You are on page 1of 13

Article

Reticular Chemistry in Action: A Hydrolytically


Stable MOF Capturing Twice Its Weight in
Adsorbed Water
Sk Md Towsif Abtab, Dalal Alezi,
Prashant M. Bhatt, ..., Umer
Samin, Mohamed Nejib Hedhili,
Mohamed Eddaoudi
mohamed.eddaoudi@kaust.edu.sa

HIGHLIGHTS
Hydrolytically stable and highly
porous chromium-based MOF
with soc topology

Cr-soc-MOF-1 can capture water


twice its weight (z200 wt %)

This water uptake is maintained


over more than 100 adsorption-
desorption cycles

Potential candidate for moisture


control, dehumidification,
desalination, etc

Hydrolytically stable and highly porous Cr-soc-MOF-1 can capture 200 wt % of


water. The exceptional water uptake is maintained over more than 100 adsorption-
desorption cycles. Cr-soc-MOF-1 adsorbent can find potential use for water
adsorption-related applications such as moisture control, dehumidification,
adsorption-based desalination, and adsorption heating and cooling pumps.

Towsif Abtab et al., Chem 4, 94–105


January 11, 2018 ª 2017 Published by Elsevier
Inc.
https://doi.org/10.1016/j.chempr.2017.11.005
Article
Reticular Chemistry in Action:
A Hydrolytically Stable MOF Capturing
Twice Its Weight in Adsorbed Water
Sk Md Towsif Abtab,1 Dalal Alezi,1 Prashant M. Bhatt,1 Aleksander Shkurenko,1 Youssef Belmabkhout,1
Himanshu Aggarwal,1 Łukasz J. Weseli nski,1 Norah Alsadun,1 Umer Samin,1 Mohamed Nejib Hedhili,2
and Mohamed Eddaoudi1,3,*

SUMMARY The Bigger Picture


Hydrolytically stable adsorbents, with notable water uptake, are of prime The ability to design and construct
importance and offer great potential for many water-adsorption-related appli- advanced adsorbent materials
cations. Nevertheless, deliberate construction of tunable porous solids with with the requisite hydrolytic
high porosity and high stability remains challenging. Here, we present the suc- stability and concurrent ultra-high
cessful deployment of reticular chemistry to address this demand: we con- porosity, affording exceptional
structed Cr-soc-MOF-1, a chemically and hydrolytically stable chromium-based adsorbed water uptake with
metal-organic framework (MOF) with underlying soc topology. Prominently, distinct S-shape adsorption-
Cr-soc-MOF-1 offers the requisite thermal and chemical stability concomitant desorption isotherms, will permit
with unique adsorption properties, namely extraordinary high porosity the successful deployment of
(apparent surface area of 4,549 m2/g) affording a water vapor uptake of made-to-order adsorbents for
1.95 g/g at 70% relative humidity. This exceptional water uptake is maintained water-adsorption-related
over more than 100 adsorption-desorption cycles. Markedly, the adsorbed wa- applications such as moisture
ter can be fully desorbed by just the simple reduction of the relative humidity at control, dehumidification,
25 C. Cr-soc-MOF-1 offers great potential for use in applications pertaining to adsorption-based desalination,
water vapor control in enclosed and confined spaces and dehumidification. and adsorption heating and
cooling pumps. Successful
INTRODUCTION positioning of these applications
The development and deployment of energy-saving technologies offer great poten- offers great potential to address
tial to sustain and address the ever-increasing worldwide energy demand, and sub- various United Nations
sequently support reduction in the carbon footprint. Credibly, continuous globaliza- Sustainable Development Goals
tion of a better quality of living leads to the consumption of excessive amounts of pertaining to affordable clean
energy, such as for indoor air-conditioning in regions with extremely high or low water, prosperous health, and
temperatures and desalination of seawater in desert areas. From the indoor air-con- reliable and sustainable energy-
ditioning perspective, adsorptive heat transformation applications, provided efficient air-conditioning
optimal water vapor (H2O) adsorbent is available, are considered to be highly en- technologies. Here, we unveil a
ergy-efficient and environmentally friendly technologies in contrast to the conven- hydrolytically stable and highly
tional compression-decompression systems.1,2 Air-conditioning devices based on porous MOF, Cr-soc-MOF-1, as a
thermally driven adsorption heat pumps (AHP) or desiccant cooling systems (DCS) unique adsorbent material for
are considered to operate with moderate consumption of electric power.3 The work- water adsorption applications.
ing principle of both AHP and DCS systems is governed by reversible exothermic Prominently, Cr-soc-MOF-1
adsorption and endothermic desorption of water in micro- or mesoporous solid ma- captures twice its weight and can
terials. Porous materials with distinct water sorption properties and remarkable wa- address many acute societal and
ter uptake are ideal for humidity control in confined and poorly ventilated spaces.4 industrial challenges.
Very high (>65%) or very low (<25%) relative humidity (RH) levels can have adverse
effects on human health and comfort levels. Materials with the potential to adsorb
a large amount of water vapor at undesired high humidity and subsequently release

94 Chem 4, 94–105, January 11, 2018 ª 2017 Published by Elsevier Inc.


it when the humidity level drops below the recommended limit are well positioned to
address the essential requirements for humidity control. Principally, energy effi-
ciency, the working humidity range of AHP and DCS and the required performance
of a humidity controller are directly correlated to the properties of water vapor ad-
sorbents.5,6 Extensive research has been devoted to the design, synthesis, and
development of new adsorbents with a water uptake capacity exceeding that of ex-
isting commercial materials (e.g., silica gels or zeolites) and offering relatively milder
regeneration conditions.

