ArticlePDF AvailableLiterature Review

Adsorption and Separation by Flexible MOFs

Wiley
Advanced Materials
Authors:

Abstract and Figures

Flexible metal–organic frameworks (MOFs) offer unique opportunities due to their dynamic structural adaptability. This review explores the impact of flexibility on gas adsorption, highlighting key concepts for gas storage and separation. Specific examples demonstrate the principal effectiveness of flexible frameworks in enhancing gas uptake and working capacity. Additionally, mixed gas adsorption and separation of mixtures are reviewed, showcasing their potential in selective gas separation. The review also discusses the critical role of the single gas isotherms analysis and adsorption conditions in designing separation experiments. Advanced combined characterization techniques are crucial for understanding the behavior of flexible MOFs, including monitoring of phase transitions, framework–guest and guest–guest interactions. Key challenges in the practical application of flexible adsorbents are addressed, such as the kinetics of switching, volume change, and potential crystal damage during phase transitions. Furthermore, the effects of additives and shaping on flexibility and the “slipping off effect” are discussed. Finally, the benefits of phase transitions beyond improved working capacity and selectivity are outlined, with a particular focus on the advantages of intrinsic thermal management. This review highlights the potential and challenges of using flexible MOFs in gas storage and separation technologies, offering insights for future research and application.
This content is subject to copyright. Terms and conditions apply.
REVIEW
www.advmat.de
Adsorption and Separation by Flexible MOFs
Irena Senkovska,* Volodymyr Bon, Antonia Mosberger, Yutong Wang, and Stefan Kaskel*
Dedicated to the 60th birthday of Prof. Omar Yaghi
Flexible metal–organic frameworks (MOFs) offer unique opportunities due to
their dynamic structural adaptability. This review explores the impact of
flexibility on gas adsorption, highlighting key concepts for gas storage and
separation. Specific examples demonstrate the principal effectiveness of
flexible frameworks in enhancing gas uptake and working capacity.
Additionally, mixed gas adsorption and separation of mixtures are reviewed,
showcasing their potential in selective gas separation. The review also
discusses the critical role of the single gas isotherms analysis and adsorption
conditions in designing separation experiments. Advanced combined
characterization techniques are crucial for understanding the behavior of
flexible MOFs, including monitoring of phase transitions, framework–guest
and guest–guest interactions. Key challenges in the practical application of
flexible adsorbents are addressed, such as the kinetics of switching, volume
change, and potential crystal damage during phase transitions. Furthermore,
the effects of additives and shaping on flexibility and the “slipping off effect”
are discussed. Finally, the benefits of phase transitions beyond improved
working capacity and selectivity are outlined, with a particular focus on the
advantages of intrinsic thermal management. This review highlights the
potential and challenges of using flexible MOFs in gas storage and separation
technologies, offering insights for future research and application.
1. Introduction
Metal–organic frameworks (MOFs) are crystalline coordination
polymers composed of metal ions or clusters coordinated by
organic ligands, facilitating the formation of inner cavities.[1]
Due to their dual inorganic/organic nature as coordination com-
pounds, the chemistry of MOFs is enormously rich. The vast
number of metal/ligand combinations grants these compounds
exceptional versatility, unlocking immense potential for a wide
I. Senkovska, V. Bon, A. Mosberger, Y. Wang, S. Kaskel
Chair of Inorganic Chemistry I
Technische Universität Dresden
Bergstrasse 66, 01069 Dresden, Germany
E-mail: irena.senkovska@tu-dresden.de;stefan.kaskel@tu-dresden.de
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adma.202414724
© 2025 The Author(s). Advanced Materials published by Wiley-VCH
GmbH. This is an open access article under the terms of the Creative
Commons Attribution License, which permits use, distribution and
reproduction in any medium, provided the original work is properly cited.
DOI: 10.1002/adma.202414724
range of applications, including storage and
separation of molecular species.[2]How-
ever, the vast diversity also turns MOFs into
a “haystack,” making it challenging for re-
searchers to find the “needle” when try-
ing to identify the best MOF for a specific
application.
Recognized early on for their poten-
tial as porous materials, MOFs have un-
dergone significant advancements since
their inception. At the early stage of de-
velopment, MOFs faced challenges with
maintaining permanent porosity, and ini-
tial structures often collapsed after remov-
ing guest molecules. However, advances in
desolvation techniques[3]and a better un-
derstanding of building principles[4]led to
the development of robust porous MOFs,
among others, with enhanced chemical sta-
bility. After this breakthrough, MOFs be-
came prominent for their ability to adsorb
and separate gases.
The primary interest in MOFs is,
nonetheless, motivated by their modular
construction and accessibility to deliberate
structural design and reticular chemistry.[5]
Access to isoreticular families of mate-
rials allowed for systematic studies of
structure–properties relationships.[6]The result of these designs
is ultrahigh porosity with surface areas of up to 7850 m2g1and
pore volume of 5.0 cm3g1.[7]MOFs still hold records for specific
surface area and pore volume among microporous adsorbents.
Such new horizons in terms of porosity and storage capacities
catapulted MOFs to high-performance adsorbents.[8]
However, the main feature distinguishing MOFs from other
porous adsorbents, such as activated carbons, is not only the crys-
tallinity of these materials and, as a consequence, the regular
and crystallographically precise pore structure. A unique feature
of some MOFs is the structural flexibility and the ability to un-
dergo stimuli-induced structural transitions.[9]Some MOF repre-
sentatives are able to switch between the crystalline phases with
different porosity induced by guest physisorption. A number of
terms have been proposed to manifest the flexibility of MOFs,
including “soft,” “flexible,” “dynamic,” “stimuli-responsive,” or
“switchable.” They all describe solid-state structural phase tran-
sitions initiated by relatively small energetic stimuli and usually
(but not necessarily) include considerable changes in the unit cell
and pore volumes.[9,10]
They are also emerging material classes, including covalent
organic frameworks (COFs)[11]and hydrogen-bonded organic
Adv. Mater. 2025, 2414724 2414724 (1 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 1. Schematic representations of the structural transitions between narrow pore phase (np) and open pore phase (op) induced by a general state
variable X: a) free energy F(V;X); b) response V(X), Reproduced with permission.[16]Copyright 2019, Elsevier Ltd.; as well as schematic representation
of the two corresponding framework states (c).
frameworks (HOFs),[12]sharing with MOFs key characteristics,
such as crystallinity, porosity, and ability to undergo structural
transitions. While this review focuses on MOFs, the general ob-
servations and dependencies discussed here are likely applicable
to these materials as well.
Although the variety of phase transitions inducing stimuli
is broad, including pressure, temperature, electric field, light
etc.,[13]we are going to focus on the structural changes caused
by the adsorption or desorption of molecules and the bene-
fits/drawbacks of them in fields of gas storage and separation.
In general, the structural transitions lead to a stepwise increase
(or decrease) of porosity between two extreme cases:[14]i) non-
porous state (closed pore phase, cp), resulting from the dense
packing of the building blocks and ii) maximal possible pore
size and volume for given framework topology (open pore phase,
op). The number of (meta)stable crystal structures (intermedi-
ate phases (ip), or narrow pore phases (np)) between these two
extremes in the given system depends on the free energy land-
scape for the given fluid/framework combination, e.g., the num-
ber of local minima and the height of activation barriers that sep-
arate them (multistability of free energy profile, Figure 1a). The
existence of the energy barrier between two minima in the free
energy landscape is the origin of hysteresis loops in the adsorp-
tion/desorption isotherms (Figure 1b).[15]
Depending on the initial state of the framework and the transi-
tion pathway, the gate-opening” and “breathing” transitions are
mainly discussed. Sometimes, the authors also use the terms in-
terchangeably. “Gate-opening” behavior is characterized by the
increase in the porosity upon the transition, whereby a less
porous structure (cp or np) expands to a more porous after a
certain threshold gas pressure (pgo) is reached. The physisorp-
tion isotherm, in this case, is characterized by the stepwise in-
crease in the adsorbed amount in a pretty narrow pressure region
(Figure 1c).
“Breathing” usually includes two consecutive transitions dur-
ing gas pressure increase: first contraction transition of the more
porous phase (op) to the less porous one (np) at characteristic
contraction pressure (pc), followed by second expansion transi-
tion (gate opening) to a more porous (large pore or op) phase at
characteristic pex. The first contraction is usually not easily iden-
tified in the isotherm (particularly if the phases have a minor
porosity difference) and have to be often detected by complemen-
tary techniques, such as in situ powder X-ray diffraction (PXRD)
(see Section 3).
The isotherms of flexible MOFs can be constructed by super-
imposing the isotherms of two or more structures with varying
adsorption capacities, adsorption enthalpies, etc. However, typi-
cally along the isotherm, the metastable states are observed. The
adsorption or desorption branch (or both) can be kinetically hin-
dered, and the energetic balance strictly depends on the num-
ber of guest molecules in the pores. Thus, the global minimum
state, among all possible configurations, changes depending on
the external guest pressure.[17]These activation barriers stem 1)
from the solid–solid phase transitions and are particularly high
if the volume change is high, but also 2) the fluid-phase activa-
tion barriers (diffusion, nucleation, etc.) may play a role. These
factors lead to hysteresis, which does not vanish even for long
equilibration times because the barriers depend exponentially on
one or more thermodynamic variables and can only be overcome
if the variable reaches a critical magnitude (such as gate open-
ing pressure, see below). Solid–solid phase transitions with high
activation barriers may be characterized as 1st order transitions
leading to steps in the isotherm. If the barriers for solid–solid
transitions are minimal, a quasi-continuous structural change is
observed, particularly when the volume changes are minor be-
cause the pores are already filled with linker functionalities.[18]
The fascination of flexible MOFs for separations originates
from the vision that an MOF may recognize a molecular stimulus
so specifically that it opens the pores only for this species, even if
it is only a minor component in a mixture of molecules, similar
to an enzyme recognizing its substrate.[19]However, the bene-
fits of the dynamic nature of MOFs for adsorptive application are
controversially discussed in the literature. Besides the benefits of
structural flexibility discussed below in detail (Sections 2.12.3),
there are some critical material-relevant, as well as technologi-
cal aspects related to flexibility. The primary concerns involve the
cyclability and the damage of the crystals in the course of mul-
tiple expansion/shrinkage cycles (Section 4.3) and the manage-
ment of the crystal volume change in the adsorption chamber
or column (Section 4.2). Since the first discovery of switchable
MOFs, the number of new MOFs showing pronounced flexibil-
ity and huge changes in pore size is increasing exponentially.
Also, some unique phenomena connected to gas-induced flex-
ibility, such as “negative gas adsorption were discovered, and,
Adv. Mater. 2025, 2414724 2414724 (2 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
despite open questions, more and more functions resulting from
pore switchability are reported. In the following, we discuss the
concepts of using flexible MOFs as efficient adsorbents, giving
some examples.
2. Impact of Flexibility and Concepts for Gas
Adsorption and Separation Using Flexible MOFs
2.1. Gas Storage Concept
Gas storage is one of the most essential applications of porous
materials, based on gas enrichment in an interfacial layer due to
solid–fluid interactions.
Volumetric and gravimetric capacities are of central impor-
tance for gas storage, although the temperature of operation and
thermal effects upon adsorption are also crucial. The aim is,
therefore, to achieve high volumetric and gravimetric capacities
at a usable temperature.[20]
The main indicator for the comparison of the storage perfor-
mance, however, is the so-called working (also usable) capacity,
reflecting the difference between the amount of gas adsorbed at
the target storage pressure and the amount that still be adsorbed
at the lowest desorption pressure, acceptable for the operation
of the system. In rigid adsorbents, the best performance can be
reached by the adjustment of the pore sizes to the operational
pressure and temperature range. The subatmospheric pressures
require (ultra)microporosity, whereas the mesoporous MOFs or
MOFs with hierarchical pore structure are beneficial for high-
pressure storage.[21]
The stepwise isotherms, characteristic for flexible MOFs,
open a new opportunity for boosting usable capacity, where the
amount of gas adsorbed would be negligible at low pressures (in
the pressure region below the lowest desorption pressure accept-
able) but rise sharply just before the pressure reaches the desired
storage pressure (Figure 2). Flexibility allows the material to close
pores at relatively high pressure and expel the gas molecules from
the pores. As a consequence, more gas can be released from the
pores in comparison to the rigid MOF with a comparable pore
size. In the pressure region above that is needed for the opening
transition, the gas uptake is equal to that of comparable micro-
porous adsorbent. Therefore, the optimal opening and closing
pressures are crucial for achieving high usable capacity. A criti-
cal aspect is hysteresis, which should be as narrow as possible to
achieve a high working capacity and reduce energy consumption.
The second benefit offered by flexible MOFs is thermal man-
agement. There are practical challenges involved in designing
systems with high capacities and managing the temperature
change associated with the adsorption and desorption of gas from
the adsorbent due to the high gas adsorption enthalpies in micro-
porous materials. In practical gas storage application systems, the
heat of adsorption (exothermic) and desorption (endothermic)
can lead to strong temperature variations,[23]which can have a
negative impact on the usable capacity. Rapid charging and dis-
charging are desirable, but they also lead to significant tempera-
ture fluctuations. For example, in the case of CH4storage, mea-
surements with commercially available Adsorbed Natural Gas
(ANG) cylinders show a temperature drop of up to 37 °C at high
discharge rates, with performance loss approaching 25% of the
Figure 2. Comparison of the typical adsorption isotherms of flexible MOF
(X, top) and rigid MOF with comparable pore size (Y, bottom). The work-
ing capacity in both cases is indicated by the arrow. Reproduced with
permission.[22]Copyright 2020, The Authors, under CC BY 4.0.
isothermal capacity. The performance loss is expected to be 15–
20% at moderate discharge rates.[24]Analysis indicates that the
thermal capacity of the vessel and external heat transfer condi-
tions significantly affect system behavior. MOFs have very low
thermal conductivity,[25]so additional efforts are required to ad-
dress thermal management challenges.
In flexible MOFs, fortunately, the enthalpy associated with the
phase transition (absorption or release of heat upon reversible
phase transitions) can counteract the thermal effect of adsorp-
tion, contributing to positive thermal management (Section 5).[26]
Net heat removed from the flexible system during the desorption
process under adiabatic conditions is the sum of the endothermic
heat associated with guest desorption, the exothermic heat gen-
erated by host shrinkage, and the exothermic heat of the guest
molecules that remain in the host framework due to changes in
host–guest interactions upon shrinkage of the host.[17]
Last but not least, flexible MOFs could potentially minimize
the “ageing” of adsorbent generated by the accumulation of im-
purities. The selectivity in adsorption (discussed in detail in
Adv. Mater. 2025, 2414724 2414724 (3 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Section 2.2) could lead to the adsorption of only the main, valu-
able components from the gas mixture.
For example, the adsorptive storage of methane often suffers
from the adsorption of impurities present in natural gas, such
as higher hydrocarbons (ethane, propane, etc.). These hydrocar-
bons can have a deleterious effect due to accumulation, decreas-
ing methane storage capacity, and affecting the long-term stabil-
ity of the adsorbent.[27]The selective adsorption of the main com-
ponent or forced desorption of the impurities supported by the
flexibility could improve long-term performance.
2.1.1. Examples of Flexible MOFs in Storage Applications
In this section, we explore the unique and critical features of flex-
ible MOFs in terms of usable capacity for gas storage and high-
light the advantages of flexible MOFs through specific examples.
Many gases have been used to study the adsorption behavior of
flexible porous materials, but most are not considered for stor-
age applications. This is partly because of well-developed stor-
age technologies, e.g., compressed Ar, O2,N
2. The gas storage
in porous solids is mainly motivated by the goal of increasing
the amount of gas within a given volume, e.g., for applications
in the mobility sector. Consequently, only a limited number of
gases, notably H2and CH4, are typically considered for practical
storage purposes.
An additional motivation is the safe storage of hazardous
gases, such as acetylene.
CH4Storage in Flexible MOFs: As the primary constituent of
natural gas, CH4offers a high calorific value, low carbon dioxide
emissions, and a high research octane number, making it an at-
tractive fuel for vehicular applications. The Advanced Research
Projects Agency-Energy (ARPA-E) under the U.S. Department of
Energy (DOE) has established several ambitious targets for on-
board CH4storage systems.[21,28]The gravimetric storage target
is set at 0.5 g(CH4) per 1 g of adsorbent. The volumetric target (de-
liverable capacity) is 263 cm3(at standard temperature and pres-
sure, STP) per mL of adsorption chamber at 298 K and 65 bar, cor-
responding to compressed methane gas at 250 bar.[28,29]Addition-
ally, accounting for up to 50% volume loss due to the low packing
densities of MOFs,[21]the volumetric storage target for the ma-
terial itself should be even higher. Over the past three decades,
MOFs have been extensively investigated and documented for
their substantial potential in CH4storage applications.[21,30]Com-
putational calculations demonstrated, however, that the usable
(deliverable) capacity of almost all conventional rigid adsorbents
would not be higher than 200 cm3methane per cm3of adsor-
bent (defined as STP volume at 298 K and 65–5.8 bar pressure
window),[21,31]which is very low compared to the DOI target. One
of the reasons is that most adsorbents exhibit Type I isotherms,
where a substantial amount of methane remains adsorbed in the
microporous frameworks at pressures around 5 - 6 bar.
Recently, flexible MOFs have gained prominence due to their
ability to markedly increase the usable storage capacity. Methane
storage in flexible MOFs has been recently summarized by For-
rest et al.,[32]therefore we will discuss only a few examples in the
following.
In 2002 and 2003, Seki and Kitagawa reported for the first
time a flexible pillared layer MOFs, [Cu2(bdc)2(4,4-bipyridine)]n
(bdc =terephthalate) and [Cu(dhbc)(4,4-bipyridine)]n(Hdhbc =
2,5-dihydroxybenzoate), showing the gate opening CH4adsorp-
tion isotherm at high pressures and 298 K (Figure 3).[33]Although
the overall methane uptake is not high, the gate opening and clos-
ing pressures are below 10 bar; thus, this reports opened up new
avenues for methane storage in flexible frameworks.[32]
In 2009, Kaneko and co-workers[34]reported methane ad-
sorption isotherm of ELM-11 ([Cu(BF4)2(4,4-bipyridine)2]n)
(Figure 37b,g), and discussed the benefits of gating isotherm for
efficient methane storage. The adsorption capacity of the MOF
at65baris55mgg
1. Taking the packing density of the mate-
rial into account, the storage capacity was calculated to be 155
cm3cm3(volume of gas/volume of storage vessel).
In 2012, Stoeck et al.[21b]synthesized a carbazole-based meso-
porous MOF (DUT-49) exhibiting a hierarchical pore system
and showing an exceptionally high specific surface area of
5476 m3g1and a large total pore volume of 2.91 cm3g1,as
well as record gravimetric total methane uptake of 0.56 g g1
(236 cm3cm3) at 298 K and 110 bar (Figure 4b). The work-
ing capacity of the MOF in the 5–65 bar range amounts to
177 cm3cm3.
Later, authors recognized that DUT-49 displays an unusual
breathing type isotherm with a negative step in the adsorption
branch, coined as “negative gas adsorption” (NGA) in methane
isotherm measured at 111 K (Figure 4c).[35]NGA is recognized
as a new, counterintuitive phenomenon based on the release of
already adsorbed gas upon contraction of metastable open pore
phase of DUT-49 to a narrow pore phase, accompanied by the
reduction of the pore volume of the structure by 1.8 cm3g1.
Krause et al.[36]discovered that by varying the length of the
linker, the number of CH4molecules adsorbed at the intersec-
tion of the open-pore and closed-pore phases increased with the
ligand length. This leads to a more favorable overall change in
adsorption enthalpy for structural contraction, thereby promot-
ing the occurrence of NGA. Then, they investigated the effects of
temperature and adsorbate on the occurrence and extent of NGA
in DUT-49.[37a]Through experiments covering a broad range
of gases and temperatures, the specific temperature ranges in
which NGA is observable for each guest molecule could be deter-
mined. These findings were further complemented by molecular
simulations to explain the absence of NGA at higher tempera-
tures and the non-monotonic behavior observed at lower temper-
atures. Although the effect disappears in the methane adsorp-
tion isotherm at room temperature, in future, it can be explored
for the separation of other gases, such as C3–C4 hydrocarbons,
which provoke NGA transitions at room temperature.
In 2015, Long and co-workers reported the [M(bdp)]n(bdp =
1,4-benzenedipyrazolate, Figure 5a)seriesofMOFs(M=Co, Fe)
displaying working capacities for CH4, very close to the high-
est working capacities among all MOF materials reported up to
now,[32]owing to their large pore volume and a high degree of
flexibility.[26]The [Co(bdp)]nshows negligible CH4uptake at low
pressures, with a noticeable step in the adsorption isotherm after
reaching 16 bar (Figure 5b). Desorption exhibits hysteresis, and
the loop closes at 7 bar, indicating a reversible structural phase
transition (Figure 5a), stable over 100 adsorption/desorption cy-
cles. The CH4usable volumetric storage capacity of [Co(bdp)]nis
197 cm3cm3for 5–65 bar pressure range at 25 °C, calculated
using the crystallographic density of the materials and geometric
Adv. Mater. 2025, 2414724 2414724 (4 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 3. a) Pillared layer structure and b) adsorption (filled circles) and desorption (open circles) isotherms of N2,CH
4,CO
2and O2at 298 K for
[Cu(dhbc)(4,4-bipyridine)]n. Reproduced with permission.[33]Copyright 2002, Wiley-VCH GmbH & Co. KGaA.
