Environ. Sci. 7 1250–80
[168] Steenhaut T, Hermans S and Filinchuk Y 2020 Green synthesis of a large series of bimetallic MIL-100 (Fe, M) MOFs New J. Chem.
44 3847–55
[169] Crickmore T S, Sana H B, Mitchell H, Clark M and Bradshaw D 2021 Toward sustainable syntheses of Ca-based MOFs Chem.
Commun. 57 10592–5
[170] Dyosiba X, Ren J, Musyoka N M, Langmi H W, Mathe M and Onyango M S 2019 Feasibility of varied polyethylene terephthalate
wastes as a linker source in metal–organic framework UIO-66 (Zr) synthesis Ind. Eng. Chem. Res. 58 17010–6
[171] Roy P K, Ramanan A and Rajagopal C 2013 Post consumer PET waste as potential feedstock for metal-organic frameworks Mater.
Lett. 106 390–2
[172] Dyosiba X, Ren J, Musyoka N M, Langmi H W, Mathe M and Onyango M S 2016 Preparation of value-added metal-organic
frameworks (MOFs) using waste PET bottles as source of acid linker Sustain. Mater. Technol. 10 10–13
[173] Sircar S, Mohr R, Ristic C and Rao M 1999 Isosteric heat of adsorption: theory and experiment J. Phys. Chem. B 103 6539–46
[174] Pan H, Ritter J A and Balbuena P B 1998 Examination of the approximations used in determining the isosteric heat of adsorption
from the Clausius–Clapeyron equation Langmuir 14 6323–7
[175] Nguyen H G T et al 2020 A reference high-pressure CH4 adsorption isotherm for zeolite Y: results of an interlaboratory study
Adsorption 26 1253–66
[176] Nguyen H G T et al 2018 A reference high-pressure CO2 adsorption isotherm for ammonium ZSM-5 zeolite: results of an
interlaboratory study Adsorption 24 531–9

28

You might also like