Metal-organic frameworks (MOFs), an emerging class of crystalline porous materials,


are considered to offer great potential in addressing many enduring challenges per-
taining to energy and environmental sustainability.7–15 Nevertheless, despite the
attractive features of MOFs, such as their extraordinary porosity and high degree
of structural tunability and stability,16–19 the degradation of some earlier examples
of highly porous MOFs in a water (moisture)-containing environment hindered their
industrial implementation. Remarkably, advances in MOF chemistry have permitted
the deployment of several strategies for the synthesis of water-stable MOFs, paving
the way for water-sorbent candidates for water adsorption-related applications.20–26
From a qualitative point of view, the adsorption properties of MOFs are obviously
quite diverse in terms of water uptake capacity and the associated relative pressure
at which the pore filling occurs. Certainly, hydrolytically stable porous materials of-
fering remarkable pore volumes are expected to exhibit large H2O adsorption ca-
pacity. The quest for hydrolytically stable and recyclable MOFs with superior total
water uptake remains a focal point of intensive research in MOF chemistry.

In this context, zirconium-based MOFs have attracted considerable attention


because of their exceptional chemical and thermal stability. Nonetheless, although
numerous high-surface-area Zr-MOFs are chemically stable in water,27–29 their struc-
ture cannot sustain the desorption of water as the open framework collapses under
the influence of capillary forces of desorbing water.30 Chromium-based MOFs are
another promising class of highly porous MOFs shown to offer remarkable water
stability. Particularly, Cr-MIL-101, the material built up from assembly of super tetra-
hedron building units based on trinuclear inorganic chromium clusters with a linear
ditopic ligand, is one of the most prominent MOFs in terms of structural robustness,
water stability, and excellent water adsorption properties with a large water uptake
capacity (>1.0 g/g).31 In view of that, Cr(III)-based clusters, as a result of their kinetic
inertness and very slow kinetics of ligand exchange of Cr(III),32 are positioned as
strong candidates to serve as the requisite robust inorganic nodes for the construc-
tion of hydrolytically stable MOFs with unparalleled water uptake. Accordingly, we
directed our research efforts to the design and construction of Cr-based MOFs
with ultra-high surface area and plausibly offering superior water uptake capacity.
1KingAbdullah University of Science and
Recently, our group reported the first aluminum MOF with underlying soc topology, Technology, Division of Physical Sciences and
Al-soc-MOF-1, obtained by the judicious selection of an expanded rectangular Engineering, Advanced Membranes & Porous
Materials Center, Functional Materials Design,
tetracarboxylate organic linker that facilitated the in situ formation of the desired
Discovery & Development Research Group,
inorganic oxo-centered trinuclear Al(III) cluster.33 Notably, Al-soc-MOF-1 offers Thuwal 23955-6900, Kingdom of Saudi Arabia
exceptional porosity characteristics and adsorption properties, namely a micropore 2King Abdullah University of Science and
volume of 2.3 cm3/g and an apparent BET surface area of 5,500 m2/g accessed Technology, Imaging and Characterization
Laboratory, Thuwal 23955-6900, Kingdom of
through a conventional activation protocol (solvent exchange and vacuum). Saudi Arabia
3Lead Contact
The exceptional pore volume of Al-soc-MOF-1 is solely associated with its micropo- *Correspondence:
rous architecture, composed of well-defined cages and channels with an average mohamed.eddaoudi@kaust.edu.sa
pore size of 14.3 Å, and offering superior gas storage capacity (but relatively low https://doi.org/10.1016/j.chempr.2017.11.005

Chem 4, 94–105, January 11, 2018 95


hydrolytic stability after 60% RH) inspired us to target its isostructural Cr analog by
using reticular chemistry and subsequently explore the potential of Cr-soc-MOF-1
in water-vapor-adsorption-related applications. Nevertheless, our countless at-
tempts at the direct synthesis of single-crystal quality Cr-soc-MOFs were unsuccess-
ful, plausibly because of the relatively high inertness of Cr(III) toward carboxylate
bonding, imposing the need to use harsher synthetic conditions, as exemplified
by the limited number of crystalline Cr-MOFs obtained directly in the open litera-
ture.32,34–36 Post-synthetic metal cluster metathesis from a template MOF with a
known structure offers the opportunity to obtain isostructural MOFs that cannot
be synthesized directly by conventional methods.37–41

Here, we report single-crystal to single-crystal transformation of a [Fe3(m3-O)(O2C)6]


molecular building block (MBB) in Fe-soc-MOF-1 into [Cr3(m3-O)(O2C)6] MBB to
form isostructural Cr-soc-MOF-1 with near-complete exchange as revealed by the
combination of inductively coupled plasma optical emission spectrometry (ICP-
OES) and X-ray photoelectron spectroscopy (XPS) analysis. Extensive water adsorp-
tion studies were carried out on the isostructural Cr-soc-MOF-1, which shows a
higher water adsorption capacity than all existing water vapor adsorbents reported
so far. Particularly, Cr-soc-MOF-1 preserves its optimal porosity even after 100 water
adsorption-desorption cycles, as demonstrated by the unaltered nitrogen adsorp-
tion and full water adsorption-desorption isotherms. Comparing the working capac-
ities in the RH ranges 35%–65%, 25%–75%, and 25%–85%, associated with the work-
ing ranges for moisture control systems, revealed that Cr-soc-MOF-1 exhibits better
performance than all solid-state porous materials. Single X-ray diffraction studies on
the hydrated Cr-soc-MOF-1 unveiled the localization of adsorbed water molecules
within the pore system of Cr-soc-MOF-1, as demonstrated by the resolution of the
highly ordered and well-defined water clusters formed in the cubic cages.