Figure 4. a) Structure and conformation of the DUT-49 framework and 9,9-([1,1-biphenyl]4,4-diyl)bis(9H-carbazole-3,6-dicarboxylate) linker in op
(left) and cp (right) structures. Guest molecules are omitted for clarity. Reproduced with permission.[35]Copyright 2016, Nature Publishing Group.
b) Total volumetric CH4adsorption isotherm of DUT-49 at 298 K (adsorption =closed symbols, desorption =open symbols). The working capacity
is given between 65 and 5 bar, and 100 and 5 bar. c) CH4adsorption and desorption isotherms in DUT-49 at 111 K. Reproduced with permission.[35 ]
Copyright 2016, Springer Nature Limited.
Adv. Mater. 2025, 2414724 2414724 (5 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 5. a) The structure of H2bdp ligand along with the crystal structures of the closed pore (left) and open pore (right) [Co(bdp)]nphases.
b,c) Total CH4adsorption isotherms for [Co(bdp)]n(b) and [Fe(bdp)]n(c) at 298 K. Filled circles represent adsorption; open circles represent desorption.
Reproduced with permission.[26]Copyright 2015, Nature Publishing Group.
pore volume of the expanded phase. High working capacity is ex-
plained by nearly complete desorption of the methane from the
pores (remind adsorbed amount less than 0.2 mmol g1at 25 °C)
below gate closing pressure.
For isoreticular [Fe(bdp)]n, opening and closing transitions
occur at 24 and 10 bar, respectively (Figure 5c). Nonetheless,
[Fe(bdp)]nmaintains a high volumetric usable capacity of 190
cm3cm3at 5–65 bar and 298 K. Additionally, [Fe(bdp)]nunder-
goes further expansion at pressures above 40 bar, forming an op
framework with nearly perfect square channels.[26]
The same research group proposed a strategy to adjust the
pressure required for structural transformation by chemical
modification of the linker, which led to five functionalized com-
pounds (Figure 6a), isostructural to [Co(bdp)]n, namely, [Co(F-
bdp)]n, [Co(p-F2-bdp)]n,[Co(o-F
2-bdp)]n,[Co(D
4-bdp)]n, and [Co(p-
Me2-bdp)]n.[38]Experimental results indicate that the functional-
ization of the ligand by introducing fluorine atoms in the phenyl
ring leads to a decrease in phase transition pressure by disrupting
edge-to-face 𝜋𝜋interactions, while methyl groups increase the
opening pressure by enhancing these interactions (Figure 6b,c).
Yang et al.[39]reported a [NiL2]n(L =4-(4-pyridyl)-biphenyl-
4-carboxylate), flexible MOF with a dia topology and sixfold in-
terpenetration, denoted as X-dia-1-Ni. The framework demon-
strates substantial flexibility, and the initially nonporous X-dia-1-
Ni underwent multiple phase transitions during CO2adsorption
at 195 K (for more crystallographic details, see Section 3.1). High-
pressure CH4adsorption studies at 298 K revealed an S-shaped
adsorption isotherm (Figure 7), with negligible CH4uptake be-
low 20 bar, followed by a sharp increase as the pores open. At 298
K, the working CH4capacity is 162 cm3g1(147 cm3cm3)at
1–65bar.
[39,40]
However, low closing pressure observed in the desorption
isotherms significantly reduces the working capacities, consid-
ering the working pressure window between 5 and 65 bar. Cobalt
doping in this system (similar to DUT-8(Ni)[41]) enables control
over the gate opening. So, the methane-induced phase transfor-
mations can be fine-tuned by using different Ni/Co ratios to en-
hance methane working capacity.[40]The hysteresis was shifted
to a higher pressure, and the amount of CH4retained at 5 bar
during desorption decreased from 60 cm3g1in X-dia-1-Ni to 19
cm3g1in X-dia-1-Ni0.89Co0.11 . Therefore, the working capacity
of X-dia-1-Ni0.89Co0.11 reached 202 cm3g1(5–65 bar) at 298 K
(Figure 7).
Kaskel and co-workers investigated high-pressure CH4adsorp-
tion at 298 K of [Zn2(BPnDC)2(bpy)]n(SNU-9, BPnDC =ben-
zophenone 4,4-dicarboxylate, bpy =4,4-bipyridine), the frame-
work reported by Park and Suh already in 2010.[42]The com-
pound demonstrates stepwise adsorption isotherm and hys-
teresis between the adsorption and desorption branches.[43]Al-
though the pore volume of this interpenetrated framework is
moderate (0.37 cm3g1), the working capacity between 100 and
5 bar at 298 K is 144 cm3cm3due to the relatively high crystal-
lographic density of the material (1.074 cm3g1in the open pore
state). In the window between 65 and 5 bar it is, however, much
lower because of the high gate opening pressure (Figure 8).
Two isostructural flexible MOFs of the MIL-53 series, namely,
MIL-53(Al)-OH and MIL-53(Al)-(OH)2, exhibit transitions from a
narrow pore state to an open-pore state during CH4adsorption
Adv. Mater. 2025, 2414724 2414724 (6 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 6. a) The H2bdp derivatives used as linker. b) CH4adsorption isotherms of [Co(bdp)]n,[Co(Fbdp)]
n, [Co(p-F2-bdp)]n, [Co(o-F2-bdp)]n,and
[Co(D4-bdp)]nat 298 K. c) N2adsorption for [Co(bdp)]nand derivatives at 77 K. Reproduced with permission.[38]Copyright 2016, American Chemical
Society.
at 298 K, and the opening pressures are 15 and 46 bar, respec-
tively, with hysteresis in the desorption isotherms leading to clo-
sure pressures of 4.1 and 5.8 bar[46](Figure 9).
The higher transition pressure of MIL-53(Al)-(OH)2is at-
tributed to the additional OH groups, which enhance the inter-
molecular interactions and impede the transition of the frame-
work (Figure 9a). At 298 K, the usable capacities for MIL-53(Al)-
OH and MIL-53(Al)-(OH)2in the 5–65 bar range are 71 and 164
cm3cm3, respectively.
The examples demonstrate the importance of the control over
phase transition pressure at desired thermodynamic conditions.
Such control can be realized by the constitutional changes of the
building blocks (metal in the cluster or substituent on the linker),
by the combination of multiple linkers in the same framework
(solid solution),[47]or by the adjustment of particle size.[48 ]
The mixed ligand strategy was demonstrated by Bolinois et al.
in MIL-53(Al), where the bdc was combined with NH2-bdc in var-
ious ratios to induce flexibility upon high-pressure methane ad-
sorption (Figure 10).[47b]
Very recently, Zhai group was able to show that by intro-
ducing a functional group (in this case, azido-group), a for-
merly rigid CAU-10 ([Al(OH)(1,3-bdc)]n, 1,3-bdc =isophthalate)
framework[49]can be transformed into a flexible one.[50 ]
So far, no MOF structure, rigid or flexible, has been able to
meet the DOE’s current targets for viable methane on-board stor-
age systems in terms of deliverable capacity. Flexible MOFs with
a potential for high deliverable capacity are a promising class of
materials that need further improvement in the pore volume and
pore size optimization.
Hydrogen Storage in Flexible MOFs: Hydrogen is widely ac-
cepted as an environmentally friendly and promising alternative
energy source. However, its extremely low critical temperature
and standard boiling point (33 and 20 K, respectively) pose sig-
nificant challenges for its storage by liquefaction or compression.
Even under high pressures and low temperatures, H2has a rela-
tively low energy density per unit volume, making storage chal-
lenging.
DOE has set the gravimetric and volumetric working capacity
targets of 5.5 wt%; 40 g L1for 2025, and 6.5 wt%; 50 g L1as an
ultimate goal.[51]
Although several MOFs have high gravimetric working ca-
pacity reaching the DOE target, simultaneously high volumetric
Adv. Mater. 2025, 2414724 2414724 (7 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 7. High-pressure CH4physisorption isotherms for X-dia-1-Ni and
X-dia-1-Ni0.89Co0.11 at 298 K. Adapted with permission.[40 ]Copyright 2023,
Wiley-VCH GmbH.
working capacity is still very challenging, and the current record
holder (NPF-200) has a working capacity of 37.2 g L1between
100 and 5 bar at 77 K.[52]
To the best of our knowledge, only a small number of MOFs
have been reported where hydrogen adsorption induces gating
due to framework flexibility at 77 K.
In 2008, Long and co-workers reported the profiles of hydrogen
high-pressure physisorption isotherms of [Co(bdp)]nbetween 50
and 87 K (Figure 11).[53]At 77 K and 40 bar, the excess adsorption
capacity amounts to 3 wt%, and the corresponding gate opening
and closing pressures are 20 and 10 bar. Unfortunately, the au-
thors reported only excess H2uptake values in their work and the
total adsorption and working capacities were not calculated.
Recently, McGuirk and co-workers[54]reported a
mixed linker approach successfully induced flexibility in
[Cd(benzimidazolate)2]n(CdIF-13) framework, which shows
negligible uptake at 77 K in the pressure range up to 100 bar
(Figure 12). The 13% of 2-methyl-5,6-difluorobenzimidazolate
(2M56DFbim) instead of benzimidazolate (b induces gate-
opening for H2. This stepped sorption profile enables a usable
H2capacity of 1.17 wt%, which is, however, far below that of
[Co(bdp)]n.
MIL-53(Al) was demonstrated to show gate opening during hy-
drogen adsorption at 77 and 87 K.[55]Under vacuum and cryo-
genic conditions, the closed pore form is thermodynamically sta-
ble for desolvated MIL-53(Al), resulting in no H2adsorption in
the low-pressure range. According to the isotherm shape, the
opening of the structure and the transition to the intermediate
pore phase occurs at the pressure of 1 bar, followed by the open-
ing transition at 3 bar at 77 K (Figure 13).
Thus, in the case of H2storage, DOI-recommended values
have not been achieved to date. However, the examples re-
ported so far demonstrate that this area of research holds sig-
nificant promise and is a highly encouraging direction for future
exploration.
Acetylene Storage: Acetylene (C2H2) is a crucial raw mate-
rial in industrial manufacturing and precision electronics. It
is typically stored in acetone, as storing it under pressures ex-
ceeding 1.5 bar can lead to polymerization and potential explo-
sion hazards.[56]However, the presence of acetone vapours poses
a challenge for applications that demand high-purity C2H2.[57]
Therefore, there is a need for a more suitable storage medium
for C2H2. Although extensive research has been conducted on
C2H2adsorption, most studies focus on atmospheric pressure,
and no materials for high-pressure storage have been reported.
For flexible adsorbents, the inflexion point of the isotherm
typically appears well below 1.0 bar. Nevertheless, even un-
der low-pressure adsorption conditions, it is possible to iden-
tify highly effective adsorbent materials. For instance, Kitagawa
and co-workers[58]utilized synchrotron X-ray powder diffraction
data to determine the structure of a [Cu2(pzdc)2(pyz)]n(pzdc =
pyrazine-2,3-dicarboxylate, pyz =pyrazine) MOF that adsorbs
C2H2molecules, discovering that C2H2adsorption can readily in-
duce significant framework distortion. This compound exhibits
a high adsorption affinity for C2H2(adsorption enthalpy 42.5 kJ
mol1), with an adsorption capacity in saturation of 42 cm3g1.
In 2009, Zhang et al. reported MAF-2 ([Cu(etz)]n,Hetz=3,5-
diethyl-1,2,4-triazole) MOF, which exhibited a sigmoid adsorp-
tion isotherm for C2H2. This facilitated the release of C2H2in
MAF-2 and increased its usable capacity. The single-crystal struc-
ture of C2H2-loaded MAF-2 revealed the formation of C2H2hex-
amers within the pores. At 298 K and pressures between 1.0 and
1.5 bar, the estimated C2H2uptake in MAF-2 reaches 17 cm3g1
(equivalent to 20 cm3cm3), which is over 40 times the usable ca-
pacity of an equal volume gas cylinder (0.5 cm3cm3)(Figure 14).
Subsequently, the same research group reported that MAF-
123-Cd[59]exhibited even higher working capacity. At 298 K,
the C2H2isotherm was nearly linear, with an uptake of only
2.23 mmol g1(54.514 cm3g1or 83.55 cm3cm3)at1bar.How-
ever, at 273 K, the isotherm displayed a clear S-shape, reaching
an uptake of 6.34 mmol g1(142.06 cm3g1or 217.85 cm3cm3)
at 1 bar, indicating that a substantial usable capacity could be
achieved through temperature swing adsorption. Based on the
extrapolated isotherm (1.0–1.5 bar), the usable storage capacity
of MAF-123-Cd was calculated to be 1.3 mmol g1(31.79 cm3g1
or 48.73 cm3cm3), which is 98 times higher than that of a stan-
dard gas cylinder (0.5 cm3cm3).
In 2019, Zeng et al.[60]investigated C2H2adsorption on flexi-
ble microporous JNU-1 ([Zn3(OH)2(btca)2]nbtca =benzotriazole-
5-carboxylate) MOF, possessing a high density of open metal
sites (Figure 15a). The isotherms display adsorption steps at
low pressures that gradually disappear with the increase in tem-
perature. Structural contraction of the pores upon the adsorp-
tion and induced-fit effect leads to a significant rise in adsorp-
tion enthalpy. The binding between C2H2and the MOF is ex-
ceptionally strong, making complete desorption possible only
under high vacuum and elevated temperatures (Figure 15b).
Therefore, in MOF design, it is crucial to balance binding
affinity with regeneration energy, which involves managing the
trade-off between thermal management and the final usable
capacity.
Moreover, flexible MOFs are highly likely to exhibit remark-
able potential for storing highly toxic gases (such as AsH3,
BF3, and PH3), which are critically important in the semi-
conductor industry. Currently, these gases are stored in cylin-
ders, requiring extremely high levels of sealing. Despite the as-
sociated risks, the necessity of using such highly toxic gases
Adv. Mater. 2025, 2414724 2414724 (8 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 8. Total methane physisorption isotherm for SNU-9 at 298 K. Reproduced with permission.[44,45]Copyright 2019, Springer Nature Singapore Pte
Ltd, Copyright 2014, American Chemical Society.
highlights the significant application value of flexible adsorbent
materials.
2.2. Mixed Gas Adsorption and Separation Concept
Separation and purification technologies account for up to 15%
of global energy consumption and nearly half of industrial energy
usage. Conventional gas separation technologies, such as distil-
lation, involve repeatedly evaporating and condensing mixtures
under harsh conditions. At the same time, liquid absorbents re-
quire heating and cooling large solvent volumes to release ab-
sorbed gases during regeneration. To reduce CO2emissions, for
example, methods like wet scrubbing of exhaust gases have been
applied in industry for many years.[61]The exhaust gases are led
through an absorber filled with amine solution, which is capable
of reacting with CO2and forming carbamates. These processes
are connected to high energy consumption, as the recovery of
absorbents is only possible through thermal desorption. Hence,
more energy-efficient and sustainable gas separation processes
are needed, and different methods and adsorbents have to be
developed.[62]
One of the most important separation problems at present
is the separation of CO2from other gases, making selective
CO2adsorption from diverse gas mixtures an important research
field.[63]CO2separation from an exhaust gas stream is one of the
most challenging questions of the carbon capture process. The
separation of these two gases in MOFs is proposed as an alter-
native to conventional materials.[63,64]Thus, it has been shown
that MOFs would be able to significantly reduce energy con-
sumption during regeneration, as the desorption of physisorbed
CO2requires less energy input as compared to recovery from
Figure 9. Comparison of gravimetric CH4sorption isotherms at different temperatures for a) MIL-53(Al)-OH and b) MIL-53(Al)-(OH)2. Reproduced
with permission.[46]Copyright 2019, American Chemical Society.
Adv. Mater. 2025, 2414724 2414724 (9 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 10. CH4excess uptakes of MIL-53(Al)–NH2, MIL-53(Al), and MIL-
53(Al) with the BDC/BDC–NH2mixed ligand at 298 K (filled symbols -
adsorption; empty symbols - desorption). Reproduced with permition.[47]
Copyright 2017, Royal Society of Chemistry.
carbamates. Comparing the so-called parasitic energy, which is
the energy penalty that is used to regenerate an adsorber or scrub-
ber after sorption, it becomes clear that MOF adsorbents pos-
sess a huge advantage in this prospect.[65]State-of-the-art amine
scrubbing technologies (with monoethanol amine; MEA) showed
a parasitic energy higher than 1000 kJ kg1, whereas rigid MOFs
such as Mg-MOF-74 or Ni-(4PyC)2show much lower values of
727 and 655 kJ kg1, respectively.[65,66]
The important indicators for efficient adsorptive gas separa-
tion include also the working capacities of each component in a
mixture, enthalpies or heats of adsorption, but also selectivities
and mass transfer (adsorption kinetic).[67]Despite extensive re-
search in the field of MOF-based gas separation/purification in
more than the last two decades, the trade-off of adsorption capac-
ity versus selectivity is still a major challenge.[68]
Figure 11. Isotherms for the excess uptake of H2on [Co(bdp)]n, showing
temperature-dependent hysteresis loops at 50, 65, 77, and 87 K. Filled and
open symbols represent adsorption and desorption curves, respectively.
Reproduced with permission.[53]Copyright 2008, American Chemical
Society.
Figure 12. H2physisorption isotherms at 77 K for CdIF-13 (red) and
[Cd(bim)1.87(2M56DFbim)0.13]n(violet), exhibiting the reduced gate open-
ing pressure threshold for the multivariate MOF whereas CdIF-13 does
not show any gate opening in this operating pressure regime. Closed cir-
cles correspond to adsorption, and open circles correspond to desorp-
tion. Reproduced with permission.[54]Copyright 2023, American Chemical
Society.
The main advantage of flexible MOFs is the selectivity of the
structural transitions against the triggering fluid. There are many
examples of MOFs and fluids where the structural transition can
be triggered by a specific fluid only at the specific conditions
(Figure 16b), or the pressure needed to provoke structural transi-
tion differs for various fluids.[10d,69]In the case of MOFs with “gate
opening” behavior, such phenomena could lead to incredibly
high (near infinite) separation selectivity if such behavior could
be translated to the case of a mixture of fluids (Figure 16a), and
some of them have been introduced as “perfect” adsorbents.[69]
It should also be pointed out that even in a mixture, in many
cases, a relatively high concentration of the “molecular opener”
(i.e., the guest stimulating the phase transition) is required, im-
plying that trace impurity separations based on flexible MOFs are
less likely achievable.
The main concern, however, for flexible MOFs is whether
the structural transition would suppress or enable the coadsorp-
tion of multiple components from the mixture (Figure 16), and
whether the separation factors predicted from the single compo-
nent isotherms can be transferred to the case of multicomponent
adsorption.
A visionary goal would be the development of switchable
MOFs in which the pores are accessible for one molecular species
only (able to provoke structural transition and enter the pores)
while the others in the mixture are kept outside (due to the un-
preferable free energy of the system containing multiple compo-
nents) (Figure 16a).
However, predicting the separation performance of flexible
MOFs based solely on single-gas physisorption isotherms is chal-
lenging, as both scenarios—exclusion of a second component or
coadsorption—are theoretically possible, depending on the free
energy profile of the MOF/adsorbate system. As a result, multi-
component adsorption measurements are the primary method
Adv. Mater. 2025, 2414724 2414724 (10 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 13. High-pressure H2physisorption isotherm of MIL-53(Al) measured at 77 K (black) and 87 K (red). Reproduced with permission.[55a]Copyright
2017, American Chemical Society.
for accurately evaluating the true separation capability of flexible
MOFs.
To evaluate the separation performance of an adsorbent, the
adsorption capacities for the fluids of interest are usually ex-
plored first. The most common method for measuring the ad-
sorption capacity of MOFs for various gases is the static adsorp-
tion method. Here, the amount of gas adsorbed is determined
in a state of equilibrium by adding a defined amount of gas to
the sample. Once equilibrium has been reached, the amount ad-
sorbed is determined either by the change in gas pressure (volu-
metric) or by the increase in sample mass (gravimetric).[70]This
method can also be used for gas mixtures, but in this case, the
composition of the gas should be controlled beside the pressure
change, and direct measurement of binary mixture equilibria is
more complicated and time-consuming.[71]
The researchers are still working to improve mathematical
models for predicting mixture adsorption. Indeed, from the ad-
sorption isotherms of individual gases, a prediction can be made
using ideal adsorbed solution theory (IAST) or Grand Canonical
Monte Carlo (GCMC) simulations.[72]However, even for rigid ad-
sorbents, the IAST cannot give information on deviations from
ideal behavior.[73]
The calculations for flexible materials were not feasible for a
long time since the contraction and expansion of the network
could not be taken into account in these calculations, and the ap-
plied models were, therefore, invalid or subjected to significant
errors.[72,74]In the past years, however, new methods, like the cou-
pling of adsorbed solution theories and the thermodynamic os-
motic ensemble, resulting in Osmotic Framework Adsorbed So-
lution Theory (OFAST)[75]or hybrid Monte Carlo and Molecular
Dynamics[76]simulations, have been developed to come up with
more accurate predictions.[76,77]The OFAST allows the prediction
of phase transition pressures upon coadsorption, but its major
drawback is that it relies on the IAST to describe adsorption in
each phase of the host material.[78]Thus, true selectivity is not
accessible by this model if the adsorption behavior deviates from
that expected by IAST.
It should also be noted that the OFAST model only deals with
the thermodynamic stability of the phases of the material at equi-
librium and yields no insight into the hystereses that are typically
observed experimentally. Therefore, the theoretical predictions
need to be verified by experimental data, leaving the most reli-
able method for determining sorption properties measuring the
mixed gas adsorption isotherms.[71]
Breakthrough experiments can be considered as essential
characterization technique for industrially relevant separation
processes. Under dynamic conditions, many new parameters
have to be considered to fully assert the suitability so that dynamic
gas sorption measurements become necessary.[70,79]A break-
through curve is a plot that reflects the concentration change of
the adsorptive in the effluent stream at the outlet of a fixed bed
adsorber, giving information not only about the adsorption capac-
ity of the adsorbent but also about the kinetics of the adsorption
processes.
Therefore, dynamic sorption experiments are one of the fre-
quently employed methods for investigating gas separation prob-
lems. The behavior of flexible, phase-changing adsorbents un-
der dynamic breakthrough conditions can undergo several phase
transitions, yielding nonconventional breakthrough profiles.[80]
The challenges for applications of flexible compounds in a bed
are, amongst others, discussed below.
There are numerous highlights in the field of separation of
diverse molecules on flexible MOFs, but due to limitations and
the need to stay focused, we will limit ourselves to a few examples
(such as some hydrocarbons, CO2/CH4and D2/H2) here.
2.2.1. Examples of Flexible MOFs for Separation
CO2/CH4: The carbon dioxide/methane separation ability of
flexible MOFs is among the characteristics that have been heavily
investigated. Significant differences in the adsorption enthalpies
of the two gases render it likely that the framework will respond
differently to each during the adsorption.
Adv. Mater. 2025, 2414724 2414724 (11 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 14. a) Semilogarithmic physisorption isotherms of C2H2(squares)
and CO2(circles) for MAF-2 at 195 K. The filled and open symbols repre-
sent adsorption and desorption, respectively. b) Volumetric C2H2uptakes
of MAF-2 and other typical materials at room temperature. Two dashed
lines represent the practical working limit of charging and discharging
pressures. The curve marked as MP is shown for a hypothetical micro-
porous adsorbent (following Langmuir isotherm) having the same uptake
as MAF-2 at 1 atm and p0. Reproduced with permission.[57]Copyright
2009, American Chemical Society.
Mixed gas CO2/CH4adsorption was intensively investigated
experimentally and supported by theoretical calculations for MIL-
53(Cr) up to 25 bar at 303 K.[81]Upon adsorption of pure CO2, the
MOF undergoes a breathing-type structural transition.
The adsorption of CH4does not provoke structural transi-
tion in MIL-53(Cr) in the investigated pressure and temperature
range, showing the isotherm typical for microporous rigid ad-
sorbent (Figure 17a). The results of mixed gas adsorption experi-
ments made clear that the composition of the employed gas mix-
ture does have a substantial effect on the flexibility and capacity
of the material. The total amount adsorbed was reduced with an
increasing content of CH4compared to the single gas adsorption
of CO2(Figure 17a). In situ Raman spectroscopic studies of the
narrow pore and open phases revealed that both gases are coad-
sorbed, showing that there is a significantly higher affinity toward
CO2, but in the presence of CH4, coadsorption is not completely
hindered.[81]
For MIL-53(Al), Coudert et al.[75a]determined for each com-
position of the CO2/CH4mixture whether breathing occurs and
transition pressures of op to np as well as np to op phase by
solving the OFAST equations numerically. Figure 17b demar-
cates the existence domains of the op (blue) and np forms (col-
orful) as well as selectivity in the (x(CO2), p) phase diagram of
the mixture adsorption. The np phase can be seen as a high-
selectivity island (with values of selectivity in the range of a few
tens), separated from the lower-selectivity background that is the
op phase. The pressure–composition phase diagram resulting
from OFAST predictions is consistent with the experimental re-
sults of Finsy et al.,[82]where two distinct selectivity mechanisms
could be identified. Up to 5 bar at 303 K and equimolar gas mix-
ture, the adsorption of CO2is dominated by strong specific inter-
actions with the np framework and average separation factors of
7. Above 6 bar, the average separation factor decreases to 4, due
to the framework opening.
Denayer and co-workers proposed a general methodology to
model the behavior of flexible MOFs under dynamic break-
through conditions.[80]Each phase was modeled as a rigid adsor-
bent using a suitable adsorption model in correspondence to the
method proposed by Coudert et al.[83]The phase diagram was
modeled by a sudden transition (SGE method) or a smoother
s-function (PDE method). The s-function varies between 0 and
1 and can be physically correlated to the (mass)fractions of the
occurring phases. The model was used to qualitatively simulate
the breakthrough curves reported by Hamon et al.[75,81]The ex-
perimental and simulated F-profiles are shown in Figure 17d.
It was shown that the op to np transformation must be explic-
itly accounted. The feed conditions trigger the MIL-53 to contract
Figure 15. a) Structure of JNU-1. b) C2H2adsorption isotherms at differ-
ent temperatures. Reproduced with permission.[60]Copyright 2019, Wiley-
VCH GmbH & Co. KGaA.
Adv. Mater. 2025, 2414724 2414724 (12 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 16. a–c) A key scientific question regarding mixed gas adsorption on flexible MOFs: How do the mixed gas adsorption isotherms for individual
components behave in the pressure range above the phase transition pressure? a) Only one component (fluid triggering the transition) is adsorbed; and
b) both components are coadsorbed, following the adsorption run of a prototypical rigid MOF with the same pore size and composition. c) Schematic
illustration of two possible scenarios. Adapted with permission.[22]used under CC BY 4.0. Copyright 2020, The authors.
from the op to the np phase in equilibrium with the feed state,
and the breakthrough profile shows an associated step. Simula-
tions could capture a double roll-up, noticed in the experimental
breakthrough profile, and reveal the fractions of both adsorbent
phases.
The selectivity for CO2and CH4mixture was investigated for
[Co(bdp)]nMOF (Figure 5a).[69]This MOF shows the gate open-
ing behavior for both gases in the single-component adsorption
experiments, but the opening transition for CH4occurs at a
much higher pressure than that for CO2(Figure 18) and suggests
that [Co(bdp)]ncould be highly selective for CO2at pressures be-
tween those corresponding phase transition pressures.
The composition of the adsorbate in the bicomponent
CO2/CH4(1:1) equilibrium adsorption experiment was moni-
tored using mass spectrometry. The hypothesis based on the
single-component isotherms could be indeed confirmed, and
[Co(bdp)]nadsorbs approximately no CH4at the examined pres-
sures and the framework is most accurately described as having
near-perfect CO2selectivity under these conditions (Figure 18a).
But, when the framework was exposed to a 6:94 molar ratio
of CO2/CH4at 58.6 bar (the region where the single compo-
nent experiments suggest the opening for both gases), the ma-
terial remained selective for CO2(adsorbing 8.5 mmol g1), but
a significant amount of CH4(2.1 mmol g1) is also coadsorbed
(Figure 18b).
DUT-8(Ni) ([Ni2(2,6-ndc)2dabco]n, 2,6-ndc =2,6-
naphthalenedicarboxylate, dabco =1,4-diazabicyclo[2.2.2]-
octane), is an interesting and widely investigated model “gate
opening” pillared layer MOF, showing selectivity in the gate
opening for a large variety of molecular species in the gas and
liquid states.[10d,84]It has been shown that the network has a
high affinity toward CO2compared to other gases, including
CH4, only showing the phase transition for CO2at 298 K up
to 60 bar.[84b]To investigate the adsorption selectivity in the
CO2and CH4mixture, the adsorption experiments were per-
formed with the starting gas composition of 75:25 v/v at 215 K.
The compositions of the adsorbed phase and gas phase were
simultaneously monitored in situ by 13C NMR spectroscopy
(Figure 19a,b).[43]To understand the influence of flexibility on
the separation performance, comparable experiments were also
conducted with the rigid version of DUT-8(Ni). The flexible
and rigid samples differ merely in the particle size. Analyzing
the observed adsorbed and gaseous species of both gases in
the resulting NMR-spectra, it was proven that the flexibility of
DUT-8(Ni) causes nearly perfect selectivity, which is not the case
for rigid MOF, showing significant uptake for both gases.
In situ 13C NMR spectroscopy was also applied to investigate
the CO2/CH4selectivity in the adsorption of a 75:25 mixture at
215 K for SNU-9.[43]It is capable of changing its pore structure
between the np and op states during the adsorption of CO2at
195 K through the formation of an intermediate (ip) phase. In
contrast, CH4does not initiate such transitions at 195 K, at least
up to a pressure of 2 bar (Figure 19c).
Interestingly, the phase transition of SNU-9 induces a signifi-
cant change in adsorbed gas composition. Stepwise pressure in-
crease upon adsorption results in a steeply increasing amount of
adsorbed CO2upon the structural np–ip–op transition. The ab-
solute amount of methane coadsorbed with CO2is relatively low
in the np state. It increases in the ip state and decreases after the
transition into the op state, thus passing through a maximum in
the ip state (Figure 19d). At higher pressures, the NMR signal
of adsorbed CH4decreased in intensity while the gaseous CH4
signal increased, indicating the desorption of CH4.Incontrastto
that, no desorption of CO2was observed, resulting in a selectivity
varying strongly with the current phase of the MOF.
Schneemann et al. investigated the coadsorption on a series
of substituted [Zn2(bdc)2(dabco)]n(D-MOF) MOFs belonging to
the pillared layer materials.[18a]D-MOF itself does not show
pronounced flexibility upon adsorption of methane or carbon
dioxide.
By exchanging bdc by DiP-bdc (2,5-diisopropoxy-1,4-
benzenedicarboxylate), however, flexibility could be unlocked.
The single-component isotherms show that the structural tran-
sition can be induced by CO2only, up to 30 bar investigated in
the experiment. In the first run of the coadsorption experiment,
the partial CO2pressure was chosen to be slightly below the np
op transition pressure and during the second experiment, the
partial CO2pressure was above the np op transition pressure.
From the first experiment (Figure 20a), it could be seen that the
adsorbed amount of CO2during the coadsorption experiments
is much higher than for the single component adsorption at
the same pressure (23.47 compared to 6.5 cm3g1), while the
CH4uptake is only slightly increased (6.5 compared to 4.6 cm3
g1), suggesting that a part of the sample has been already
Adv. Mater. 2025, 2414724 2414724 (13 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 17. a) Total adsorbed amount of CO2/CH4mixtures on MIL-53(Cr) at 303 K. Red diamonds: pure CO2, pink squares: 75/25 CO2/CH4, green cir-
cles: 50/50 CO2/CH4, blue triangles: 25/75 CO2/CH4and black crosses: pure CH4; Reproduced with permission.[81]Copyright 2009, American Chemical
Society. b) Calculated CO2/CH4selectivity upon adsorption of a mixture in MIL-53(Al), as a function of the total pressure and mixture composition. The
narrow pore phase corresponds to the central “island,” with high selectivity, while the op phase has lower selectivity. Reproduced with permission.[75 ]
Copyright 2009, American Chemical Society. c) Calculated adsorption isotherms for a CO2/CH4mixture in MIL-53(Al) at 304 K as a function of pressure
for pure CO2(top panel), pure CH4(bottom panel), and an equimolar mixture of the two (middle panel). Black dashed lines: total adsorbed quantity;
red lines: quantity of CO2; blue lines: quantity of CH4. Reproduced with permission.[75]Copyright 2009, American Chemical Society. d) Experimental
breakthrough curve of binary CO2/CH4mixtures on MIL-53(Cr) at 303 K at 1.0 MPa (left). Reproduced with permission.[81]Copyright 2009, American
Chemical Society; and breakthrough profiles simulated by partial differential equations (PDEs). Reproduced with permission.[80]Coryright 1016, Royal
SocietyofChemistry.
Adv. Mater. 2025, 2414724 2414724 (14 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 18. a) Multicomponent adsorption experiments for CO2/CH4mix-
tures in [Co(bdp)]nshow near-perfect CO2selectivity at 6.7, 13.9, and
25.3 bar, under equilibrium CO2/CH4molar ratios of 46:54, 42:58, and
43:57, respectively. b) Multicomponent adsorption experiment performed
under a CH4-rich atmosphere (with an equilibrium CO2/CH4molar ratio
of 6:94) shows that [Co(bdp)]nadsorbs only a small amount of CH4at
this ratio, leading to a selectivity of 61 ±4. (a,b) Purple diamonds rep-
resent the overall amount of gas adsorbed by [Co(bdp)]n(y-axis) from a
CO2/CH4mixture at a given equilibrium pressure (x-axis). Each purple
diamond is paired with a corresponding red and blue star: red stars rep-
resent the CO2adsorbed from the mixture (y-axis) at the equilibrium par-
tial pressure of CO2(x-axis), and blue stars represent the CH4adsorbed
from the mixture (y-axis) at the equilibrium partial pressure of CH4(x-axis).
Single-component isotherms of CO2(red circles) and CH4(blue circles)
are shown for reference. Reproduced with permission.[69]Copyright 2018,
American Chemical Society.
transformed to op state upon experiment. In the second exper-
iment (Figure 20b), it can be clearly seen that the CO2uptake
coincides with the CO2uptake from the single-component
experiment and the CH4is coadsorbed.[85]
The adsorption of CO2/CH4mixture on JUK-8
(([Zn(oba)(pip)]n,oba=4,4-oxybis(benzenedicarboxylate),
pip =4-pyridyl-functionalized benzene-1,3-dicarbohydrazide)
was intensively investigated by Roztocki et al. by a combination
of different techniques.[86]JUK-8 possesses a gating type of
flexibility, where the structural transition from cp to op phase
is induced by CO2(Figure 21a). In situ 13C NMR studies at
195 K point on the coadsorption of both gases in the open pore
phase at 5 bar for 1:1 CO2/CH4ratio. At temperatures above
280 K, however, the near-perfect selectivity is achieved, and
only CO2seems to be adsorbed from a CO2/CH4(75:25 v/v)
mixture.
Recently, CO2/CH4[22]separation was extensively investigated
for gate opening ELM-11 by Hiraide et al. for application in pres-
sure vacuum swing adsorption systems. The CO2adsorption
isotherm at 298 K shows two steps. The first step (cp to np transi-
tion) occurs at the relative pressure of 103(0.75 bar). The second
step (np to op transition) was observed at p/p00.3 (19.2 bar)
(Figure 22c). Methane adsorption isotherm of ELM-11 at 303 K
indicates the opening transition at 40 bar and closing transition
upon the desorption is finished at 20 bar).[34]Thus, the open-
ing of the cp phase to the np phase can be provoked by carbon
dioxide but not by methane, and the operation window between
1 and 19 bar should guarantee high selectivity (Figure 22). At
pressures above 20 bar, the compound is expected to adsorb both
components.
To tune the separation performance of flexible MOFs, the
building blocks of the MOF can be adjusted, such as metal in
the cluster or substituent on the linker, as discussed above, but
also the temperature can be adapted since the temperature be-
longs to important thermodynamic parameters and strongly in-
fluences the free energy landscape of the given system.[22,37]In
2020, Dong et al. demonstrated that the separation performance
can be optimized through tuning gate-opening pressure by tem-
perature (Figure 23) on the example of ternary mixture separation
in flexible NTU-65 ([Cu(L1)2SiF6)]n,L1 =1,4-di(1H-imidazol-1-
yl)benzene).[88]
The authors showed that by varying the adsorption tempera-
ture, the uptake was strongly influenced, allowing for high up-
takes of CO2and C2H2but decreasing the amount adsorbed
of C2H4when optimized. Using the obtained information from
static single gas adsorption, the separation of the ternary mixture
(C2H4/C2H2/CO290/1/9 v/v/v) was possible at the optimized
temperature of 263 K yielding high purity >99.99 % for C2H4
after only one cycle.
Light Hydrocarbon Separation (C2–C3): Light hydrocarbons
are not only important energy resource but also important raw
materials for fine chemicals. The similarity of boiling points,
kinetic diameters, dipole moments, and other physical prop-
erties makes the separation and purification of light hydrocar-
bons challenging, making the applications of flexible MOFs very
attractive.[89]
The adsorption of methane, ethane, propane, ethylene, and
propylene was studied using ZIF-7 ([Zn(bim)2]n, bim =benzimi-
dazolate) at 298 K.[90]Except for methane, all the gases show gat-
ing isotherms, with distinct differences between the gate opening
and closing pressures, leaving a window of selective uptake oper-
ation. It could be shown in breakthrough experiments that ethane
is selectively adsorbed over ethylene in their mixtures, which re-
sults in the direct production of pure ethylene.
It makes ZIF-7 a perfect candidate for the separation of olefins
from paraffins since, in contrast to most microporous materials,
the paraffin is selectively adsorbed.
Adv. Mater. 2025, 2414724 2414724 (15 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 19. CO2and CH4single-component physisorption isotherms for a) DUT-8(Ni) and b) SNU-9 measured volumetrically at 195 K. Signal intensities
for adsorbed 13CO2and 13 CH4measured by in situ 13C NMR spectroscopy for c) DUT-8(Ni) and d) SNU-9 pressurized with a 13 CO2/13CH4mixture
(75:25) at 215 K. A full adsorption (filled symbols and desorption (empty symbols) cycle up to 7.6 bar is shown. No signal of adsorbed CH4is detectable
for DUT-8(Ni). Reproduced with permission.[43]Copyright 2019, American Chemical Society.
Figure 20. Depiction of the excess single component and coadsorption measurements for [Zn2(DiP-bdc)2(dabco)]n:CO
2(red), CH4(black), total
(green). Reproduced with permission.[18a]Copyright 2016, Royal Society of Chemistry.
Adv. Mater. 2025, 2414724 2414724 (16 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 21. a) CO2physisorption isotherm of JUK-8 at 195 K and the crystal structures of corresponding cp and op phases. b) 13C NMR spectrum of
the 13CO2/13 CH4gas mixture at 5.5 bar and 195 K. The signal at 10 ppm (grey bar) corresponds to gaseous methane, at 4 ppm to the adsorbed
methane, at 127 ppm to the gaseous carbon dioxide (orange bar), and the broad signal between 90 and 180 ppm to the adsorbed carbon dioxide.
c) Single component methane and carbon dioxide physisorption isotherms at 298 K. d) Isothermal multicomponent adsorption experiments for CO2/CH4
(75:25 v/v) mixtures at 293 K. Reproduced with permission.[86]Copyright 2021, The Authors. Published by American Chemical Society. This publication
is licensed under CC-BY 4.0.
The potential of [Cu(dhbc)2(4,4-bipy)]nfor C1–C3 hydrocar-
bon hydrocarbon separation was tested at 298 K.[91]The single-
component isotherms of each C2–C3 hydrocarbon display gating
and differ in gate opening pressures, which inversely correlate
well with the latent heat of vaporization of the corresponding hy-
drocarbon. C3H4has the lowest gate opening pressure, which in-
dicates that the flexible framework can highly selectively adsorb
C3H4at a very low pressure. Several breakthrough experiments
were performed using equimolar C3H4/C3H6/C3H8mixtures at
298 K. C3H6and C3H8break first, while C3H4breaks through af-
ter some period of time. C3H4can be effectively separated from
the C3H4/C3H6/C3H8mixtures in a nearly pure form with gas
phase concentrations of more than 99.9%.
[Co(vttf)]n(vttf =2,2-[1,2-bis(4-benzoate)1,2-
ethanediylidene]bis-1,3-benzodithiole)) reported by Kitagawa
and co-workers, exhibits exclusive gate opening for ethylene,
potentially enabling the discriminatory adsorption of it over
ethane.[92]In the close pore phase, the compound is nonporous
and features crosslinking via the coordination of tetrathiafulva-
lene sulfur atoms with the axial sites of the paddle wheels. The
framework is not responsive to ethane, but the ethylene is able
to coordinate cobalt and, therefore, induces a phase transition,
displacing the tetrathiafulvalene linkers and yielding an open
structure. Once open, however, the framework adsorbs both
ethylene and ethane, resulting in only modest selectivities.