RESULTS AND DISCUSSION


Synthesis and Characterization of Cr-soc-MOF-1
In our initial effort to isolate the isostructural chromium-soc-MOF analog, countless
attempts to isolate the reaction conditions that consistently allow the in situ forma-
tion of the desired trinuclear chromium (III) MBB were unsuccessful. Accordingly, we
directed our research efforts to an alternative route on the basis of transmetalation of
the synthesized Fe-soc-MOF into the required Cr-soc-MOF. A solvothermal reaction
between the tetratopic ligand 3,300 ,5,500 -tetrakis(4-carboxyphenyl)-p-terphenyl
(H4TCPT) and FeCl3,6H2O in acidic solution containing a mixture of N,N0 -dimethyl-
formamide (DMF) and acetonitrile (CH3CN) afforded yellow-orange cube-shaped
homogeneous crystals of Fe-soc-MOF-1 (1). Single-crystal X-ray diffraction (SCXRD)
analysis and the powder X-ray diffraction (PXRD) pattern (Figure S1) confirmed that 1
is isostructural with the reported Al-soc-MOF-1 with the formula [Fe3(m3-O)
(H2O)2(TCPT)1.5Cl] (Data S1 and Table S1). Subsequently, the reaction of the yel-
low-orange colored crystals of Fe-soc-MOF-1 with CrCl2 in DMF under an inert at-
mosphere at 115 C for a period of 24 hr afforded dark green Cr-soc-MOF-1 single
crystals (2) with nearly complete exchange of iron by chromium as supported by
ICP-OES analysis, which revealed that Cr-soc-MOF-1 contained 98% chromium.
Furthermore, the single-crystal structure of Cr-soc-MOF-1 was analyzed by SCXRD
studies (Data S2 and Table S2). As envisioned, the structure of Cr-soc-MOF-1
showed well-defined 1D infinite channels and cubic-shaped cages constructed by
six TCPT4 ligands, which occupy the faces of the cage, and eight inorganic trinu-
clear Cr(III) clusters located on the vertices of the cuboidal cage (Figure 1). The struc-
ture of the Cr-soc-MOF-1 analog differed slightly from that of Al-soc-MOF-1 with

96 Chem 4, 94–105, January 11, 2018


Figure 1. Select Fragments from the Crystal Structure of Cr-soc-MOF-1
(A) The m 3 -oxygen-centered trinuclear Cr(III) carboxylate clusters and the deprotonated organic linker (TCPT 4).
(B) Representation of the well-defined channels and cages found in Cr-soc-MOF-1.
Color code: C, gray; O, red; Cl, pink; Cr, green. Hydrogen atoms are omitted for clarity.

regard to the size of the 1D channels and the cages: the channels in Cr-soc-MOF-1
are wider with an estimated dimension of 17 Å taking van der Waals (vdW) radii into
consideration, which is approximately at the borderline of micro- and mesoporous
materials. This is also reflected in the shape of the nitrogen adsorption isotherm,
as opposed to the typical type I isotherm for Al analogs (Figure S7). The size of
the cage in the case of Cr-soc-MOF-1 is slightly smaller than the corresponding
size of the Al analog.

The oxidation state of all chromium ions in Cr-soc-MOF-1 is (+III), as verified by XPS
studies (Figure 2B). The charge of the cationic trinuclear cluster, [Cr3(m3-O) (O2C)6],
is balanced by one Cl anion per cluster as supported by XPS data (Figure S2). The
efficiency of the metathesis was also confirmed by energy-dispersive X-ray spectros-
copy (EDS) elemental mapping analysis, which indicated that Cl, Cr, and O were uni-
formly distributed on the crystal surface (Figure 2A).

The purity of compound 2 was confirmed by similarities between the experimental


and calculated PXRD patterns (Figure 2C). The Cr-soc-MOF-1 crystals, washed
with DMF, were subjected to exchange with acetonitrile and acetone for 24 hr and
then used for further experiments. The high thermal stability of Cr-soc-MOF-1 was
confirmed by variable-temperature PXRD studies (Figure S3). Delightfully, variable
humidity PXRD analysis (Figure S4) showed that Cr-soc-MOF-1 is an exceptionally
stable material on prolonged exposure to moisture. To the best of our knowledge,
this is a rare highly porous MOF with high stability in the presence of moisture. As a
first test of water stability, Cr-soc-MOF-1 was soaked in liquid water at room

Chem 4, 94–105, January 11, 2018 97


Figure 2. Structural Characterization of Cr-soc-MOF-1
(A) Energy-dispersive X-ray spectroscopy elemental mapping analysis of Cr-soc-MOF-1.
(B) High-resolution X-ray photoelectron spectroscopy spectrum of the Cr 2p core level of the Cr-soc-MOF-1 sample, the binding energies of the
components of the Cr 2p doublet, and their corresponding satellites are characteristic of the Cr 3+ oxidation state of chromium.
(C) Experimental and calculated PXRD patterns for Cr-soc-MOF-1, indicating the phase purity of the sample.
(D) Nitrogen adsorption isotherm at 77 K on Cr-soc-MOF-1.

temperature for 1 day without any alteration in structure or performance. This un-
precedented result was demonstrated by the unchanged PXRD patterns of the pris-
tine sample and the after soaking the crystals in liquid water (Figure S5).

As in the case of the Al-soc-MOF-1, the guest solvents in the pore system of Cr-soc-
MOF-1 were easily and fully removed by a traditional activation approach (vacuum
and heating). The optimal porosity of Cr-soc-MOF-1 was obtained by heating at
120 C under a vacuum as supported by nitrogen and argon sorption isotherms (Fig-
ures 2D and S6). The calculated apparent BET surface area and pore volume were
4,549 m2/g and 2.1 cm3/g, respectively, in good agreement with the theoretical
values (PVtheo 2.2 cm3/g). Unlike Fe-soc-MOF-1, exchange of Fe by Cr led to the
full access to the pore volume similar to the parent Al analog (Figure S7). Exploration
of oxygen and methane adsorption at 90 K and 112 K (Figures S8 and S9) confirmed
attainment of the optimal porosity for Cr-soc-MOF-1. Similar to Al-soc-MOF-1 (the
current benchmark material for methane gravimetric uptake and delivery), CH4

98 Chem 4, 94–105, January 11, 2018


Figure 3. Water Adsorption Study of Cr-soc-MOF-1
(A) Water adsorption (solid spheres) and desorption (empty circles) isotherms at 298 K for activated
Cr-soc-MOF-1.
(B) 100 cycles of water uptake profile versus relative humidity of the Cr-soc-MOF-1 at 298 K.

storage studies showed that Cr-soc-MOF-1 exhibits one of the highest CH4 total
volumetric and gravimetric uptakes (Figure S10). Analysis of CH4 sorption data
showed that it exhibits a volumetric and gravimetric working capacity of
187.1 cm3/cm3 and 37.1 wt %, respectively, at a pressure range of 5–80 bar and
298 K.