The adsorption ability of triply interpenetrated, flexible SD-
65 ([Zn(NO2ip)(dpe)]n,NO
2ip =5-nitroisophthalate, dpe =1,2-
di(4-pyridyl)ethylene) was studied for seven C4hydrocarbons, in-
cluding 1,3-butadiene, trans-2-butene, cis-2-butene, n-butane, 1-
butene, isobutene, and isobutane at 298 K.[93]Theuptakeofall
butenes and butanes was negligible at 1 bar, but the 1,3-butadiene
shows a gate-opening sorption profile with a gate-opening pres-
sure of 60 kPa, gate-closing pressure of 50 kPa and 40 cm3g1
uptake at 101 kPa. Therefore, SD-65 can separate 1,3-butadiene
from C4 hydrocarbon mixtures and readily release adsorbed 1,3-
butadiene under the PSA conditions.
H2/D2: Deuterium, constituting 0.016% of total hydrogen
occurring in nature, is a scientifically and industrially relevant
molecule, but its separation from H2is challenging due to its
similar physical properties.
Adv. Mater. 2025, 2414724 2414724 (17 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 22. a) Single-component adsorption isotherms of CO2and CH4at 298 K on ELM-11 (filled symbols: experimental adsorption data, open symbols:
experimental desorption data, lines: adsorption data simulated by GCMC). b) Total and component adsorption isotherms of an equimolar mixture of CO2
and CH4at 298 K for ELM-11 simulated by GCMC. The abscissae correspond to the partial pressures of CO2.Reproduced with permition.[22]Copyright
2020, The Authors, Published by Springer Nature Limited under CC BY 4.0. c) CO2sorption isotherms of ELM-11 measured at different temperatures.
Green, 195 K; light blue, 253 K (p0=1.96 MPa); blue, 273 K (p0=3.47 MPa); red, 298 K (p0=6.4 MPa). Open and filled symbols indicate sorption
and desorption branches, respectively, in each successive run. Reproduced with permission.[87]Copyright 2016, American Chemical Society. d) CH4
adsorption/desorption isotherm of ELM-11 at 303 K. Adapted with permission.[34]Copyright 2009, Elsevier Inc.
Physisorption of H2and D2at different temperatures has been
studied in detail for MIL-53(Al) (Figure 24).[55,94]The experiments
show that the selectivity for D2over H2is strongly related to the
state of the pore structure of MIL-53(Al). At temperatures below
150 K, the solvent-free MIL-53(Al) forms the closed pore phase
(unit cell volume V=857 Å3), inaccessible for adsorptives.[95]A
two-step transition is observed upon adsorption of D2at the boil-
ing point temperature, where the structure transforms from the
cp to the intermediate pore phase (ip2, V=1349 Å3) and fur-
ther to the open pore phase (V=1531 Å3)(Figure24b). Upon
hydrogen adsorption, however, the framework transforms only
to the ip2 phase at the corresponding boiling point temperature
(Figure 24a). During desorption, additional intermediate phases
are formed, which is evident from the multiple steps in the des-
orption branch. In general, 130 different structures of the MIL-53
family are deposited in the Cambridge Structural Database,[96]
showing the complexity of the free energy landscape of this
compound.
The selectivity of H2/D2adsorption was studied by thermal
desorption spectroscopy (TDS), at which the MOF was sub-
jected to the 10 mbar of 1:1 mixture at 25 K. The experiment
shows no outstanding selectivity because both gases transform
the MOF to the thermodynamically stable ip2 phase at these
conditions.
Nevertheless, optimizing the conditions can lead to an in-
creased selectivity of up to 10.5 at 10 mbar and 40 K. At these
conditions, the breathing of the network is in progress, leading to
the optimized environment for quantum sieving (Figure 24d–f).
The ability to separate hydrogen isotopes was also investigated
for the gate-opening DUT-8(Ni) (Figure 25).[97]Low-temperature
adsorption of H2and D2gas, each at their standard boiling
points, reveals pressure-dependent responsivity toward D2.The
adsorbed amount of H2is barely above the detection limit with
a slight increase up to 1 mmol g1at1bar.Thus,DUT-8(Ni)re-
mains in the cp phase under the H2atmosphere. In stark con-
trast, after little to no uptake at low pressure, there is steep ad-
sorption of D2at 0.24 bar, the gate-opening pressure resulting
in a huge saturation uptake of 41.2 mmol g1at 1 bar. The des-
orption branch indicates a hysteresis, as expected for a first-order
structural transition.
TDS after exposure of the sample to H2/D2(1:1) isotope mix-
ture shows that the D2uptake is approximately ten times higher
than for H2, and the average selectivity is 9.9. The D2uptake in-
creases with increasing total pressure of the mixture to 0.8 bar
Adv. Mater. 2025, 2414724 2414724 (18 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 23. Illustration of the temperature optimization for ternary mixture separation by flexible MOFs. At low temperatures, all three gases can be
adsorbed, which leads to the coadsorption of target gas and low productivity. At optimized temperature, two impurities are selectively adsorbed, and
pure target gas is obtained in high yield. At high temperatures, only one type of impurity is adsorbed, therefore, the product is not pure. Reproduced
with permission.[88]Copyright 2020, Wiley-VCH GmbH.
and shows the highest value of 9.44 mmol g1together with the
best selectivity of 11.6. Nevertheless, the experiments show that
both gases are coadsorbed at the chosen conditions.
2.3. How to Maximize Selectivity in Flexible MOFs
Analyzing the experimental examples of various mixed adsorp-
tion experiments, we propose a hypothesis, answering the ques-
tion posed in Figure 16: whether the structural flexibility would
suppress or enable the coadsorption of multiple components
from the mixture, what is the underlying reason, and what are
the prerequisites for highest selectivity? Analyzing the cases re-
ported, it can be seen that both scenarios are possible: near-
perfect selectivity as well as coadsorption of both components.
Obviously, the reason is hidden in the isotherms runs for
the individual components, or more precisely, in the over-
lay of two isotherms and thermodynamic stability of the
framework/gas_𝛼/gas_𝛽system at experimental condition cho-
sen for coadsorption/separation experiments.
It is important to consider the hysteretic behavior of the
isotherms when designing the separation processes. The hystere-
sis occurs due to the activation barriers separating the phases
with energetic minima and energy penalty associated with the
nucleation and increasing interfacial area upon the phase trans-
formation.
The desorption branch of the gating isotherms is closer
to thermodynamic equilibrium, while the adsorption branch
reflects the system accessing metastable states. Thus, the
kinetic barriers determine the opening pressure of the
framework, while the closing pressure determines the
thermodynamic stability of the framework filled with the
adsorbate.
In the case of breathing, the contraction of the op phase to
the np phase is characterized by a much lower barrier than the
opening (cp to op, or np to op transitions upon the desorption).
Thus, in this case, the adsorption branch upon the first breath-
ing transition is closer to thermodynamic equilibrium, while the
desorption branch reflects the system accessing metastable state.
Thus, the kinetic barriers determine the expansion pressure
of the framework, while the contraction pressure determines
the thermodynamic stability of the framework filled with the
adsorbate.
Below some particular cases are considered.
Adv. Mater. 2025, 2414724 2414724 (19 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 24. Hydrogen (red) and deuterium (blue) physisorption isotherms of MIL-53(Al): a) H2at 20 and 25 K, b) D2at 23 and 25 K, c,d) H2at 40 K, e) D2
at 40 K. (c,f) TDS of D2/H2(1:1) mixed gas from MIL-53(Al) exposed at 10 mbar. Reproduced with permission.[55]Copyright 2017, American Chemical
Society; Reproduced with permission.[94]Copyright 2020, American Chemical Society.
2.3.1. Case I: Gate Pressure MOF
The gate opening pressure (pgo) for adsorptive 𝛼is lower than the
gate closing pressure (pgc) for adsorptive 𝛽(Figure 26a). In this
case, three potential pressure ranges exist.
1) Below p(𝛼)go , where the adsorbed amount is neglectable due
to the thermodynamic (p<p(𝛼)gc) or kinetic (p(𝛼)gc <p<
p(𝛼)go stability of the cp phase.
2) Between p(𝛼)gc and p(𝛽)gc is the region of nearly perfect se-
lectivity, where the free energy of the open phase is favor-
able only if the pores contain the adsorptive 𝛼. The partial
pressure of component 𝛼in the experiment should, however,
be larger than p(𝛼)go to initiate the structural transition. (For
example, see the behavior of DUT-8(Ni)[43](Figure 19a,b)or
[Co(bdp)]n[69](Figure 18a)inCO
2/CH4adsorption.)
3) Above p(𝛽)gc, the selectivity is expected to drop to the value
characteristic for the virtually rigid framework with the same
Figure 25. a) H2and D2isotherms for DUT-8(Ni) at 20.3 and 23.3 K. b) TDS spectra for DUT-8(Ni) and D2/H2selectivities (SD2/H2) according to the
exposed pressure. Measurements were performed under a H2/D2(1:1) isotope mixture at 23.3 K. Reproduced with permission.[97]Copyright 2022, The
Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under CC BY-NC 4.0.
Adv. Mater. 2025, 2414724 2414724 (20 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 26. Hypothetical single component isotherms for a,b) hypothetical gate-opening or c) breathing MOF (op to np transition) (upper row) and
the expected selectivity in the case of an equimolar mixture of the components (middle row). The adsorption is shown as a solid, colored line, and the
desorption is shown as a dashed line of the corresponding color. The grey dashed lines represent the isotherms of the hypothetical “rigid” MOFs with
comparable porosity. The star marks one of the optimal operation pressures. The bottom row shows the thermodynamically stable state of the MOF in
the presence of each gas in the corresponding chemical potential range.
pore size and chemistry. (For example, see the behavior of
[Co(bdp)]n[69](Figure 18b)inCO
2/CH4adsorption at low CO2
concentration.)
2.3.2. Case II: Gate Pressure MOF
The gate closing pressure (pgc) for adsorptive 𝛽is lower than the
gate opening pressure (pgo) for adsorptive 𝛼(Figure 26b).
The pressure regions defining the adsorption selectivity are the
same as in Case I, but the window of the highest selectivity is
much narrower. Moreover, the pressure needed for the frame-
work opening (exemplarily marked by the star in Figure 26b)is
necessarily outside the selectivity window, making the separation
not feasible. Thus, the adsorption selectivity is expected to be sim-
ilar to that of a hypothetical rigid MOF, because the open pore
phase at pressures above p(𝛽)gc is thermodynamically stable in
the presence of both individual components. Case II is expected
for high activation barriers (broad hysteresis) as well as for guests
with similar adsorption enthalpies.
It should be mentioned that if the physisorption isotherms of
the individual components are measured in the limited pressure
range only (below p(𝛽)go), the picture can pretend Case I, leading
to a wrong assessment of the expected selectivity.
2.3.3. Case III: Breathing MOF
The breathing transition usually involves two transformations:
the contraction of the more porous phase in the presence of
a certain amount of adsorbate and the expansion (gate open-
ing) to the more porous phase at higher pressures upon adsorp-
tion. Upon desorption, the framework contracts to the np phase
first and, at lower pressure, expands to the initial, guest-free
framework.
For simplicity, we only consider the case where the pressure
of the first transition (contraction) at characteristic contraction
pressure of 𝛼component (p(𝛼)c), is much lower than the contrac-
tion pressure for the 𝛽component (p(𝛽)c) and the pressures of
the gate opening (p(𝛼)go and p(𝛽)go and closing (p(𝛼)gc and p(𝛽)gc)
in the second transitions (gating).
The activation barrier of the contraction transition is usually
very low. The expansion upon the desorption, however, is kinet-
ically hindered, resulting in the hysteresis between adsorption
and desorption branches and pex <pc.
If the contraction pressure for adsorptive 𝛼(p(𝛼)c)islower
than the expansion pressure for adsorptive 𝛽(p(𝛽)ex), then
(Figure 26c):
1) At pressures below (p(𝛼)ex) the selectivity is approaching that
of the op phase. (For example, see CO2/CH4separation on
MIL-53(Al)[75]at low pressure, Figure 17.)
2) In the pressure range between p(𝛼)ex and p(𝛽)ex , where the
narrow pore phase is thermodynamically stable only in the
presence of adsorbate 𝛼, the preferable adsorption of compo-
nent 𝛼is expected. (For example, see CO2/CH4separation on
MIL-53(Al)[75a,81]at moderate pressure, Figure 17.)
3) At pressures above (p(𝛽)ex) and below p(𝛼)gc, np phase can be
stabilized by each component, (𝛼or 𝛽), therefore, the selectiv-
ity is approaching that of the virtual, rigid np phase.
Adv. Mater. 2025, 2414724 2414724 (21 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
If we would take into consideration the second gating phase
transition of the breathing (the np to op transition), situations
similar to that described in Cases I and II are feasible, as well as
more complicated cases, depending on the values of pc(𝛼), pc(𝛽),
pex(𝛼), pex (𝛽),p
go(𝛼), pgo (𝛽),p
gc(𝛼), and pgc (𝛽).
But considering the fact, that in the breathing system, the
closed pore phase (unporous state) does not exist, the maximum
selectivity of breathing MOF is expected always to be lower than
that of gate-opening MOF.
Thus, to achieve optimal separation performance, the transi-
tion pressures and hysteresis widths of the individual adsorp-
tion/desorption isotherms must be considered when specifying
the separation conditions. The use of the adsorption branch only
(particularly if the gate opening conditions for the second compo-
nent are not fulfilled) will lead to an overestimation of the sepa-
ration performance. Breathing MOFs are expected to have lower
selectivity in comparison to gate-opening MOFs due to the in-
termediate narrow pore phase and the absence of the highly se-
lective closed pore phase.[37b]The highest selectivity for the par-
ticular MOF can be achieved if the adsorption–desorption hys-
tereses of single components are clearly separated on the pres-
sure/concentration axis.
Hence, in order to predict selectivity, it is essential to report
isotherm adsorption and desorption data, ideally in a digital for-
mat based on IUPAC recommendations.[98]Such an approach
will give a wider community the opportunity to screen and eval-
uate potential applications if flexible MOFs in a variety of fields
with AI tools and machine learning and validate the hypothesis
outlined here. Moreover, it would reduce ambiguity in adsorption
data using IUPACs established and well-defined units for the ad-
sorbed amount (mmol g1) and improve reproducibility.
3. Advanced Characterization Techniques
3.1. Monitoring the Framework Transitions
In this section, we emphasize the importance of a mechanis-
tic understanding of the guest-induced transitions of the flexi-
ble host structure at the atomic level, which is important to un-
derstand the thermodynamics of the transition and follow the
changes in porosity and pore accessibility.
In the case of “gate-opening” MOFs, the crystal structures of
the guest-free and guest-filled phases can be determined from
ex situ measurements of solvated and degassed structures. How-
ever, such results should be interpreted with caution, as guest
molecules may influence the opening degree and do not con-
tain information about the transition pathway.[99]It is, therefore,
recommended to carry out in situ X-ray diffraction measure-
ments under conditions close to those used in the gas separa-
tion studies. In the case of minor volume changes (e.g., subnet-
work displacement or linker rotation[9]), in situ experiments can
be performed on the single crystals. Single crystal X-ray diffrac-
tion (SCD) allows to follow the changes in the framework and, in
many cases, to determine the position and occupancy of the guest
molecules in the pores. Such experiments can be carried out us-
ing laboratory single-crystal X-ray diffractometers equipped with
customized cells, adapted to specific gas loading conditions.[100]
In all other cases, in situ PXRD can be used.
In some cases, unexpected mechanisms behind the selectivity
can be revealed, for example, self-accelerated and selective CO
sorption in [Cu(aip)(H2O)]n(aip 5-azidoisophthalate) reported
by Kitagawa and co-workers.[101]In situ PXRD studied in parallel
to N2and CO adsorption at 120 K (Figure 27) demonstrates the
mechanism of stepwise CO sorption, involving the formation of
aCu
2+CO bond, and inducing a global structural transforma-
tion with the expansion of the squeezed paths. This expansion
promotes additional CO adsorption in the center of the channel,
which is the so-called self-accelerating gas adsorption.
Barbour and co-workers developed and optimized a cell for
in situ single crystal X-ray diffraction under static gas loading
conditions.[100]SCD was used to locate the preferable adsorption
sites in [NiNbOF5(pyrazine)2]n(NbOFFIVE-1-Ni) one of the
state-of-the-art frameworks for direct CO2capture from air.[102]
Interestingly, the CO2molecule occupies an energetically favor-
able position, where the electropositive carbon of the CO2is sur-
rounded by four electronegative fluorine centers from four dis-
tinct (NbOF5)2pillars and the electronegative oxygen atoms of
the CO2are encaged by pyrazine hydrogens.
The cascade of the phase transitions was observed during
the desolvation of the adamantoid framework with a compo-
sition [NiL2]n(L =4-(4-pyridyl)-biphenyl-4-carboxylate).[39]All
structures were solved and refined from the single crystal X-
ray diffraction data collected on the single crystal with CH2Cl2
in the pores at different stages of degassing. As a result, three
different structures, a1,a2,anda3, all crystallizing in the same
I41cd space group, were visualized. Desolvation of the structure
leads to reductions in cell volume (9528, 8637, and 7441 Å3for
a1–a3, respectively) and solvent-accessible void volume (49%,
43%, and 33% for a1a3, respectively) induced by the changes
in N–Ni–N/C–Ni–C bond angles: 93.0°/101.5°, 90.7°/101.4°,and
88.0°/102.4°for a1a3, respectively (Figure 28). The heating of
a1–a3 phases in vacuum yields the completely desolvated and
dense X-dia-1-Ni-c1 structure, showing only 2% of the solvent-
accessible void. In the same work, the authors followed the
CO2physisorption by PXRD, indicating the reversibility of the
transitions. Moreover, multistep isotherm was observed in high-
pressure methane physisorption, which served as a valid precon-
dition for the gas separation studies.
In the follow-up manuscripts, authors developed the strategy
for controlling “gate opening pressure by utilizing a multivariate
approach and mixing Ni and Co in the structure.[40]Here, in situ
PXRD were collected upon high-pressure methane physisorp-
tion, indicating the phase transition of the framework. Finally, the
mixed-metal MOF with a composition X-dia-1-Ni0.89Co0.11 was
tested in C2H6/C2H4separation and indicated high inverse se-
lectivity toward C2H6with 9.1 times higher uptake compared to
C2H4. In situ PXRD experiments were conducted on X-dia-1-Ni
at 263 K, indicating the phase transition at 40 kPa upon C2H6ad-
sorption, whereas no changes were observed upon the physisorp-
tion of C2H4in the same conditions (Figure 29).
As discussed above, [Co(bdp)]ncompound is a promising can-
didate in terms of gas storage and separation. The structural
flexibility was investigated by in situ PXRD in parallel to high-
pressure physisorption of pure CO2and CO2/CH4mixture at 298
K(Figure 30).
Tanaka and co-workers studied the rearrangement of ELM-11
framework upon CO2adsorption by time-resolved in situ PXRD
Adv. Mater. 2025, 2414724 2414724 (22 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 27. In situ PXRD/N2and CO sorption measurements on [Cu(aip)]n.a)N
2sorption isotherms at 120 K (insert: crystal structure of NbOFFIVE-
1-Ni). b) PXRD patterns measured at a–d points shown in the N2sorption isotherms in Figure 27a. The calculated pattern for the dried [Cu(aip)]nis
shown at the bottom. c) CO sorption isotherms at 120 K. Adsorption and desorption branches are shown in solid and open circles, respectively. d) PXRD
patterns measured at a–j points of the CO sorption isotherm shown in Figure 27c. The c patterns for the dried [Cu(aip)]nand CO-adsorbed phases are
shown at the bottom and top, respectively. Reproduced with permission.[101]Copyright 2014, American Association for the Advancement of Science.
at different temperatures and threshold pressures to evaluate
the switching kinetics in the powdered sample (Figure 31).[22]
The authors conducted experiments at different incrementally
increasing CO2pressures all above pgo, showing dependence be-
tween the rate constant and the ppgo pressure difference, where
pis the CO2pressure introduced.
To study hydrogen isotope separation, neutron powder diffrac-
tion became a useful technique because of the high elastic scat-
tering length of deuterium. In such a case, both the information
about the host and guests can be extracted. Oh and co-authors
demonstrated that the physisorption of H2and D2in MIL-53(Al)
at cryogenic temperatures proceeds in different ways.[94]While
hydrogen shows a one-step isotherm, two steps are observed
in deuterium isotherm, measured at the boiling point. Neu-
tron powder diffraction (NPD), conducted at defined gas load-
ings, indicates contraction in the case of hydrogen. The same ex-
periment with deuterium indicates the contraction and reopen-
ing of the structure and shines a light on the mechanism of
selective isotope adsorption. Due to nuclear quantum effects,
deuterium shows higher enthalpy of adsorption, which is suf-
ficient to open the structure and keep it in an open pore state.
Besides NPD experiments, proving the structural changes, au-
thors conducted in situ quasi elastic neutron scattering experi-
ments, which showed that the mobility of the hydrogen in this
clamped mode is less than that when the molecule is in the open
structure.
In summary, the in situ scattering techniques provide unique
structural information on the dynamics of the host framework.