Water Adsorption Studies of Cr-soc-MOF-1


In order to develop a suitable porous material for water adsorption-related applica-
tions, three criteria have been considered and targeted: (1) pore filling or condensa-
tion of water into the pore system of the porous solid must exhibit a steep uptake
isotherm at a specific RH, depending on the nature of the targeted application, (2)
a high water uptake capacity for the requisite maximum delivery of water and a spon-
taneous adsorption-desorption process for the desired energy efficiency, and (3) a
highly reproducible cycling performance of the material toward water adsorption-
desorption, an essential criterion for prospective deployment in industrial
applications.42

On the basis of these criteria and considering the combined exceptional porosity
and extremely high stability of Cr-soc-MOF-1, the water adsorption properties
were extensively evaluated experimentally at ambient temperature (Figure 3). As de-
picted in the water adsorption isotherm (Figure 3A), the amount of water adsorbed

Chem 4, 94–105, January 11, 2018 99


gradually increases with increasing RH up to 55%, followed by a steep uptake of wa-
ter between 60% and 75% RH. Evidently, Cr-soc-MOF-1 offers an exceptional
maximum water uptake of 1.95 g (195 wt %) of adsorbed water per gram of sorbent
at 75% RH, with an S-shaped-like form of the adsorption isotherm.

To the best of our knowledge, this is the highest value of water adsorbed at saturation
among all MOFs, carbons, and inorganic materials. The pore volume calculated from
the water vapor adsorption at 298 K and at 0.95 p/p0 (ca. 1.95 cm3/g) is in excellent
agreement with the calculated and experimental pore volume from nitrogen adsorption
at 77 K, an unprecedented feature for MOFs with such high porosity and attesting to the
evident outstanding stability of Cr-soc-MOF-1 to water vapor. Interestingly, only one
such example of high water-cycling stability was reported by Kitagawa and co-workers
for Cr-MIL-100-F,43 but with gradual uptake and much less total adsorption uptake of
water vapor (0.8 g/g versus 1.95 g/g for Cr-soc-MOF-1).

Markedly, the association of this extremely high water uptake with a distinct
S-shaped water adsorption isotherm at 298 K positions Cr-soc-MOF-1 as a suitable
adsorbent candidate for moisture control applications. Accordingly, we opted to
check the durability and recyclability of Cr-soc-MOF-1 by performing multiple water
adsorption and desorption tests at 298 K, by simply altering the moisture levels be-
tween 25% and 85% RH. Delightfully, the water uptake remained unaltered
throughout 100 adsorption-desorption cycles (Figure 3B) performed between 25%
and 85% RH. We reviewed the N2 sorption isotherm and the water adsorption-
desorption isotherms of the sample after 100 water adsorption-desorption cycles
and confirmed that the uptake and the shape of the N2 and water isotherms were
preserved (Figures S11 and S12). Figure S13 shows that the characteristics of the wa-
ter adsorption isotherm for Cr-soc-MOF-1 are preserved, independently of the car-
rier gas used (N2 or air).

Further, we evaluated the water adsorption performance of Cr-soc-MOF-1 in com-


parison with the best-performing existing materials for water adsorption and indoor
moisture control applications. Interestingly, the water uptake plotted for selected
best materials against their associated pore volume determined using the nitrogen
isotherm at 77 K (Figure 4 and Table S4)42–47 showed that Cr-soc-MOF-1 largely out-
performs all the materials reported to date in terms of total water adsorption capac-
ity at ambient conditions. Noticeably, the pore volume determined from the water
vapor adsorption isotherm for Cr-soc-MOF-1 matches perfectly the corresponding
pore volume derived theoretically from the crystal structure and experimentally us-
ing the nitrogen adsorption isotherm (77 K). Analysis of the water adsorption equi-
librium working capacity (from equilibrium water vapor adsorption isotherms) for
different RH ranges (35%–65%, 25%–75%, and 25%–85%), relevant to indoor mois-
ture control at room temperature, revealed that Cr-soc-MOF-1 outperformed all the
materials reported so far for water adsorption (Figure 5). Particularly, this compara-
tive analysis showed that Cr-soc-MOF-1 displays a 300% higher working capacity at
room temperature than the current benchmark material, Y-shp-MOF-5,4 for moisture
control in the 25%–85% RH range.

Evidently, the exceptional water adsorption features of Cr-soc-MOF-1 are a direct


result of combining the requisite structural characteristics in a single adsorbent,
namely hydrolytically stability, ultra-high micropore volume, and the proper pore
system (shape, size, and functionality), conferring the observed energy barrier dur-
ing adsorption and desorption steps as reflected in the S-shaped adsorption
isotherm at room temperature.

100 Chem 4, 94–105, January 11, 2018


Figure 4. Correlation between Pore Volume and Water Uptake Capacity for Cr-soc-MOF-1 and
the Best-Performing Materials

In order to understand and gain better insights on the observed adsorption-desorp-


tion behavior, we further analyzed the crystal structure of hydrated Cr-soc-MOF-1 (3)
(Data S3 and Table S3) and appositely localized the guest water molecules within the
MOF pore system (Figure 6A). Closer examination of the adsorbed guest molecules
revealed significant disorder of the water molecules in the channels and well-or-
dered 114 water molecule clusters within the cages (Figure 6B), indicating that the
water adsorption is plausibly governed by water molecule nucleation within the
pore system, corroborating the noted energy barrier during the adsorption step, fol-
lowed by the formation of the observed large water clusters. Purportedly, the initial
water adsorption is promoted by the formation of eight water heptamers, located at
the vertices of the cubic cage (Figure 6C, red), with O,,,O distances in the range of
2.808(9)–2.829(8) Å and further connected by single water molecules at the edges
(Figure 6C, yellow) with hydrogen bonds of 2.709(8) and 2.803(9) Å. Additional ad-
sorbed water molecules form water hexamers (2.78(2)–2.84(1) Å) in a boat conforma-
tion, centered at each face of the cubic cages (Figure 6C, blue), and interact more
strongly with the heptamers (2.826(9) Å) than with the water molecules at the
edges (2.89(2) and 2.94(3) Å). The aforementioned 104 adsorbed water molecules
(8 3 heptamers in the corner, 12 3 1 water at the edge, 6 3 hexamers at the face)
constitute the first layer of the adsorbed water molecules within the cubic cage.
The second adsorbed water layer encompasses eight additional water molecules,
also arranged in cubic fashion (Figure 6C, green), and bridged with neighboring hex-
amers (2.74(1) Å). The third water layer consists of a disordered water dimer (Fig-
ure 6C, pink), where the interaction between the two water molecules is relatively
stronger than with the previous neighboring layer (2.79(2) versus 2.92(1) Å). The
self-assembled water cluster, consisting of 114 water molecules sustained by
hydrogen bonds, is further expanded via additional hydrogen bonds with the other
adsorbed water molecules within the channels through the shared cage-channel
windows.

Notably, the presence of a well-defined water assembly and/or network within the
Cr-soc-MOF-1 pore system supports the distinctive water adsorption and desorp-
tion properties, as reflected by the noticeably open hysteresis indicating that the

Chem 4, 94–105, January 11, 2018 101


Figure 5. Working Capacity at Different Relative Humidity Ranges (35%–65%, 25%–75%, and
25%–85%) Relevant to Indoor Moisture Control at Room Temperature

water desorption process entails the need to surmount the energy barrier required
for the dissociation of the hosted water hydrogen-bonded network. Significantly, the
relatively higher energy of dissociation during the water desorption step inflicts the
observed shift in the water desorption branch to a relatively lower pressure, a unique
behavior not common for microporous materials. Evidently, the associated energy
barriers to both the water adsorption and desorption steps on Cr-soc-MOF-1 af-
forded the resulting interesting S-shaped adsorption-desorption isotherms as
recently observed with Y-shp-MOF-5 adsorbent.4

In summary, we have successfully synthesized the first Cr-MOF with underlying soc
topology, Cr-soc-MOF-1, via a post-synthetic modification approach. The resultant
Cr-soc-MOF-1 exhibits a rare combination of high porosity, high thermal and chem-
ical stability, and high water vapor adsorption capacity. It preserves its optimal
porosity even after soaking in liquid water, a feature rarely observed for highly micro-
porous MOFs, with an apparent surface area and pore volume around 4,500 m2/g
and 2.1 cm3/g, respectively. Furthermore, Cr-soc-MOF-1 is a hydrolytically stable
MOF with better water loading and uptake (1.95 g/g) than other MOFs reported
so far. To the best of our knowledge, the reported Cr-soc-MOF-1 outperforms the
existing MOFs in terms of total and working capacity, reversibility, and cyclic perfor-
mance, particularly for indoor moisture control. The occurrence of S-shaped water
adsorption-desorption isotherms is mainly governed by the formation and dissocia-
tion of water clusters during the adsorption and desorption steps, respectively.
Another salient feature of Cr-soc-MOF-1 is that the adsorbed water molecules can
be completely desorbed at room temperature just by the simple reduction of the
RH without heating, suggesting an energy-efficient and cost-effective recycling pro-
cess. Practically, Cr-soc-MOF-1 meets the required criteria for its conceivable
deployment in real applications such as water vapor removal in enclosed and
confined spaces and dehumidification.

EXPERIMENTAL PROCEDURES
Synthesis of Fe-soc-MOF-1 (1)
A solution of FeCl3$6H2O in DMF (0.1 M, 0.3 mL, 0.03 mmol) was added to a mixture
of H4TCPT (7.1 mg, 0.01 mmol), DMF (1 mL), and acetonitrile (1 mL) in a 20 mL

102 Chem 4, 94–105, January 11, 2018


Figure 6. Crystal Structure of the Hydrated Cr-soc-MOF-1 (3)
(A) Packing diagram along the c axis with a hydrogen bonding network shown.
(B) The ordered 114 water molecules cluster in the cubic cage.
(C) The cluster decomposition: water heptamers at the vertices of the cube (red), a single water molecule at the edges (yellow), hexamers at the faces
(blue), inner cube (green), and a dimer (pink). Hydrogen atoms are omitted for clarity.

scintillation vial. Dilute nitric acid solution in DMF (3.5 M, 1.5 mL) was added to the
reaction mixture followed by sonication. The clear orange-yellow solution was sub-
sequently placed in a preheated oven at 115 C for 3 days to give pure small orange-
yellow cube-shaped crystals (Figure S14). The crystals were washed four to five
times with DMF. Fourier transform infrared spectroscopy (FTIR) (4,000–650 cm1):
3,360 (br), 2,971 (w), 1,652 (s), 1,593 (s), 1,400 (s), 1,254 (w), 1,185 (w), 1,092(m),
1,046 (s), 1,016 (w), 859 (w), 836 (w), 782 (s), 701 (w).