The obtained structures can be used not only for the evaluation
of geometrical porosity but also for more complex calculations,
such as GCMC, osmotic energy, and molecular dynamics simu-
lations. In some particular cases, information about the frame-
work dynamics and guests can be gained. However, much more
information on the guest’s behavior upon adsorption can be
Adv. Mater. 2025, 2414724 2414724 (23 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 28. Crystal structures of NiL2. a) Structure of the ligand. b) Adamantanoid cage in NiL2. c) Sixfold interpenetrated dia nets in NiL2. d) Rectangular
channels viewed along the c-axis. e) Single crystal structures of the porous (a1a3) and nonporous (c1) phases of X-dia-1-Ni. f) Crystal morphology and
PXRD pattern of a1. g) Crystal morphology and synchrotron PXRD (𝜆=0.8262 Å) pattern of c1. Reproduced with permission.[39]Copyright 2018,
Wiley-VCH GmbH & Co. KGaA.
derived from spectroscopic techniques, in particular, one that can
differentiate between adsorptive and adsorbate.
3.2. Monitoring the Guest Behavior
The monitoring of the guest molecules u the adsorption process
in situ is crucial for understanding the role of the guests for the
flexibility and adsorption process.
The main drawback of volumetric coadsorption techniques,
usually used to measure mixed gas adsorption isotherms, is the
limited temperature and pressure range, which is indeed quite
close to the real-world application for CO2/N2,CO
2/CH4,and
alkane/alkene separation. However, it technically does not cover
the cryogenic temperatures and low-pressure range typical for hy-
drogen isotope separation and has certain limitations in terms of
the gas mixture composition. In order to analyze the selectivity
at cryogenic temperatures, TDS can be used, where the gas mix-
ture is adsorbed first, followed by applying an ultrahigh vacuum
to remove unadsorbed gas. The desorbed gases are then moni-
tored by mass-spectroscopy upon heating, which allows for the
estimation of the interaction strength and quantification of the
species desorbed.[94,97,104]
In situ NMR is an advanced technique that differentiates and
quantifies the fluid in the gas and adsorbed phases, assuming
that all gases contain an NMR active nucleus. NMR offers the
advantage of not only enabling quantification but also providing
insights into the mobility of molecules.
Brunner and co-workers established a high-pressure in situ
cell connected to the homemade gas mixing unit.[97]The instru-
mentation was used to evaluate the coadsorption of CO2/CH4
mixture in DUT-8(Ni), SNU-9, and JUK-8 (as discussed in Sec-
tion 2.2.1).[43,86]The same setup was used for the analysis of
krypton/xenon coadsorption on DUT-8(Ni) at 280 K. 129Xe NMR
experiments on rigid and flexible DUT-8(Ni) at 283 and 237
K using a mixture of xenon and krypton indicates a clear de-
crease in the chemical shift pointing on the partial replacement
of xenon by krypton, i.e., krypton coadsorption. Analysis of the
spectra suggests a slightly higher selectivity for xenon adsorbed
in the flexible version of DUT-8(Ni). This can be explained by the
Adv. Mater. 2025, 2414724 2414724 (24 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 29. a) C2H6physisorption isotherm at 263 K. b) Corresponding PXRD patterns collected in situ. X-dia-1-Ni undergoes a reversible transformation
from an np phase to an lp phase upon ethane dosing. c) C2H4physisorption isotherm at 263 K. d) Corresponding in situ PXRD patterns. In situ PXRD
patterns indicate that the desolvated phase adsorbs C2H4,butC
2H6is not adsorbed until phase transformation occurs. Reproduced with permission.[103]
Copyright 2024, American Chemical Society.
overlapping the hysteresis in Xe in Kr isotherms in the analyzed
pressure range; namely, once the gate opening” pressure for
xenon is reached and the framework is open, coadsorption of
krypton will not change the thermodynamics, op phase remains
energetically favorable phase.
3.3. Combination of Different Techniques
The most efficient way to study the gas separation in flexible MOF
is to apply multiple in situ techniques, as shown by Roztocki et al.
for JUK-8 in the mixed gas CO2/CH4adsorption (Figure 32).[86]
In order to explain the observed performance, in situ PXRD was
conducted at 195 K, indicating the phase transition from the JUK-
8_cp to the JUK-8_op phase. To characterize the selectivity in the
low-temperature regime, in situ 13C NMR coadsorption experi-
ment using a 13CO2/13 CH4mixture (molar ratio 1:1) at 195 K
was conducted. The analysis of the spectra indicates that only
minor amount of methane coadsorbs on JUK-8op@CO2, even
at low temperatures and high pressure. In situ IR spectroscopy
conducted under CO2loading suggests the changes in the host
structure, namely, changes in the O–C–O asymmetric stretching
region of the oba2carboxylate linkers. In addition, two signals
from the adsorbed CO2(at 2340 and 2376 cm1) were observed
in contrast to one signal observed for gaseous CO2(2345 cm1).
In summary, the application of in situ diffraction techniques
is crucial for understanding the behavior of the host structure
and guest molecules upon the physisorption of the gas mix-
tures. The results provide important information about the host
structure under defined conditions, which is essential for under-
standing the material performance and may be used as input
for in silico studies. Spectroscopic techniques, applied at desired
gas pressure/temperature conditions, can clearly differentiate be-
tween adsorbed and uadsorbed guest molecules and provide di-
rect proof of the separation performance.[55,97,105]
4. Challenges to Solve for the Application of
Flexible Adsorbents
4.1. Kinetics of Switching
When considering flexible MOFs as adsorbents, not only the
working capacity and selectivity play a crucial role. The kinetics
of switching between the phases have to be taken into account
as well. The adsorption kinetics, in this case, is not only depen-
dent on the crystal surface barriers and gas diffusion kinetics but
arises from the framework bistability and is coupled to the phys-
ical properties of the guest molecules with characteristic barriers
of fluid nucleation, diffusion, and adsorption processes.[106]Al-
though the importance of kinetics has been known for over 15
years,[107]the kinetics of the adsorption process in flexible MOFs
is still underexplored.[108]
The most frequently applied method to analyze the kinetics of
gas-induced phase transition in an ensemble of crystals (powder
Adv. Mater. 2025, 2414724 2414724 (25 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 30. Powder X-ray diffraction data for [Co(bdp)]ndosedwitha)pureCO
2and b) a 50:50 mixture of CO2/CH4over a range of pressures. In both data
sets, the abrupt appearance or disappearance of peaks indicates discrete phase changes, whereas gradually shifting peaks indicate gradual framework
expansion/contraction. Reproduced with permission.[69]Copyright 2018, American Chemical Society.
Figure 31. In situ time-resolved PXRD/CO2physisorption on ELM-11: a) fractions of the open phase and d) for the closed phase at 40.8 kPa and 241 K,
at 41.0 kPa and 264 K, and at 40.8 kPa and 273K. The numbers in (a) and (d) denote the temperature. b) Fractions of the open phase, and e) for the closed
phase at 227 K as a function of the CO2pressure. The numbers in (b) and (e) denote the CO2gas pressure in kPa. The curves after 4.15s in (a–e) were
obtained by fitting the KJMA equation. c) Relationship between the rate coefficients and the pressure difference between the CO2gas pressure, p,and
the gate-opening pressure, pgo. The error bar represents the standard deviation of the value obtained using the least-square fitting of the KJMA equation
to the experimental data (n3, n: number of experimental points used for fitting). Reproduced with permission.[22]Copyright 2020, The Authors, under
CC BY 4.0.
Adv. Mater. 2025, 2414724 2414724 (26 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 32. Evaluation of JUK-8 behavior in CO2/CH4adsorption: a) PXRD patterns collected upon CO2physisorption. b) Evolution of unit cell volume
upon CO2physisorption, and c) in situ 13C NMR in parallel to CO2/CH4(50:50) physisorption at 195 K. Reproduced with permission.[86 ]Copyright
2021, The Authors. Published by American Chemical Society. This publication is licensed under CC-BY 4.0.
or pellets) is to monitor the pressure drop in the adsorption cell.
This method is also applied in the industry to evaluate adsor-
bents, particularly for pressure swing adsorption, with high cy-
cling rates. However, this method does not analyze the intrinsic
switching but the macroscopic gas uptake of the sample, which
also involves mass transport and is therefore highly dependent
on the reactor volume and sample amount.
In such a way, Li and co-workers investigated the rate
of adsorption coupled with the structural transition in
a flexible RPM3_Zn ([Zn2(bpdc)2(bpee)]n, bpdc =4,4-
biphenyldicarboxylate; bpee =1,2-bipyridylethene).[109]The
compound exhibits a significant induction period for opening by
N2and Ar at low temperatures (Figure 33). Such a long induction
period is not observed for H2or O2at comparable pressures
and temperatures, suggesting the rate of opening is strongly
influenced by the gas–surface interaction rather than external
stress.
The induction period leads to severe mass transfer limitations
for adsorption and overestimation of the gate-opening pressure.
The authors reviewed several adsorption rate models and found
that none adequately describes the experimental rate. Statisti-
cally, the rate data are best described by a compressed exponen-
tial function. The resulting fitted parameters exceed the expec-
tations for adsorption but fall within those expected for phase
transition.[109]
By treating adsorption as a phase transition, the generalized
Avrami theory of phase transition kinetics was used to describe
adsorption in flexible hosts.
Gläser and co-workers[110]investigated the kinetics of the n-
butane and iso-butane gas mixture adsorption on gate pressure
Cu-IHMe-pw ([Cu2(H-Me-trz-Ia)2])n[111]MOF. The uptake curves
reveal complex interactions on the outer surface of MOF parti-
cles. The overall rate of adsorption-induced structural transition
depends on several factors, including degree of pressure rise,
temperature, particles size, and the subsequent diffusion rate
into newly opened pores. With the aid of a kinetic model based
on the linear driving force (LDF) approach, both rates of diffu-
sion and structural transition were studied independently of each
other. The authors claim that the overall velocity of gas uptakes
in flexible MOFs is predominantly determined by the rate of the
structural transition.
It could also be shown that the overall gas uptake is slower
for the gas mixture as compared to the single-gas uptake of n-
butane with the same partial pressure step, although much faster
than compared to the bare iso-butane adsorption (half coverage
is reached after 40 min for the mixture, 1 min for n-butane and
1000 min for iso-butane individually) (Figure 34). From the evo-
lution of the gas phase composition, it could be seen that n-
butane is predominantly adsorbed at the beginning, reducing
the total gas-phase fraction to 33%, resulting in a total sepa-
ration factor of a maximum of 10 (at time 200 min). Beyond
that, iso-butane continuously enters the opened framework, ex-
Figure 33. Nitrogen adsorption on RPM3_Zn at 77 K shows a pronounced
effect of (fictives) exposure time (dt): (A) dt=5s,(B)dt=20 s, and (D, E)
dt=360 s collected by ASAP 2020 from Micromeritics. (C) Collected using
a fresh sample and Autosorb-1 MP from Quantachrome Instruments. The
allowed time varied from 6 min to 3.7 h for adsorption. A history effect
is also observed in which the number of data points affects the uptake
(e.g., screening experiment D with 4 points versus full isotherm E with 45
points). Times (from the previous data point) for select points are Di =
67 h, Dii =8.5 h, Ei =62 h, Eii =20 h, and Eiii =14 h. The total times for
adsorption (in hours) are A B(1) <C (16) <D (80) <E (135). Prior
to the sharp rise, curves C–E follow the A and B curves. Reproduced with
permission.[109]Copyright 2011, American Chemical Society.
Adv. Mater. 2025, 2414724 2414724 (27 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 34. Kinetic of gas uptake (top), gas phase composition (bottom, left axis), and adsorption selectivity (bottom, right axis) in dependence of time
for an equimolar mixture of n-butane and iso-butane on Cu-IHMe-pw for a pressure jump of 0–40 kPa. Additionally, the overall gas phase compositions
and sorption kinetics are modeled by the “GO” model and a mass balance. Within the bottom graph, a selectivity larger than 1 indicates a preference
for n-butane, as indicated by the additional colored ribbons. It should be noted that a 50:50 molar mixture was aimed for, but actual results show slight
deviations with 48.8 to 51.2. Reproduced with permission.[110]Copyright 2024, The authors. Licensee MDPI, Basel, Switzerland. Distributed under the
terms and conditions of the CC BY 4.0.
changes the adsorbed n-butane and incorporates itself within
the framework. This leads to a subsequent increase in the gas
phase fraction of n-butane and a final separation factor of 0.9 af-
ter 10 000 min (7 days) for n-butane.
In recent years, new methods have been proposed to inves-
tigate the kinetic of phase transitions, involving in situ (single
crystal) XRD, optical, scanning electron microscopy (SEM), and
atomic force microscopy (AFM).
The phase transformation of a [Zn2(1,4-ndc)2(dabco)]n(1,4-
bdc =1,4-benzenedicarboxylate) crystal was investigated in situ
by AFM (with a time resolution of 40 s in DMF solutions with
various concentrations of the biphenyl as a guest molecule) by
the group of Kitagawa.[112]It was found that the lattice structure
of the liquid–solid interface quickly changed (within 10 min)
in response to the biphenyl concentration change. It should be
pointed out that in many cases, it is not only the analytical instru-
ment limiting the temporal resolution but also the realization of a
stepwise stimulus increase that poses challenges because of mass
transfer limitations and diffusional broadening of a pressure or
concentration pulse peak.
Phase changes were monitored on 1μm single crystals of
MIL-53(Al) by in situ environmental SEM.[113]
Adv. Mater. 2025, 2414724 2414724 (28 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 35. SEM images of MIL-53(Al) obtained at 0, 6, 12, 16, and 20 s of
beam irradiation; bottom right, particle width and length as a function of
observation time. Reproduced with permission.[113]Copyright 2014, Else-
vier Inc.
The deformation of single crystals was observed over several
seconds upon electron beam irradiation, which caused partial
desorption of the toluene from the pores. The phase transitions
were assigned through the detected variations in crystal shape
(Figure 35), which was further supported by TEM measurements.
Bon and co-workers demonstrated that the phase transition
kinetic in the MIL-53(Al) system is significantly controlled by
the particle size (Figure 36).[114]Samples with the average par-
ticle size of 1.2 and 28 μm were compared upon adsorption of
n-butane at 298 K. It was also found that the contraction transi-
tion (op–np) is the fastest transition for both particle size regimes
and the reopening of the structure proceeds much slower. Hence,
the activation barrier for opening the narrow, butane-filled pore
is larger, probably due to the diffusion limitations of the guest.
The CO2-induced gate opening of ELM-11 was investigated by
time-resolved in situ PXRD (Figures 31 and 37), and a theoreti-
cal kinetic model of this process was developed to gain atomistic
insight into the transition dynamics.[115]
The developed model consists of the differential pressure from
the gate opening (indicating the ease of structural transition)
and reaction model terms (indicating the transition propagation
within the crystal).
The reaction model of ELM-11 is an autocatalytic reaction with
two pathways for CO2penetration into the framework. How-
ever, gas adsorption analyses of MIL-53(Al), possessing different
mechanisms of phase transitions (breathing), indicate that the
kinetics of the adsorption-induced structural transition is highly
dependent on framework structure (Figure 38). The authors con-
firm the differences in the kinetics of op to np and np to op tran-
sitions, consistent with the work of Bon et al.[114]
The rate of structural transition of gating in ELM-11 investi-
gated by time-resolved in situ synchrotron Pin orderXRD mea-
surements shows that the structural transition started immedi-
ately after the introduction of CO2at 40.8 kPa and 273 K and
was accomplished in 10 s.[22]The MOF accommodating CO2
responded quickly to the decrease in gas pressure at 273 K: the
structural transition was completed in 5 s when the CO2pres-
sure was decreased at the rate of 2.4 kPa s1. Moreover, the rate
of phase transition increases as the temperature decreases under
the same CO2pressure. At 227K the rate of structural transition
increased with increasing CO2pressure, and the phase transition
was completed within a few seconds at the highest gas pressure.
Furthermore, these data were found to obey the Kolmogorov–
Johnson–Mehl–Avrami (KJMA) equation.[22 ]
Kaskel and co-workers investigated the phase transition kinet-
ics of DUT-8(Ni) using three different techniques.[106]The newly
designed microfluidic breakthrough apparatus allowed observa-
tion of transformation times for individual crystals varying with
respect to Ni/Co ratios. Comparison with the in situ PXRD data
led to the conclusion that the macroscopically observed switch-
ing rate is mainly governed by variations in the induction period.
The transformation of individual crystals is much faster than that
of an ensemble of crystals, which is, therefore, not easily observ-
able. Crystals with a higher activation barrier (corresponding to
a higher gate opening pressure) show a longer induction period,
leading to an overall slower adsorption kinetic of the ensemble.
However, the individual crystal transformation rate is below 1 s,
which was the temporal limit of resolution in this study. The ex-
act and intrinsic individual crystal transformation rate remains
uncovered, and novel methods approaching a temporal resolu-
tion in the ms or s regime are still to be developed in order to
shed light on this aspect.
4.2. Volume Change of the Crystals Associated with Phase
Transition
Although network flexibility gives rise to significant advantages
for gas storage and separation applications, it also poses many
challenges. The structural phase transitions are typically associ-
ated with macroscopic changes in crystal size.[116]This aspect has
been beneficially used for actuators and sensors.[117]
ELM11 shows two steps upon the CO2adsorption, with a
28% expansion in the interlayer distance in the first transition
and a 56% expanded layer structure compared to the initial struc-
ture in the second one.[118]The macroscopic changes in the pow-
der bed upon CO2adsorption have been reported already in
2006.[119]
Kaskel and co-workers studied volume expansion and force ex-
erted by flexible MOFs through expansion for MIL-53(Al).[117b]
The effect of the packing density on the mechanical response was
Adv. Mater. 2025, 2414724 2414724 (29 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 36. In situ time-resolved PXRD measured during the adsorption of n-butane at 298 K on MIL-53(Al): a) reopening of the crystals with average
crystal size of 28 ±18 μm, and b) breathing transition in the crystals with average crystal size of 1.2 ±0.5 μm. Reproduced with permission.[114]Copyright
2022, Royal Society of Chemistry.
also evaluated. Three different regimes were identified according
to the packing density. The pressure gained from the opening
step was found to be higher than that needed to compress the
empty framework and is specific for the stimulating fluid.
The macroscopic expansion, however, creates challenges in ad-
sorptive storage and separation applications, especially in the ad-
sorption bed. In most applications, the adsorption column is sev-
eral meters high, and thus, the adsorbent at the bottom of the bed
experiences significant gravitational force due to the high weight
of the packed bed above. This lower part of the bed has to exert a
significant force against the upper part to open its pores and ac-
complish the volume change. In extreme cases, pore expansion
may be completely suppressed in such a real macroscopic appli-
cation. Another aspect is that the adsorber column has to pro-
vide enough dead volume for expansion, which may limit volu-
metric efficiency, which is often overlooked in academic research.
Moreover, the expanding adsorbent will most likely induce a pres-
sure drop in the column.[46a]Hence, MOFs showing minimal vol-
ume changes or flexibility based on the internal movement of the
building blocks are more favorable.
4.3. Crystal Damage Associated with Phase Transition
The huge volume change associated with the mechanical
and thermal stress often leads to the damage of the crystals
(adsorption-induced milling).[120]The magnitude of the effect de-
pends on the structure and chemical composition of the MOF
itself, as well as on the particle size, adsorptive, and adsorption
temperature.
MIL-53(Al) and ELM-11 were proven to be stable in
cyclic experiments with butane at 298 K, withstanding 100
adsorption/desorption cycles without significant changes in
performance.[120]Also repeated pressure swings between 0.5 and
Adv. Mater. 2025, 2414724 2414724 (30 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 37. a) XRD pattern at 42 kPa CO2and b) crystal structure of ELM-11 in the open pore phase. c) Colormap of time resolved XRD patterns of
ELM-11 at 248 K. CO2was introduced 5 s after the start of the measurement with the constant flow rate of 0.8 kPa s1, until 42 kPa pressure was
reached. e) Time evolution of fraction transformed, calculated from the peak intensity ratio from the XRD patterns in (a) and (f) at regular intervals of
time. f) XRD pattern at 0 kPa and g) crystal structure of ELM-11 in the closed pore phase. Reproduced with permission.[115 ]Copyright 2023, The Authors,
under CC BY 4.0.
2.0 MPa at 298 and 323 K upon six CO2adsorption/desorption
cycles for MIL-53(Al) did not cause the deterioration in the ad-
sorption capacity.[121]For ELM-11, the reproducibility of the gat-
ing was demonstrated in 50 cycles of CH4physisorption at
303 K.[34]
In contrast, for DUT-8(Ni) and SNU-9, the multiple adsorp-
tion/desorption stress upon butane adsorption at room tempera-
ture leads to the reduction of crystallite size, causing changes in
the switching behavior in the initial 10 physisorption runs, and a
characteristic shift of the “gate-opening” pressure to higher val-
uesisobserved.