Synthesis of Cr-soc-MOF-1 (2)


The Fe-soc-MOF-1 crystals were washed with acetonitrile quickly three to four times
to remove surface DMF. Finally, washing was accomplished with acetone one to two
times, and the compound was transferred to an inert atmosphere glove box. Inside
the glove box, about 25 mg of Fe-soc-MOF-1 was weighed in a 20 mL scintillation
vial. In another vial, 150 mg of CrCl2 was dissolved in 3 mL of DMF, resulting in a
light-sky-blue clear solution. The solution was then added carefully to the former
vial. An immediate color change to green was observed. The vial was capped, taken
out of the glovebox, and incubated at 115 C for 20 hr. The vial was then allowed to
cool to room temperature. The dark-green supernatant solution was removed, and
the resulting green crystals of Cr-soc-MOF-1 (Figure S14) were washed four to five
times with DMF. FTIR (4,000–650 cm1): 3,334 (br), 2,971 (w), 1,654 (s), 1,598 (s),
1,553 (w), 1,400 (s), 1,253 (w), 1,186 (w), 1,088 (m), 1,045 (s), 1,016 (w), 879 (w),
861 (w), 780 (s), 701 (m).

Water Sorption Experiments


Water sorption experiments were carried out with a VTI-SA vapor sorption analyzer
from TA Instruments (New Castle, DE, United States). The water vapor activity was
controlled automatically by mixing wet vapor feed with a dry N2 line; hence, N2
acts as a carrier gas for water vapor. The sample ‘‘dry mass’’ was measured under
N2 and was at equilibrium before water vapor was introduced into the chamber.
The adsorption isotherms, obtained at equilibrium, were collected within a range
of 0%–90% RH. The acetone-exchanged sample was activated at 120 C before
the sorption experiment for 8 hr. The maximum equilibrium time for each RH was

Chem 4, 94–105, January 11, 2018 103


maintained at 2 hr. The cycles were carried out between 25% and 85% RH with a
maximum equilibrium time of 3 hr.

DATA AND SOFTWARE AVAILABILITY


The single-crystal X-ray diffraction data of Fe-soc-MOF-1 (1), Cr-soc-MOF-1 (2), and
hydrated Cr-soc-MOF-1 (3) have been deposited in the Cambridge Crystallographic
Data Centre under accession numbers CCDC: 1542715, 1542714, and 1566535,
respectively.

SUPPLEMENTAL INFORMATION
Supplemental Information includes Supplemental Experimental Procedures, 14 fig-
ures, 4 tables, 1 scheme, and 3 data files and can be found with this article online at
https://doi.org/10.1016/j.chempr.2017.11.005.

ACKNOWLEDGMENTS
Research reported in this publication was fully supported by the King Abdullah Uni-
versity of Science and Technology.

AUTHOR CONTRIBUTIONS
S.M.T.A., D.A., and M.E. conceptualized the design and the construction of the re-
ported MOF materials; S.M.T.A. and D.A. carried out the synthesis and characteriza-
tion of the materials; Ł.J.W. and M.E. conceptualized the design and the synthesis of
the ligand; A.S. and H.A. conducted the crystallographic experiments; A.S. and M.E.
interpreted and analyzed the crystallographic experiments; N.A., U.S., and S.M.T.A.
conducted and interpreted ICP-OES and EDS data; M.N.H. conducted and analyzed
XPS data; P.M.B., H.A., M.E., and Y.B. contributed to conceptualizing, designing,
conducting, and interpreting adsorption experiments; S.M.T.A., D.A., P.M.B.,
A.S., Y.B., and M.E. wrote the manuscript; and Y.B. and M.E. supervised the whole
work. All authors proofread, commented on, and approved the final version of this
manuscript for submission.

Received: August 16, 2017


Revised: September 18, 2017
Accepted: November 13, 2017
Published: January 11, 2018

REFERENCES AND NOTES


1. Henninger, S.K., Schmidt, F.P., and Henning, 5. Saha, B.B., Akisawa, A., and Kashiwagi, T. 9. McKinlay, A.C., Morris, R.E., Horcajada, P.,
H.M. (2010). Water adsorption characteristics (2001). Solar/waste heat driven two-stage Férey, G., Gref, R., Couvreur, P., and Serre, C.
of novel materials for heat transformation adsorption chiller: the prototype. Renew. (2010). BioMOFs: metal–organic frameworks
applications. Appl. Therm. Eng. 30, 1692–1702. Energy 23, 93–101. for biological and medical applications.
Angew. Chem. Int. Ed. 49, 6260–6266.
2. Jänchen, J., and Stach, H. (2012). Adsorption 6. Henninger, S.K., Jeremias, F., Kummer, H., and
properties of porous materials for solar thermal Janiak, C. (2012). MOFs for use in adsorption 10. Zhou, H.C., Long, J.R., and Yaghi, O.M. (2012).
energy storage and heat pump applications. heat pump processes. Eur. J. Inorg. Chem. Introduction to metal–organic frameworks.
Energy Procedia 30, 289–293. 2012, 2625–2634. Chem. Rev. 112, 673–674.

11. Shekhah, O., Belmabkhout, Y., Chen, Z.,


3. Henninger, S.K., Habib, H.A., and Janiak, C. 7. Furukawa, H., Cordova, K.E., O’Keeffe, M., and Guillerm, V., Cairns, A., Adil, K., and Eddaoudi,
(2009). MOFs as adsorbents for low Yaghi, O.M. (2013). The chemistry and M. (2014). Made-to-order metal-organic
temperature heating and cooling applications. applications of metal-organic frameworks. frameworks for trace carbon dioxide removal
J. Am. Chem. Soc. 131, 2776–2777. Science 341, 1230444. and air capture. Nat. Commun. 5, 4228.
4. AbdulHalim, R.G., Bhatt, P.M., Belmabkhout, 8. Cook, T.R., Zheng, Y.R., and Stang, P.J. (2013). 12. Nugent, P., Belmabkhout, Y., Burd, S.D.,
Y., Shkurenko, A., Adil, K., Barbour, L.J., and Metal–organic frameworks and self-assembled Cairns, A.J., Luebke, R., Forrest, K., Pham, T.,
Eddaoudi, M. (2017). A fine-tuned metal– supramolecular coordination complexes: Ma, S., Space, B., Wojtas, L., et al. (2013).
organic framework for autonomous indoor comparing and contrasting the design, Porous materials with optimal adsorption
moisture control. J. Am. Chem. Soc. 139, synthesis, and functionality of metal–organic thermodynamics and kinetics for CO2
10715–10722. materials. Chem. Rev. 113, 734–777. separation. Nature 495, 80–84.