[120]The reason for such behavior is particle size-
dependence t flexibility.[48]
The comprehensive investigations show the existence of three
main particle size regimes in the case of DUT-8(Ni): Several mi-
crometer large particles show pronounced gating behavior, while
500 nm–1 μm sized particles reveal a suppressed opening (sta-
bilization of the metastable cp phase). Particles with sizes below
500 nm have suppressed closing ability (stabilization of the op
phase).[48]
Abylgazina et al.[48b]demonstrated that the changes in the gate
opening pressure with particle size are not linear and follow a log-
arithmic dependence on the facet extension. Moreover, the steep-
ness of the adsorption branch in the gating region decreases with
decreasing crystal size, pointing to the significantly broader acti-
vation energy distribution in the small grains compared to the
large crystallites. The threshold of the particle size, at which the
compound experiences rigidification and loss of crystallinity, de-
pends in the DUT-8(M) (M =metal) system, on the metal type in
the paddle wheel.[48a,122]The effects of crystal size and morphol-
ogy are critically discussed in a recent review.[48]
4.4. Change in the Flexibility Characteristics due to the Additives
and Shaping
When building an adsorber for gas separations, the adsorbing
material is usually pelletized to improve the gas flow, reduce pres-
sure drop, and improve overall performance. For MOF materials,
this poses a significant challenge since many of the well-known
frameworks cannot withstand the stress of shaping into pellets
and, as a consequence, important properties, like sorption ca-
pacity and selectivity, are reduced.[82,123]In flexible MOFs, loss or
massive impact on the flexibility characteristics can be expected
in addition to mechanical deterioration.[124]
Adv. Mater. 2025, 2414724 2414724 (31 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 38. Atomic structure, a schematic of the structural transition, and measurement and calculation results for a–c) the gate opening of ELM-11 at
248 K, d–f ) op np transition of MIL-53(Al) at 223 K, and g–i) nplp transition of MIL-53(Al) (at 195 K). Reproduced with permission.[115 ]Copyright
2023, The Authors, under CC BY 4.0.
MIL-53(Al), however, shows stable performance after shap-
ing while maintaining the textural properties and characteristic
framework dynamics.[125]It was demonstrated that pellets with
5 or 10 wt% of binder reached mechanical resistances compara-
ble with some of the carbon molecular sieves pellets. All exper-
imental insights revealed that the reversible breathing effect—
characteristic for MIL-53—was preserved even though the sam-
ples were shaped into cylindrical extrudates. In situ X-ray diffrac-
tion studies under humid conditions confirmed the similarity
in phase transition kinetics for extrudates and powders. More-
over, the phase transition of the MIL-53 extrudates upon ad-
sorption of carbon dioxide was confirmed by high-pressure ad-
sorption isotherms of CH4and CO2on MIL-53 powders and
extrudates, which exhibited similar capacities for both gases
(Figure 39).
Mechanical stress can potentially induce phase transition in
flexible MOFs.[126]In a recent study the group of Song shows
the influence of mechanical pressures on MIL-53(Al). High me-
chanical pressures and temperatures support the interactions be-
tween CO2, and the hydroxyl groups and the Al centers of the net-
work and increase the sorption capacity because of chemisorptive
behavior.[127]
Kundu et al. tested the pelletization of MIL-53(Al)-OH and
MIL-53(Al).[46a]Although no decrease in the uptake or changes
in the isotherms’ shape were observed, both samples could not
withstand the stress of deformation, resulting in the breaking of
Adv. Mater. 2025, 2414724 2414724 (32 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 39. High-pressure methane (red) and carbon dioxide (black) ph-
ysisorption isotherms at 303 K on MIL-53 with 3 wt% of methyl cellulose as
binder (dashed lines) in comparison to the adsorption isotherms on MIL-
53 powder (symbols). Reproduced with permission.[125]Copyright 2019,
Wiley-VCH GmbH & Co. KGaA.
the pellets during CH4adsorption. Similar results were shown by
the group of Denayer,[82]who also investigated the pelletization
of MIL-53(Al).
The application of binders to create flexible MOFs containing
composites can also potentially lead to changes in the flexible
behavior.
Along these lines, the Watanabe group was able to show the
influence of polymer matrix (polyvinylpyrrolidone, PVP) on the
switching behavior of ELM-11.[124]Later on, similar experiments
were also conducted with an interpenetrated [Cu2(bdc)2(bpy)]n
framework (referred to as JG-MOF).[128]Interestingly, the ma-
terial shows a very distinct change in the gate-opening behav-
ior, depending not only on the amount of PVP added but also
on the sample preparation approach. Increasing binder amount
causes a “slacking” of the gate opening. The term describes an
external force-induced shift from a stepwise to a sequential tran-
sition, leading to a smearing of the originally steep adsorption
branch originating from the phase transition.[124]This effect was
explained through free energy analysis, showing a nonsynchro-
nized transformation within a single MOF particle. Furthermore,
the slacking could be modulated by the composite preparation
method as it impacts the force inflicted on the material. For
higher external forces, more pronounced slacking was observed.
In line with theoretical calculations, similar experiments on the
less-expanding interpenetrated JG-MOF confirmed this theory.
Due to the reduced expansion of the networks, less force was
exerted in the pellet, thus minimizing slacking.[128]Therefore,
developing methods to reduce the stress inflicted by binders is
crucial.[124]
4.5. Slipping Off Effect
To use flexible MOFs for separation, the set partial pressure
of the stimulus must exceed gate opening pressure to induce
the phase transition and enable the separation. Thus, in break-
through applications, no pure gases can be obtained because the
Figure 40. Comparison of typical breakthrough curves (blue) and
isotherms (purple) for rigid microporous adsorbents (top) and “gating”-
type flexible MOFs (bottom). The gate opening and gate closing pressures
are marked by A and B, respectively.The breakthrough curves for gating ad-
sorbents have a “stepped” feature not seen in rigid adsorbents, defined by
a plateau in the effluent fraction (represented by a “step height”) before the
ultimate breakthrough. Reproduced with permission.[77a]Copyright 2017,
American Chemical Society.
target gas exhausts until the gate opening pressure is reached.
This phenomenon was named the “slipping-off” effect.[129]A
comparison of a typical breakthrough curve profile of a rigid
adsorbent and a representative “stepped” breakthrough curve
of a “gating”-type flexible MOF with “slipping-off” is shown in
Figure 40.
Such an effect was reported several times for the breakthrough
curves of flexible materials. For example, the breakthrough
curves collected for CH4/CO2on MIL-53(Cr) exhibit distinct
changes in the slope of effluent CO2concentration with time.[81]
An even more pronounced slipping-off breakthrough curve was
observed for the separation of xylenes on MIL-53(Al)[33a]and sep-
aration of CH4and CO2on CID-5, where the partial pressure of
CO2dropped below the gate opening pressure of the adsorbents,
while it was passing the column.[130]
Breakthrough measurements performed on ELM-11 also un-
covered such a breakthrough step, which was discussed by So-
Adv. Mater. 2025, 2414724 2414724 (33 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 41. a) Pictures of the column system used by Tanaka and co-workers in order to show the slipping-off effect and its solution b) by combining
a column of ELM-11 and HKUST-1. c,d) Breakthrough curves of CO2and CH4for the system depicted in (a) and (b), respectively. Reproduced with
permission.[22]Copyright 2020, The Authors, under CC BY 4.0.
tomayor and Lastoskie in 2017 and has also been later on found
in [Co(bdp)]n.[69,77a]
Since this effect is very relevant for practical applications, more
than one solution has been proposed.
i) One possible solution is to overpressure the feed gas so that
the partial pressure of the opening gas remains above the
gate pressure along the entire length of the column. This ap-
proach will, however, incur an energy penalty in the work
input required to pressurize the feed gas.[77a]
ii) Another way is to directly influence the adsorption condi-
tions by increasing or decreasing the temperature to shift
the gate opening pressure. Further functionalization of the
linker shows an impact if, afterward, one of the phases will
be further stabilized with this functionality.[69]
iii) An advanced approach was demonstrated by Kitagawa and
co-workers,[130]who used a solid solution of two MOFs, CID-
5 and CID-6, to cleanly separate methane from carbon diox-
ide and ethane. CID-5 has more structural flexibility than
CID-6 and exhibits selective gated adsorption for CO2and
C2H6over CH4, whereas CID-6 has permanent microp-
orosity and nonselectively adsorbs all three gases. In break-
through experiments with a 60:40 vol% CH4/CO2mixture
and a 90:10 vol% CH4/C2H6mixture on CID-5, the CH4
fraction in the effluent was with 90% well below the target
due to the “slipping-off”. Pure methane could also not be ob-
tained for the same gas mixtures and CID-6, because of coad-
sorption. However, a solid solution of CID-5 and CID-6 with
gated adsorption characteristics for both CO2and C2H6,pre-
pared by substituting 10% of the nitroisophthalate ligands
Adv. Mater. 2025, 2414724 2414724 (34 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
of CID-5 with the methoxyisophthalate ligands of CID-6,
cleanly separated methane from both carbon dioxide and
ethane for retention times of 8 and 25 min, respectively, in
breakthrough column experiments. Hence, by tuning the
gate opening pressures via framework composition, it was
possible to optimize the gas separation performance under
dynamic conditions.[77a,130]
iv) In 2020, a novel approach was demonstrated by Tanaka and
co-workers by successively operating columns containing
flexible and rigid MOF as a guard bed (Figure 41).[129]Signif-
icant improvement of selectivity for CO2was achieved when
connecting a column containing flexible ELM-11 to a column
loaded with rigid microporous HKUST-1. In this setup, al-
most the entire CO2could be adsorbed. The slipping-off CO2
was collected by the HKUST-1 column and thereby allowed
for very high purity of the exhaust gas while maintaining a
high capacity for CO2.
v) The MOFs showing smooth, quasi-second-order phase tran-
sitions can also be beneficial in overcoming the slipping-off
problem.[18b]
5. Beyond Improving Working Capacity and
Selectivity: Intelligent Materials and Use of
Switchability
5.1. Thermal Management
A very important problem is the change in the temperature of
the adsorption bed, leading to a reduction of adsorption capac-
ity upon adsorption and an increase of adsorption capacity upon
desorption or even preventing the complete removal of guest
molecules.[131]
Engineering solutions to this problem can be found in adsor-
ber design, which allows for better heat transfer. The trade-off,
however, remains between achieving high selectivity and capac-
ity while managing the inherent challenges associated with these
properties.[129]Developing novel materials capable of satisfying
both requirements is a key focus of ongoing research.
One potential solution was proposed to use phase change ma-
terials (PCMs) as latent heat storing additives. PCMs are able
to absorb and store some of the heat released during adsorp-
tion through a phase transition.[131]The flexible MOF itself is,
in a certain sense, a PCM (Figure 42).[132]This leads to an intrin-
sic thermal management upon physisorption that can be mon-
itored in calorimetric measurements.[26]The intrinsic thermal
management capability of flexible MOFs should be useful for
storage applications as well as for developing highly efficient PSA
systems.[118]Typical gas physisorption enthalpies are between 5
and 60 kJ per mol of gas,[61]depending on adsorptive and sur-
face chemistry, as well as the pore sizes of the adsorbent. The
structural transition enthalpies in MOFs are connected with the
thermal effects ranging between 6 and 31 kJ per mol of MOF.[37]
In [Co(bdp)]n,differential enthalpies of CH4and CO2adsorp-
tion reveals significant reductions in heat released upon adsorp-
tion during the discrete, endothermic structural phase changes
relative to the regions between these phase changes (Figure 43).
For comparison, the isostructural rigid MOFs [Ni(bdp)]nand
[Zn(bdp)]n, display differential enthalpies of CO2adsorption of
Figure 42. When adiabatic gas adsorption is considered, the temperature
rise of the system for the flexible MOF, ΔTX, is smaller than that for the
conventional adsorbent, ΔTY, and the resulting decrease in the adsorption
amount can be suppressed because of the smaller net heat of adsorption
of the flexible MOF owing to its intrinsic thermal management capability.
Reproduced with permission.[22]Copyright 2020, The Authors, under CC
BY 4.0.
20 kJ mol1at zero coverage.[133]Notably, these values are very
close to those observed for [Co(bdp)]nat the pressures between
the CO2-induced phase transition. During the first phase transi-
tion, [Co(bdp)]nshows values in the range of 24 to 26 kJ mol1,
which are significantly lower in magnitude than what might be
expected. During the second and third CO2-induced phase tran-
sitions, the magnitude of adsorption enthalpy plummets dra-
matically, reaching values as small as 11 and 16 kJ mol1,
respectively (Figure 43). Thus, the structural phase changes of
[Co(bdp)]n, because of their endothermic nature, can substan-
tially mitigate the amount of heat that must be dissipated during
adsorption.
For ELM-11, the molar integral heats of adsorption in the
phases with increasing porosity are estimated by GCMC simu-
lations to be 40.8, 38.3, and 33.6 kJ per mol CO2, respectively.
The transition enthalpies were determined to be 21.9 kJ per mol
Adv. Mater. 2025, 2414724 2414724 (35 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Figure 43. Differential enthalpies of CO2adsorption in [Co(bdp)]nare
shown in purple (standard errors are shown as black bars) as a function
of CO2loading. Local minima in adsorption enthalpy correspond to re-
gions in which [Co(bdp)]nundergoes an endothermic structural expan-
sion, which offsets some of the heat released upon CO2adsorption and
provides intrinsic thermal management. The single-component CO2ad-
sorption isotherm (red circles) is provided for comparison. Reproduced
with permission.[69]Copyright 2018, American Chemical Society.
CO2for the gate closing at higher pressures and 25.5 kJ per mol
CO2for the gate closing at lower pressure, which suggests that
both closings are endothermic processes and that more heat per
mole of CO2is removed from the system for the gate closing at
lower pressure than for that at the higher pressures.
Hiraide et al.,[129]in their study on ELM-11, demonstrated that
during the adsorption of an equimolar CO2/CH4gas mixture at
500 kPa and 298 K, the host material undergoes an endothermic
expansion with an associated enthalpy change of 55.7 J g1.Con-
currently, the exothermic enthalpy change due to the adsorption
of the gas mixture is 135.3 J g1. The net heat effect is decreased
to 79.6 J g1, indicating that 41% of the exothermic heat is offset
by the endothermic process.
Thus, the structural transitions in flexible adsorbents offer ob-
vious advantages for thermal management compared to rigid
adsorbents.
6. Conclusion
Flexible MOFs are emerging materials for advanced applications
in energy storage and gas separation thanks to their unique ad-
sorption properties and responsive behavior. The flexibility of
the frameworks essentially contributes to achieving exceptionally
high selectivity in separation processes and improves deliverable
storage capacity. These materials offer great potential for tackling
challenging (isotope) separations that would be difficult or even
impossible utilizing rigid structures. The responsivity of MOFs
can be fine-tuned by adjusting the metal type, linker substituents,
or by employing mixed metal or mixed linker strategies.
When designing separation processes, it is crucial to consider
not only the single gas adsorption isotherms of both components
involved but also the desorption branches (as discussed in Sec-
tion 4), as this is essential for accurately predicting or leveraging
the enhanced separation capabilities of flexible MOFs. Addition-
ally, the temperature-dependent behavior of these flexible frame-
works must be carefully considered, as it plays a key role in their
performance under different conditions.
For column-based applications, MOFs with minimal volume
changes are generally preferable to ensure stable operation, but
it has been shown, that also flexible MOFs can be shaped while
maintaining their storage capacity and flexibility, allowing for
greater adaptability in industrial applications. Additionally, their
intrinsic thermal management capabilities further enhance their
suitability for demanding energy storage and gas separation pro-
cesses. With their versatile and tunable properties, flexible MOFs
are poised to revolutionize these two fields in the future.
Acknowledgements
The DGF is acknowledged for financial support in the frame of the research
unit (FOR 2433). Y.W. gratefully acknowledges funding from the Alexan-
der von Humboldt Foundation. V.B. acknowledges BMBF for funding
within project Nos. 05K22OD1 (“TIMESWITCH”) and 05K22OD2 (“TO-
MOPORE”).
Open access funding enabled and organized by Projekt DEAL.
Conflict of Interest
The authors declare no conflict of interest.
Keywords
adsorption selectivity, breathing, flexible metal–organic frameworks, gate
opening, gas separation, gas storage, soft porous crystals
Received: September 28, 2024
Revised: November 28, 2024
Published online:
[1] S. R. Batten, N. R. Champness, X.-M. Chen, J. Garcia-Martinez, S.
Kitagawa, L. Öhrström, M. O’Keeffe, M. P. Suh, J. Reedijk, Pure Appl.
Chem. 2013,85, 1715.
[2] X. Zhang, Z. Chen, X. Liu, S. L. Hanna, X. Wang, R. Taheri-Ledari, A.
Maleki, P. Li, O. K. Farha, Chem.Soc.Rev.2020,49, 7406.
[3] a) K. R. Wright, K. Nath, A. J. Matzger, Angew. Chem., Int. Ed. 2022,
61, 202213190; b) J. Ma, A. P. Kalenak, A. G. Wong-Foy,A. J. Matzger,
Angew. Chem., Int. Ed. 2017,56, 14618; c) A. J. Howarth, A. W. Peters,
N. A. Vermeulen, T. C. Wang, J. T. Hupp, O. K. Farha, Chem. Mater.
2017,29, 26.
[4] a)K.O.Kirlikovali,S.L.Hanna,F.A.Son,O.K.Farha,ACS Nanosci.
Au 2023,3, 37; b) B. Pramanik, R. Sahoo, M. C. Das, Coord. Chem.
Rev. 2023,493, 215301.
[5] a) O. M. Yaghi, J. Am. Chem. Soc. 2016,138, 15507; b) H. Jiang, D.
Alezi, M. Eddaoudi, Nat. Rev. Mater. 2021,6, 466; c) O. M. Yaghi,
M. J. Kalmutzki, C. S. Diercks, Introduction to Reticular Chemistry.
Metal–Organic Frameworks and Covalent Organic Frameworks,Wiley-
VCH, Weinheim, Germany 2019.d)R.Freund,S.Canossa,S.M.
Cohen, W. Yan, H. Deng, V. Guillerm, M. Eddaoudi, D. G. Madden,
D. Fairen-Jimenez, H. Lyu, L. K. Macreadie, Z. Ji, Y. Zhang, B. Wang,
F. Haase, C. Wöll, O. Zaremba, J. Andreo, S. Wuttke, C. S. Diercks,
Angew. Chem., Int. Ed. 2021,60, 23946.
[6] a)M.R.Ryder,B.Civalleri,J.C.Tan,Phys. Chem. Chem. Phys. 2016,
13, 9079. b) E. Mahmoud, Crystal s 2020,10,700; c) W. Fan, X. Zhang,
Adv. Mater. 2025, 2414724 2414724 (36 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Z. Kang, X. Liu, D. Sun, Coord. Chem. Rev. 2021,443, 213968; d) F.-F.
Lu, X.-W. Gu, E. Wu, B. Li, G. Qian, J. Mater.Chem. A 2023,11, 1246;
e) S. Krause, J. D. Evans, V. Bon, I. Senkovska, S. Ehrling, U. Stoeck,
P. G. Yot, P. Iacomi, P. Llewellyn, G. Maurin, F.-X. Coudert, S. Kaskel,
J. Phys. Chem. C 2018,122, 19171.
[7] I. M. Hönicke, I. Senkovska, V. Bon, I. A. Baburin, N. Bönisch, S.
Raschke, J. D. Evans, S. Kaskel, Angew. Chem., Int. Ed. 2018,57,
13780.
[8] a) W. Xu, O. M. Yaghi, ACS Cent. Sci. 2020,6, 1348; b) W. Song, Z.
Zheng, A. H. Alawadhi, O. M. Yaghi, Nat. Water 2023,1, 626; c) C.
Charalambous, E. Moubarak, J. Schilling, E. Sanchez Fernandez, J.-Y.
Wang, L. Herraiz, F. McIlwaine, S. B. Peh, M. Garvin, K. M. Jablonka,
S. M. Moosavi, J. Van Herck, A. Y. Ozturk, A. Pourghaderi, A.-Y. Song,
G. Mouchaham, C. Serre, J. A. Reimer, A. Bardow, B. Smit, S. Garcia,
Nature 2024,632, 89; d) J.-B. Lin, T. T. T. Nguyen, R. Vaidhyanathan,
J. Burner, J. M. Taylor, H. Durekova, F.Akhtar, R. K. Mah, O. Ghaffari-
Nik, S. Marx, N. Fylstra, S. S. Iremonger, K. W. Dawson, P. Sarkar, P.
Hovington, A. Rajendran, T. K. Woo, G. K. H. Shimizu, Science 2021,
374, 1464.
[9] a) S. Horike, S. Shimomura, S. Kitagawa, Nat. Chem. 2009,1, 695;
b) A. Schneemann, V. Bon, I. Schwedler, I. Senkovska, S. Kaskel, R.
A. Fischer, Chem. Soc. Rev. 2014,43, 6062; c) S. Kitagawa, Flexible
Metal–Organic Frameworks: Structural Design, Synthesis and Proper-
ties, Royal Society of Chemistry, Cambridge, UK 2024.
[10] a) S. Krause, N. Hosono, S. Kitagawa, Angew. Chem., Int. Ed. 2020,
59, 15325; b) J. H. Lee, S. Jeoung, Y. G. Chung, H. R. Moon, Coord.
Chem. Rev. 2019,389, 161; c) S. Seth, S. Jhulki, Mater. Horiz. 2021,8,
700; d) I. Senkovska, V. Bon, L. Abylgazina, M. Mendt, J. Berger, G.
Kieslich, P. Petkov, J. Luiz Fiorio, J. O. Joswig, T. Heine, L. Schaper, C.
Bachetzky, R. Schmid, R. A. Fischer, A. Pöppl, E. Brunner, S. Kaskel,
Angew. Chem., Int. Ed. 2023,62, 202218076.
[11] a) X. Liu, Z. Wang, Y. Zhang, N. Yang, B. Gui, J. Sun, C. Wang, J. Am.
Chem. Soc. 2024,146, 11411; b) A. Mauri, R. Vismara, M. Moroni,
E. Roldán-Molina, J. A. R. Navarro, S. Galli, Small Sci. 2024,4,
2400277.
[12] a) J. Li, B. Chen, Chem. Sci. 2024,15, 9874; b) Y. He, S. Xiang, B.
Chen, J. Am. Chem. Soc. 2011,133, 14570; c) C. Jiang, J.-X. Wang,
D. Liu, E. Wu, X.-W. Gu, X. Zhang, B. Li, B. Chen, G. Qian, Angew.
Chem., Int. Ed. 2024,63, 202404734.
[13] Z. Liu, L. Zhang, D. Sun, Chem. Commun. 2020,56, 9416.
[14] J. D. Evans, V. Bon, I. Senkovska, H.-C. Lee, S. Kaskel, Nat. Commun.
2020,11, 2690.
[15] a) S. Watanabe, H. Sugiyama, H. Adachi, H. Tanaka, M. T. Miyahara,
J. Chem. Phys. 2009,130; b) H. Sugiyama, S. Watanabe, H. Tanaka,
M. T. Miyahara, Langmuir 2012,28, 5093.
[16] H. Tanaka, M. T. Miyahara, Curr. Opin. Chem. Eng. 2019,24, 19.
[17] R. Goeminne, S. Krause, S. Kaskel, T. Verstraelen, J. D. Evans, J. Am.
Chem. Soc. 2021,143, 4143.
[18] a) A. Schneemann, Y. Takahashi, R. Rudolf, S.-I. Noro, R. A. Fischer,
J. Mater. Chem. A 2016,4, 12963; b) C. Wang, X.-W. Zhang, X.-X.
Chen, W.-X. Zhang, J.-P. Zhang, J. Am. Chem. Soc. 2024,146, 13886.
[19] D.-D. Zhou, J.-P. Zhang, Acc. Chem. Res. 2022,55, 2966.
[20] D. P. Broom, Adsorption 2024,30, 1565.
[21] a) I. Senkovska, S. Kaskel, Microporous Mesoporous Mater. 2008,112,
108; b) U. Stoeck, S. Krause, V. Bon, I. Senkovska, S. Kaskel, Chem.
Commun. 2012,48, 10841; c) Y. He, W. Zhou, G. Qian, B. Chen,
Chem.Soc.Rev.2014,43, 5657; d) C.-C. Liang, Z.-L. Shi, C.-T. He, J.
Tan, H.-D. Zhou, H.-L. Zhou, Y. Lee, Y.-B. Zhang, J. Am. Chem. Soc.
2017,139, 13300; e) C. M. Simon, J. Kim, D. A. Gomez-Gualdron,
J. S. Camp, Y. G. Chung, R. L. Martin, R. Mercado, M. W. Deem,
D. Gunter, M. Haranczyk, D. S. Sholl, R. Q. Snurr, B. Smit, Energy
Environ. Sci. 2015,8, 1190.
[22] S. Hiraide, Y. Sakanaka, H. Kajiro, S. Kawaguchi, M. T. Miyahara, H.
Tana k a , Nat. Commun. 2020,11, 3867.
[23] a) H. Demir, M. Mobedi, S. Ülkü, Int. Commun. Heat Mass Transfer.
2009,36, 372; b) J. Wieme, S. Vandenbrande, A. Lamaire, V. Kapil, L.
Vanduyfhuys, V. Van Speybroeck, ACS Appl. Mater. Interfaces 2019,
11, 38697.
[24] K. J. Chang, O. Talu, Appl. Therm. Eng. 1996,16, 359.
[25] a) M. M. Mohseni, M. Jouyandeh, S. M. Sajadi, A. Hejna, S.
Habibzadeh, A. Mohaddespour, N. Rabiee, H. Daneshgar, O.
Akhavan, M. Asadnia, M. Rabiee, S. Ramakrishna, R. Luque, M. R.
Saeb, Chem. Eng. J. 2022,449, 137700; b) H. Babaei, K. R. Meihaus,
J. R. Long, Chem. Mater. 2023,35, 6220; c) C. Yuan, S. S. Sørensen,
T. Du, Y. Song, M. M. Smedskjaer, Int. J. Heat Mass Transfer 2024,
233, 126015.
[26] J. A. Mason, J. Oktawiec, M. K. Taylor, M. R. Hudson, J. Rodriguez,
J. E. Bachman, M. I. Gonzalez, A. Cervellino, A. Guagliardi, C. M.
Brown, P. L. Llewellyn, N. Masciocchi, J. R. Long, Nature 2015,527,
357.
[27] X. Zhang, X. Wang, F. Gao, Y. Chen, H. Liu, P. Zhou, Z. Kang, Y.Wang,
W. Fa n , Mater. Adv. 2024,5, 3135.
[28] a) MOVE (Mobility Optimized Vehicle Energy) program by
ARPA-E: https://arpa-e.energy.gov/sites/default/files/documents/
files/MOVE_ProgramOverview.pdf (accessed: September 2024). b)
ISO 15403-1:2016 standard: https://www.iso.org/obp/ui/#iso:std:
iso:15403:-1:ed- 1:v1 (accessed: September 2024). c) M. Beckner, A.
Dailly, Appl. Energy 2015,149, 69.
[29] A. Granja-DelRío, I. Cabria, Int. J. Hydrogen Energy 2024,54, 665.
[30] a) Y. Peng, V. Krungleviciute, I. Eryazici, J. T. Hupp, O. K. Farha,
T. Yildirim, J. Am. Chem. Soc. 2013,135, 11887; b) J. Jiang, H.
Furukawa, Y.-B. Zhang, O. M. Yaghi, J. Am. Chem. Soc. 2016,138,
10244; c) M. Zhang, W. Zhou, T. Pham, K. A. Forrest, W. Liu, Y. He,
H.Wu,T.Yildirim,B.Chen,B.Space,Y.Pan,M.J.Zaworotko,J.Bai,
Angew. Chem., Int. Ed. 2017,56, 11426; d) A. Dailly, M. Beckner,
in Nanoporous Materials for Gas Storage, Green Energy and Tech-
nology (Eds: K. Kaneko, F. Rodríguez-Reinoso), Springer, Singapore
2019.
[31] D. A. Gómez-Gualdrón, C. E. Wilmer, O. K. Farha, J. T. Hupp, R. Q.
Snurr, J. Phys. Chem. C 2014,118, 6941.
[32] K. A. Forrest, G. Verma, Y. Ye, J. Ren, S. Ma, T. Pham, B. Space, Chem.
Phys. Rev. 2022,3, 021308.
[33] a) R. Kitaura, K. Seki, G. Akiyama, S. Kitagawa, Angew. Chem., Int.
Ed. 2003,42, 428; b) K. Seki, Phys.Chem.Chem.Phys.2002,4,
1968.
[34] H. Kanoh, A. Kondo, H. Noguchi, H. Kajiro, A. Tohdoh, Y. Hattori,
W.-C. Xu, M. Inoue, T. Sugiura, K. Morita, H. Tanaka, T. Ohba, K.
Kaneko, J. Colloid Interface Sci. 2009,334,1.
[35] S. Krause, V. Bon, I. Senkovska, U. Stoeck, D. Wallacher, D. M.
Többens, S. Zander, R. S. Pillai, G. Maurin, F.-X. Coudert, S. Kaskel,
Nature 2016,532, 348.
[36] S. Krause, J. D. Evans, V. Bon, I. Senkovska, P. Iacomi, F. Kolbe, S.
Ehrling, E. Troschke, J. Getzschmann, D. M. Többens, A. Franz, D.
Wallacher, P. G. Yot, G. Maurin, E. Brunner, P. L. Llewellyn, F.-X.
Coudert, S. Kaskel, Nat. Commun. 2019,10, 3632.
[37] a) S. Krause, J. D. Evans, V. Bon, I. Senkovska, F.-X. Coudert, D. M.
Többens, D. Wallacher, N. Grimm, S. Kaskel, Faraday Discuss. 2021,
225, 168; b) S. Rahman, A. Arami-Niya, X. Yang, G. Xiao, G. Li, E. F.
May, Commun. Chem. 2020,3, 186.
[38] M. K. Taylor, T. Runˇ
cevski, J. Oktawiec, M. I. Gonzalez, R. L.
Siegelman, J. A. Mason, J. Ye, C. M. Brown, J. R. Long, J. Am. Chem.
Soc. 2016,138, 15019.
[39] Q. Y. Yang, P. Lama, S. Sen, M. Lusi, K. J. Chen, W. Y. Gao, M.
Shivanna, T. Pham, N. Hosono, S. Kusaka, J. J. Perry, S. Ma, B.
Space, L. J. Barbour, S. Kitagawa, M. J. Zaworotko, Angew. Chem.,
Int. Ed. 2018,57, 5684.
[40] S.-M. Wang, M. Shivanna, P. Lama, Q.-Y. Yang, L. J. Barbour, M. J.
Zaworotko, ChemSusChem 2023,16, 202300069.
Adv. Mater. 2025, 2414724 2414724 (37 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
[41] S. Ehrling, M. Mendt, I. Senkovska, J. D. Evans, V. Bon, P. Petkov, C.
Ehrling, F. Walenszus, A. Pöppl, S. Kaskel, Chem. Mater. 2020,32,
5670.
[42] H. J. Park, M. P. Suh, Chem. Commun. 2010,46, 610.
[43] M. Sin, N. Kavoosi, M. Rauche, J. Pallmann, S. Paasch, I. Senkovska,
S. Kaskel, E. Brunner, Langmuir 2019,35, 3162.
[44] V. Bon, I. Senkovska, S. Kaskel, in Nanoporous Materials for Gas Stor-
age, Green Energy and Technology (Eds: K. Kaneko, F. Rodríguez-
Reinoso), Springer, Singapore 2019.
[45] V. Bon, I. Senkovska, D. Wallacher, D. M. Többens, I. Zizak, R.
Feyerherm, U. Mueller, S. Kaskel, Inorg. Chem. 2014,53, 1513.
[46] a) T. Kundu, B. B. Shah, L. Bolinois, D. Zhao, Chem. Mater. 2019,
31, 2842; b) J. Yang, X. Yan, T. Xue, Y. Liu, RSC Adv. 2016,6, 55266;
c) S. Biswas, T. Ahnfeldt, N. Stock, Inorg. Chem. 2011,50, 9518.
[47] a) M. Bonneau, C. Lavenn, J.-J. Zheng, A. Legrand, T. Ogawa, K.
Sugimoto, F.-X. Coudert, R. Reau, S. Sakaki, K.-i. Otake, S. Kitagawa,
Nat. Chem. 2022,14, 816; b) L. Bolinois, T. Kundu, X. Wang, Y. Wang,
Z.Hu,K.Koh,D.Zhao,Chem. Commun. 2017,53, 8118.
[48] a) L. Abylgazina, I. Senkovska, S. Ehrling, V. Bon, P. St Petkov,
J. D. Evans, S. Krylova, A. Krylov, S. Kaskel, CrystEngComm 2021,
23, 538; b) L. Abylgazina, I. Senkovska, R. Engemann, S. Ehrling,
T. E. Gorelik, N. Kavoosi, U. Kaiser, S. Kaskel, Front. Chem. 2021,
9, 674566; c) S. Ehrling, H. Miura, I. Senkovska, S. Kaskel, Trends
Chem. 2021,3, 291.
[49] H. Reinsch, M. A. van der Veen, B. Gil, B. Marszalek, T. Verbiest, D.
de Vos, N. Stock, Chem. Mater. 2013,25, 17.
[50] H.-P. Li, J.-W. Wang, Z. Dou, L.-Z. Wu, Y. Wang, Y. Liang, Q.-G. Zhai,
Chem. Eng. J. 2024,492.
[51] a) U.S. Department of Energy’s (DOE) technical targets for
onboard hydrogen storage systems in light-duty vehicles:
https://www.energy.gov/eere/fuelcells/doe-technical-targets-
onboard-hydrogen-storage-light-duty-vehicles (accessed: Septem-
ber 2024). b) M. D. Allendorf, Z. Hulvey, T. Gennett, A. Ahmed,
T. Autrey, J. Camp, E. Seon Cho, H. Furukawa, M. Haranczyk, M.
Head-Gordon, S. Jeong, A. Karkamkar, D.-J. Liu, J. R. Long, K. R.
Meihaus, I. H. Nayyar, R. Nazarov, D. J. Siegel, V. Stavila, J. J. Urban,
S.P.Veccham,B.C.Wood,Energy Environ. Sci. 2018,11, 2784.
[52] X. Zhang, R.-B. Lin, J. Wang, B. Wang, B. Liang, T. Yildirim, J. Zhang,
W. Zhou, B. Chen, Adv. Mater. 2020,32, 1907995.
[53] H. J. Choi, M. Dinc˘
a, J. R. Long, J. Am. Chem. Soc. 2008,130, 7848.
[54] A. Halder, R. A. Klein, S. Shulda, G. A. McCarver, P. A. Parilla, H.
Furukawa, C. M. Brown, C. M. McGuirk, J. Am. Chem. Soc. 2023,
145, 8033.
[55] a) J. Y. Kim, L. Zhang, R. Balderas-Xicohtencatl, J. Park, M. Hirscher,
H. R. Moon, H. Oh, J. Am. Chem. Soc. 2017,139, 17743; b) R. A.
Pollock, J.-H. Her, C. M. Brown, Y. Liu, A. Dailly, J. Phys. Chem. C
2014,118, 18197.
[56] H. Li, L. Li, R.-B. Lin, W. Zhou, Z. Zhang, S. Xiang, B. Chen, Energy-
Chem 2019,1, 100006.
[57] J.-P. Zhang, X.-M. Chen, J. Am. Chem. Soc. 2009,131, 5516.
[58] R. Matsuda, R. Kitaura, S. Kitagawa, Y. Kubota, R. V. Belosludov, T.
C. Kobayashi, H. Sakamoto, T. Chiba, M. Takata, Y. Kawazoe, Y. Mita,
Nature 2005,436, 238.
[59] C.-T. He, Z.-M. Ye, Y.-T. Xu, D.-D. Zhou, H.-L. Zhou, D. Chen, J.-P.
Zhang, X.-M. Chen, Chem. Sci. 2017,8, 7560.
[60] H. Zeng, M. Xie, Y. L. Huang, Y. Zhao, X. J. Xie, J. P. Bai, M. Y. Wan,
R. Krishna, W. Lu, D. Li, Angew. Chem., Int. Ed. 2019,58, 8515.
[61] D. M. D’Alessandro, B. Smit, J. R. Long, Angew. Chem., Int. Ed. 2010,
49, 6058.
[62] M. R. Abdul Hamid, Y. Qian, R. Wei, Z. Li, Y. Pan, Z. Lai, H.-K. Jeong,
J. Membr. Sci. 2021,640, 119802.
[63] a) L. Wang, Y. Li, Y. Wang, J. Yang, L. Li, J. Li, Sep. Purif. Technol. 2020,
251, 117311; b) O. T. Qazvini, L. K. Macreadie, S. G. Telfer, Chem.
Mater. 2020,32, 6744; c) O. T. Qazvini, V.-J. Scott, L. Bondorf, M.
Ducamp, M. Hirscher, F.-X. Coudert, S. G. Telfer, Chem. Mater. 2021,
33, 8886.
[64] a) Y. Takakura, S. Sugimoto, J. Fujiki, H. Kajiro, T. Yajima, Y. Kawajiri,
ACS Sustainable Chem. Eng. 2022,10, 14935; b) S. Hiraide, H.
Tanaka, N. Ishikawa, M. T. Miyahara, ACS Appl Mater. Interfaces
2017,9, 41066; c) R. Haldar, M. Inukai, S. Horike, K. Uemura, S.
Kitagawa, T. K. Maji, Inorg. Chem. 2016,55, 4166.
[65] J. M. Huck, L.-C. Lin, A. H. Berger, M. N. Shahrak, R. L. Martin, A. S.
Bhown, M. Haranczyk, K. Reuter, B. Smit, Energy Environ. Sci. 2014,
7, 4132.
[66] a) L. C. Lin, A. H. Berger, R. L. Martin, J. Kim, J. A. Swisher, K.
Jariwala, C. H. Rycroft, A. S. Bhown, M. W. Deem, M. Haranczyk,
B. Smit, Nat. Mater. 2012,11, 633; b) Z. Hu, Y. Wang, B. B. Shah, D.
Zhao, Adv. Sustainable Syst. 2019,3, 1970002.
[67] a) P. Pullumbi, F. Brandani, S. Brandani, Curr. Opin. Chem. Eng.
2019,24, 131; b) Y.-S. Wang, X.-J. Zhang, Y.-Q. Ba, T.-Y. Li, G.-P.
Hao, A.-H. Lu, Research 2022,2022,https://doi.org/10.34133/2022/
9780864.
[68] a) R.-B. Lin, S. Xiang, W. Zhou, B. Chen, Chem 2020,6, 337; b) C.
Jiang, X. Wang, Y. Ouyang, K. Lu, W. Jiang, H. Xu, X. Wei, Z. Wang,
F. Dai, D. Sun, Nanoscale Adv. 2022,4, 2077; c) X. Zhao, Y. Wang, D.
S. Li, X. Bu, P. Feng, Adv. Mater. 2018,30, 1705189; d) X. Zhang, Y.
Li, J.-R. Li, Trends Chem. 2024,6, 22.
[69] M. K. Taylor, T. Runcevski, J. Oktawiec, J. E. Bachman, R. L.
Siegelman, H. Jiang, J. A. Mason, J. D. Tarver, J. R. Long, J. Am.
Chem. Soc. 2018,140, 10324.
[70] J. U. Keller, F. Dreisbach, H. Rave, R. Staudt, M. Tomalla, Adsorption
1999,5, 199.
[71] D. Shade, B. W. S. Bout, D. S. Sholl, K. S. Walton, Ind. Eng. Chem.
Res. 2022,61, 2367.
[72] K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch,
Z.R.Herm,T.H.Bae,J.R.Long,Chem. Rev. 2012,112,724.
[73] O. Talu, Adv. Colloid Interface Sci. 1998,76–77, 227.
[74] S. Hiraide, Y. Sakanaka, Y. Iida, H. Arima, M. T. Miyahara, S.
Watanabe, Proc. Natl. Acad. Sci. USA 2023,120, 2305573120.
[75] a) F. X. Coudert, C. Mellot-Draznieks, A. H. Fuchs, A. Boutin, J. Am.
Chem. Soc. 2009,131, 11329; b) F.-X. Coudert, Phys. Chem. Chem.
Phys. 2010,12, 10904.
[76] S. M. J. Rogge, R. Goeminne, R. Demuynck, J. J. Gutiérrez-Sevillano,
S. Vandenbrande, L. Vanduyfhuys, M. Waroquier, T. Verstraelen, V.
Van S pe ybroe ck, Adv. Theory Simul. 2019,2, 1800177.
[77] a) F. J. Sotomayor, C. M. Lastoskie, Langmuir 2017,33, 11670; b) H.
Tanaka, S. Hiraide, A. Kondo, M. T. Miyahara, J. Phys. Chem. C 2015,
119, 11533.
[78] G. Fraux, A. Boutin, A. H. Fuchs, F.-X. Coudert, Adsorption 2018,24,
233.
[79] a) S. Sircar, Ind. Eng. Chem. Res. 2006,46, 2917; b) A. Rajendran, V.
Kariwala, S. Farooq, Chem. Eng. Sci. 2008,63, 2696.
[80] T.R.C.VanAssche,G.V.Baron,J.F.M.Denayer,Dalton Trans. 2016,
45, 4416.
[81] L. Hamon, P. L. Llewellyn, T. Devic, A. Ghoufi, G. Clet, V. Guillerm,
G. D. Pirngruber, G. Maurin, C. Serre, G. Driver, W. van Beek, E.
Jolimaitre, A. Vimont, M. Daturi, G. Ferey, J. Am. Chem. Soc. 2009,
131, 17490.
[82] V. Finsy, L. Ma, L. Alaerts, D. E. De Vos, G. V. Baron, J. F. M. Denayer,
Microporous Mesoporous Mater. 2009,120, 221.