104 Chem 4, 94–105, January 11, 2018


13. Belmabkhout, Y., Mouttaki, H., Eubank, J.F., (2015). Ultra-tuning of the rare-earth fcu-MOF nanoporous chromium(III)-based solids: MIL-53
Guillerm, V., and Eddaoudi, M. (2014). Effect of aperture size for selective molecular exclusion or CrIII(OH)${O2CC6H4CO2}$
pendant isophthalic acid moieties on the of branched paraffins. Angew. Chem. Int. Ed. {HO2CC6H4CO2H}x$H2Oy. J. Am. Chem.
adsorption properties of light hydrocarbons in 54, 14353–14358. Soc. 124, 13519–13526.
HKUST-1-like tbo-MOFs: application to
methane purification and storage. RSC Adv. 4, 25. Cadiau, A., Adil, K., Bhatt, P.M., Belmabkhout, 36. Férey, G., Mellot-Draznieks, C., Serre, C.,
63855–63859. Y., and Eddaoudi, M. (2016). A metal-organic Millange, F., Dutour, J., Surblé, S., and
framework–based splitter for separating Margiolaki, I. (2005). A chromium
14. Belmabkhout, Y., Pillai, R.S., Alezi, D., Shekhah, propylene from propane. Science 353, terephthalate-based solid with unusually large
O., Bhatt, P.M., Chen, Z., Adil, K., Vaesen, S., 137–140. pore volumes and surface area. Science 309,
De Weireld, G., Pang, M., et al. (2017). Metal- 2040–2042.
organic frameworks to satisfy gas upgrading 26. Kim, H., Yang, S., Rao, S.R., Narayanan, S.,
demands: fine-tuning the soc-MOF platform Kapustin, E.A., Furukawa, H., Umans, A.S., 37. Zhang, Z., Zhang, L., Wojtas, L., Nugent, P.,
for the operative removal of H2S. J. Mater. Yaghi, O.M., and Wang, E.N. (2017). Water Eddaoudi, M., and Zaworotko, M.J. (2012).
Chem. A 5, 3293–3303. harvesting from air with metal-organic Templated synthesis, postsynthetic metal
frameworks powered by natural sunlight. exchange, and properties of a porphyrin-
15. Liu, Y., Eubank, J.F., Cairns, A.J., Eckert, J., Science 356, 430–434. encapsulating metal–organic material. J. Am.
Kravtsov, V.C., Luebke, R., and Eddaoudi, M. Chem. Soc. 134, 924–927.
(2007). Assembly of metal–organic frameworks 27. Deria, P., Gómez-Gualdrón, D.A., Bury, W.,
(MOFs) based on indium-trimer building Schaef, H.T., Wang, T.C., Thallapally, P.K., 38. Brozek, C.K., and Dinca, M. (2014). Cation
blocks: a porous MOF with soc topology and Sarjeant, A.A., Snurr, R.Q., Hupp, J.T., and exchange at the secondary building units of
high hydrogen storage. Angew. Chem. Int. Ed. Farha, O.K. (2015). Ultraporous, water stable, metal-organic frameworks. Chem. Soc. Rev. 43,
46, 3278–3283. and breathing Zirconium-based metal–organic 5456–5467.
frameworks with ftw topology. J. Am. Chem.
16. Cadiau, A., Belmabkhout, Y., Adil, K., Bhatt, Soc. 137, 13183–13190. 39. Liu, T.-F., Zou, L., Feng, D., Chen, Y.-P.,
P.M., Pillai, R.S., Shkurenko, A., Martineau- Fordham, S., Wang, X., Liu, Y., and Zhou, H.-C.
Corcos, C., Maurin, G., and Eddaoudi, M. 28. Gutov, O.V., Bury, W., Gomez-Gualdron, D.A., (2014). Stepwise synthesis of robust metal–
(2017). Hydrolytically stable fluorinated metal- Krungleviciute, V., Fairen-Jimenez, D., organic frameworks via postsynthetic
organic frameworks for energy-efficient Mondloch, J.E., Sarjeant, A.A., Al-Juaid, S.S., metathesis and oxidation of metal nodes in a
dehydration. Science 356, 731–735. Snurr, R.Q., Hupp, J.T., et al. (2014). Water- single-crystal to single-crystal transformation.
stable zirconium-based metal–organic J. Am. Chem. Soc. 136, 7813–7816.
17. Bhatt, P.M., Belmabkhout, Y., Assen, A.H., framework material with high-surface area and
Weseli nski, Ł.J., Jiang, H., Cadiau, A., Xue, gas-storage capacities. Chemistry 20, 12389– 40. Lian, X., Feng, D., Chen, Y.-P., Liu, T.-F.,
D.-X., and Eddaoudi, M. (2017). Isoreticular rare 12393. Wang, X., and Zhou, H.-C. (2015). The
earth fcu-MOFs for the selective removal of H2S preparation of an ultrastable mesoporous
from CO2 containing gases. Chem. Eng. J. 324, 29. Feng, D., Gu, Z.-Y., Li, J.-R., Jiang, H.-L., Wei, Z., Cr(III)-MOF via reductive labilization. Chem.
392–396. and Zhou, H.-C. (2012). Zirconium- Sci. 6, 7044–7048.
Metalloporphyrin PCN-222: mesoporous
18. Alezi, D., Spanopoulos, I., Tsangarakis, C., metal–organic frameworks with ultrahigh 41. Lalonde, M., Bury, W., Karagiaridi, O., Brown,
Shkurenko, A., Adil, K., Belmabkhout, Y., stability as biomimetic catalysts. Angew. Z., Hupp, J.T., and Farha, O.K. (2013).