[83] F.-X. Coudert, M. Jeffroy, A. H. Fuchs, A. Boutin, C. Mellot-
Draznieks, J. Am. Chem. Soc. 2008,130, 14294.
[84] a) N. Klein, C. Herzog, M. Sabo, I. Senkovska, J. Getzschmann,
S. Paasch, M. R. Lohe, E. Brunner, S. Kaskel, Phys. Chem. Chem.
Phys. 2010,12, 11778; b) N. Klein, H. C. Hoffmann, A. Cadiau, J.
Getzschmann, M. R. Lohe, S. Paasch, T. Heydenreich, K. Adil, I.
Senkovska, E. Brunner, S. Kaskel, J. Mater. Chem. 2012,22, 10303;
c) N. Kavoosi, T. Savchenko, I. Senkovska, M. Maliuta, V. Bon, A.
Adv. Mater. 2025, 2414724 2414724 (38 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Eychmüller, S. Kaskel, Microporous Mesoporous Mater. 2018,271,
169.
[85] A. Schneemann, Y. Jing, J. D. Evans, T. Toyao, Y. Hijikata, Y. Kamiya,
K. I. Shimizu, N. C. Burtch, S. I. Noro, Dalton Trans. 2021,50, 10423.
[86] K. Roztocki, M. Rauche, V. Bon, S. Kaskel, E. Brunner, D. Matoga,
ACS Appl. Mater. Interfaces 2021,13, 28503.
[87] M. Ichikawa, A. Kondo, H. Noguchi, N. Kojima, T. Ohba, H. Kajiro,
Y. Hattori, H. Kanoh, Langmuir 2016,32, 9722.
[88] Q. Dong, X. Zhang, S. Liu, R. B. Lin, Y. Guo, Y. Ma, A. Yonezu, R.
Krishna, G. Liu, J. Duan, R. Matsuda, W. Jin, B. Chen, Angew. Chem.,
Int. Ed. 2020,59, 22756.
[89] B. R. Barnett, M. I. Gonzalez, J. R. Long, Trends Chem. 2019,1, 159.
[90] C. Gücüyener, J. van den Bergh, J. Gascon, F. Kapteijn, J. Am. Chem.
Soc. 2010,132, 17704.
[91] L. Li, R. Krishna, Y. Wang, J. Yang, X. Wang, J. Li, J. Mater. Chem. A
2016,4, 751.
[92] S. Sen, N. Hosono, J.-J. Zheng, S. Kusaka, R. Matsuda, S. Sakaki, S.
Kitagawa, J. Am. Chem. Soc. 2017,139, 18313.
[93] K. Kishida, Y. Okumura, Y. Watanabe, M. Mukoyoshi, S. Bracco, A.
Comotti, P. Sozzani, S. Horike, S. Kitagawa, Angew. Chem., Int. Ed.
2016,55, 13784.
[94] J. Y. Kim, J. Park, J. Ha, M. Jung, D. Wallacher, A. Franz, R. Balderas-
Xicohténcatl, M. Hirscher, S. G. Kang, J. T. Park, I. H. Oh, H. R.
Moon, H. Oh, J. Am. Chem. Soc. 2020,142, 13278.
[95] a) Y. Liu, J.-H. Her, A. Dailly, A. J. Ramirez-Cuesta, D. A. Neumann,
C. M. Brown, J. Am. Chem. Soc. 2008,130, 11813; b) M. Mendt, B.
Jee, N. Stock, T. Ahnfeldt, M. Hartmann, D. Himsl, A. Pöppl, J. Phys.
Chem. C 2010,114, 19443.
[96] C.R.Groom,I.J.Bruno,M.P.Lightfoot,S.C.Ward,Acta Cryst. B
2016,72, 171.
[97] L. Bondorf, J. L. Fiorio, V. Bon, L. Zhang, M. Maliuta, S. Ehrling, I.
Senkovska, J. D. Evans, J. O. Joswig, S. Kaskel, T. Heine, M. Hirscher,
Sci. Adv. 2022,8, 7035.
[98] a) J. D. Evans, V. Bon, I. Senkovska, S. Kaskel, Langmuir 2021,
37, 4222; b) IUPAC project, focussing on standardized reporting
of gas adsorption isotherms: https://iupac.org/project/2021-016-
1-024 (accessed: September 2024).
[99] V. Bon, N. Klein, I. Senkovska, A. Heerwig, J. Getzschmann, D.
Wallacher, I. Zizak, M. Brzhezinskaya, U. Mueller, S. Kaskel, Phys.
Chem. Chem. Phys. 2015,17, 17471.
[100] T. Jacobs, G. O. Lloyd, J.-A. Gertenbach, K. K. Müller-Nedebock, C.
Esterhuysen, L. J. Barbour, Angew. Chem., Int. Ed. 2012,51, 4913.
[101] H. Sato, W. Kosaka, R. Matsuda, A. Hori, Y. Hijikata, R. V.
Belosludov, S. Sakaki, M. Takata, S. Kitagawa, Science 2014,343,
167.
[102] P. M. Bhatt, Y. Belmabkhout, A. Cadiau, K. Adil, O. Shekhah, A.
Shkurenko, L. J. Barbour, M. Eddaoudi, J. Am. Chem. Soc. 2016,138,
9301.
[103] S.-M. Wang, M. Shivanna, S.-T. Zheng, T. Pham, K. A. Forrest, Q.-Y.
Yang, Q. Guan, B. Space, S. Kitagawa, M. J. Zaworotko, J. Am. Chem.
Soc. 2024,146, 4153.
[104] a) E. Tal-Gutelmacher, D. Eliezer, E. Abramov, Mater. Sci. Eng., A
2007,445, 625; b) L. Zhang, S. Jee, J. Park, M. Jung, D. Wallacher,
A. Franz, W. Lee, M. Yoon, K. Choi, M. Hirscher, H. Oh, J. Am.
Chem. Soc. 2019,141, 19850; c) J. Y. Kim, L. Zhang, R. Balderas-
Xicohténcatl, J. Park, M. Hirscher, H. R. Moon, H. Oh, J. Am. Chem.
Soc. 2017,139, 17743; d) M. Liu, L. Zhang, M. A. Little, V. Kapil, M.
Ceriotti, S. Yang, L. Ding, D. L. Holden, R. Balderas-Xicohténcatl, D.
He, R. Clowes, S. Y. Chong, G. Schütz, L. Chen, M. Hirscher, A. I.
Cooper, Science 2019,366, 613.
[105] L. Zhang, S. Jee, J. Park, M. Jung, D. Wallacher, A. Franz, W. Lee,
M. Yoon, K. Choi, M. Hirscher, H. Oh, J. Am. Chem. Soc. 2019,141,
19850.
[106] a) H. Miura, V. Bon, I. Senkovska, S. Ehrling, N. Bonisch, G. Mader,
S. Grunzner, A. Khadiev, D. Novikov, K. Maity, A. Richter, S. Kaskel,
Adv. Mater. 2023,35, 2207741; b) J. D. Evans, V. Bon, I. Senkovska,
H. C. Lee, S. Kaskel, Nat. Commun. 2020,11, 2690.
[107] D. Tanaka, K. Nakagawa, M. Higuchi, S. Horike, Y. Kubota, T. C.
Kobayashi, M. Takata, S. Kitagawa, Angew. Chem., Int. Ed. 2008,47,
3914.
[108] L. Schaper, R. Schmid, Commun. Chem. 2023,6, 233.
[109] a) S. Sircar, H. Wu, J. Li, A. D. Lueking, Langmuir 2011,27, 14169;
b) A. D. Lueking, C.-Y. Wang, S. Sircar, C. Malencia, H. Wang, J. Li,
Dalton Trans. 2016,45, 4242.
[110] H. Preißler-Kurzhöfer, M. Lange, J. Möllmer, O. Erhart, M. Kobalz,
H. Krautscheid, R. Gläser, Nanomaterials 2024,14, 241.
[111] M. Kobalz, J. Lincke, K. Kobalz, O. Erhart, J. Bergmann, D. Lässig,
M. Lange, J. Möllmer, R. Gläser, R. Staudt, H. Krautscheid, Inorg.
Chem. 2016,55, 3030.
[112] N. Hosono, A. Terashima, S. Kusaka, R. Matsuda, S. Kitagawa, Nat.
Chem. 2019,11, 109.
[113] B. Seoane, S. Sorribas, Á. Mayoral, C. Téllez, J. Coronas, Microporous
Mesoporous Mater. 2015,203, 17.
[114] V. Bon, N. Busov, I. Senkovska, N. Bönisch, L. Abylgazina,
A. Khadiev, D. Novikov, S. Kaskel, Chem. Commun. 2022,58,
10492.
[115] Y. Sakanaka, S. Hiraide, I. Sugawara, H. Uematsu, S. Kawaguchi, M.
T. Miyahara, S. Watanabe, Nat. Commun. 2023,14, 6862.
[116] A. V. Neimark, F. X. Coudert, C. Triguero, A. Boutin, A. H. Fuchs, I.
Beurroies, R. Denoyel, Langmuir 2011,27, 4734.
[117] a) J. Troyano, D. Maspoch, Chem. Commun. 2023,59, 1744; b) P.
Freund, I. Senkovska, B. Zheng, V. Bon, B. Krause, G. Maurin, S.
Kaskel, Chem. Commun. 2020,56, 7411.
[118] S. Hiraide, H. Tanaka, N. Ishikawa, M. T. Miyahara, ACS Appl. Mater.
Interfaces 2017,9, 41066.
[119] A. Kondo, H. Noguchi, S. Ohnishi, H. Kajiro, A. Tohdoh, Y. Hattori,
W. C. Xu, H. Tanaka, H. Kanoh, K. Kaneko, Nano Lett. 2006,6,
2581.
[120] V. Bon, N. Kavoosi, I. Senkovska, S. Kaskel, ACS Appl. Mater. Inter-
faces 2015,7, 22292.
[121] J. Kim, W. Y. Kim, W.-S. Ahn, Fuel 2012,102, 574.
[122] M. Maliuta, I. Senkovska, V. Romaka, M. Roslova, Z. Huang, P.
Petkov, V. Bon, S. Kaskel, CCS Chem 2023,5, 2225.
[123] D. Bazer-Bachi, L. Assié, V. Lecocq, B. Harbuzaru, V. Falk, Powder
Technol. 2014,255, 52.
[124] S. Hiraide, H. Arima, H. Tanaka, M. T. Miyahara, ACS Appl Mater.
Interfaces 2021,13, 30213.
[125] M. Kriesten, J. Vargas Schmitz, J. Siegel, C. E. Smith, M. Kaspereit,
M. Hartmann, Eur. J. Inorg. Chem. 2019,2019, 4700.
[126] K. Titov, Z. Zeng, M. R. Ryder, A. K. Chaudhari, B. Civalleri, C. S.
Kelley, M. D. Frogley, G. Cinque, J. C. Tan, J. Phys. Chem. Lett. 2017,
8, 5035.
[127] J. Liu, B. Li, V. Martins, Y. Huang, Y. Song, J. Phys. Chem. C 2024,
128, 8007.
[128] H. Arima, S. Hiraide, M. T. Miyahara, S. Watanabe, ACS Appl Mater.
Interfaces 2023,15, 36975.
[129] S. Hiraide, Y. Sakanaka, H. Kajiro, S. Kawaguchi, M. T. Miyahara, H.
Tana k a , Nat. Commun. 2020,11, 3867.
[130] S. Horike, Y. Inubushi, T. Hori, T. Fukushima, S. Kitagawa, Chem. Sci.
2012,3, 116.
[131] J. F. Horstmeier, A. G. Lopez, D. W. Agar, Int. J. Greenhouse Gas Con-
trol 2016,47, 364.
[132] S. Horike, S. Shimomura, S. Kitagawa, Nat. Chem. 2009,1, 695.
[133] V. Colombo, C. Montoro, A. Maspero, G. Palmisano, N. Masciocchi,
S.Galli,E.Barea,J.A.R.Navarro,J. Am. Chem. Soc. 2012,134,
12830.
Adv. Mater. 2025, 2414724 2414724 (39 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
www.advancedsciencenews.com www.advmat.de
Irena Senkovska obtained her diploma in chemistry from Ivan Franko National University of Lviv,
Ukraine, in 1995, and earned her Ph.D. in natural sciences from Ulm University,Germany, in 2004.
Since 2005, she has been a senior scientist at the Chair of Inorganic Chemistry at Dresden University
of Technology. Her research focuses on the design, synthesis, and application of metal–organic frame-
works, mainly focusing on flexibility phenomena and adsorption processes.
Volodymyr Bon received his Ph.D. from the Institute of General and Inorganic Chemistry, National
Academy of Sciences of Ukraine in 2008 and his Venia Legendi from TU Dresden in 2024. His research
interests include crystal engineering, synthesis and crystallographic characterization of novel crys-
talline porous solids for adsorption-related applications. A particular focus is the development of
advanced in situ X-ray diffraction/adsorption characterization techniques in large scale facilities. He
is currently coordinating a subgroup focused on the synthesis of new flexible MOFs and the in-depth
study of stimuli-induced switching phenomena.
Antonia Mosberger obtained her B.Sc. in 2021 from the University of Bayreuth, working under the
supervision of Prof. Jürgen Senker.After relocating to Dresden, she completed her Master’s studies
in the research group of Prof. Stefan Kaskel, focusing on developing new metal–organic frameworks
(MOFs) for gas storage and separation applications. Since early 2024, she has continued her research
as a Ph.D. candidate, broadening her focus to include flexible MOFs for enhanced gas separation.
YutongWang obtained his Ph.D. degree (2022) in chemistry under the supervision of Prof. Daofeng
Sun from the China University of Petroleum (East China). Then he joined the group of Prof. Stefan
Kaskel at TU Dresden as a postdoctoral researcher under the support of the Humboldt Fellowship. His
research interests focus on the structural control of flexible functionalized metal–organic frameworks
for hydrocarbon adsorption and separation.
Adv. Mater. 2025, 2414724 2414724 (40 of 40) © 2025 The Author(s). Advanced Materials published by Wiley-VCH GmbH
Article
Full-text available
Due to their similar boiling points, separation of benzene and cyclohexane mixtures is among the current challenging processes faced by the petrochemical industry. As recently assessed, the soft imine‐based covalent organic framework [(TAM)(BDA)2] (COF‐300; TAM = tetrakis(4‐aminophenyl)methane, BDA = terephthaldehyde) possesses higher affinity for benzene than cyclohexane in both static conditions at 298 K and dynamic conditions in the range of 298–348 K. As shown in this contribution, in situ powder X‐ray diffraction while dosing benzene and cyclohexane vapors in the range of 0.01–4.74 bar on the narrow‐pore form of COF‐300 confirmed the coherent switchability of its framework, unveiling the progressive formation of different intermediate‐ and large‐pore forms. In addition, a basket of otherwise inaccessible key crystallochemical details—“on/off” structural‐feature changes cooperating to adsorption, primary adsorption sites, and host–guest and guest–guest interactions—was successfully retrieved. Overall, these findings allowed to shed light on the framework dynamics underneath the previously observed selectivity toward benzene over cyclohexane, completing this case of study and providing relevant information for the design of new‐generation adsorbents for this applicative context.
Article
Full-text available
Reducing carbon dioxide (CO2) emissions urgently requires the large-scale deployment of carbon-capture technologies. These technologies must separate CO2 from various sources and deliver it to different sinks1,2. The quest for optimal solutions for specific source–sink pairs is a complex, multi-objective challenge involving multiple stakeholders and depends on social, economic and regional contexts. Currently, research follows a sequential approach: chemists focus on materials design³ and engineers on optimizing processes4,5, which are then operated at a scale that impacts the economy and the environment. Assessing these impacts, such as the greenhouse gas emissions over the plant’s lifetime, is typically one of the final steps⁶. Here we introduce the PrISMa (Process-Informed design of tailor-made Sorbent Materials) platform, which integrates materials, process design, techno-economics and life-cycle assessment. We compare more than 60 case studies capturing CO2 from various sources in 5 global regions using different technologies. The platform simultaneously informs various stakeholders about the cost-effectiveness of technologies, process configurations and locations, reveals the molecular characteristics of the top-performing sorbents, and provides insights on environmental impacts, co-benefits and trade-offs. By uniting stakeholders at an early research stage, PrISMa accelerates carbon-capture technology development during this critical period as we aim for a net-zero world.
Article
Full-text available
Flexible behavior is one of the most fascinating features of hydrogen-bonded organic frameworks (HOFs), which represent an emerging class of porous materials that are self-assembled via H-bonding between organic building units. Due to their unique flexibility, HOFs can undergo structural changes or transformations in response to various stimuli (physical or chemical). Taking advantage of this unique structural feature, flexible HOFs show potential in multifunctional applications such as gas storage/separation, molecular recognition, sensing, proton conductivity, biomedicine, etc. While some other flexible porous materials have been extensively studied, the dynamic behavior of HOFs remains relatively less explored. This perspective highlights the inherent flexible properties of HOFs, discusses their different flexible behaviors, including pore size/shape changes, interpenetration/stacking manner, H-bond breaking/reconstruction, and local dynamic behavior, and highlights their potential applications. We believe that this perspective will not only contribute to HOF chemistry and materials science, but will also facilitate the ongoing extensive research on dynamic porous materials.
Article
Full-text available
The development of porous materials with flexible‐robust characteristics shows some unique advantages to target high performance for gas separation, but remains a daunting challenge to achieve so far. Herein, we report a carboxyl‐based hydrogen‐bonded organic framework (ZJU‐HOF‐8a) with flexible‐robust porosity for efficient purification of natural gas. ZJU‐HOF‐8a features a four‐fold interpenetrated structure with dia topology, wherein abundant supramolecular entanglements are formed between the adjacent subnetworks through weak intermolecular hydrogen bonds. This structural configuration could not only stabilize the whole framework to establish the permanent porosity, but also enable the framework to show some flexibility due to its weak intermolecular interactions (so‐called flexible‐robust framework). The flexible‐robust porosity of ZJU‐HOF‐8a was exclusively confirmed by gas sorption isotherms and single‐crystal X‐ray diffraction studies, showing that the flexible pore pockets can be opened by C3H8 and n‐C4H10 molecules rather by C2H6 and CH4. This leads to notably higher C3H8 and n‐C4H10 uptakes with enhanced selectivities than C2H6 over CH4 under ambient conditions, affording one of the highest n‐C4H10/CH4 selectivities. The gas‐loaded single‐crystal structures coupled with theoretical simulations reveal that the loading of n‐C4H10 can induce an obvious framework expansion along with pore pocket opening to improve n‐C4H10 uptake and selectivity, while not for C2H6 adsorption. This work suggests an effective strategy of designing flexible‐robust HOFs for improving gas separation properties.
Article
Full-text available
The advancement of separation technology is not only conducive to lowering energy consumption, but also opens an avenue to obtain the world's key resources. Metal–organic frameworks (MOFs), a novel type of porous materials, have unique advantages in separation and purification compared with traditional adsorbents and separation technologies, especially in the separation of petroleum-based platform compounds, owing to their advantages of easy structure regulation and functionalization. Significant advancements have been achieved in the separation of binary component petroleum-based platform compounds based on MOFs. However, industrial gas mixtures often exist in multiple components. In contrast to the rapid development of binary mixture separations, the separation of ternary and even multi-component mixtures is much more difficult and rarely achieved with a single material. This review briefly introduces the adsorptive separation technology and its mechanism based on MOFs. Next, recent advances in MOFs for the separation of multiple-component petroleum-based platform compounds (C2H2/C2H4/C2H6/CO2, C3H4/C3H6/C3H8, butane/butene/isobutene/1,3-butadiene, trans-2-pentene/1-pentene/isoprene, n-hexane/2-methylpentane/3-methylpentane/2,2-dimethylbutane/2,3-dimethylbutane, o-xylene/m-xylene/p-xylene/ethylbenzene, etc.) are presented to guide more complex chemical separation processes. Furthermore, inherent obstacles and future development prospects from academia to eventual industrial implementation are presented. Based on adsorptive separation technology, we are committed to exploring an alternative separation route that is energy-efficient and environmentally friendly and strives to achieve the high-efficiency separation of multi-component petroleum-based platform compounds.
Article
Full-text available
Single gas sorption experiments with the C4-hydrocarbons n-butane, iso-butane, 1-butene and iso-butene on the flexible MOFs Cu-IHMe-pw and Cu-IHEt-pw were carried out with both thermodynamic equilibrium and overall sorption kinetics. Subsequent static binary gas mixture experiments of n-butane and iso-butane unveil a complex dependence of the overall selectivity on sorption enthalpy, rate of structural transition as well as steric effects. A thermodynamic separation favoring iso-butane as well as kinetic separation favoring n-butane are possible within Cu-IHMe-pw while complete size exclusion of iso-butane is achieved in Cu-IHEt-pw. This proof-of-concept study shows that the structural flexibility offers additional levers for the precise modulation of the separation mechanisms for complex mixtures with similar chemical and physical properties with real selectivities of >10.