O0 Keeffe, M., Eddaoudi, M., and Trikalitis, P.N. Chem. Int. Ed. 124, 10453–10456. Transmetalation: routes to metal exchange
(2016). Reticular chemistry at its best: directed within metal-organic frameworks. J. Mater.
assembly of hexagonal building units into the 30. Mondloch, J.E., Katz, M.J., Planas, N., Chem. A 1, 5453–5468.
awaited metal-organic framework with the Semrouni, D., Gagliardi, L., Hupp, J.T., and
intricate polybenzene topology, pbz-MOF. Farha, O.K. (2014). Are Zr6-based MOFs water 42. Furukawa, H., Gándara, F., Zhang, Y.-B., Jiang,
J. Am. Chem. Soc. 138, 12767–12770. stable? Linker hydrolysis vs. capillary-force- J., Queen, W.L., Hudson, M.R., and Yaghi, O.M.
driven channel collapse. Chem. Commun. 50, (2014). Water adsorption in porous metal–
19. Bhatt, P.M., Belmabkhout, Y., Cadiau, A., Adil, 8944–8946. organic frameworks and related materials.
K., Shekhah, O., Shkurenko, A., Barbour, L.J., J. Am. Chem. Soc. 136, 4369–4381.
and Eddaoudi, M. (2016). A fine-tuned 31. Ehrenmann, J., Henninger, S.K., and Janiak, C.
fluorinated MOF addresses the needs for trace (2011). Water adsorption characteristics of MIL- 43. George, A., Ryotaro, M., and Kitagawa, S.
CO2 removal and air capture using 101 for heat-transformation applications of (2010). Highly porous and stable coordination
physisorption. J. Am. Chem. Soc. 138, 9301– MOFs. Eur. J. Inorg. Chem. 2011, 471–474. polymers as water sorption materials. Chem.
9307. Lett. 39, 360–361.
32. Park, J., Feng, D., and Zhou, H.-C. (2015).
20. Canivet, J., Fateeva, A., Guo, Y., Coasne, B., Structure-assisted functional anchor 44. Canivet, J., Bonnefoy, J., Daniel, C., Legrand,
and Farrusseng, D. (2014). Water adsorption in implantation in robust metal–organic A., Coasne, B., and Farrusseng, D. (2014).
MOFs: fundamentals and applications. Chem. frameworks with ultralarge pores. J. Am. Chem. Structure-property relationships of water
Soc. Rev. 43, 5594–5617. Soc. 137, 1663–1672. adsorption in metal-organic frameworks. New
J. Chem. 38, 3102–3111.
21. Burtch, N.C., Jasuja, H., and Walton, K.S. 33. Alezi, D., Belmabkhout, Y., Suyetin, M., Bhatt,
(2014). Water stability and adsorption in metal– P.M., Weseli nski, Ł.J., Solovyeva, V., Adil, K., 45. Pires, J., Pinto, M.L., Carvalho, A., and de
organic frameworks. Chem. Rev. 114, 10575– Spanopoulos, I., Trikalitis, P.N., Emwas, A.-H., Carvalho, M.B. (2003). Assessment of
10612. et al. (2015). MOF crystal chemistry paving the hydrophobic-hydrophilic properties of
way to gas storage needs: aluminum-based microporous materials from water adsorption
22. Wang, C., Liu, X., Keser Demir, N., Chen, J.P., soc-MOF for CH4, O2, and CO2 storage. J. Am. isotherms. Adsorption 9, 303–309.
and Li, K. (2016). Applications of water stable Chem. Soc. 137, 13308–13318.
metal-organic frameworks. Chem. Soc. Rev. 45, 46. Ko, N., Choi, P.G., Hong, J., Yeo, M., Sung, S.,
5107–5134. 34. Sonnauer, A., Hoffmann, F., Fröba, M., Kienle, Cordova, K.E., Park, H.J., Yang, J.K., and Kim, J.
L., Duppel, V., Thommes, M., Serre, C., Férey, (2015). Tailoring the water adsorption
23. Xue, D.-X., Belmabkhout, Y., Shekhah, O., G., and Stock, N. (2009). Giant pores in a properties of MIL-101 metal-organic
Jiang, H., Adil, K., Cairns, A.J., and Eddaoudi, chromium 2,6-naphthalenedicarboxylate frameworks by partial functionalization.
M. (2015). Tunable rare Earth fcu-MOF open-framework structure with MIL-101 J. Mater. Chem. A 3, 2057–2064.
platform: access to adsorption kinetics driven topology. Angew. Chem. Int. Ed. 48, 3791–
gas/vapor separations via pore size 3794. 47. Rieth, A.J., Yang, S., Wang, E.N., and Dinc
a, M.
contraction. J. Am. Chem. Soc. 137, 5034–5040. (2017). Record atmospheric fresh water capture
35. Serre, C., Millange, F., Thouvenot, C., Noguès, and heat transfer with a material operating at
24. Assen, A.H., Belmabkhout, Y., Adil, K., Bhatt, M., Marsolier, G., Louër, D., and Férey, G. the water uptake reversibility limit. ACS Cent.
P.M., Xue, D.-X., Jiang, H., and Eddaoudi, M. (2002). Very large breathing effect in the first Sci. 3, 668–672.

Chem 4, 94–105, January 11, 2018 105

You might also like